content
stringlengths
1
15.9M
\section{ISOGAL Observations} The ISOGAL observations were made in the two filters mentioned, as rasters covering squares of 15 $\times$ 15 arcmin$^2$ orientated in $\ell$,$b$. They were centered at $\ell$=+1.03$^{\circ}$, $b$=--3.83$^{\circ}$, which includes the globular cluster NGC\,6522 itself, and at $\ell$=+1.37$^{\circ}$, $b$=--2.63$^{\circ}$ in Sgr I. Each position on the sky was observed for a total of 22s on average. Two rasters of each field were made with the LW2 filter and one with LW3 (Table 1). The first three digits of each identification indicate the ISO revolution number ($\simeq$ day of flight). The second LW2 observation in each case is almost simultaneous with that in LW3 (within $\sim$ 30 min). The images are shown in Figs 1--4. The 6$''$ pixel field of view was utilised in all cases. \begin{table} \caption{Journal of observations} \begin{tabular}{lllll} Field & Filter & Identification & Julian date \\ NGC\,6522 & LW2 & 47101493 & 2450509 \\ NGC\,6522 & LW2 & 84001115 & 2450877 \\ NGC\,6522 & LW3 & 84001116 & 2450877 \\ Sgr I & LW2 & 47101494 & 2450509 \\ Sgr I & LW2 & 83800913 & 2450875 \\ Sgr I & LW3 & 83800914 & 2450875 \\ \end{tabular} \end{table} \begin{figure} \epsfxsize=8cm \epsffile[0 0 332 342]{fig1.ps} \caption{ISOCAM 7\,$\mu$m (LW2) image of the field in the NGC\,6522 Baade Window, covering 15 $\times$ 15 arcmin$^2$. Increasing $\ell$ is to the left and increasing $b$ is upwards. The NGC\,6522 cluster is visible near the middle of the bottom edge (Image 84001115). The circle has a radius 0.02$^{\circ}$.} \end{figure} \begin{figure} \epsfxsize=8cm \epsffile[0 0 332 342]{fig2.ps} \caption{ISOCAM 15\,$\mu$m (LW3) image of the field in the NGC\,6522 Baade Window, covering 15 $\times$ 15 arcmin$^2$. Increasing $\ell$ is to the left and increasing $b$ is upwards (Image 84001116).} \end{figure} \begin{figure} \epsfxsize=8cm \epsffile[0 0 332 342]{fig3.ps} \caption{ISOCAM 7\,$\mu$m (LW2) image of the field in the Sgr I Baade Window, covering 15 $\times$ 15 arcmin$^2$. Increasing $\ell$ is to the left and increasing $b$ is upwards (Image 83800913).} \end{figure} \begin{figure} \epsfxsize=8cm \epsffile[0 0 332 342]{fig4.ps} \caption{ISOCAM 15\,$\mu$m (LW3) image of the field in the Sgr I Baade Window, covering 15 $\times$ 15 arcmin$^2$. Increasing $\ell$ is to the left and increasing $b$ is upwards (Image 83800914).} \end{figure} Reduction of the science processed data (SPD) from version 6.32 of the OLP (Off-line Processing) pipeline was carried out with the CIA\footnote{The ISOCAM data presented in this paper was analysed using `CIA', a joint development by the ESA Astrophysics Division and the ISOCAM Consortium. The ISOCAM Consortium is led by the ISOCAM PI, C. Cesarsky, Direction des Sciences de la Mati\`{e}re, C.E.A., France.} package (CIA version 3.0) The data were first corrected for dark current, using the default method of subtracting a `model' dark frame. Following this, the cosmic-ray hits were removed from the data cube using the multi-resolution median method. At this stage, two copies of the data cube were made and the individual copies were treated by two different methods for simulating the time behaviour of the pixels of the ISOCAM detectors: the `vision' method and the `IAS model transient correction', also known as the `inversion' method (Abergel et al., 1998). The vision method does not correct much for the transient behaviour but is useful in removing the memory remnants (electronic ghosts) of previously observed strong sources. From this stage onwards in the reduction procedure we thus had two sets of data, corresponding to the two methods of stabilization. For each data set the images at each raster pointing were averaged. The data were then flat-fielded with the flats generated from the raster observations themselves. Subsequent to the flat fielding, the individual images were mosaiced together after correcting for the field of view distortion using the `projection' method. The two individual rasters thus obtained were then in units of ADU/gain/sec. They were converted into mJy/pixel units within CIA. The conversion factors were: \[ F({\rm mJy}) = (\rm ADU/gain/sec)/2.33 \] for LW2 and \[ F({\rm mJy}) = (\rm ADU/gain/sec)/1.97. \] \noindent for LW3 (Blommaert, 1998). These are correct for a $F_{\lambda}$ $\propto$ $\lambda^{-1}$ power-law spectrum at wavelengths 6.7\,$\mu$m and 14.3\,$\mu$m respectively. Source extraction was performed on each pair of reduced images using a point spread function fitting routine (Alard et al., in preparation). The vision-treated point sources were cross-identified with the inversion-treated ones and the final catalogue of point sources was built with the inversion photometry for those sources found in both vision- and inversion-treated images. This procedure ensured that most false sources were dropped while the better photometry of the inversion-treated images was retained. Conversion to magnitudes was then carried out using the formulae \[ [7] = 12.38 - 2.5 \; {\rm log} \; F_{\rm LW2}({\rm mJy}) \] and \[ [15] = 10.79 - 2.5 \; {\rm log} \; F_{\rm LW3}({\rm mJy}) \] where the zero point has been chosen to get zero magnitude for a Vega model flux at the respective wavelengths mentioned earlier. We have limited the extracted catalogue to sources with fluxes greater than 5mJy in both LW2 and LW3. This corresponds to [7] = 10.64 and [15] = 8.99. The reliability of the ISOCAM data is affected at the faint end by crowding and noise. These effects can be investigated by constructing histograms of source counts vs mag. Figure 5 is a histogram of the LW3 observations in the NGC\, 6522 field. Also included is the expected distribution, based on averaged [7] magnitudes and a relation [15]$_{\rm expected}$ = 1.56 $\times$ [7] -- 5.0 (This approximation is a line passing through the [15], [7]--[15] values of (9.0, 0.0) and (3.96, 1.8) in the colour-magnitude diagrams (Figures 8 and 9)). It appears that the NGC\,6522 detections at 15\,$\mu$m are reliable to a depth of [15] $\sim$ 9.0 mag. Figure 6 is a similar diagram for the Sgr I field. \begin{figure} \epsfxsize=8cm \epsffile[28 322 539 786]{fig5.ps} \caption{Total observed (solid line) and expected (dotted line) distributions of all sources at 15\,$\mu$m (LW3) in the NGC\,6522 field, demonstrating that the LW3 detections are reliable to $\sim$ 9.0 mag (see text for explanation). The dashed histogram is the distribution of sources detected only at 15\,$\mu$m. The latter may be spurious, especially near the faint limit.} \end{figure} \begin{figure} \epsfxsize=8cm \epsffile[28 322 539 786]{fig6.ps} \caption{Observed (solid line) and expected (dotted line) distributions for the Sgr I field. See caption of Fig.\ 5 for further details.} \end{figure} From analysis of repeated observations, the rms dispersion of the ISOGAL photometry is estimated to be generally less than 0.2 magnitudes, except towards the fainter magnitude ranges, where it rises to $\sim$0.4 mag (Ganesh et al., in preparation). See also section 3.3.3 of this paper. A small uncertainty remains in the absolute photometry due to the crowded nature of the ISOGAL fields. It is felt that the behaviour of the inversion method of transient correction is not yet completely understood and it is hoped that the uncertainty can be reduced in the near future. It may be noted that, with $\sim$ 45 and $\sim$ 35 pixels per LW2 source, respectively, the density of sources in our two fields is close to the confusion limit in both. The astrometry of the extracted sources was improved by using the newly available DENIS $K$ band observations of these fields. A systematic shift of $0.8''$ (with a rms dispersion of 0.6$''$) in right ascension and $5.1''$ (0.7$''$) in declination was found for the NGC\,6522 observation. The corresponding numbers for the Sgr I observation were found to be $2.0''$ (0.9$''$) and $6.1''$ (0.8$''$) respectively. The positions given in Tables 2 and 3 include the corrections. \subsection{NGC\,6522} Sources were extracted from the LW2 and LW3 images and cross-correlated to form a working table. The total number detected was 497. Of these, 182 or 37\% were detected at 15\,$\mu$m. The remainder were detected only in LW2, usually because they were too faint to be seen at 15\,$\mu$m. Sources were detected in both LW2 exposures in 363 cases. Non-detections in LW2 are often clearly due to blending with strong sources in their neighbourhoods. In some cases the reliability of the detection is questionable and needs to be confirmed when deeper data become available. Fifty-four sources were detected only in the first LW2 observation and 66 only in the second. These are for the most part close to the detection limit and their existence requires confirmation. Some 14 sources were detected at 15\,$\mu$m only. Each was investigated visually in the three images and comments on individual sources were included in the working table. Because of the small uncertainty in the zero-point of the photometry that has already been mentioned, we have decided to publish only those sources with [7] $\leq$ 7.5 for the time being (see table 2). The full table will be published when it is felt that the performance of the camera in crowded fields is better understood. The lack of a strong density enhancement in the neighbourhood of NGC\,6522 indicates that few sources are likely to be members of the cluster itself. Circles with radius 0.02 degree about the cluster centre are shown on figs.\ 1 and 2. \setcounter{table}{1} \begin{table*} \begin{minipage}{18cm} \caption{Bright sources detected by ISO in the NGC\,6522 field} \begin{tabular}{lllllll} No. & Name & [7]\#1 & [7]\#2 & [7]$_{avg}$ & [15] & Cross-identifications and comments.\\ 1 & ISOGAL-PJ180228.2-300038 & 6.75 & & 6.75 & & near raster edge - poor astrometry \\ 12 & ISOGAL-PJ180234.8-295957 & 4.88 & 5.05 & 4.96 & 3.64 & TLED9 400:d GF FW \\ 14 & ISOGAL-PJ180235.2-295855 & 6.46 & 6.53 & 6.50 & 5.78 & *H \\ 26 & ISOGAL-PJ180239.4-295636 & 7.46 & & 7.46 & 7.16 & \\ 41 & ISOGAL-PJ180242.9-300335 & & 7.48 & 7.48 & 5.86 & \\ 46 & ISOGAL-PJ180243.5-300250 & 6.77 & 6.56 & 6.66 & 6.35 & Unresolved? (See 49) \\ 85 & ISOGAL-PJ180251.2-300013 & 7.59 & 7.39 & 7.49 & 6.61 & BMB7 7 \\ 110 & ISOGAL-PJ180256.1-295533 & 6.77 & 6.64 & 6.71 & 5.77 & *I \\ 120 & ISOGAL-PJ180257.0-300417 & 6.23 & 6.76 & 6.50 & 5.54 & *G BMB18 6 \\ 128 & ISOGAL-PJ180257.6-295124 & 6.95 & 6.65 & 6.80 & 6.32 & \\ 147 & ISOGAL-PJ180259.6-300253 & 6.92 & 7.13 & 7.03 & 5.28 & *C BMB28 7 FW \\ 162 & ISOGAL-PJ180301.1-300141 & 6.23 & 6.49 & 6.36 & 5.51 & *B \\ 170 & ISOGAL-PJ180302.3-295806 & 7.18 & 7.23 & 7.21 & 7.04 & \\ 180 & ISOGAL-PJ180303.4-295729 & 6.03 & 6.03 & 6.03 & 5.11 & TLE 228 270d GF (3) BMB39 9 150 \\ 192 & ISOGAL-PJ180305.3-295515 & 6.84 & 6.87 & 6.86 & 5.52 & *F BMB46 7 FW (Unresolved in LW2?) \\ 205 & ISOGAL-PJ180306.3-295203 & 7.50 & 7.31 & 7.40 & 5.75 & \\ 218 & ISOGAL-PJ180307.8-300034 & 7.26 & 7.16 & 7.21 & 5.84 & BMB52 9 FW \\ 224 & ISOGAL-PJ180308.5-300524 & 6.39 & 6.25 & 6.32 & 4.82 & TLE 403 335d GF BMB54 7 FW \\ 235 & ISOGAL-PJ180309.8-295505 & 7.23 & 7.35 & 7.29 & 7.01 & BMB61 7 \\ 245 & ISOGAL-PJ180311.5-295745 & 6.75 & 6.93 & 6.84 & 5.80 & TLE 238 290: (2) BMB63 7 FW \\ 254 & ISOGAL-PJ180312.5-300428 & 6.99 & 7.10 & 7.04 & 5.98 & BMB64 7 \\ 256 & ISOGAL-PJ180313.4-300055 & 7.42 & 7.33 & 7.38 & 6.35 & BMB67 8 \\ 270 & ISOGAL-PJ180315.5-300737 & 7.42 & 7.25 & 7.33 & 6.57 & BMB70 6.5 - at edge in LW2 (unresolved?) \\ 299 & ISOGAL-PJ180318.4-295346 & 7.23 & 7.07 & 7.15 & 5.29 & *D BMB86 9 FW \\ 333 & ISOGAL-PJ180322.4-300255 & 7.43 & 7.43 & 7.43 & 7.08 & B24 M6 BMB93 6 FW \\ 382 & ISOGAL-PJ180328.4-295544 & 7.34 & 7.17 & 7.25 & 5.54 & *E BMB119 9 (unresolved?)\\ 389 & ISOGAL-PJ180329.4-295939 & 7.36 & 7.40 & 7.38 & 5.99 & B57 M7 BMB120 7 FW \\ 410 & ISOGAL-PJ180331.3-300100 & 6.77 & 6.65 & 6.71 & 6.33 & BMB129 6.5 GC member? \\ 432 & ISOGAL-PJ180334.1-295957 & 7.03 & 6.92 & 6.97 & 5.56 & B81 M8 BMB142 8 FW \\ 441 & ISOGAL-PJ180335.7-300300 & 7.33 & 7.45 & 7.39 & 5.74 & GC member? \\ 445 & ISOGAL-PJ180336.9-300147 & 7.12 & 7.02 & 7.07 & 6.03 & BMB152 9 V (Unresolved?) GC member?\\ 476 & ISOGAL-PJ180346.1-295911 & 6.91 & 7.08 & 7.00 & 5.21 & *A B138 M7 BMB179 7 FW \\ 491 & ISOGAL-PJ180350.9-295617 & 6.29 & 6.33 & 6.31 & 5.26 & TLE136 270d (1) BMB194 6.5 FW \\ \end{tabular} \vspace{1mm} Notes to tables 2 and 3: The ``Name" column follows the IAU nomenclature. The letter ``P" denotes provisional\\ 1 Noted by Lloyd Evans (1976) as semiregular \\ 2 Noted by Lloyd Evans (1976) as small amplitude in $I$\\ 3 =IRAS 17598-2957\\ Key: TLE: LPVs in Lloyd Evans (1976)\\ GF: Miras with $JHKL$ photometry by Glass and Feast (1982)\\ \*A etc: stars selected for blink examination\\ B: Stars classified by Blanco (1986)\\ BMB: stars classified by BMB (1984), followed by M sub-class\\ FW: $JHK$ photometry in Frogel \& Whitford (1987)\\ \end{minipage} \end{table*} \subsection{Sgr I} A total of 696 sources were detected. Of these, 287 or 41\% were detected at 15\,$\mu$m . Only 20 sources were detected solely at 15\,$\mu$m. At 7\,$\mu$m, 517 sources were detected twice. Eighty-eight sources were only seen in the first LW2 exposure and 71 were only seen in the second. As in the case of the NGC\,6522 field, only those sources with [7] $\leq$ 7.5 are given in Table 3. \setcounter{table}{2} \begin{table*} \begin{minipage}{180mm} \caption{Bright sources detected in the Sgr I field by ISO} \begin{tabular}{lllllll} No & Name & [7]\#1 & [7]\#2 & [7]$_{avg}$ & [15] & Cross-identifications and comments\\ 32 & ISOGAL-PJ175836.4-290802 & 7.27 & 7.14 & 7.21 & 6.28 & \\ 45 & ISOGAL-PJ175839.1-290522 & 7.29 & 7.23 & 7.26 & 6.54 & \\ 61 & ISOGAL-PJ175841.8-290352 & 6.84 & 6.88 & 6.86 & 5.41 & *D \\ 65 & ISOGAL-PJ175842.0-290649 & 7.19 & 7.06 & 7.12 & 6.73 & Unresolved? \\ 69 & ISOGAL-PJ175842.6-291028 & & 7.44 & 7.44 & 6.59 & at edge of raster; outside [7]\#1 observation; unresolved? \\ 78 & ISOGAL-PJ175843.9-290711 & 6.61 & 6.06 & 6.33 & 5.28 & TLE65 Unresolved? \\ 112 & ISOGAL-PJ175848.0-291002 & 7.32 & 7.34 & 7.33 & 7.26 & \\ 117 & ISOGAL-PJ175848.7-290743 & 7.41 & 7.55 & 7.48 & 6.58 & \\ 129 & ISOGAL-PJ175850.3-290454 & 7.43 & 7.52 & 7.47 & 6.57 & \\ 167 & ISOGAL-PJ175854.8-285831 & 6.93 & 6.76 & 6.84 & 6.94 & \\ 168 & ISOGAL-PJ175854.9-290117 & 7.40 & 7.53 & 7.46 & 7.76 & \\ 170 & ISOGAL-PJ175855.1-290627 & 6.95 & 7.03 & 6.99 & 5.53 & *M Unresolved? \\ 180 & ISOGAL-PJ175855.9-285845 & 7.34 & 7.42 & 7.38 & 6.86 & \\ 187 & ISOGAL-PJ175856.4-290049 & 6.81 & 6.65 & 6.73 & 5.37 & *F \\ 189 & ISOGAL-PJ175856.6-290213 & 7.30 & 7.35 & 7.32 & 7.24 & \\ 200 & ISOGAL-PJ175857.4-291215 & 7.14 & 7.25 & 7.20 & 6.11 & \\ 206 & ISOGAL-PJ175857.8-290114 & 6.60 & 6.71 & 6.65 & 5.34 & *J \\ 247 & ISOGAL-PJ175901.1-285821 & 6.13 & 5.79 & 5.96 & 4.46 & TLE79 Unresolved? \\ 291 & ISOGAL-PJ175904.7-290744 & 7.01 & 7.03 & 7.02 & 5.65 & *N \\ 303 & ISOGAL-PJ175905.7-290235 & 6.99 & 7.17 & 7.08 & 6.01 & \\ 328 & ISOGAL-PJ175907.3-291024 & 7.17 & 7.31 & 7.24 & 6.68 & \\ 338 & ISOGAL-PJ175908.3-290855 & 7.53 & 7.38 & 7.46 & 6.40 & \\ 351 & ISOGAL-PJ175909.5-290903 & 7.34 & 7.40 & 7.37 & 7.27 & \\ 352 & ISOGAL-PJ175909.5-290825 & 7.52 & 7.45 & 7.48 & 7.31 & \\ 363 & ISOGAL-PJ175910.5-290129 & 4.66 & 5.08 & 4.87 & 3.37 & TLE53 Unresolved? \\ 365 & ISOGAL-PJ175910.6-290456 & 7.22 & 7.27 & 7.24 & 6.45 & \\ 375 & ISOGAL-PJ175911.1-290315 & 6.66 & 6.75 & 6.71 & 5.34 & *E Unresolved? \\ 400 & ISOGAL-PJ175913.6-285812 & 7.37 & 7.34 & 7.36 & 7.10 & \\ 401 & ISOGAL-PJ175913.7-291113 & 5.03 & 4.97 & 5.00 & 4.77 & \\ 402 & ISOGAL-PJ175913.7-285852 & 7.04 & 6.88 & 6.96 & 5.75 & TLE87 \\ 405 & ISOGAL-PJ175914.0-290950 & 7.24 & 7.05 & 7.14 & 5.58 & *G \\ 410 & ISOGAL-PJ175914.4-291335 & 7.24 & 7.18 & 7.21 & 6.80 & \\ 416 & ISOGAL-PJ175914.8-291129 & 7.46 & 7.35 & 7.40 & 6.41 & \\ 421 & ISOGAL-PJ175915.5-290133 & 6.79 & 6.82 & 6.80 & 6.07 & appears unresolved. \\ 434 & ISOGAL-PJ175916.8-290010 & 7.27 & 7.22 & 7.24 & 7.38 & \\ 446 & ISOGAL-PJ175917.7-285839 & 7.47 & 7.40 & 7.44 & & \\ 452 & ISOGAL-PJ175918.2-291433 & 6.69 & 6.63 & 6.66 & 6.50 & \\ 453 & ISOGAL-PJ175918.2-290123 & 5.99 & 6.15 & 6.07 & 5.83 & \\ 455 & ISOGAL-PJ175918.5-290504 & 7.34 & 7.25 & 7.29 & 6.58 & \\ 457 & ISOGAL-PJ175918.5-290607 & 7.28 & 7.36 & 7.32 & 5.77 & \\ 478 & ISOGAL-PJ175920.7-291505 & 7.41 & 7.33 & 7.37 & 7.07 & \\ 488 & ISOGAL-PJ175922.0-291229 & 6.36 & 6.55 & 6.46 & 6.03 & \\ 511 & ISOGAL-PJ175923.4-290215 & 6.18 & 6.56 & 6.37 & 5.39 & TLE54 \\ 514 & ISOGAL-PJ175923.7-291236 & 6.76 & 6.91 & 6.83 & 5.64 & TLE39 \\ 522 & ISOGAL-PJ175924.4-291358 & 7.24 & 7.24 & 7.24 & 6.21 & \\ 524 & ISOGAL-PJ175924.6-291236 & & 7.47 & 7.47 & 6.15 & \\ 529 & ISOGAL-PJ175925.3-290336 & 7.05 & 7.22 & 7.13 & 5.70 & *L \\ 540 & ISOGAL-PJ175926.4-290707 & 7.15 & 6.90 & 7.03 & 5.84 & \\ 541 & ISOGAL-PJ175926.5-290217 & 7.31 & 7.37 & 7.34 & 5.60 & *H \\ 550 & ISOGAL-PJ175927.4-290309 & 7.75 & 7.17 & 7.46 & 5.91 & \\ 556 & ISOGAL-PJ175927.9-290530 & 7.44 & 7.47 & 7.46 & 6.61 & \\ 576 & ISOGAL-PJ175929.7-290318 & 6.35 & 6.68 & 6.51 & 5.47 & TLE55 \\ 581 & ISOGAL-PJ175930.7-290950 & 6.77 & 7.00 & 6.88 & 6.83 & \\ 585 & ISOGAL-PJ175931.1-290858 & 6.93 & 7.18 & 7.05 & 5.67 & *K \\ 601 & ISOGAL-PJ175932.8-290734 & 7.25 & 7.33 & 7.29 & 6.59 & \\ 609 & ISOGAL-PJ175934.0-290236 & 5.41 & 5.63 & 5.52 & 5.53 & unresolved? \\ 613 & ISOGAL-PJ175934.4-290702 & 7.43 & 7.49 & 7.46 & 6.40 & \\ 626 & ISOGAL-PJ175936.1-290915 & 6.94 & 6.97 & 6.96 & 6.75 & \\ 633 & ISOGAL-PJ175937.0-290834 & 7.38 & 7.50 & 7.44 & 6.56 & \\ 681 & ISOGAL-PJ175945.2-290442 & 6.99 & 7.26 & 7.12 & 5.54 & *C \\ 684 & ISOGAL-PJ175946.1-290318 & 7.03 & 7.44 & 7.24 & 6.97 & TLE57 Unresolved? \\ 686 & ISOGAL-PJ175947.0-290227 & 7.27 & & 7.27 & & at edge of raster \\ 689 & ISOGAL-PJ175948.2-290350 & 6.90 & 7.21 & 7.05 & 6.37 & TLE56 \\ 692 & ISOGAL-PJ175949.3-290247 & 7.36 & 7.48 & 7.42 & & \\ \end{tabular} \end{minipage} \end{table*} \section{Correlations with other Catalogues} The fields observed by ISOGAL overlap completely or in part with areas surveyed in other ways. \subsection{Spectroscopic information (NGC\,6522 only)} The spectroscopic survey of BMB (1984) covers M6 and later M-type giants. Spatially, more than 80\% of our field is included and almost all their stars in the overlap area ($\sim$ 112) were detected. Cross-correlation of source positions was performed by means of transparent overlays. The sources BMB20 and 21 may both correspond to our no.\ 118. The following BMB sources that we do {\it not} detect are of spectral type M6: BMB 41, 56, 62, 82, 100, and 118. We also did not see BMB\,13 (M5). These are all at the early end of the range. More interesting is the deeper survey of Blanco (1986) which covers a much smaller portion of the ISOGAL field but includes earlier spectral types, from M1 onwards. Table 4 summarizes our detections as a function of spectral type for the part of the field coincident with the deep spectroscopic survey. It is clear that our survey cuts off between spectral classes M3 and M4 (III), where the changeover from majority non-detections to majority detections occurs. It must also be expected that some brighter objects from the foreground will have been included. In fact, there are eight objects in the overlap region that were detected in the ISOGAL programme but were not classified as having spectra in the range M1--M9 by Blanco (1986). Their $I$-band counterparts appear quite bright. They were examined on $I$ and $V$ plates (Lloyd Evans, 1976) and were found to be probably non-variable and to have colours corresponding to K or early M types. Only one was detected at 15\,$\mu$m. The others are almost certainly foreground stars. \subsection{Photometric information} \subsubsection{NGC\,6522} Only one object in the IRAS Point-source Catalog falls within our survey area. This is IRAS\,17598-2957, which coincides with our no.\ 180, and has [15] = 5.11 (IRAS 12\,$\mu$m flux = 0.87 Jy). Star 224 has [15] = 4.82 and star 12 has [15] = 3.64, so it is surprising that these objects were not also detected by IRAS. The cause may be variability or crowding of sources. A number (56) of our NGC\,6522 stars have been observed by Frogel and Whitford (1987) or Tiede, Frogel and Terndrup (1995) on the CTIO/CIT near-infrared system. Some of the variables also have photometry by Glass and Feast (1982). Fig.\ 7 shows the $(J-K)_0$, $K_0$ diagram of the ISOGAL objects that were measured on the CIT/CTIO system. {\it I}-band photographic photometry was included in the BMB (1984) M-star survey. Sharples, Walker and Cropper (1990) have pointed out that the stars with {\it I} mag $<$ 11.8 have a significantly lower velocity dispersion than fainter ones, indicating that they probably belong to the foreground disc. A similar effect is seen amongst K giants in the same field (Sadler, Terndrup and Rich, 1996). There are 9 BMB stars with {\it I} $<$ 11.8 amongst the ISO detections, including two Miras (TLE\,238 and TLE\,136) and a possible semiregular variable (BMB\,18). \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig7.ps} \caption{$(J-K)_0$, $K_0$ diagram for ISOGAL objects in the NGC\,6522 field that have photometry by Frogel and Whitford (1987) or Tiede, Frogel and Terndrup (1995) on the CIT/CTIO system. The NGC\,6522 Miras are indicated by heavy points. Other high mass-loss objects specially examined (see below, section 6) are crossed. Also shown as circled points are Miras in our Sgr I field, taken from Glass et al.\ (1995). The tip of the RGB is estimated to occur at $K$ $\sim$ 8.2. Note that the NGC\,6522 Mira brightest at $K$ (= TLE\,D9) appears unusually blue in the Frogel and Whitford photometry. It was considerably redder when measured by Glass and Feast (1982).} \end{figure} \subsubsection{Sgr I} There are two IRAS sources within the ISOGAL field, viz no.\ 363 = IRAS 17559-2901 = TLE 53 (IRAS 12\,$\mu$m flux = 1.7 Jy) and no.\ 247 = IRAS 17558-2858 = TLE 79 (IRAS 12\,$\mu$m flux = 1.19 Jy). \subsection{Mira-type Variability} The spectra of Mira variables change by several sub-types around the cycle, so that a direct correspondence between, for example, period and spectral sub-type is not to be expected. As will be seen, several stars with positions around those of the less-luminous Miras in the colour-magnitude diagrams have been examined for large-amplitude variability with negative results. It therefore appears that our inventory of the Mira variables in these two fields is complete. \setcounter{table}{3} \begin{table} \caption{ISOGAL detections in NGC\,6522 as a function of spectral type for stars classified in a deep objective-prism survey by Blanco (1986), falling within the area of mutual overlap.} \begin{tabular}{lll} M class & Detected & Not detected \\ 1 & 0 & 38 \\ 2 & 5 & 15 \\ 3 & 7 & 9 \\ 4 & 6 & 5 \\ 5 & 6 & 2 \\ 6 & 13 & 0 \\ 6.5 & 6 & 0 \\ 7 & 5 & 0 \\ 8 & 1 & 0 \\ \end{tabular} \end{table} \subsubsection{NGC\,6522} Six stars from the ISOGAL field were detected by Lloyd Evans (1976) as Mira variables. Four of these are amongst the most luminous at 15\,$\mu$m, while one, of relatively short period for a ``long-period variable" is near the limit of detection. The IRAS source corresponds to TLE 228. Table 5 shows the cross-identifications. The relatively short-period star (115d) TLE\,395 is only just detectable at 15\,$\mu$m. \setcounter{table}{4} \begin{table} \caption{Stars within the ISOGAL NGC\,6522 field known to be long-period variables} \begin{tabular}{llll} ISOGAL & Lloyd & Period & Differ- \\ no.\ & Evans no.\ & days & ences$^1$ \\ 12 & D9 & 400: & -0.17 \\ 180$^2$ & 228 & 270 & 0.00 \\ 224 & 403 & 335 & 0.14 \\ 245 & 238 & 290: & -0.18 \\ 258 & 395 & 115 & 0.27 \\ 491 & 136 & 270 & -0.04 \\ \end{tabular} Note: A colon denotes uncertainty in the period $^1$Between the two measurements in LW2. See also table 6. $^2$ = IRAS\, 17598-2957 \end{table} The two ISOGAL 7\,$\mu$m exposures were compared by magnitude range (in the first measurement) and the average differences and their rms values were found for stars that appear in both. The result is shown in Table 6. \begin{table} \caption{Average and r.m.s values of the differences in photometric results between the two 7\,$\mu$m exposures of stars in NGC\,6522.} \begin{tabular}{llll} Range in & No.\ of & Average & RMS \\ 7\,$\mu$m mag & stars & difference & difference \\ 5--6 & 1 & -0.17 & 0.17 \\ 6--7 & 13 & -0.008 & 0.21 \\ 7--8 & 43 & 0.097 & 0.18 \\ 8--9 & 113 & 0.111 & 0.22 \\ 9--10 & 160 & 0.097 & 0.17 \\ 10--11 & 33 & -.006 & 0.15 \\ \end{tabular} \end{table} \subsubsection{Sgr I} Nine stars within the ISOGAL field are known long-period variables (see Table 7). These have been followed at {\it JHKL} by Glass et al (1995), who list their mean magnitudes and periods. \begin{table} \caption{Stars within the ISOGAL Sgr I field known to be long-period variables} \begin{tabular}{llll} ISOGAL & Lloyd & Period & Differ- \\ no. & Evans no.\ & days$^1$ & ences$^2$ \\ 514 & 39 & 336: & -0.15 \\ 363$^3$ & 53 & 480 & -0.42 \\ 511 & 54 & 293 & -0.38 \\ 576 & 55 & 330: & -0.33 \\ 689 & 56 & 235 & -0.31 \\ 684 & 57 & 153 & -0.41 \\ 78 & 65 & 237 & 0.55 \\ 247$^4$ & 79 & 383 & 0.34 \\ 402 & 87 & 308 & 0.16 \\ \end{tabular} Notes: A colon denotes uncertainty in the period. $^1$Periods taken from Glass et al (1985) except for ISOGAL no.\ 576 which is from Lloyd Evans (1976). $^2$Between the two measurements in LW2. See also table 8. $^3$ = IRAS17559-2901 $^4$ = IRAS17558-2858 \end{table} The two ISOGAL 7\,$\mu$m exposures were again compared by magnitude range. The result is shown in Table 8. \begin{table} \caption{Average and r.m.s values of the differences in photometric results between the two 7\,$\mu$m exposures of stars in Sgr I.} \begin{tabular}{llll} Range in & No.\ of & Average & RMS \\ 7\,$\mu$m mag & stars & difference & difference \\ 4--5 & 1 & 0.06 & 0.06 \\ 5--6 & 3 & -0.010 & 0.34 \\ 5--6 & 1$^1$ & -0.22 & 0.22 \\ 6--7 & 16 & -0.010 & 0.15 \\ 6--7 & 11$^1$ & -0.001 & 0.14 \\ 7--8 & 78 & -0.030 & 0.17 \\ 7--8 & 76$^1$ & -0.020 & 0.14 \\ 8--9 & 157 & -0.020 & 0.17 \\ 9--10 & 219 & -0.017 & 0.17 \\ 10--11 & 43 & -0.12 & 0.24 \\ \end{tabular} $^1$When known variables are omitted. \end{table} \subsubsection{Photometric consistency} If the known variables are omitted the average and rms differences are somewhat reduced. In the case of NGC\,6522, the known variables show only small differences and do not materially affect the rms values. The error in the repeatibility of a single 7\,$\mu$m observation in either field is thus about 0.14 mag, where we have divided the RMS difference columns in tables 6 and 8 by a factor of $\sqrt{2}$. \subsection{OH/IR Catalogues} No known OH/IR sources fall within our fields (of 0.063 square degrees each). The density of known OH/IR sources in the part of the sky occupied by our two fields is only 1--2 per square degree (Sevenster et al., 1997), though it is higher at the Galactic Centre ($\sim$10 per square degree for a survey of similar depth). The deepest surveys of the Central region indicate that the density of OH/IR stars reaches several hundred per square degree or about 1/3 the number of known large-amplitude variables (Glass et al., 1999). From their $H_K$ colours, the Central region sources do not necessarily possess optically thick dust shells at $K$. However, it is not clear whether a deeper OH survey of the NGC\,6522 and Sgr I fields would yield detections from the objects with moderate dust shells reported here. \section{The ISOGAL [7]--[15], [15] colour-magnitude diagrams} Stars detected at both wavelengths are shown in the [7]--[15], [15] colour-magnitude diagrams (Figs 8 and 9). The fact that these and similar diagrams discussed below show well-defined sequences implies that they are not significantly contaminated by foreground stars. \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig8.ps} \caption{ISOCAM colour-magnitude diagram for the 15 $\times$ 15 arcmin$^2$ field in the NGC\,6522 Window. Note that many sources are detected only at 7\,$\mu$m and do not appear here. Known Mira variable stars are indicated by heavy points and some other stars chosen for investigation as possible high mass-loss objects are crossed. The [7]-[15] colours make use only of the second LW2 exposure, which was nearly simultaneous with the LW3.} \end{figure} \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig9.ps} \caption{ISOCAM colour-magnitude diagram for the 15 $\times$ 15 arcmin$^2$ field in the Sgr I Window. See caption of fig.\ 8 for further details.} \end{figure} It is instructive to see how the [7]--[15] colours change with spectral type, bearing in mind that the few Miras may vary between late types. This is shown for the NGC\,6522 field in Fig.\ 10, which includes all our sources that have been classified by BMB (1984) or Blanco (1986) and have been detected in both bands. Note that the fraction of stars satisfying both these criteria is almost zero for the early M-types and almost one for the later. Except for M sub-types 6, 6.5 and 7, the numbers are small. This selection effect arises from the fact that many stars fall below our limit in the 15\,$\mu$m band. \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig10.ps} \caption{[7]--[15] colours vs spectral type for all spectrally classified stars that were detected at both wavelengths. The heavy dots are known Mira variables. Other high mass-loss objects are crossed. See text for discussion of selection effects. In calculating the colours only the second LW2 exposure was used.} \end{figure} \subsection{Mira [7]--[15] colours} The [7]--[15] colours of the Mira variables are noticeably displaced from the extrapolation of the sequence formed by the remainder of the late M-stars in the ISOCAM colour-magnitude diagrams (Figs.\ 8 and 9). For Miras, the relative flux in the 7\,$\mu$m-band must be greater or that in the 15\,$\mu$m-band must be weaker or both. O-rich Mira spectra have strong peaks due to the 9.7\,$\mu$m and 18\,$\mu$m silicate dust emission features. These are seen clearly in ISO-SWS spectra (Onaka et al., 1998). The ISOCAM LW3 band (12--18 $\mu$m) is situated longward of the 9.7\,$\mu$m silicate band but is strongly influenced for about half of its width by the 18\,$\mu$m feature. So far as is known, the LW3 band is not affected significantly by (gaseous) molecular absorption, the most conspicuous feature being the CO$_2$ band at 15.0\,$\mu$m, which appears in some O-rich Miras (Onaka et al, 1997) with small equivalent width. In many sources an extension of the 10\,$\mu$m peak to 13\,$\mu$m, attributed to Al$_2$O$_3$, may also contribute to the ISOCAM 15\,$\mu$m band. \section{$K$--[15] colours} The $K$--[15] colour is primarily a measure of the infrared excess emitted by circumstellar dust shells. The $K$-band flux originates in the stellar photosphere and is only moderately affected ($\leq$ 0.1--0.2 mag overall) by CO and H$_2$O absorption bands. As mentioned above, the ISOCAM LW3 band (12--18 $\mu$m) is strongly influenced for about half of its width by the 18\,$\mu$m silicate feature. Here we consider those sources detected at 15\,$\mu$m which also have ground-based near-infrared observations available from Frogel and Whitford (1987). Spectral classifications by BMB (1984) or Blanco (1986) are also available for these objects. Figure 11 shows their 15\,$\mu$m mag vs $K$--[15] colour and Fig.\, 12 shows the 15\,$\mu$m vs M-subtype diagram. \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig11.ps} \caption{15\,$\mu$m mag vs $K_0$--[15] for a subset of stars in the NGC\,6522 field detected at 15\,$\mu$m and observed (not simultaneously) at $K$ by Frogel and Whitford (1987). Mira variables are indicated by heavy points. Other high-mass-loss stars are crossed (see below). For convenience, the Miras in Sgr I (with $K$ photometry from Glass et al., 1995) are also plotted (circled points).} \end{figure} There is a well-defined sequence of increasing 15\,$\mu$m luminosity with $K_0$--[15] colour. Figs 11 and 12, as well as the other C-M diagrams, show that there is a steady increase of mass-loss from at least the mid-M giants to the Miras. It is also clear, for example by adding up the fluxes from each 15\,$\mu$m magnitude band, that the flux from non-Miras exceeds that from the Miras. It is therefore also likely that at least half of the mass being returned from the stars to the interstellar medium is from the non-Miras. As pointed out by Omont et al (1999c, in preparation), $K_0$ $\sim$ 8.2 corresponds to the tip of the red giant branch (RGB)(Tiede, Frogel and Terndrup, 1995). This point corresponds (Fig.\ 11) to $K_0$--[15] $\sim$ 0 or [15] $\sim$ 8.2. Thus stars brighter than [15] $\sim$ 8 are on the AGB. \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig12.ps} \caption{15\,$\mu$m mag vs M subtype for a subset of stars in the NGC\,6522 field with 15\,$\mu$m detections and spectroscopic information available. Symbols as in previous figure.} \end{figure} \section{High mass-loss objects with little or no variation} Five of the known long-period variables in the NGC\,6522 field are near the top of the ISOCAM colour-magnitude diagram (Fig.\ 8), but are interspersed with other stars that appear similarly luminous and somewhat redder. Ten of the latter stars, brighter than [15] = 5.7, were selected for detailed investigation. They are denoted by letters in Table 2 and are visible as red stars on $V$ and $I$ plates (Lloyd Evans, 1976). They have been re-examined for variability. Although BMB (1984) assigned very late spectral types and suggested variability, none of the ten stars is, in fact, a large-amplitude variable. However, variability with small amplitude cannot be excluded in about half of them. They therefore represent a type of late M-star with high mass loss but small or zero amplitude of variation. The colours of four of these stars (A, C, D and F) are shown in the $(J-K)_0$, $K_0$ diagram, Fig.\ 7, where they appear to be similar to 200--300 day Miras. One of the sample (star B) may be mainly at maximum, with occasional faint episodes, while stars A and G could be variables with occasional bright episodes superimposed on a constant background. Such variability in these stars is consistent with differences at 7\,$\mu$m between the two observations, especially for star G (where the change amounts to $\sim$ 0.53 mag). For Sgr I, as in the case of the NGC\,6522 field, 12 objects with 15\,$\mu$m mags similar to the known Miras and somewhat redder colours were checked for variability by Lloyd Evans on the plate material. In all cases, these objects could be identified with very red stars. Only two were found to be fairly certain variables of low amplitude (F and H). Both these stars spend most of their time at maximum, with two fainter episodes separated by about 250 days for F and a single one for H. The positions of these objects in the [7]--[15], [15] diagram (Figs. 8, 9 ) suggest that they form a continuation of the general sequence of late M stars and are not similar to the Mira variables. Unfortunately, only four of them, BMB\,28, BMB\,46, BMB\,86 and BMB\,179, have been measured in the near-infrared. Their $K_0$--[7] indicies are near-zero (Fig.\ 14). Two of these non-Mira high mass-loss objects have been classified spectroscopically by BMB (1984) as M9, a spectral type that is usually associated with large-amplitude variability. It is known that the onset of Mira-type behaviour occurs at later types with increasing metallicity. The existence of this class of stars may therefore be a consequence of super metal richness. Alternatively, variability may have ceased temporarily but the dust shells have not yet dissipated. Sloan and Price (1995) and Sloan, LeVan and Little-Marenin (1996) have shown that certain irregular and semi-regular AGB variables have dust excesses in IRAS spectra and are associated with the appearance of the 13\,$\mu$m dust feature. By obtaining mid-infrared spectra for our objects or by obtaining more precise variability information it will be possible to decide if they are identifiable with this category. They may also be similar to the `red' O-rich SRb variables of Kerschbaum and Hron (1996). \section{The $K_0$--[7] colours} The 56 objects of NGC\,6522 with $JHK$ photometry and spectral classifications are plotted in figure 13, which shows their $K_0$--[7] colours and spectral types. \begin{figure} \epsfxsize=8cm \epsffile[28 295 539 786]{fig13.ps} \caption{$K_0$--[7] colours vs spectral type for NGC\,6522 field stars that have $K_0$ photometry by Frogel and Whitford (1987) and spectral classifications by BMB (1984) or Blanco (1986). This colour index is predominantly negative for unreddened M giants, except for some of the latest spectral types. Note the $K$ and 7\,$\mu$m photometry was not simultaneous, and that averaged 7\,$\mu$m mags were used when available. Miras are shown as heavy dots and other high mass-loss objects are crossed.} \end{figure} These colours are unexpectedly negative in most cases, especially for the earlier M-subtypes. While part of this could be due to the small uncertainty in the absolute 7\,$\mu$m photometry in crowded fields, ISO SWS spectra of late-type O-rich M stars frequently show a broad absorption shortward of the strong silicate dust peak at 10$\mu$m. This region is known to be affected by the SiO fundamental at $\sim$ 7.9\,$\mu$m (Cohen et al., 1995). Carefully calibrated spectrophotometry of $\beta$ Peg (M2.5\,II--III) shows absorption at around 6\,$\mu$m attributed to the $\nu_2$ band of water vapour centred at 6.25\,$\mu$m. Spectra of O-rich M-type giants that also demonstrate these features are presented by Tsuji et al.\ (1998), who envisage that they arise in a warm absorbing layer somewhat above the photosphere. Approximate calculations based on the Tsuji et al.\ spectra yielded $K_S$--[7] $\sim$ --0.1 for $\beta$ Peg but slightly positive $K_{\bf S}$--[7] for SW Vir (M7III)\footnote{Spectrophotometry shows that $K$ -- $K_{\it DENIS}$ should have values of 0.02 to 0.07 for Miras because the $K_S$ band does not include the first overtone band of CO.}. The presence of significant amounts of dust could be an influence in the latter case. The depth of the CO fundamental band in late M-type giants is known to have a similar effect on the $L-M$ colours, which tend to have low or negative values. Figure 13 shows that in the latest M-type giants the $K_0$--[7] colour approaches zero or may even become positive. It is known that the overtone SiO absorption band at 3.95--4.1\,$\mu$m is strong in semiregular variables but in Miras its strength varies with time and it can become weak (Aringer et al., 1995; Rinsland and Wing, 1982). The optical depth of the circumstellar dust in the stars that we are discussing is very low in the earlier spectral types, rising somewhat towards the Miras but never becoming high. The emission at 9.7 and 18\,$\mu$m is probably dominated by particles at a fairly uniform temperature around 1000\,K, where condensation of silicate dust grains first becomes possible in the stellar wind. Thus the 15\,$\mu$m fluxes are largely a measure of the mass of grains present and ultimately of the mass-loss rate from the star. The near-infrared $K$ flux will not be affected by {\it emission} from the dust but will be reduced slightly by the {\it absorption} it causes at shorter wavelengths. The opacity of silicate dust at 7\,$\mu$m is very low (see e.g. Schutte and Tielens, 1989) so that the trend of $K$--[7] colour with spectral type that we observe could partly be caused by increasing extinction at $K$ and decreasing SiO and possibly H$_2$O molecular band strengths at 5.5--8\,$\mu$m. However, dust emission at 7\,$\mu$m is certainly responsible for the largest values observed in LPVs, as shown by Groenewegen (private communication) for various dust models. \subsection{The $K_0$--[7], [7] colour-magnitude diagram} In the $K_0$--[7], [7] colour-magnitude diagram (Fig.\ 14) we include the Miras from both fields. It is seen that almost all of the Miras are brighter and redder at 7\,$\mu$m than the other red giants. A diagram involving the $K_0$--[7] colour will therefore be the best criterion for detection of large-amplitude LPVs. \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig14.ps} \caption{The $K_0$--[7], [7] colour-magnitude diagram, including Miras from NGC\,6522 (heavy dots) and Sgr I (circled). The Miras are almost completely distinguishable from other stars by their bright 7\,$\mu$m mags. Note that the $K$ and 7\,$\mu$m photometry was not simultaneous and that averaged 7\,$\mu$m mags were used when available. Crosses denote other high mass-losing objects.} \end{figure} \section{Absence of very luminous sources} Let us recall that, for the class of sources that we detect, the $K$ bolometric magnitude correction is practically constant ($\sim$ 3.0; Groenewegen, 1997). With a distance to the Galactic Centre of 8.5\,kpc, $M_{\rm Bol}$ $\sim$ $K_0$ - 11.65. The brightest stars in our fields therefore have bolometric magnitudes $>$ --5.9. Neither field includes any luminous dust-enshrouded AGB sources of the type found in the Large Magellanic Cloud by Wood et al.\ (1992), which reach luminosities of $M_{\rm Bol}$ $\sim$ --7.5. \section{Mid-infrared period-colour relation for Miras} The Mira variables show a clear period-colour relation with very moderate scatter in the mid-infrared (Fig.\ 15). The [7] and [15] photometry is almost simultaneous, so that such scatter as exists is not attributable to variation between measurements. Following the discussion in section 7, this relation confirms that the mass-loss rate in Miras is directly related to their periods (see e.g., Whitelock et al., 1994). The colour-magnitude diagrams for each of our two fields do not appear to contain any stars brighter than the known Miras. It is therefore likely that the census of stars at the long-period end of the the range is complete and that there is no evidence for long-period variables in our fields with periods longer than those already known, i.e. $\sim$ 700 d (see Glass et al., 1995). \begin{figure} \epsfxsize=8cm \epsffile[28 298 539 786]{fig15.ps} \caption{Mid-infrared period-colour relation for Mira variables, based on the Miras in the NGC\,6522 (points) and Sgr I fields (circled points). The full regression line (solid) follows the equation [7]--[15] = 2.10 log $P$ -- 4.05. The rms differences between the fit and the observed points is 0.12 mag, rather better than expected from the photometric errors. If the shortest-period point is omitted, the fit (dashed line) becomes [7]--[15] = 2.52 log $P$ -- 5.11, with rms difference = 0.07. Only the second 7\,$\mu$m exposure was used in forming the colours.} \end{figure} \section{Star Counts} Determination of the spatial density distributions of the various stellar populations in the central Galaxy, and especially the bulge, is one of the key scientific goals of the ISOGAL project. A detailed analysis, covering several fields, and with careful consideration of completeness, extinction, etc, is in preparation. For the present we consider simply the differential number counts between the two fields at magnitudes above the completeness limit. At [15] this covers the AGB above the RGB-tip, while at [7] the limit extends approximately one mag below the RGB tip, to [7]$\sim 9.5$, using $K$ -- [15] and $K$ -- [7] colours from Figs 11 and 14 respectively. The ratio of the surface density of sources is, within the sampling errors, identical for both RGB and AGB stars, indicating that there are no steep population gradients apparent. It is also identical at both [7] and [15] microns. The surface density ratio is $$\rm \rho(Sgr\;I)/\rho(NGC\,6522) = 1.85.$$ This count ratio corresponds to an exponential scale height of 2.0$^{\circ}$, or $\sim 280$\,pc. That is, the inner bulge minor axis scale height is the same as that of the Galactic disk. A more interesting comparison is with the scale height derived from analysis of the COBE/DIRBE surface brightness observations of the inner Galaxy. These have been analysed most completely by Binney, Gerhard, and Spergel (1997), whose best-fit model has a bulge density distribution which is a truncated power-law: $$f_b = f_0{{{\rm e}^{-a^2/a_m^2}}\over{(1+a/a_0)^{1.8}}}$$ with $a_m$=1.9kpc, $a_0$=100pc, $f_0$ a normalisation constant, and $a = (z/\xi)$ for ($x = y = 0$), where $z$ is the minor axis distance and $\xi$ an axial ratio, with $\xi = 0.6$ best fitting the data. This density profile, which interpolates between the central luminosity spike, which follows an $R^{-1.8}$ luminosity profile, and the outer bulge, which follows an $R^{-3.7}$ profile, is also used in the models of Kent, Dame and Fazio (1991) and the kinematic analysis of Ibata and Gilmore (1995). It is known to provide an acceptable description of the bulge from latitudes of 4$^{\circ}$ to at least 12$^{\circ}$. It remains to be tested at intermediate latitudes. Direct comparison of the ISOGAL data with this function is of specific interest, since the fields surveyed here are included in the COBE/DIRBE analysis, and are of low reddening. The COBE/DIRBE data are however integrated light, and so must have a statistical correction for the foreground disk. This can be a complex function of wavelength and spatial resolution, especially at low latitudes (see Unavane et al., 1998, and Unavane and Gilmore, 1998) for a more complete description). Our ISOGAL source counts are however strongly biased against the foreground disk, being dominated by sources in the central Galaxy. Thus, the ISOGAL observations test directly the analysis of the integrated light. Our present analysis however does not consider line of sight depth effects in the bulge itself. With the adopted vertical axis ratio $\xi=0.6$ the predicted count ratio in the ISOGAL data is $\rm \rho(Sgr\;I)/\rho(NGC\,6522) = 2.04$, somewhat larger than the observed value. Interestingly, near-exact agreement with the model above follows with an axis ratio $\xi =1.0$, which predicts $\rm \rho(Sgr\;I)/\rho(NGC\,6522) = 1.81$. This systematic discrepancy between the model and our data is in agreement with the small systematic residuals emphasised to exist by Binney et al.\ (1997) and shown in their figure 2. \section{Conclusions} Most of the detected objects are late-type M stars, with a cut-off for those earlier than about M3--M4. There is a continuous sequence of increasingly mass-losing objects from the mid-M-type giants to the long-period Miras, which are the most luminous stars in the field. There appears to be no component of dust-enshrouded very long-period OH/IR stars or similar objects faint even at $K$ in these fields. The upper limit of luminosities remains at $M_{\rm Bol}$ $\sim$ --5.7 and the upper limit of periods remains at about 700 days, as determined from near-infrared studies (Glass et al., 1995). There is a group of late-type M-stars on the AGB which are not large-amplitude variables but may be irregular or of small amplitude, that have luminosities similar to Mira variables in the 200--300 day period range and show redder [7]--[15] but bluer $K_0$--[7] colours. The ISOCAM 7\,$\mu$m band is almost certainly affected by molecular absorption in ordinary M-giant stars. However, Mira variables are brighter than other stars possibly in part because of reduced SiO and H$_2$O absorption. The results from these fields should form a template for analyses of more heavily obscured regions about which little is known from visible-light studies. \section{Acknowlegments} We would like to acknowledge the Les Houches 1998 summer school for access to the JUN98 version of CIA. We acknowledge useful discussions with K.S. Krishnaswamy, TIFR. Eric Copet is thanked for his help with the Unix scripts and Martin Groenewegen for his comments on dust at 7\,$\mu$m. ISG wishes to acknowledge the hospitality of IAP during part of this work. SG and MS acknowlege receipts of fellowships from the Ministere des Affaires Etrang\`{e}res, France, and ESA respectively.
\section{Introduction} In the near future, a large part of the high energy experimental program will be devoted to testing the Cabibbo--Kobayashi--Maskawa (CKM) picture of quark mixing and $CP$ violation by directly measuring the sides and (some) angles of the unitarity triangle. If the value of $\sin(2\beta)$, the $CP$ asymmetry in $B\to J/\psi K_S$, is not too far from the CDF central value~\cite{sin2beta}, then searching for new physics at the $B$ factories will require a combination of several precision measurements. Particularly important are $|V_{ub}|$ and $|V_{td}|$, which are the least precisely known elements of the CKM matrix. The latter will be measured hopefully in the upcoming run of the Tevatron from the ratio of $B_s$ and $B_d$ mixing, and will not be discussed here. This talk is motivated by trying to understand what the chances are to \vspace{-9pt} \begin{itemize} \itemsep=-4pt \item \ldots reduce (conservative) error in $|V_{cb}|$ below $\sim5\%$? Although sometimes a smaller error is quoted already (e.g., by the Particle Data Group), it is hard to bound model independently the possible quark-hadron duality violation in the inclusive, and the size of $1/m_{c,b}^2$ corrections in the exclusive determination. \item \ldots determine $|V_{ub}|$ with less than $\sim10\%$ error? The inclusive measurements require significant cuts on the available phase space; the exclusive measurements require knowledge of the form factors. \item \ldots reduce the theoretical uncertainties in the $B\to X_s\gamma$ photon spectrum? The effect of the experimental cut on the photon energy needs to be better understood, and there are subtleties in the OPE beyond leading order. \end{itemize} \vspace{-9pt} \noindent The theoretical reliablility of inclusive measurements can be competitive with the exclusive ones (or even better in some cases). For example, for the determination of $|V_{cb}|$ model dependence enters at the same order of $\Lambda_{\rm QCD}^2/m^2$ corrections from both the inclusive semileptonic $\bar B\to X_c e\bar\nu$ width and the $\bar B\to D^* e\bar\nu$ rate near zero recoil. Inclusive $B$ decay rates can be computed model independently in a series in $\Lambda_{\rm QCD}/m_b$, using an operator product expansion (OPE)~\cite{CGG,incl,MaWi}. The $m_b\to\infty$ limit is given by $b$ quark decay. For most quantities of interest, this result is known including the order $\alpha_s$ and the dominant part of the order $\alpha_s^2$ corrections. Observables which do not depend on the four-momentum of the hadronic final state (e.g., total decay rate and lepton spectra) receive no correction at order $\Lambda_{\rm QCD}/m_b$ when written in terms of $m_b$, whereas differential rates with respect to hadronic variables (e.g., hadronic energy and invariant mass spectra) also depend on $\bar\Lambda/m_b$, where $\bar\Lambda$ is the $m_B-m_b$ mass difference in the $m_b\to\infty$ limit. At order $\Lambda_{\rm QCD}^2/m_b^2$, the corrections are parameterized by two hadronic matrix elements, usually denoted by $\lambda_1$ and $\lambda_2$. The value $\lambda_2 \simeq 0.12{\rm GeV}^2$ is known from the $B^*-B$ mass splitting. For inclusive $b\to q$ decay, corrections to the $m_b\to\infty$ limit are expected to be under control in parts of phase space where several hadronic final states are allowed to contribute with invariant masses satisfying $m_{X_q}^2 \gtrsim m_q^2 + {\rm (few\ times)}\Lambda_{\rm QCD} m_b$.\footnote{However, it is not true that several such states are required to contribute; e.g., $\bar B\to X_c e\bar\nu$ decay in the small velocity limit can be computed reliably, even though it is saturated by $D$ and $D^*$ only~\cite{SVlimit,SVdual}.} Such observables which ``average" sufficiently over different hadronic final states may be predicted reliably. (We are just beginning to learn quantitatively what sufficient averaging is.) The major uncertainty in these predictions is from the values of the quark masses and $\lambda_1$, or equivalently, the values of $\bar\Lambda$ and $\lambda_1$. These quantities can be extracted from heavy meson decay spectra, which is the subject of a large part of this talk. An important theoretical subtlety is related to the fact that $\bar\Lambda$ cannot be defined unambiguously beyond perturbation theory~\cite{renormalon}, and its value extracted from data using theoretical expressions valid to different orders in the $\alpha_s$ may vary by order $\Lambda_{\rm QCD}$. However, these ambiguities cancel~\cite{rencan} when one relates consistently physical observables to one another, i.e., the formulae used to determine $\bar\Lambda$ and $\lambda_1$ contain the same orders in the $\alpha_s$ perturbation series as those in which the so extracted values are applied to make predictions. \section{The HQET parameters $\bar\Lambda$ and $\lambda_1$} The shape of the lepton energy~\cite{gremmetal,Volo,GK} or hadronic invariant mass~\cite{FLSmass1,FLSmass2,GK} spectrum in semileptonic $\bar B\to X_c\,\ell\,\bar\nu$ decay, and the photon spectrum in $\bar B\to X_s\gamma$~\cite{AZ,LLMW,FLS,Bauer,KaNe} can be used to measure the heavy quark effective theory (HQET) parameters $\bar\Lambda$ and $\lambda_1$. Testing our understanding of the $\bar B\to X_c e\bar\nu$ spectra is important to assess the reliability of the inclusive determination of $|V_{cb}|$, and especially that of $|V_{ub}|$. Understanding the photon spectrum in $\bar B\to X_s\gamma$ is important to evaluate how precisely the total rate can be predicted in the presence of an experimental cut on the photon energy~\cite{CLEObsg}. Studying these distributions is also useful to establish the limitations of models which were built to fit the lepton energy spectrum in semileptonic decay. The theoretical predictions are known to order $\alpha_s^2\beta_0$, where $\beta_0=11-2n_f/3$ is the coefficient of the one-loop $\beta$-function. This part of the order $\alpha_s^2$ piece usually provides a reliable estimate of the full order $\alpha_s^2$ correction, and it is straightforward to compute using the method of Smith and Voloshin~\cite{SmVo}. In some cases the full $\alpha_s^2$ correction is known. The order $\Lambda_{\rm QCD}^3/m_b^3$ terms have also been studied~\cite{GK,FLSmass2,Bauer}, and are used to estimate the uncertainties. The OPE for semileptonic (or radiative) $B$ decay does not reproduce the physical spectrum point-by-point in regions of phase space where the hadronic invariant mass of the final state is restricted. For example, the maximum electron energy for a particular hadronic final state $X$ is $E_e^{\rm (max)}=(m_B^2-m_X^2)/2m_B$, so comparison with experimental data near the endpoint can only be made after sufficient smearing, or after integrating over a large enough region. The minimal size of this region was estimated to be around $300-500\,$MeV \cite{MaWi}. The hadron mass spectrum cannot be predicted point-by-point without additional assumptions, but moments of it are calculable. In general, higher moments of a distribution or moments over smaller regions are less reliable than lower moments or moments over larger regions. The strategy is to find observables which are sensitive to $\bar\Lambda$ and $\lambda_1$, but the deviations from $b$ quark decay are small, so that the contributions from operators with dimension greater than 5 are not too important. \subsection{$\bar B\to X_{\lowercase{c}} {\lowercase{e}}\bar\nu$ decay spectra} Last year the CLEO Collaboration measured the first two moments of the hadronic invariant mass-squared ($s_H$) distribution, $\langle s_H - \overline{m}_D^2 \rangle$ and $\langle (s_H - \overline{m}_D^2)^2 \rangle$, subject to the constraint $E_e > 1.5\,$GeV~\cite{CLEOparams}. Here $\overline{m}_D = (m_D+3m_{D^*})/4$. Each of these measurements give an allowed band on the $\bar\Lambda - \lambda_1$ plane. These bands are almost perpendicular, so they give the fairly small intersection region shown in Fig.~1. The central values at order $\alpha_s$ are \cite{CLEOparams} \begin{equation}\label{cleonums} \bar\Lambda = (0.33 \pm 0.08)\, {\rm GeV}\,, \qquad \lambda_1 = -(0.13 \pm 0.06)\,{\rm GeV}^2\,. \end{equation} The unknown order $\Lambda_{\rm QCD}^3/m_b^3$ corrections not included in this result introduce a large uncertainty, especially for the second moment. As a result, the allowed range is much longer in the direction of the first moment band than perpendicular to it; see Fig.~2. Similar information on $\bar\Lambda$ and $\lambda_1$ can be obtained from the lepton energy spectrum, $d\Gamma/dE_e$. It has been measured both by demanding only one charged lepton tag, and using a double tagged data sample where the charge of a high momentum lepton determines whether the other lepton in the event comes directly from semileptonic $B$ decay (primary) or from the semileptonic decay of a $B$ decay product charmed hadron (secondary). The single tagged data has smaller statistical errors, but it is significantly contaminated by secondary leptons below about $1.5\,$GeV. Therefore, Ref.~\cite{gremmetal} considered the observables \noindent \begin{minipage}[t]{8cm} \vspace{0.1cm} \centerline{\epsfxsize=8cm\epsffile{cleobands.eps}} \small{FIG.~1.~~Bands in the $\bar\Lambda - \lambda_1$ plane defined by the measured first and second moments of the hadronic mass-squared and lepton energy distributions. Note that $M_X^2 \equiv s_H$. (From Ref.~\cite{CLEOparams}.)} \vspace{0.4cm} \end{minipage} \hfill \begin{minipage}[t]{9cm} \vspace{0.1cm} \centerline{\epsfxsize=9cm\epsfysize=7cm\epsffile{bauer.eps}} \small{FIG.~2.~~Estimates of the theoretical uncertainty in $\bar\Lambda$ and $\lambda_1$ due to unknown $1/m_b^3$ terms from the shape of the electron spectrum (solid ellipse), from hadronic mass moments (long dashed ellipse), and the $\bar B\to X_s\gamma$ spectrum (short dashed ellipse; see Sec.~IIB). Only the relative size and orientation of the ellipses are meaningful, not their position. (From Refs.~\cite{FLSmass2} and \cite{Bauer}.)} \vspace{0.5cm} \end{minipage} \addtocounter{figure}{2} \begin{equation}\label{R12def} R_1 = {\displaystyle \int_{1.5\,{\rm GeV}} E_\ell\, {d\Gamma\over dE_\ell}\, dE_\ell \over \displaystyle \int_{1.5\,{\rm GeV}} {d\Gamma\over dE_\ell}\, dE_\ell }\,, \qquad R_2 = {\displaystyle \int_{1.7\,{\rm GeV}} {d\Gamma\over dE_\ell}\, dE_\ell \over \displaystyle \int_{1.5\,{\rm GeV}} {d\Gamma\over dE_\ell}\, dE_\ell }\,. \end{equation} Using the CLEO data~\cite{RoyPhD}, the central values $\bar\Lambda = 0.39\,$GeV and $\lambda_1 = -0.19\,{\rm GeV}^2$ were obtained~\cite{gremmetal}, which is in good agreement with Eq.~(\ref{cleonums}). However, following Ref.~\cite{Volo}, CLEO determined the first two moments of the spectrum, $\langle E_e \rangle$ and $\langle (E_e - \langle E_e \rangle)^2 \rangle$, without any restriction on $E_e$, using the double tagged data and an extrapolation to $E_e < 0.6\,$GeV. The result of this analysis is also plotted in Fig.~1, and yields quite improbable values for $\bar\Lambda$ and $\lambda_1$. The extrapolation to $E_e < 0.6\,$GeV introduces unnecessary model dependence, so this result may be less reliable than the hadronic invariant mass analysis. The extracted values of $\bar\Lambda$ and $\lambda_1$ are very sensitive to small systematic effects; and it seems problematic for the exclusive models used for the extrapolation to simultaneously reproduce the inclusive lepton spectrum and the $B$ semileptonic branching fraction~\cite{Ryd}. There are a number of points to emphasize regarding these results: \vspace{-8pt} \begin{enumerate} \itemsep=-4pt \item Taking $\bar\Lambda$ from Eq.~(\ref{cleonums}) gives a determination of $|V_{cb}|$ from semileptonic $B$ width with $\sim3\%$ uncertainty. \item The inclusive $|V_{cb}|$ seems to be slightly larger than the exclusive (maybe not ``significantly" but ``consistently"). Can the theoretical uncertainties be reduced by combining the exclusive and inclusive determinations of $|V_{cb}|$? \item Since the bands from the lepton energy spectrum and the first moment of the hadronic mass-squared spectrum are almost parallel, an independent constraint on the $\bar\Lambda - \lambda_1$ plane is needed. (See Section~II~B.) \item Models do not seem to do well for $\bar B\to D^{**}e\bar\nu$ and $\bar B\to D^{(*)}\pi e\bar\nu$ --- relying on them may be dangerous. Composition of semileptonic $B$ decay seems to be not really understood yet. (See next subsection.) \end{enumerate} \vspace{-8pt} \subsubsection{Semileptonic $B$ decays to excited charmed mesons} In the heavy quark symmetry limit~\cite{HQS}, the spin and parity of the light degrees of freedom in a heavy meson are conserved. In the charm sector, the ground state is the $(D,\, D^*)$ doublet of heavy quark spin symmetry with spin-parity of the light degrees of freedom $s_\ell^{\pi_\ell} = \frac12^-$. The four lightest excited states, sometimes referred to as $D^{**}$, are the $(D_0^*,\, D_1^*)$ doublet with $s_\ell^{\pi_\ell} = \frac12^+$ and the $(D_1,\, D_2^*)$ doublet with $s_\ell^{\pi_\ell} = \frac32^-$. The $D_1$ and $D_2^*$ have been observed with masses near $2.42$ and $2.46\,$GeV, respectively, and width around $20\,$MeV. States in the $s_l^{\pi_l}=\frac12^+$ doublet can decay into $D^{(*)}\pi$ in an $s$-wave, and so they are expected to be much broader than the $D_1$ and $D_2^*$ which can only decay in a $d$-wave. (An $s$-wave decay for the $D_1$ is forbidden by heavy quark spin symmetry \cite{IWprl}.) The first observation of the $D_1^*$ state with mass and width about $2.46\,$GeV and $290\,$MeV, respectively, was reported at this Conference~\cite{CLEObroad}. $\bar B\to D_1 e\bar\nu$ and $\bar B\to D_2^* e\bar\nu$ account for sizable fractions of semileptonic $B$ decays, and are probably the only three-body semileptonic $B$ decays (other than $\bar B\to D^{(*)} e\bar\nu$) whose differential decay distributions will be precisely measured. ALEPH and CLEO measured recently, with some assumptions, ${\cal B}(\bar B\to D_1 e\bar\nu_e) = (6.0\pm1.1) \times 10^{-3}$~\cite{AlephCleo}, while $\bar B\to D_2^* e\bar\nu_e$ has not been seen yet. Heavy quark symmetry implies that in the $m_Q\to\infty$ limit ($Q=c,b$), matrix elements of the weak currents between a $B$ meson and an excited charmed meson vanish at zero recoil. However, in some cases at order $\Lambda_{\rm QCD}/m_Q$ these matrix elements are not zero. Since most of the phase space for semileptonic $B$ decay to excited charmed mesons is near zero recoil, $1 < v \cdot v' \lesssim 1.3$, $\Lambda_{\rm QCD}/m_Q$ corrections can be very important. The matrix elements of the weak currents between $B$ mesons and $D_1$ or $D_2^*$ mesons are conventionally parameterized in terms of a set of eight form factors $f_i$ and $k_i$~\cite{LLSW}. At zero recoil only $f_{V_1}$, defined by \begin{equation}\label{formf1} \langle D_1(v',\epsilon)|\, \bar c\,\gamma^\mu\,b\, |B(v)\rangle = \sqrt{m_{D_1}\,m_B}\, [ f_{V_1}\, \epsilon^{*\mu} + (f_{V_2} v^\mu + f_{V_3} v'^\mu)\, (\epsilon^*\cdot v) ] \,, \end{equation} can contribute to the matrix elements. In the $m_Q\to\infty$ limit, $f_i$ and $k_i$ are given in terms of a single Isgur-Wise function, $\tau(w)$~\cite{IWsr}. Heavy quark symmetry does not fix $\tau(1)$, since $f_{V_1} = (1-w^2)\,\tau(w) + {\cal O}(1/m_Q)$, and so $f_{V_1}(1) = 0$ in the infinite mass limit independent of the value of $\tau(1)$. At order $1/m_Q$ several new Isgur-Wise functions occur, together with the parameter $\bar\Lambda'$, which is the analog of $\bar\Lambda$ for the $(D_1,\, D_2^*)$ doublet. $\bar\Lambda' - \bar\Lambda \simeq 0.39\,$GeV follows from the measured meson masses~\cite{LLSW}. At order $1/m_Q$, $f_{V_1}(1)$ is no longer zero, but it can be written in terms of $\bar\Lambda'-\bar\Lambda$ and the Isgur-Wise function $\tau(w)$ evaluated at zero recoil~\cite{LLSW}, \begin{equation}\label{fV1} \sqrt{6}\, f_{V_{1}}(1) = - {4 (\bar\Lambda'-\bar\Lambda) \over m_c}\, \tau (1)\,. \end{equation} This relation means that at zero recoil heavy quark symmetry gives some model independent information about the $1/m_Q$ corrections, similar to Luke's theorem~\cite{Luke} for the decay into the ground state $(D,\, D^*)$ doublet. Since the allowed kinematic range for $\bar B\to D_1 e\bar\nu_e$ and $\bar B\to D_2^* e\bar\nu_e$ decay are fairly small ($ 1 < w \lesssim 1.3$), and there are some constraints on the $1/m_Q$ corrections at zero recoil, it is useful to consider the decay rates expanded in powers of $w-1$. The general structure of the expansion of $d\Gamma/dw$ is elucidated schematically below, \begin{eqnarray}\label{schematic} {d\Gamma_{D_1}^{(\lambda=0)}\over dw} &\sim& \sqrt{w^2-1}\, \Big[ ( 0 + 0\,\varepsilon + \varepsilon^2 + \ldots ) + (w-1)\, ( 0 + \varepsilon + \ldots ) + (w-1)^2\, ( 1 + \varepsilon + \ldots ) + \ldots \Big] \,, \nonumber\\* {d\Gamma_{D_1}^{(|\lambda|=1)}\over dw} &\sim& \sqrt{w^2-1}\, \Big[ ( 0 + 0\,\varepsilon + \varepsilon^2 + \ldots ) + (w-1)\, ( 1 + \varepsilon + \ldots ) + (w-1)^2\, ( 1 + \varepsilon + \ldots ) + \ldots \Big] \,, \nonumber\\* {d\Gamma_{D_2^*}^{(|\lambda|=0,1)}\over dw} &\sim& (w^2-1)^{3/2} \Big[ ( 1 + \varepsilon + \ldots ) + (w-1) ( 1 + \varepsilon + \ldots ) + \ldots \Big] \,. \end{eqnarray} Here $\lambda$ is the helicity of the $D_1$ or $D_2^*$, and $\varepsilon^n$ denotes a term of order $(\Lambda_{\rm QCD}/m_Q)^n$. The zeros in Eq.~(\ref{schematic}) are consequences of heavy quark symmetry. \begin{table}[t] \begin{tabular}{c|ccc} ~~~Approximation~~~ & $R=\Gamma_{D_2^*}\big/\Gamma_{D_1}$ & $\tau(1)\, \bigg[\displaystyle {6.0\times10^{-3} \over {\cal B}(\bar B\to D_1\,e\,\bar\nu_e)} \bigg]^{1/2}$ & $\Gamma_{D_1+D_2^*+D_1^*+D_0^*}\big/\Gamma_{D_1}$~~~ \\[8pt] \hline B$_\infty$ & $1.65$ & $1.24$ & $4.26$ \\ B$_1$ & $0.52$ & $0.71$ & $2.55$ \\ B$_2$ & $0.67$ & $0.75$ & $2.71$ \\ \end{tabular} \vspace{4pt} \caption[6]{Predictions for $\Gamma_{D_2^*}/\Gamma_{D_1}$, $\tau(1)$, and $\Gamma_{D_1+D_2^*+D_1^*+D_0^*}/\Gamma_{D_1}$. (From Ref.~\cite{LLSW}.)} \end{table} Table~I summarizes some of the more important predictions. The first row ($B_\infty$) shows the infinite mass limit. Approximations $B_1$ and $B_2$ are two different ways of treating unknown $1/m_Q$ corrections, and may give an indication of the uncertainty at this order. One of the most interesting predictions is that the $B\to D_1$ decay rate should be larger than $B\to D_2^*$, contrary to the infinite mass limit, but in agreement with the data. A number of other predictions for different helicity amplitudes, factorization of $\bar B\to (D_1,\, D_2^*)\pi e\bar\nu$ decay, sum rules, etc., are presented in Ref.~\cite{LLSW} (see also \cite{MNfact}). The main points are: \vspace{-8pt} \begin{enumerate} \itemsep=-4pt \item At zero recoil, order $1/m_Q$ contributions to semileptonic $B\to D_1,\, D_2^*$ decays (any excited charmed meson with $+$ parity) are determined by the $m_Q\to\infty$ Isgur-Wise function and known hadron mass splittings. \item Decay spectra can be predicted near zero recoil, including the $1/m_Q$ corrections, with reasonable assumptions. \item Test heavy quark symmetry for $B$ decays to excited charmed mesons, where the $1/m_Q$ terms are sometimes the leading contributions. Large $1/m_Q$ corrections to some predictions can be checked against data. \item Better understanding of inclusive$\,=\sum\,$exclusive in semileptonic $B$ decay. If semileptonic $B$ decays into the four $D^{**}$ states indeed account for less than 2\% of the $B$ width, then about another 2\% of $B$ decays must be semileptonic decays to higher mass excitations or nonresonant channels. \item Constrain the validity of models constructed to fit the decays to the ground state final state (most predict, for example, larger $B\to D_2^*$ than $B\to D_1$ semileptonic rate). \end{enumerate} \vspace{-8pt} \subsection{$\bar B\to X_{\lowercase{s}} \gamma$ photon spectrum} Comparison of the measured weak radiative $\bar B\to X_s\gamma$ decay rate with theory is an important test of the standard model. In contrast to the decay rate itself, the shape of the photon spectrum is not expected to be sensitive to new physics, but it can nevertheless provide important information. First of all, studying the photon spectrum is important for understanding how precisely the total rate can be predicted in the presence of an experimental cut on the photon energy~\cite{CLEObsg}, which is needed for a model independent interpretation of the resulting decay rate. Secondly, moments of the photon spectrum may be used to measure the HQET parameters $\bar\Lambda$ and $\lambda_1$~\cite{AZ,LLMW,FLS,Bauer,KaNe}. An important observable is \begin{equation}\label{moment1} \overline{(1 - x_B)} \Big|_{x_B > 1 - \delta} = {\displaystyle \int_{1-\delta}^1 dx_B\, (1-x_B)\, \frac{d\Gamma}{dx_B} \over \displaystyle \int_{1-\delta}^1 dx_B\, \frac{d\Gamma}{dx_B} } \,, \end{equation} where $x_B = 2E_\gamma / m_B$ is the rescaled photon energy. The parameter $\delta = 1 - 2E_\gamma^{\rm min}/m_B$ corresponds to the experimental lower cut on $E_\gamma$, and it has to satisfy $\delta > \Lambda_{\rm QCD}/m_B$; otherwise nonperturbative effects are not under control. It is straightforward to show that~\cite{LLMW} \begin{equation}\label{beauty} \overline{(1 - x_B)} \Big|_{x_B > 1-\delta} = {\bar\Lambda\over m_B} + \bigg(1-{\bar\Lambda\over m_B}\bigg)\, \int_{1-\delta}^1 dx_b\, (1-x_b)\, \frac1{\Gamma_0}\, \frac{d\Gamma}{dx_b} - {\bar\Lambda\over m_B}\, \delta(1-\delta)\, \frac1{\Gamma_0}\, \frac{d\Gamma}{dx_b} \bigg|_{x_b = 1-\delta} + \ldots \,, \end{equation} where $x_b = 2E_\gamma / m_b$, and $d\Gamma/dx_b$ is the photon spectrum in $b$ quark decay. It has recently been computed away from the endpoint ($x_b = 1$) to order $\alpha_s^2\beta_0$~\cite{LLMW}. $\Gamma_0 = G_F^2\,|V_{tb}V_{ts}^*|^2\,\alpha_{\rm em}\,C_7^2\, m_b^5 / 32\pi^4$ is the contribution of the tree level matrix element of $O_7 (\sim \bar s_L \sigma_{\mu\nu} F^{\mu\nu} b_R)$ to the $\bar B\to X_s\gamma$ decay rate. All terms but the first one on the right-hand-side of Eq.~(\ref{beauty}) have perturbative expansions which begin at order $\alpha_s$. The ellipses denote contributions of order $(\Lambda_{\rm QCD}/m_B)^3$, $\alpha_s(\Lambda_{\rm QCD}/m_B)^2$, and $\alpha_s^2$ terms not enhanced by $\beta_0$, but it does not contain contributions of order $(\Lambda_{\rm QCD}/m_B)^2$ or additional terms of order $\alpha_s(\Lambda_{\rm QCD}/m_B)$. Terms in the operator product expansion proportional to $\lambda_{1,2}/m_b^2$ enter precisely in the form so that they are absorbed in $m_B$ in Eq.~(\ref{beauty})~\cite{AZ}. A determination of $\bar\Lambda$ is straightforward using Eq.~(\ref{beauty}). The left hand side is directly measurable, while the quantities entering on the right-hand-side are presented in Ref.~\cite{LLMW}. The CLEO data in the region $E_\gamma > 2.1\,$GeV~\cite{CLEObsg} yields the central values $\bar\Lambda_{\alpha_s^2\beta_0} \simeq 270\,$MeV and $\bar\Lambda_{\alpha_s} \simeq 390\,$MeV. Uncertainties due to the unknown order $\Lambda_{\rm QCD}^3/m_b^3$ terms in the OPE give the short dashed ellipse in Fig.~2~\cite{Bauer}, whose major axis is roughly perpendicular to those from $\bar B\to X_c e\bar\nu$. This is why it is important to determine $\bar\Lambda$ and $\lambda_1$ from both analyses. The potentially most serious uncertainty is from both nonperturbative and perturbative terms that are singular as $x_b\to 1$ and sum into a shape function~\cite{shapefn} that modifies the spectrum near the endpoint. For sufficiently large $\delta$ these effects are not important. They have been estimated in Refs.~\cite{Bauer,KaNe} using phenomenological models. Whether these effects are small in a certain range of $\delta$ can be tested experimentally by checking if the extracted value of $\bar\Lambda$ is independent of $\delta$. This would also improve our confidence that the total decay rate in the region $x_B>1-\delta$ can be predicted model independently. There have been other important developments for the $\bar B\to X_s\gamma$ decay rate. The next-to-leading order computation of the total rate has been completed~\cite{Misiak,match,fourquark,matrixel}, reducing the theoretical uncertainties significantly. It was realized by Voloshin that beyond leading order there are terms in the OPE for the decay rate which are suppressed only by $\Lambda_{\rm QCD}^2/m_c^2$ instead of $\Lambda_{\rm QCD}^2/m_b^2$~\cite{1/mc2}. In fact, there is a series of contributions of the form $(\Lambda_{\rm QCD}^2/m_c^2)\, (\Lambda_{\rm QCD} m_b/m_c^2)^n$, which gives for $n=0$ a calculable correction of $\delta\Gamma/\Gamma = - (C_2/9C_7) (\lambda_2/m_c^2) \simeq 2.5\%$. It is also known that there are uncalculable contributions suppressed by $\alpha_s$ but not by $\Lambda_{\rm QCD}/m_b$ from photon coupling to light quarks, for which there is no OPE~\cite{KLP}. While no correction larger than a few percent has been identified, a better understanding of the nonperturbative contributions would be desirable once the four-quark operators are included. \section{$|V_{\lowercase{ub}}|$ from the $\bar B\to X_{\lowercase{u}} {\lowercase{e}}\bar\nu$ hadron mass spectrum} The traditional method for extracting $|V_{ub}|$ involves a study of the electron energy spectrum in inclusive semileptonic $B$ decay. Electrons with energies in the endpoint region $m_B/2 > E_e > (m_B^2-m_D^2)/2m_B$ (in the $B$ rest frame, and neglecting the pion mass) must arise from $b \to u$ transition. Since the size of this region is only about $300\,$MeV, at the present time it is not known how to make a model independent prediction for the spectrum in this region. Another possibility for extracting $|V_{ub}|$ is based on reconstructing the neutrino momentum. The idea is to reconstruct $p_{\bar\nu}$ and infer the invariant mass-squared of the hadronic final state, $s_H = (p_B - q)^2$, where $q = p_e + p_{\bar\nu}$. Semileptonic $B$ decays satisfying $s_H < m_D^2$ must come from $b \to u$ transition~\cite{FLW,DiUr,oldmass}. The first analyses of LEP data utilizing this idea have been performed recently~\cite{Vubmass}. Both the invariant mass region, $s_H < m_D^2$, and the electron endpoint region, $m_B/2 > E_e > (m_B^2-m_D^2)/2m_B$, receive contributions from hadronic final states with invariant masses between $m_\pi$ and $m_D$. However, for the electron endpoint region the contribution of states with masses nearer to $m_D$ is strongly suppressed kinematically. In fact, in the ISGW model~\cite{ISGW} the electron endpoint region is dominated by the $\pi$ and the $\rho$ with higher mass states making a small contribution, and this region includes only of order 10\% of the $\bar B\to X_u e\bar\nu$ rate. The situation is very different for the low invariant mass region, $s_H < m_D^2$. Now all states with invariant masses up to $m_D$ contribute without any preferential weighting towards the lowest mass ones. In this case the ISGW model suggests the $\pi$ and the $\rho$ comprise only about a quarter of the $B$ semileptonic decays to states with $s_H < m_D$, and only of order 10\% of the $\bar B\to X_u e\bar\nu$ rate is excluded from this region. Consequently, it is much more likely that the first few terms in the OPE provide an accurate description of $B$ semileptonic decay in the region $s_H < m_D^2$ than in the region $E_e > (m_B^2-m_D^2)/2m_B$. Combining the invariant mass constraint with a modest cut on the electron energy will not destroy this conclusion. Let us first consider the contribution of dimension three operators in the OPE to the hadronic mass-squared spectrum. This is equivalent to $b$ quark decay, and implies a result for $d\Gamma / dE_0\, ds_0$ (where $E_0 = p_b \cdot (p_b-q)/m_b$ and $s_0 = (p_b-q)^2$ are the energy and invariant mass of the strongly interacting partons arising from the $b$ quark decay) that is straightforward to calculate to order $\alpha_s^2\beta_0$~\cite{FLW}. Even at this leading order in the OPE there are important nonperturbative effects that come from the difference between $m_b$ and $m_B$. The most significant effect comes from $\bar\Lambda$, and including only it ({\it i.e.}, neglecting $\lambda_{1,2}$), the hadronic invariant mass $s_H$ is related to $s_0$ and $E_0$ by~\cite{FLSmass1} \begin{equation} s_H = s_0 + 2 \bar\Lambda E_0 + \bar\Lambda^2 + \ldots \,. \end{equation} Changing variables from $(s_0, E_0)$ to $(s_H, E_0)$ and integrating $E_0$ over the range \begin{equation} \sqrt{s_H} - \bar\Lambda < E_0 < {1\over 2m_B}\, (s_H - 2\bar\Lambda m_B + m_B^2), \end{equation} gives $d\Gamma/ds_H$, where $\bar\Lambda^2<s_H<m_B^2$. Feynman diagrams with only a $u$ quark in the final state contribute at $s_0 = 0$, which corresponds to the region $\bar\Lambda^2 < s_H < \bar\Lambda m_B$. Although $d\Gamma/ds_H$ is integrable in perturbation theory, it has a double logarithmic singularity at $s_H = \bar\Lambda m_B$. At higher orders in perturbation theory, increasing powers of $\alpha_s \ln^2[(s_H - \bar\Lambda m_B) / m_B^2]$ appear in the invariant mass spectrum. Therefore, $d\Gamma/ds_H$ in the vicinity of $s_H = \bar\Lambda m_B$ is hard to predict reliably even in perturbation theory. (In the region $s_H \lesssim \bar\Lambda m_B$ nonperturbative effects are also important.) The behavior of the spectrum near $s_H = \bar\Lambda m_B$ becomes less important for observables that average over larger regions of the spectrum, such as $d\Gamma/ds_H$ integrated over $s_H<\Delta^2$, with $\Delta^2$ significantly greater than $\bar\Lambda m_B$. Fig.~3 shows the fraction of $\bar B\to X_u e\bar\nu$ events in the region $s_H < \Delta^2$ as a function of $\Delta^2$ for three different values of $\bar\Lambda$. For a certain value of $\bar\Lambda$, Fig.~3 together with Eq.~(\ref{Vub}) can be used to extract $|V_{ub}|$ from data, up to the nonperturbative effects discussed next. \begin{figure}[t] \centerline{\epsfysize=7cm\epsffile{flw.eps}} \caption[]{Fraction of $\bar B\to X_u e\bar\nu$ decays with $s_H < \Delta^2$, for $\bar\Lambda = 0.2\,$GeV (dotted curve), $0.4\,$GeV (solid curve), and $0.6\,$GeV (dashed curve). Note that $m_D^2 = 3.5\,{\rm GeV}^2$. Nonperturbative effects not in $\bar\Lambda$ are neglected. (From Ref.~\cite{FLW}.) } \end{figure} In the low mass region, $s_H \lesssim \bar\Lambda m_B$, nonperturbative corrections from higher dimension operators in the OPE are very important. Just as in the case of the electron endpoint region in semileptonic $B$ decay or the photon energy endpoint region in radiative $B$ decay, the most singular terms can be identified and summed into a shape function. These shape functions depend on the same infinite set of matrix elements. Since $\bar\Lambda m_b\approx2\,{\rm GeV}^2$ is not too far from $m_D^2$, it is necessary to estimate the influence of the nonperturbative effects on the fraction of $B$ decays with $s_H < m_D^2$. It is difficult to estimate this model independently, but upper bounds can be derived on the fraction of $\bar B\to X_u e\bar\nu$ events with $s_H > \Delta^2$ assuming that the shape function is positive~\cite{FLW}. In the ACCMM model \cite{ACCMM} with reasonable parameters, the shape function causes a small (i.e., $\sim4\%$ with $\bar\Lambda=0.4\,$GeV, and perturbative QCD corrections neglected) fraction of the events to have $s_H>m_D^2$ \cite{FLW}. This suggests that sensitivity to unknown higher dimension operators in the OPE will probably not give rise to a large uncertainty in $|V_{ub}|$ if it is determined from the hadronic invariant mass spectrum in the region $s_H<m_D^2$. If experimental resolution forces one to consider a significantly smaller region, then the sensitivity to higher dimension operators increases rapidly. In summary, to extract $|V_{ub}|$ from $d\Gamma/ds_H$ with small theoretical uncertainty, one needs to: \vspace{-9pt} \begin{enumerate} \itemsep=-4pt \item move the experimental cut $\Delta$ as close to $m_D$ as possible; \item determine $\bar\Lambda$ (at order $\alpha_s^2$) with $\lesssim 50\,$MeV uncertainty. \end{enumerate} \vspace{-9pt} \noindent Then a determination of $|V_{ub}|$ with $\sim 10\%$ theoretical uncertainty seems feasible. \section{Upsilon expansion} The main uncertainties in the theoretical predictions for inclusive $B$ decay rates, e.g., $\bar B\to X_u e\bar\nu$, arise from the $m_b^5$ dependence on the $b$ quark mass and the bad behavior of the series of perturbative corrections when it is written in terms of the pole mass. In fact, only the product of these quantities is unambiguous, but perturbative multi-loop calculations are most comfortably done in terms of the pole mass. Of course, one would like to eliminate the quark mass altogether from the predictions in favor of a physical observable. Here we present a new method of eliminating $m_b$ in terms of the $\Upsilon(1S)$ meson mass~\cite{upsexp} (instead of $m_B$ and $\bar\Lambda$ discussed in Sec.~II). Let us consider the inclusive $\bar B\to X_u e\bar\nu$ decay rate~\cite{LSW}. At the scale $\mu = m_b$, \begin{equation}\label{bupole} \Gamma(B\to X_u e\bar\nu) = {G_F^2 |V_{ub}|^2\over 192\pi^3}\, m_b^5\, \bigg[ 1 - 2.41 {\alpha_s\over\pi}\, \epsilon - 3.22 {\alpha_s^2\over\pi^2} \beta_0\, \epsilon^2 - 5.18 {\alpha_s^3\over\pi^3} \beta_0^2\, \epsilon^3 - \ldots - {9\lambda_2 - \lambda_1 \over 2m_b^2} + \ldots \bigg] . \end{equation} The variable $\epsilon \equiv 1$ denotes the order in the modified expansion. The complete order $\alpha_s^2$ calculation was done recently~\cite{TvR}, and the result is about 90\% of the $\alpha_s^2\beta_0$ part. In comparison, the expansion of the $\Upsilon(1S)$ mass in terms of $m_b$~\cite{Upsmass} has a different structure, \begin{equation}\label{upsmass} {m_\Upsilon \over 2m_b} = 1 - {(\alpha_s C_F)^2\over8} \bigg\{ 1 \epsilon + {\alpha_s\over\pi} \bigg[\bigg( \ell + \frac{11}6\bigg) \beta_0 - 4 \bigg] \epsilon^2 + \bigg({\alpha_s\beta_0\over2\pi}\bigg)^2 \bigg( 3\ell^2 +9\ell +2\zeta(3)+\frac{\pi^2}6+\frac{77}{12}\bigg) \epsilon^3 + \ldots \bigg\}\, , \end{equation} where $\ell=\ln[\mu/(m_b\alpha_s C_F)]$ and $C_F=4/3$. In this expansion we assigned to each term one less power of $\epsilon$ than the power of $\alpha_s$, because as we will sketch below, this is the consistent way of combining Eqs.~(\ref{bupole}) and (\ref{upsmass}). It is also convenient to choose the same renormalization scale $\mu$. The prescription of counting $[\alpha_s(m_b)]^n$ in $B$ decay rates as order $\epsilon^n$, and $[\alpha_s(m_b)]^n$ in $m_\Upsilon$ as order $\epsilon^{n-1}$ is called the upsilon expansion. Note that it combines different orders in the $\alpha_s$ perturbation series in Eqs.~(\ref{bupole}) and (\ref{upsmass}). The theoretical consistency of the upsilon expansion was shown at large orders for the terms containing the highest possible power of $\beta_0$, and to order $\epsilon^2$ including non-Abelian contributions. An explicit calculation using the Borel transform of the static quark potential~\cite{ugo} shows that the coefficient of the order $\alpha_s^{n+2}$ term in Eq.~(\ref{upsmass}) of the form $(\ell^n+ \ell^{n-1}+ \ldots+1)$ exponentiates to give $\exp(\ell) = \mu/(m_b\alpha_s C_F)$, and corrects the mismatch of the power of $\alpha_s$ between the two series. This is also needed for the cancellation of the renormalon ambiguities in the energy levels as given by $2m_b$ plus the potential and kinetic energies~\cite{andre,Beneke}. The infrared sensitivity of Feynman diagrams can be studied by introducing a fictitious infrared cutoff $\lambda$. The infrared sensitive terms are nonanalytic in $\lambda^2$, such as $(\lambda^2)^{n/2}$ or $\lambda^{2n}\ln\lambda^2$, and arise from the low-momentum part of Feynman diagrams. Diagrams which are more infrared sensitive, i.e., have contributions $(\lambda^2)^{n/2}$ or $\lambda^{2n}\ln\lambda^2$ for small values of $n$, are expected to have larger nonperturbative contributions. Linear infrared sensitivity, i.e., terms of order $\sqrt{\lambda^2}$, are a signal of $\Lambda_{\rm QCD}$ effects, quadratic sensitivity, i.e., terms of order $\lambda^2\ln \lambda^2$ are a signal of $\Lambda_{\rm QCD}^2$ effects, etc. From Refs.~\cite{SAZ} and \cite{Beneke} follows that the linear infrared sensitivity cancels in the upsilon expansion to order $\epsilon^2$ (probably to all orders as well, but the demonstration of this appears highly non-trivial). Substituting Eq.~(\ref{upsmass}) into Eq.~(\ref{bupole}) and collecting terms of a given order in $\epsilon$ gives~\cite{upsexp} \begin{equation}\label{buups} \Gamma(\bar B\to X_u e\bar\nu) = {G_F^2 |V_{ub}|^2\over 192\pi^3}\, \bigg({m_\Upsilon\over2}\bigg)^5\, \bigg[ 1 - 0.115\epsilon - 0.035_{\rm BLM} \epsilon^2 - 0.005_{\rm BLM} \epsilon^3 - {9\lambda_2 - \lambda_1 \over 2(m_\Upsilon/2)^2} + \ldots \bigg] , \end{equation} where the BLM~\cite{BLM} subscript indicates that only the corrections proportional to the highest power of $\beta_0$ have been kept. The complete order $\epsilon^2$ term is $-0.041 \epsilon^2$~\cite{TvR}. The perturbation series, $1 - 0.115\epsilon - 0.035_{\rm BLM}\epsilon^2 - 0.005_{\rm BLM}\epsilon^3$, is far better behaved than the series in Eq.~(\ref{bupole}), $1 - 0.17\epsilon - 0.13_{\rm BLM}\epsilon^2 - 0.12_{\rm BLM}\epsilon^3$, or the series expressed in terms of the $\overline{\rm MS}$ mass, $1+0.30\epsilon+0.19_{\rm BLM}\epsilon^2+0.05_{\rm BLM}\epsilon^3$. The uncertainty in the decay rate using Eq.~(\ref{buups}) is much smaller than that in Eq.~(\ref{bupole}), both because the perturbation series is better behaved, and because $m_\Upsilon$ is better known (and better defined) than $m_b$. The relation between $|V_{ub}|$ and the total semileptonic $\bar B\to X_u e\bar\nu$ decay rate is~\cite{upsexp} \begin{equation}\label{Vub} |V_{ub}| = (3.06 \pm 0.08 \pm 0.08) \times 10^{-3} \left( {{\cal B}(\bar B\to X_u e\bar\nu)\over 0.001} {1.6\,{\rm ps}\over\tau_B} \right)^{1/2} , \end{equation} The first error is obtained by assigning an uncertainty in Eq.~(\ref{buups}) equal to the value of the $\epsilon^2$ term and the second is from assuming a $100\,$MeV uncertainty in Eq.~(\ref{upsmass}). The scale dependence of $|V_{ub}|$ due to varying $\mu$ in the range $m_b/2< \mu <2m_b$ is less than 1\%. The uncertainty in $\lambda_1$ makes a negligible contribution to the total error. Of course, it is unlikely that ${\cal B}(\bar B\to X_u e\bar\nu)$ will be measured without significant experimental cuts, for example, on the hadronic invariant mass (see Sec.~III), but this method should reduce the uncertainties in such analyses as well. The $\bar B\to X_c e\bar\nu$ decay depends on both $m_b$ and $m_c$. It is convenient to express the decay rate in terms of $m_\Upsilon$ and $\lambda_1$ instead of $m_b$ and $m_c$, using Eq.~(\ref{upsmass}) and \begin{equation}\label{mbmc} m_b - m_c = \overline{m}_B - \overline{m}_D + \bigg( {\lambda_1\over 2\overline{m}_B} - {\lambda_1\over 2\overline{m}_D} \bigg) + \ldots \,, \end{equation} where $\overline{m}_B = (3m_{B^*}+m_B)/4=5.313\,$GeV and $\overline{m}_D = (3m_{D^*}+m_D)/4=1.973\,$GeV. We then find \begin{equation}\label{bcups} \Gamma(\bar B\to X_c e\bar\nu) = {G_F^2 |V_{cb}|^2\over 192\pi^3} \bigg({m_\Upsilon\over2}\bigg)^5\, 0.533 \times \big[ 1 - 0.096\epsilon - 0.029_{\rm BLM}\epsilon^2 - (0.28\lambda_2 + 0.12\lambda_1)/{\rm GeV}^2 \big] \,, \end{equation} where the phase space factor has also been expanded in $\epsilon$. For comparison, the perturbation series in this relation when written in terms of the pole mass is $ 1- 0.12\epsilon- 0.06\epsilon^2 -\ldots$~\cite{LSW}. Equation~(\ref{bcups}) implies~\cite{upsexp} \begin{equation}\label{Vcb} |V_{cb}| = (41.6 \pm 0.8 \pm 0.7 \pm 0.5) \times 10^{-3} \times \eta_{\rm QED} \left( {{\cal B}(\bar B\to X_c e\bar\nu)\over0.105}\, {1.6\,{\rm ps}\over\tau_B}\right)^{1/2} , \end{equation} where $\eta_{\rm QED}\sim1.007$ is the electromagnetic radiative correction. The uncertainties come from assuming an error in Eq.~(\ref{bcups}) equal to the $\epsilon^2$ term, a $0.25\,{\rm GeV}^2$ error in $\lambda_1$, and a $100\,$MeV error in Eq.~(\ref{upsmass}), respectively. The second uncertainty can be removed by determining $\lambda_1$, as discussed in Sec.~II. Other applications, such as for nonleptonic decays, exclusive semileptonic decays and $\bar B\to X_s\gamma$ photon spectrum were studied in Refs.~\cite{upsexp,LLMW}. \begin{figure}[bt] \centerline{\epsfysize=6.8cm\epsffile{fig4.eps}} \caption[]{Prediction in the upsilon expansion at order $\epsilon$ (thick dashed curve) and $(\epsilon^2)_{\rm BLM}$ (thick solid curve) for $\overline{(1 - x_B)} |_{x_B > 1-\delta}$ defined in Eq.~(\ref{moment1}). The thin curves show the $O_7$ contribution only. (From Ref.~\cite{LLMW}.) } \end{figure} The most important uncertainty in this approach is the size of nonperturbative contributions to $m_\Upsilon$ other than those which can be absorbed into the $b$ quark mass. By dimensional analysis the size of this correction is of order $a^3\Lambda_{\rm QCD}^4$, where $a\sim1/(m_b \alpha_s)$ is the Bohr radius of the $\Upsilon$. Quantitative estimates, however, vary in a large range, and it is preferable to constrain such effects from data. The upsilon expansion yields parameter free predictions for $\overline{(1 - x_B)} |_{x_B > 1-\delta}$ defined in Eq.~(\ref{moment1}). The analog of Eq.~(\ref{beauty}) is~\cite{LLMW} \begin{equation}\label{upsbeauty} \overline{(1 - x_B)} \Big|_{x_B > 1-\delta} = 1 - {m_\Upsilon\over 2m_B} \left[ 1 + 0.011\epsilon + 0.019(\epsilon^2)_{\rm BLM} - \langle 1-x_b \rangle \Big|_{x_b > (2m_B/m_\Upsilon)(1-\delta)} \right] , \end{equation} For $E_\gamma>2.1\,$GeV this relation gives 0.111, whereas the central value from the CLEO data is around 0.093. Fig.~4 shows the prediction for $\overline{(1 - x_B)} |_{x_B > 1-\delta}$ as a function of $\delta$, both at order $\epsilon$ and $(\epsilon^2)_{\rm BLM}$. The perturbation expansion is very well behaved. In Eq.~(\ref{upsbeauty}) nonperturbative contributions to $m_\Upsilon$ other than those which can be absorbed into the $b$ quark mass have been neglected. If the nonperturbative contribution to $\Upsilon$ mass, $\Delta_\Upsilon$, were known, it could be included by replacing $m_\Upsilon$ by $m_\Upsilon-\Delta_\Upsilon$. For example, $\Delta_\Upsilon = +100\,$MeV increases $\overline{(1 - x_B)}$ by 7\%, so measuring $\overline{(1 - x_B)}$ with such accuracy will have important implications for the physics of quarkonia as well as for $B$ physics. \section{Conclusions} To conclude, let me emphasize the main points, and indicate what data would be important and useful in my opinion to address them: \vspace{-9pt} \begin{itemize} \itemsep=-4pt \item Experimental determination of $\bar\Lambda$ and $\lambda_1$ from the semileptonic $\bar B\to X_c e\bar\nu$ lepton energy and hadron mass spectra will reduce the theoretical uncertainties in $|V_{cb}|$ and $|V_{ub}|$. \hfil\break (Need: Double tagged lepton spectrum with smaller errors.) \item Photon energy spectrum in $\bar B\to X_s\gamma$ gives complimentary information on $\bar\Lambda$ and $\lambda_1$, even in the presence of an experimental cut on the photon energy. \hfil\break (Need: Spectrum with cut on $E_\gamma$ lowered; even a few hundred MeV can reduce the uncertainties significantly.) \item To distinguish $\bar B\to X_u e\bar\nu$ from $\bar B\to X_c e\bar\nu$, cutting on the hadronic invariant mass is theoretically cleaner than cutting on the lepton energy. \hfil\break (Need: Experimental cut on hadron mass as close to $m_D$ as possible, and a precise determination of $\bar\Lambda$.) \item The upsilon expansion is equivalent to using a short distance $b$ quark mass, but it eliminates $m_b$ altogether from the theoretical predictions in favor of $m_\Upsilon$ in a simple and consistent manner. It raises several interesting theoretical questions, and has many important applications. \end{itemize} \vspace{-9pt} \acknowledgements I thank Adam Falk and Alan Weinstein for inviting me to give this talk, and Christian Bauer for providing Fig.~2. Fermilab is operated by Universities Research Association, Inc., under DOE contract DE-AC02-76CH03000.
\section{INTRODUCTION}\label{intro.sec} The Casimir effect originally suggested in 1948 has been generally regarded as the contribution of a nontrivial geometry on the vacuum fluctuations of quantum electromagnetic fields~\cite{Casimir,Boyer}. The change in the vacuum fluctuations caused by the change of geometry appears as a shift of the vacuum energy and a resulting vacuum pressure. For a standard example, when we insert two perfectly conducting parallel plates into the free space $R^3$, the plates are attracted towards each other~\cite{Casimir}, although being uncharged. This attractive force is experimentally confirmed by Sparnaay in 1958~\cite{Sparnaay} and recently more precise measurements have been provided~\cite{Lamoreaux}. The dynamical Casimir effect suggests that the nonuniform accelerative relative motion of the boundaries (perfectly conducting plates or mirrors) excites the electromagnetic field and promotes virtual photons from the vacuum into real photons ~\cite{DCE_Moore,DCE_FD,DCE_Sassaroli,DCE_Dodonov,DCE_some}. The works on the dynamical Casimir effect are pioneered by Moore~\cite{DCE_Moore} and have progressed by many authors~\cite{DCE_FD,DCE_Sassaroli,DCE_Dodonov,DCE_some}. Moore studied the quantum theory of a massless scalar field in the one-dimensional cavity bounded by moving mirrors, and evaluated the number of photons created by the exciting effect of the moving mirrors. In his approach, the boundary condition on the scalar field is replaced with the simple equation, referred to as the Moore's equation, which describes the constraint on the conformal transformation of the coordinate. His approach has been popularly used to investigate the problems relating to the $(1+1)$-dimensional dynamical Casimir effect. For a well-known example Fulling and Davies calculated the energy-momentum tensor with the Moore's equation, and showed the existence of the radiation from the moving mirrors~\cite{DCE_FD}. The dynamical Casimir effect occurs even in the adiabatic approximation. Indeed, we can hardly handle the configurations except for the adiabatic deformations. Here, {\it adiabatic} means the absence of mixings among the different energy levels of the system during the modulation of the mirror separation. In other words the relative velocity of the mirrors is much smaller than the velocity of light. Especially Sassaroli {\it et al.} succeeded in evaluating the number of photons produced by the adiabatic motion of the mirrors in $1+3$ dimensions~\cite{DCE_Sassaroli}. They used the Bogolubov transformation among the creation and annihilation operators of photon in order to describe the particle production. The similar phenomena of the particle production also have been predicted in a variety of general-relativistic situations~\cite{GRPCreation,HawkingRadiation,Davis}. Such phenomena include the Hawking radiation from black holes~\cite{HawkingRadiation}, the domain-wall activity in cosmology, and the high-speed collision of atomic nuclei~\cite{Davis}. Although these phenomena are interesting, the dynamical Casimir effect has not yet been experimentally confirmed. If moving mirrors create radiation, the mirrors experience a radiation-reaction force. Several authors have discussed this subject within the adiabatic approximation. Dodonov {\it et al.} showed the existence of the additional negative frictional force besides the static Casimir force in the one-dimensional cavity by using Moore's equation~\cite{DCE_Dodonov}. The advantage of Moore's approach is the properties: the theory does not need to possess the Hamiltonian or the Lagrangian to describe the time evolution of the field. However, it seems difficult to apply Moore's approach to study the Casimir effects and its backreaction in $1+2$ or $1+3$ dimensions because the boundary condition of the one-dimensional space plays a crucial role in his approach. In this paper we present an effective-theoretical approach to studying the Casimir effects in $1+1$ dimensions. Our approach, making use of the action, is considered to be applicable to study the Casimir effects and its back reaction also in the higher dimensions. In general the existence of the moving boundaries (mirrors) makes it difficult to construct the Hamiltonian or the Lagrangian describing the system, since the relative motion of the boundaries (mirrors) mixes one energy level of the system with the others. However, we note that the adiabatic motion allows us to neglect this boundary effect: we do not need the boundaries. So we replace the spatial configuration $D^1$ into $S^1$ in the adiabatic approximation. The motion of the cavity size is described by varying the radius of $S^1$ in time. Furthermore, we can naturally identify the size of space $S^1$ with the scale factor of the Robertson-Walker-type metric. That is, the mirror separation is described by the scale factor. The time evolution of the scale factor can be regarded as the space-time $R\times S^1$ with gravity. For the sake of the replacement from $D^1$ into $S^1$, we can study the Casimir effects from the viewpoint of the effective theory. The construction of the model with the replacement is very simple and general, so that it is easy to apply our approach to more realistic models in the higher dimensions by replacing the space $D^1 \times R^n$ into $S^1\times R^n$. To check the validity of our replacement, we construct a scalar model and calculate the Casimir effects. As is usual our model makes use of the conformal symmetry property of the two-dimensional theory of massless fields. In our model of the cavity-system the classical action is constructed by the classical kinetic term of the mirror separation and the Polyakov action. The Polyakov action describes the massless scalar field minimally coupling to the two-dimensional gravity. The classical action is simple and general, so the structure of the model, e.g., symmetry, is easily visible. We carry out the path integral on the scalar field, and obtain the effective action for the mirror separation. The calculation of the path integral is rather complicated; however, it can be exactly performed. The effective action consists of the classical kinetic term of the mirror separation and the quantum correction terms. The quantum correction takes a well-known form, which consists of the static Casimir energy term and the conformal anomaly term. The conformal anomaly term represents the back reaction of the dynamical Casimir effect. The effective action finally leads to the dynamical vacuum pressure depending on the relative velocity of the mirrors. Our approach also gives an explanation for the origins of the Casimir effects in terms of the effective theory: the Casimir effects are caused by the change of field configuration in the vacuum instead of the existence of the boundaries. The paper is organized as follows. In Sec. \ref{model.sec} we provide the general description of our model and the definition of the effective action. In spite of the simplicity of our model, the calculation of the effective action is rather complicated. We show the calculation in detail in the following two sections. In Sec. \ref{partition.sec} the Casimir energy is shown to be derived from the partition function part in the effective action. In Sec. \ref{conf.sec} the conformal anomaly part in the effective action is calculated, and obtained the back-reaction term of the dynamical Casimir effect. In Sec. \ref{back.sec} the back reaction of both the Casimir effects in our model is investigated, and the dynamical vacuum pressure is derived. Section \ref{summary.sec} is devoted to conclusions and discussions. In Appendix A the conformal anomaly is induced by means of the Fujikawa method~\cite{Fujikawa} and in Appendix B another path-integral calculation on the Casimir energy are shown. \section{SCALAR MODEL FOR CASIMIR EFFECTS}\label{model.sec} The steps for constructing our model are as follows: For the purpose of describing the Casimir effects in the one-dimensional cavity and the reaction received by the moving mirrors, we consider a massless scalar field in the one-dimensional finite space with two boundaries, i.e., one-dimensional disk $D^1$ [see Fig. \ref{SpaceTime.eps}(a)]. That is, we consider the scalar field between two moving ``mirrors.'' The size of $D^1$ is a dynamical variable, and we assume that the size receives all the back reaction of the Casimir effects. The motion of the boundaries generally mixes the energy levels of the system. However, when the motion of the mirror separation is adiabatic, there are no transitions among the energy levels~\cite{DCE_Sassaroli}. Because of this absence of the transitions we can neglect the existence of the boundaries. This implies that each adiabatic Hamiltonian in the space $D^1$ is the same as that in the space $S^1$ except for the overall factor. We replace the spatial configuration $D^1$ with $S^1$ in the adiabatic approximation [see Fig. \ref{SpaceTime.eps}(b)]. In the space $S^1$ the scalar field is required to satisfy the periodic boundary condition rather than the fixed boundary condition. Accordingly the energy levels of the adiabatic oscillation modes in the replaced system are two times as those in the original system. We can naturally regard the size of $S^1$ as the scale factor of the Robertson-Walker-type metric. We define the Robertson-Walker-type metric on the space-time $R\times S^1$: \begin{eqnarray} ds^2 &=& -dt^2 + D^2(t)dx^2 \qquad (0\le x \le a), \end{eqnarray} where a dimensional constant $a$ is the standard space size and the scale factor $D(t)$ is the dimensionless magnification rate. It should be noticed that the mirror separation is replaced with the scale factor of the metric. \begin{figure}[htbp]% \begin{center}% \includegraphics{SpaceTime.eps \caption{Space times for the $(1+1)$-dimensional Casimir effects. (a) One-dimensional space with two boundaries (one-dimensional disk $D^1$) as the cavity between two moving ``mirrors.'' The scalar field satisfies the fixed boundary condition on the edges. (b) The space $S^1$ which is adiabatically equivalent for the scalar field to the geometrical configuration (a). The periodic boundary condition is imposed on the field. }% \label{SpaceTime.eps}% \end{center}% \end{figure} With the help of this replacement, the model in the two-dimensional gravity is applicable to our model. The mirror separation has finite reduced mass $m$ and classically obeys free motion. Then the classical action of our model to describe the system consists of both the classical kinetic term for the scale factor and the Polyakov action, \begin{eqnarray} S[D,\phi] &\equiv& \frac{1}{2\pi} \int dt \frac{m}{2} a^2 \dot{D}^2(t) \;+\; \frac12S_{\rm Polyakov}[g_{\mu\nu}(D),\,\phi], \end{eqnarray} where \begin{eqnarray} S_{\rm Polyakov}[g_{\mu\nu},\,\phi] &=& -\frac1{2\pi}\int d^2x\: \sqrt{-g}\:\frac12\:g^{\mu\nu} \partial_{\mu}\phi\partial_{\nu}\phi. \label{Polyakov.eq} \end{eqnarray} The Polyakov action is invariant under both the general coordinate transformation and the Weyl transformation. This property is referred to as the conformal symmetry. We can always rewrite the metric into the conformal flat form by the general coordinate transformation: \begin{eqnarray} ds^2 &=& -dt^2 + D^2(t)dx^2 \;=\; - C(\eta)\left(d\eta^2 - dx^2\right) \;=\; g_{\mu\nu} \; dx^\mu dx^\nu, \label{RWMetric.eq} \end{eqnarray} where we have introduced a new coordinate $\eta$ such that $d\eta\equiv dt/D(t)$ and $C(\eta)\equiv D^2[t(\eta)]$. After performing the Weyl transformation $g_{\mu\nu}\to C^{-1}(\eta) g_{\mu\nu}$, we have the $D(t)$-independent flat metric, \begin{eqnarray} ds^2 &=& -d\eta^2 + dx^2 \;=\; \eta_{\mu\nu} \; dx^\mu dx^\nu. \label{FlatMetric.eq} \end{eqnarray} This implies that any deformation of the space size does not affect the classical action. But once we quantize the scalar field, the conformal anomaly appears in general. The quantum effects lead to the motion of the scale factor, i.e., the motion of the mirror separation. We use a path-integral formulation to evaluate the motion of $D(t)$ as the back reaction of the Casimir effects. We use the background field method, in which the metric is treated as a classical field and the scalar field is quantized. We obtain the effective action for $D(t)$ by integrating out the scalar field. The effective action for the metric, $S_{\rm eff}[D]$, is given by \begin{eqnarray} e^{iS_{\rm eff}[D]} &\equiv& \int{\cal D}\phi \; e^{iS[D,\,\phi]}\\ &=& e^{i \frac{1}{2\pi} \int dt \frac{m}{2} a^2 \dot{D}^2(t) \;+\; i\frac12\Gamma[g_{\mu\nu}(D)]},\\ e^{i\Gamma[g_{\mu\nu}]} &\equiv& \int{\cal D}\phi \; e^{iS_{\rm Polyakov}[g_{\mu\nu},\,\phi]}. \label{EffAction1.eq} \end{eqnarray} In order to calculate the effective action for the evolving metric (\ref{RWMetric.eq}), we perform the conformal transformation on the effective action (\ref{EffAction1.eq}) from the evolving metric (\ref{RWMetric.eq}) to the flat metric (\ref{FlatMetric.eq}): $g_{\mu\nu}\to e^{2\alpha} g_{\mu\nu} = C^{-1}(\eta) g_{\mu\nu} = \eta_{\mu\nu}$. By means of the Fujikawa method~\cite{Fujikawa} this conformal transformation picks up the conformal anomaly as a Jacobian factor from the path-integral measure in the effective action (\ref{EffAction1.eq}): \begin{eqnarray} e^{i\Gamma[g_{\mu\nu}]} &=& \exp\left[-\frac{i}{2}\int d^2x\,\alpha(x) \sum_n\varphi^{{\dagger}}_n(x)\,\varphi_n(x) \right] e^{i\Gamma[\eta_{\mu\nu}]} \quad , \label{EffAction2.eq} \end{eqnarray} where the parameter of the conformal transformation $\alpha(x)$ is chosen as $\alpha(x) = - \frac12 \ln C(\eta)$. $\{\varphi_n(x)\}$ is a complete set which consists of the eigenfunctions of the Hamiltonian (see Appendix A). The first exponential factor in Eq. (\ref{EffAction2.eq}) is the conformal anomaly, and the second factor is the partition function for the free scalar field in the space $S^1$. \section{CASIMIR ENERGY IN SPACE $S^1$}\label{partition.sec} We will see that $\Gamma[\eta_{\mu\nu}]$ induces the Casimir energy as the vacuum energy by evaluating the partition function for the free scalar field. Let us calculate the Euclidean partition function \begin{eqnarray} Z_E &\equiv& e^{-\Gamma_E[\eta_{\mu\nu}]} \;=\; \int{\cal D}{\phi}\, \exp \left[ -\frac1{2\pi} \int_{-\infty}^{\infty} dx^2 \int_0^a dx^1 \frac12 \, \partial\phi \, \partial\phi \right], \label{partition.eq} \end{eqnarray} where we have defined the imaginary time variable $x^2 \equiv i\eta$, and have used the Euclidean inner product $\partial\phi \partial\phi \equiv \delta^{\mu\nu}\partial_\mu\phi \partial_\nu\phi$. Since the free Lagrangian is quadratic in terms of $\phi$, this integration can be performed formally, and obtains \begin{eqnarray} \ln Z_E &=& -\frac12 {\rm Tr} \ln (\partial^2) \;=\; -\frac12 \int d^2x \; \langle x | \ln\partial^2 | x \rangle. \end{eqnarray} In the momentum representation the spatial component of the momentum is discretized in the form $(2\pi n / a)$ for arbitrary integers $n$ due to the compactness of the space, \begin{eqnarray} \ln Z_E &=& -\frac12 \, \frac1{(2\pi)^2} \int d^2x \int \frac{dk}{2\pi} \, \frac1a \, \sum_n \; \ln\left[ k^2 + \left(2\pi n / a\right)^2 \, \right] \;\equiv\; -\frac1{(2\pi)^2}\int d^2x f_{{\rm bare}}, \label{Fbare.eq} \end{eqnarray} where $f_{{\rm bare}}$ is a bare Euclidean free-energy density for the massless field. Since the integration over $k$ makes $f_{{\rm bare}}$ divergent, we introduce mass $M$ of the scalar field to regularize $f_{{\rm bare}}$~\cite{Ganguly}, then the integrand is changed as \begin{eqnarray} \ln\left[k^2 \:+\: \left(2\pi n / a\right)^2\,\right] &\to& \ln\left[k^2 \:+\: \left(2\pi n / a\right)^2 \:+\: M^2\right]. \end{eqnarray} Employing the indefinite integral of $M$, we can write \begin{eqnarray} f_{{\rm bare}} &=& \frac 12\int \frac{dk}{2\pi}\frac 1a \int dM^2 \sum_n \frac 1{k^2 + \left(2\pi n / a\right)^2 + M^2}. \end{eqnarray} The sum over $n$ can be performed in the expression \begin{eqnarray} f_{{\rm bare}} &=& \frac 12 \int \frac{dk}{2\pi} \int d\omega_k \left(1 + 2\sum_{n=1}^{\infty} e^{-na\omega_k} \right), \label{Fbare2.eq} \end{eqnarray} where we have employed $k$ and $\omega_k\equiv\sqrt{k^2 + M^2}$ as independent parameters instead of using $k$ and $M$, and have used the identity \begin{eqnarray} \sum_{n=-\infty}^{\infty}\frac 1{n^2+b^2} &=& \frac{\pi}b \coth\pi b \;=\; \frac{\pi}b \: \left[1 + \frac 2{e^{2\pi b} -1}\right] \;=\; \frac{\pi}b \: \left[1 + 2\sum_{n=1}^{\infty} e^{-2\pi nb}\right]. \end{eqnarray} Since the first term of Eq. (\ref{Fbare2.eq}) indicates the contribution of infinite volume of space time and clearly diverges, we renormalize it as a cosmological term. The second term is relevant for the free-energy density, namely, renormalized free-energy density, \begin{eqnarray} f_{{\rm reno}} &\equiv& -\frac1a \int \frac{dk}{2\pi} \: \sum_{n=1}^{\infty} \: \frac 1n \: e^{-na\omega_k}. \end{eqnarray} With the identity \begin{eqnarray} e^{-na\omega_k} &=& \frac{1}{\sqrt{\pi}} \int_0^{\infty} dt\;t^{-1/2} \exp \left[ -t - \frac{\left(na\omega_k\right)^2}{4t} \right], \end{eqnarray} we perform the integration over $k$, and obtain \begin{eqnarray} f_{{\rm reno}} &=& -\frac{M}{\pi a} \: \sum_{n=1}^{\infty} \: \frac1n K_{-1} (naM). \label{Freno1.eq} \end{eqnarray} Here $K_\nu(z)$ is the modified Bessel function \begin{eqnarray} K_{\nu}(z) &=& \frac 12\left(\frac z2\right)^{\nu} \int_0^{\infty} \frac{dt}{t^{\nu+1}} \exp \left[-t-\frac{z^2}{4t}\right]. \label{Bess.eq} \end{eqnarray} The free-energy density for the massless field is obtained by taking the limit $M \to 0$. In this limit we can use the property of the Bessel function, $K_{-1}(z) \approx 1/z$ for small $z$, and the free-energy density (\ref{Freno1.eq}) becomes \begin{eqnarray} f_{{\rm reno}} &=& -\frac 1{\pi a^2}\sum_{n=1}^{\infty}\frac 1{n^2} \;=\; -\frac{\pi}{6a^2}. \label{Freno2.eq} \end{eqnarray} The Euclidean partition function is derived by substituting Eq. (\ref{Freno2.eq}) into Eq. (\ref{Fbare.eq}). After performing the spatial integration, and going back to the Minkowski space with $x^2 = i\eta$, we obtain \begin{eqnarray} \Gamma[\eta_{\mu\nu}] &=& \frac{1}{i} \ln Z \;=\; \frac 1{2\pi} \int_{-\infty}^{\infty}dt \; \frac 1{12} \frac 1{aD(t)}, \label{PartitionFuncResult.eq} \end{eqnarray} where we have used the relation $d\eta = dt/D(t)$. It should be noticed that $-1/(12aD)$ is the Casimir energy in $1+1$ dimensions, and is caused not by the existence of the boundary but by the compactness of the space. \section{CONFORMAL ANOMALY IN SPACE-TIME $R\times S^1$}\label{conf.sec} In this section the effective action for the metric $\Gamma[g_{\mu\nu}]$ is derived by evaluating the conformal anomaly in the space-time $R\times S^1$. The conformal anomaly is formally expressed by the first exponent in the right-hand side of Eq. (\ref{EffAction2.eq}). This anomaly part appears when the space-size $S^1$ is varying with time. Then the anomaly part is considered to describe the back-reactional terms of the dynamical Casimir effect. In the Euclidean space time with the metric $ds^2 = \rho(x^2) \: \left[\left(dx^1\right)^2 + \left(dx^2\right)^2\right]$ the Jacobian induced from the conformal transformation $g_{\mu\nu} \to e^{2\alpha} g_{\mu\nu}$ is \begin{eqnarray} J_E &\equiv& \exp \left[-\frac 12\int_0^adx^1 \int_{-\infty}^{\infty} dx^2 \, \alpha(x^2) \sum_{n,k} \varphi^{\dagger}_{n,k}(x)\, \varphi_{n,k}(x) \right], \label{JacobianE1.eq} \end{eqnarray} where $\{\varphi_{n,k}(x)\}$ is a complete set of the eigenfunctions of the Hamiltonian operator, \begin{eqnarray} \hat{H} &=& -\frac12\,\frac1{\sqrt{\rho}}\,\partial\partial\,\frac1{\sqrt{\rho}}, \qquad \hat{H} \varphi_{n,k}(x) \;=\; \lambda_{n,k}^2\varphi_{n,k}(x). \label{Hamiltonial.eq} \end{eqnarray} This Jacobian will be evaluated by using the eigenfunctions $\varphi_{n,k}(x)$ which satisfy the periodic boundary condition in the space $S^1$. The factor $j(x) \equiv \sum_{n,k} \varphi^{\dagger}_{n,k}(x)\varphi_{n,k}(x)$ in the Jacobian (\ref{JacobianE1.eq}) has a divergence due to the infinite degrees of freedom of the space-time points. In order to regularize this divergence we introduce a cutoff parameter $M$ and insert the cutoff function $\exp(-\lambda_{n,k}^2 / M^2)$ into $j(x)$: \begin{eqnarray} && j(x) \;\equiv\; \sum_{n,k} \varphi^{{\dagger}}_{n,k}(x) \varphi_{n,k}(x)\nonumber\\ &\;\to\;& j(x) \;\equiv\; \lim_{M\to\infty} \sum_{n,k} \varphi^{{\dagger}}_{n,k}(x) e^{-\lambda_{n,k}^2 / M^2} \varphi_{n,k}(x) = \lim_{{M\to\infty}} \sum_{n,k} \varphi^{{\dagger}}_{n,k}(x) e^{-\hat{H} / M^2} \varphi_{n,k}(x). \nonumber \end{eqnarray} When we take $\varphi_{n,k}(x)=\frac1{\sqrt{a}}e^{ikx^2}e^{i(2\pi n / a) x^1}$ as the eigenfunction, we obtain \begin{eqnarray} j(x) \;=\; \lim_{M\to\infty} \frac1a \sum_{n=-\infty}^{\infty} \int_{-\infty}^{\infty} \frac{dk}{2\pi} \exp \left[ \frac 1{2 M^2} \left( -\frac{k^2+\left(2\pi n / a\right)^2}{\rho} + 2\frac {ik}{\sqrt\rho}\partial_2\frac 1{\sqrt\rho} + \frac 1{\sqrt\rho}\partial_2^2\frac 1{\sqrt\rho} \right) \right]. \label{jFunc1.eq} \end{eqnarray} Here we should note that $j(x)$ is independent of $x^1$. Redefining $k \to Mk$, we can write Eq. (\ref{jFunc1.eq}) with a dimensionless parameter $k$ as \begin{eqnarray} j(x^2) &=& \lim_{M\to\infty} \frac{M}{a} \sum_{n=-\infty}^{\infty} \int_{-\infty}^{\infty}\frac{dk}{2\pi} \exp\left[ - \frac{{k}^2+\left(2\pi n / Ma\right)^2}{2 \rho } + \frac{i{k}}{M} \frac1{\sqrt\rho} \partial_2 \frac1{\sqrt\rho} + \frac1{2M^2} \frac1{\sqrt\rho} \partial_2^2 \frac1{\sqrt\rho} \right].\nonumber\\ &&\label{jFunc2.eq} \end{eqnarray} The second and the third terms in the exponent in Eq. (\ref{jFunc2.eq}) are understood as operators, e.g., \begin{eqnarray} \frac1{\sqrt\rho} \partial \frac1{\sqrt\rho} &=& -\frac{\partial\rho}{2\rho^2} + \frac1{\rho}\partial . \end{eqnarray} After expanding the integrand in terms of $M^{-1}$, the order $M^2$ terms in Eq. (\ref{jFunc2.eq}) under integrating over $k$ and summation over $n$, denoted as ${\cal O}(M^{2})$, diverge with the limit on $M$. Notice that $\sum_{n=-\infty}^{\infty}\exp\left[-\frac{\left(2\pi n/Ma\right)^2}{2\rho }\right]$ gives the contribution of ${\cal O}(M)$. The part of ${\cal O}(M^{2})$, however, is renormalizable by adding a bare cosmological term to the starting Lagrangian~\cite{Fujikawa,PL}. In this expansion the terms in Eq. (\ref{jFunc2.eq}) including only one operator $\frac{i{k}}{M}\frac1{\sqrt\rho} \partial_2 \frac1{\sqrt\rho}$ become ${\cal O}(M)$ because of the existence of the dumping factor, $\exp(-{k}^2/2\rho)$. The part of ${\cal O}(M)$ in Eq. (\ref{jFunc2.eq}) becomes zero for symmetric integration on the odd function. Then the next reading terms of ${\cal O}(M^{0})$ in Eq. (\ref{jFunc2.eq}) remain under the limit on $M$. The terms of ${\cal O}(M^{0})$ in Eq. (\ref{jFunc2.eq}) consist of two kinds of contributions. One comes from the operator $\frac1{2M^2} \frac1{\sqrt\rho} \partial_2^2 \frac1{\sqrt\rho}$ in Eq. (\ref{jFunc2.eq}), becoming \begin{eqnarray*} && \frac1{Ma} \sum_{n=-\infty}^{\infty} \int_{-\infty}^{\infty}\frac{dk}{2\pi} \; \sum_{m=1}^{\infty} \frac1{m!} \left[ - \frac{k^2 + \left(2\pi n / Ma \right)^2}{2 \rho} \right]^{m-1} \\ && \qquad\qquad \times \left\{ \left( \frac{m^3}{6} + \frac{m^2}{4} - \frac{m}{24} \right) \rho^{-3} (\partial_2\rho)^2 - \frac{m^2}4 \rho^{-2} \partial_2^2 \rho \right\}, \end{eqnarray*} and another comes from the two operators of $\frac{i{k}}{M} \frac1{\sqrt\rho} \partial_2 \frac1{\sqrt\rho}$, being \begin{eqnarray*} && \frac1{Ma} \sum_{n=-\infty}^{\infty} \int_{-\infty}^{\infty}\frac{dk}{2\pi} \; \sum_{m=1}^{\infty} \frac{(ik)^2}{m!} \left[ - \frac{k^2 + \left(2\pi n / Ma \right)^2}{2 \rho} \right]^{m-2} \\ && \qquad\qquad \times \left\{ \left( \frac{m^4}{8} - \frac{m^2}{4} + \frac{m}{8} \right) \rho^{-4}(\partial_2\rho)^2 - \left( \frac{m^3}{6}-\frac{m^2}{4}+\frac{m}{12} \right) \rho^{-3}\partial_2^2\rho \right\}. \end{eqnarray*} After performing the integration over $k$, $j(x^2)$ becomes \begin{eqnarray} j(x^2) &=& \lim_{M\to\infty} \left[ F(\rho) \left(\partial_2\rho\right)^2 + G(\rho) \partial_2^2\rho \right], \end{eqnarray} where $F(\rho)$ and $G(\rho)$ are given by \begin{eqnarray*} F(\rho) &\equiv& \frac{\rho^{-5/2}}{\sqrt{2\pi}Ma} \sum_{n=-\infty}^{\infty} \exp\left[ - \frac12 \left( \frac{2\pi n}{Ma\sqrt{\rho}} \right)^2 \right] \left\{ \frac5{32} - \frac5{48} \left( \frac{2\pi n}{Ma\sqrt{\rho}} \right)^2 + \frac1{96} \left( \frac{2\pi n}{Ma\sqrt{\rho}} \right)^4 \right\},\\ G(\rho) &\equiv& \frac{\rho^{-3/2}}{\sqrt{2\pi}Ma} \sum_{n=-\infty}^{\infty} \exp\left[ - \frac12 \left( \frac{2\pi n}{Ma\sqrt{\rho}} \right)^2 \right] \left\{ - \frac18 + \frac1{24} \left( \frac{2\pi n}{Ma\sqrt{\rho}} \right)^2 \right\}. \end{eqnarray*} Under the limit on $M$ we obtain \begin{eqnarray} \lim_{M\to\infty} F(\rho) &=& \frac{\rho^{-2}}{2\pi} \frac{1}{12}, \qquad \lim_{M\to\infty} G(\rho) \;=\; - \frac{\rho^{-1}}{2\pi} \frac{1}{12}, \end{eqnarray} with the help of the definition of the Jacobi $\theta$ function and its property: \begin{eqnarray*} \theta(u,\tau) &\equiv& \sum_{l=-\infty}^{+\infty} \exp\left( 2\pi iul + i\pi\tau l^2 \right), \\ &\theta(0,i\tau)& \;=\; \frac1{\sqrt\tau} \: \theta\left(0,\,\frac{i}{\tau}\right). \end{eqnarray*} In order to evaluate the effective action (\ref{EffAction2.eq}) with the Euclidean metric $ds^2 = \rho(x^2) \:\times\: \left[\left(dx^1\right)^2 + \left(dx^2\right)^2\right]$, we have to choose the parameter of the conformal transformation as $\alpha(x^2) = -\frac12\ln \rho(x^2)$. Then the Jacobian factor (\ref{JacobianE1.eq}) becomes \begin{eqnarray} \ln J_E[\rho] &=& \frac{1}{96 \pi} \int_0^a dx^1 \, \int_{-\infty}^\infty dx^2 \, \rho^{-2} (\partial_2 \rho)^2. \end{eqnarray} Now we continue back to the Minkowski Jacobian with time evolving metric (\ref{RWMetric.eq}): \begin{eqnarray} \frac{1}{i} \ln J[D] &=& - \frac{1}{2\pi} \int dt \; \frac{a}{12} \frac{{\dot D}^2}D, \label{JacobianM2.eq} \end{eqnarray} where we have used the relations between the Euclidean parameters and the Minkowski ones: $x^2=i\eta$, $\rho(x^2) = C(\eta)$, and we note that $d\eta=dt/D(t), C(\eta)=D(t)^2, \int dx=a$. On the other hand, the well-known Polyakov-Liouville action~\cite{PL}, which is the conformal anomaly in the space-time $R^2$, brings the same result as Eq. (\ref{JacobianM2.eq}), shown as follows. The Polyakov-Liouville action is given by the general form: \begin{eqnarray} S_{\rm PL} &=& - \frac{1}{96\pi} \int d^2x \: \sqrt{-g} \: \int d^2x' \: \sqrt{-g'} R(x)\Box^{-1}(x,x') \: R(x'), \end{eqnarray} where $R(x)$ is the Ricci curvature. With the form of the metric, $ds^2 = -C(\eta)\:(d\eta^2 -dx^2)$, \begin{eqnarray} S_{\rm PL} &=& - \frac{1}{96\pi} \int_{0}^{a}dx\int d\eta \: C \: \ln C \: \Box \ln C, \label{PL.eq} \end{eqnarray} and the Ricci curvature is $R(x)=-\Box\ln C$. With the relations, $d\eta=dt/D(t)$ and $C(\eta)=D(t)^2$, we come back to the Robertson-Walker-type metric $ds^2 = -dt^2+D(t)^2dx^2$, and obtain the Ricci curvature in terms of $D(t)$: \begin{eqnarray} R(x) &=& -\Box\ln C \;=\; \frac{2\ddot{D}}{D}. \label{Ricci} \end{eqnarray} Here we use $\Box = g^{\mu\nu}\partial_{\mu}\partial_{\nu} = -\frac{1}{C}\partial_{\eta}^2$, and the relation $\partial_{\eta}=D\partial_{t}$. By substituting Eq. (\ref{Ricci}) into Eq. (\ref{PL.eq}), $S_{\rm PL}$ is modified as \begin{eqnarray} S_{\rm PL} &=& \frac{1}{24\pi} \int_{0}^{a}dx \int dt \: \ddot{D}\ln D. \label{CA.eq} \end{eqnarray} This result is consistent with the well-known fact that the regulated trace of the stress tensor is proportional to the curvature. After the partial integration, Eq. (\ref{CA.eq}) is found to be the same as our result (\ref{JacobianM2.eq}), which is the case of $R\times S^1$. Finally, combining the partition function (\ref{PartitionFuncResult.eq}) and the Jacobian factor (\ref{JacobianM2.eq}) gives the effective action for the space size $D(t)$ as \begin{eqnarray} \Gamma[D] &\equiv& \Gamma[g_{\mu\nu}] \;=\; \frac{1}{i} \ln J + \frac{1}{i} \ln Z \;=\; \frac 1{2\pi} \int_{-\infty}^{\infty} dt \left( - \frac{1}{12}\frac{\dot D^2}{D} + \frac{1}{12}\frac{1}{D} \right), \label{EffActionResult1.eq} \end{eqnarray} where we have redefined $aD \to D$. \section{BACK REACTION OF THE DYNAMICAL\\ CASIMIR EFFECT}\label{back.sec} The semiclassical effective action for the motion of the boundaries is obtained as \begin{eqnarray} S_{\rm eff} &=& \frac1{2\pi} \int dt \left( \frac{m}{2} \dot{D}^2 \;-\; \frac{\kappa}{24} \frac{{\dot{D}}^2}D \;+\; \frac{\kappa}{24} \frac{1}{D} \right), \label{SemiClAction.eq} \end{eqnarray} where $\kappa$ is the number of species of scalar fields. The second and the third terms come from the effective action (\ref{EffActionResult1.eq}). In the first term we adopted the same redefinition $aD \to D$ as that in Eq. (\ref{EffActionResult1.eq}). In this action the second term is the back-reaction term of the dynamical Casimir effect, and the third term is the static Casimir energy. This action leads to the equation of motion given by \begin{eqnarray} \left( m - \frac{\kappa}{12}\frac{1}{D} \right) \ddot{D} &=& - \frac{\kappa}{24} \left( \frac{\dot{D}}{D} \right)^2 - \frac{\kappa}{24} \frac{1}{D^2}. \label{EquationOfMotion.eq} \end{eqnarray} This equation is integrable, and the resulting relation is given by \begin{eqnarray} \left( \frac m2 - \frac{\kappa}{24}\frac 1D \right) \dot{D}^2 - \frac{\kappa}{24}\frac 1D &=& E, \label{EnergyDef.eq} \end{eqnarray} where $E$ is an integral constant. The left-hand side is the Hamiltonian of this system, thus $E$ is the energy of this system. Here it should be noticed that the semiclassical condition $m \gg 1/D(t)$ and the adiabatic condition $\dot{D}(t) \ll 1$ lead to the validity condition $|E| \ll m$. Combining the equation of motion (\ref{EquationOfMotion.eq}) and the description of the energy (\ref{EnergyDef.eq}), we obtain the mutual dynamical force between the mirrors (boundaries), namely {\it the dynamical Casimir force}, \begin{eqnarray} F_{\rm dyn} \;\equiv\; m\ddot{D} &=& - \frac{\kappa}{24} \frac{1}{D^2} \; \frac{1 + \dot{D}^2} {1 - \frac{\kappa}{12}\frac{1}{mD}} \;=\; - \frac{\kappa}{24} \frac{1}{D^2} \; \frac{1 + \frac{2E}{m}} {\left(1 - \frac{\kappa}{12}\frac{1}{mD}\right)^2}. \label{DynamicalForce.eq} \end{eqnarray} The dynamical Casimir force depends on the relative velocity of the mirrors. When the reduced mass $m$ is much larger than the scales $E$ and $1/D$, or equivalently, the velocity $\dot{D}$ is regarded as zero, the dynamical Casimir force (\ref{DynamicalForce.eq}) is approximately equal to the static one: \begin{eqnarray} F_{\rm static} \;\equiv\; - \frac{\partial}{\partial D} \left( - \frac{\kappa}{24} \frac{1}{D} \right) &=& - \frac{\kappa}{24} \frac{1}{D^2}. \label{StaticForce.eq} \end{eqnarray} The ratio of the dynamical force $F_{\rm dyn}$ to the static one $F_{\rm static}$ is given by \begin{eqnarray} F_{\rm dyn}/F_{\rm static} &=& \frac{1 + \dot{D}^2} {1 - \frac{\kappa}{12}\frac{1}{mD}} \;=\; 1 \:+\: \frac{\kappa}{12}\frac{1}{mD} \:+\: \dot{D}^2 \:+\: \cdots. \label{DynamicalRatio.eq} \end{eqnarray} Here the $\dot{D}^2$ term in the expansion is known as the negative-frictional-like-force~\cite{DCE_Dodonov}. Since $\dot{D}^2 \geq 0$, we conclude that the dynamical force $F_{\rm dyn}$ is always attractive and stronger than the static one $F_{\rm static}$ for $D > \frac{\kappa}{12}\frac{1}{m}$. \section{CONCLUSION AND DISCUSSIONS}\label{summary.sec} In this paper we presented an effective theoretical approach to studying the Casimir effects in $1+1$-dimensions within the adiabatic approximation. The point of our investigation was the replacement of the spatial configuration: $D^1 \rightarrow S^1$. We constructed the effective action of the scalar field model, and checked the validity of this replacement. In our model the quantum correction to the classical kinetic term of the mirror separation was calculated by the path-integral formalism. The resultant quantum correction naturally contains both the ordinary Casimir energy term and the back-reaction term of the dynamical Casimir effect. The semiclassical effective action (\ref{SemiClAction.eq}) was constructed of the classical kinetic term of the mirror separation and these resultant quantum corrections. From the action (\ref{SemiClAction.eq}), we have obtained the dynamical vacuum pressure. The pressure ({\it dynamical Casimir force}) includes the back-reactional force of the dynamical Casimir effect. The dynamical Casimir force was confirmed to be attractive and always stronger than the static Casimir force. The dynamical Casimir force depends on the relative velocity of the mirrors, and it is reduced to the static one when the velocity goes to zero. The perturbative expansion of the resultant dynamical Casimir force (\ref{DynamicalRatio.eq}) includes the term for the negative frictional force which agrees with the result of Dodonov {\it et al.}~\cite{DCE_Dodonov}. Although this means that our result is not entirely new, our approach reproduces the reliable result, thus it can be said that we have presented a unique effective theoretical approach to the problem. Several easier derivations of the static Casimir energy in the Hamiltonian formulation are known, but our method needs a more complex calculation to obtain the Casimir energy. Our approach, however, describes both the static and the dynamical Casimir effects together, and is applicable to more realistic models in the higher dimensions by replacing the space $D^1 \times R^n$ into $S^1\times R^n$. Furthermore, the existence of the action makes it easy for us to compare our model with others. For example, our model has a correspondence to the Callan, Giddings, Harvey, Strominger (CGHS) model which describes the two-dimensional dilaton black hole~\cite{CGHS}. The back reaction discussed in this paper is comparable to the back reaction of the Hawking radiation from the CGHS black hole~\cite{RST}. In the CGHS model the Hawking radiation is represented by the conformal anomaly in the energy-momentum tensor~\cite{CGHS}, and the back reaction of the radiation, which is described by the Polyakov-Liouville action, appears as the decrease in the black-hole mass~\cite{RST}. Our classical kinetic term in the semi-classical effective action (\ref{SemiClAction.eq}) corresponds to the kinetic term of the dilaton in the CGHS model. Some comments are in order. The quantum correction (\ref{EffActionResult1.eq}) does not include the third derivative of the dynamical variable. This looks different from the results evaluated by Fulling and Davies~\cite{DCE_FD}. They calculated the energy-momentum tensor in (1+1)-dimensional system of two relatively moving mirrors~\cite{DCE_FD} as well as that in (1+1)-dimensional system of a single non-uniformly accelerating mirror~\cite{DCE_FD,B&D}. Both energy-momentum tensors include the third derivative of the dynamical variables. Our result for the system of two mirrors does not need to coincide with their result for the system of a single mirror since the forms of the conformal anomaly for two systems are different. The result for the system of a single mirror is due to the Unruh-like effect rather than due to the dynamical Casimir effect. On the other hand, the energy-momentum tensor derived from Eq. (\ref{EffActionResult1.eq}) coincides with their result for the system of two mirrors under a certain transformation of the dynamical variable. In the semiclassical effective action (\ref{SemiClAction.eq}), the contribution from the dynamical Casimir effect generated a negative-definite kinetic term of the mirror separation. Such a kinetic term also appeared in the analysis of the CGHS model~\cite{RST}. The following point should be noted: there is a positive-definite classical kinetic term, and the negative-definite term gives only a slight correction. This holds in the case where the mass scale of the mirrors $m$ is much greater than the scale of the Casimir energy $\sim D^{-1}$. On the other hand, if the mirror separation $D(t)$ is smaller than the inverse of the mirror mass $m^{-1}$, our result (\ref{DynamicalForce.eq}) shows that the dynamical Casimir force $F_{\rm dyn}$ becomes repulsive. However, our semiclassical treatment becomes unsuitable at that time. When the motion of the mirror separation obeys the quantum mechanics, this repulsive force might be realized. We will leave this problem to subsequent developments. \begin{flushleft} {\Large\bf ACKNOWLEDGMENTS} \end{flushleft} We would like to thank Professor S. Uehara, Professor M. Harada, and Professor A. Nakayama for their useful discussions and suggestions. We thank Professor V. A. Miransky for discussions. We also appreciate helpful comments of Professor A. Sugamoto, Professor H. Funahashi, Dr. T. Itoh, Dr. A. Takamura, Dr. S. Sugimoto, Dr. Y. Ishimoto, and Dr. S. Yamada. We are also grateful to Professor R. Sch\"utzhold for his interest in our work and enlightening discussions on related matters. He suggested that no particle creation occurs in our model, and that this work is related to the dynamical back reaction of the static Casimir effect. One of us (Y.N.) is indebted to the Japan Society for the Promotion of Science (JSPS) for its financial support. The work is supported in part by a Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture (No. 03665). \section*{APPENDIX A: Fujikawa method} \setcounter{equation}{0} \renewcommand{\theequation}{A\arabic{equation}} In this appendix we briefly explain the derivation of the expression (\ref{EffAction2.eq}) from the definition of the effective action (\ref{EffAction1.eq}). This derivation is based on the evaluation of the conformal anomaly by the Fujikawa method~\cite{Fujikawa}. In order to perform the path integration of Eq. (\ref{EffAction1.eq}), we make a Wick rotation by introducing an imaginary time variable $x^2 \equiv i\eta$. Then the Euclidean metric corresponding to the Minkowski one (\ref{RWMetric.eq}) becomes \begin{eqnarray} ds^2 &=& \rho(x^2) \: [(dx^1)^2 + (dx^2)^2]. \label{rho.eq} \end{eqnarray} The Euclidean effective action is \begin{eqnarray} e^{-\Gamma_E [g_{\mu\nu}]} &=& \int{\cal D}\phi\:\exp \left[ - \frac1{2\pi} \int d^2x \: \sqrt{g} \: \frac12 \: g^{\mu\nu} \: \partial_{\mu}\phi \: \partial_{\nu}\phi \right]. \label{Ea.eq} \end{eqnarray} By introducing $\tilde\phi\equiv\sqrt[4]{g}\:\phi$ and changing the measure ${\cal D}\phi$ into the invariant form under the general coordinate transformation ${\cal D}\tilde\phi$, Eq. (\ref{Ea.eq}) becomes \begin{eqnarray} e^{-\Gamma_E [g_{\mu\nu}]} &=& \int{\cal D}{\tilde\phi} \: \exp \left[ -\frac1{2\pi} \int d^2x \: \frac12 \partial\left(\frac{\tilde\phi}{\sqrt\rho}\right) \partial\left(\frac{\tilde\phi}{\sqrt\rho}\right) \right]. \label{Etea.eq} \end{eqnarray} Here we have used a notation $\partial\phi\,\partial\phi \equiv \partial_1\phi\,\partial_1\phi + \partial_2\phi\,\partial_2\phi$. We perform a mode expansion of the field $\tilde\phi(x)$ in terms of a complete set $\{\varphi_n(x)\}$: \begin{eqnarray} \tilde\phi(x) &=& \sum_n \: a_n \varphi_n(x) \;\equiv\; \sum_n \: \langle x | n \rangle \: a_n, \end{eqnarray} where we have chosen ${\varphi_n(x)}$ as an eigenfunction of the Hamiltonian operator, \begin{eqnarray} \hat{H} &=& -\frac12\,\frac1{\sqrt{\rho}}\,\partial\partial\,\frac1{\sqrt{\rho}}, \qquad \hat{H} \varphi_n(x) \;=\; \lambda_n^2 \, \varphi_n(x). \label{Ham.eq} \end{eqnarray} Here $\varphi_n(x)$ satisfies the normalization $\int d^2x\:\varphi^{{\dagger}}_m(x)\:\varphi_n(x) \:=\: \delta_{mn}$. Now we note that the measure ${\cal D}\tilde\phi$ is expressed by the mode coefficients $a_n$ as \begin{eqnarray} {\cal D}\tilde\phi &=& \prod_x \: {\cal D}\tilde\phi(x) \;=\; \left[\,\det\langle x|n\rangle\,\right] \: \prod_n \: da_n \;=\; \prod_n \: da_n. \end{eqnarray} Under the Weyl transformation $g_{\mu\nu}\to e^{2\alpha(x)}g_{\mu\nu}$ the mode coefficients of the field $\tilde\phi(x)$, $a_n$, are transformed as an infinitesimal form, \begin{eqnarray} \tilde\phi(x) &\to& \tilde\phi^{\prime}(x) \;\equiv\; \sum_n \, a^{\prime}_n \, \varphi_n(x), \nonumber\\ a^{\prime}_n &=& a_n \:+\: \sum_m \frac12 \int d^2x \: \alpha(x) \varphi^{{\dagger}}_n(x) \: \varphi_m(x)a_m \;\equiv\; \sum_m C_{nm} \, a_m. \end{eqnarray} Then the measure is transformed as \begin{eqnarray} {\cal D}\tilde\phi^{\prime} &=& \prod_n da^{\prime}_n \;=\; \left[\det (C_{nm})\right] \: \prod_l da_l \nonumber\\ &=& \exp \left[ {\rm Tr}\ln \left( \delta_{nm} \;+\; \frac12\int d^2x\:\alpha(x) \varphi^{{\dagger}}_n(x) \: \varphi_m(x) \right) \right] \; \prod_l da_l \nonumber\\ &=& \exp \left[+\frac12\int d^2x\:\alpha(x) \sum_n\varphi^{{\dagger}}_n(x)\:\varphi_n(x) \right] \; {\cal D}\tilde{\phi}. \end{eqnarray} This gives the Jacobian of the conformal transformation. By the Weyl transformation chosen $\alpha(x)=-\frac12\ln\rho(x)$ for $\tilde{\phi} \to \tilde{\phi}^{\prime} = \tilde{\phi} / \sqrt{\rho}$, the effective action (\ref{Etea.eq}) becomes \begin{eqnarray} e^{-\Gamma_E [g_{\mu\nu}]} &=& \exp\left[- \frac12 \int d^2x \, \alpha(x) \sum_n \varphi^{{\dagger}}_n(x) \, \varphi_n(x) \right] \nonumber\\ & & \times \int{\cal D}{\tilde\phi^\prime} \, \exp\left[- \frac1{2\pi} \int d^2x \, \frac12 \, \partial\tilde\phi^\prime \, \partial\tilde\phi^\prime \right], \label{EffActionE3.eq} \end{eqnarray} where the second factor equals to the partition function of the free scalar field in the flat space time. Finally, we can arrive at our destination (\ref{EffAction2.eq}) from the description (\ref{EffActionE3.eq}) by the inverse Wick rotation. \section*{APPENDIX B: Another path-integral Calculation of the Casimir energy} \setcounter{equation}{0} \renewcommand{\theequation}{B\arabic{equation}} In this appendix we give another partition-functional derivation of the Casimir energy by means of the point-splitting ansatz and the Feynman prescription: In the path-integral method the partition function part in (\ref{EffAction2.eq}) can be also evaluated by using the point-splitting ansatz and the Feynman's renormalization prescription. Employing the ansatz of point splitting to (\ref{partition.eq}), \begin{eqnarray} Z_E &=& \int{\cal D}{\phi}\, \exp \left[ -\frac{1}{2\pi}\int d^2x \frac{1}{2\pi}\int d^2x^{\prime} \frac{1}{2} \phi(x)A(x,x^{\prime})\phi(x^{\prime}) \right] \nonumber\\ &=& \exp \left[-\frac1{2}{\rm Tr}\ln A \right], \end{eqnarray} where $\int d^2x \equiv \int_{-\infty}^{\infty} dx^2 \int_0^a dx^1$, and $ A(x,x^{\prime}) \equiv \delta^{\mu\nu}\partial_{\mu}\partial^{\prime}_{\nu} \delta^{(2)}(x-x^{\prime})$. The two-dimensional Dirac delta function in the integral representation is \begin{eqnarray} \delta^{(2)}(x-x^{\prime}) &=& \frac1{a} \,\sum_{n=-\infty}^{+\infty} \int \frac{dk}{2\pi} \, e^{-ik(x^2-x^{\prime 2})} \, e^{i\frac{2\pi n}{a}(x^1-x^{\prime 1})} \; . \end{eqnarray} Now we come back to Minkowski space and introduce mass $M$ of the scalar field to regularize the integral, \begin{eqnarray} \frac{1}{i} \ln Z &=& - \frac{1}{2} \, \frac{1}{(2\pi)^2}\int d^2x \int d^2x^{\prime} \, \delta^{(2)}(x-x^{\prime}) \, \frac1a \, \sum_n \, e^{i\frac{2\pi n}{a}(x-x^{\prime})} \nonumber\\ && \times \; \int dM^2 \int \frac{dk}{2\pi} \, \frac{e^{-ik(\eta-\eta^{\prime})}} {-k^2 + \left(2\pi n / a\right)^2 + M^2} \;. \end{eqnarray} With the $i\epsilon$ prescription, we perform the integral in the complex $k$ plane, applying the residue theorem, \begin{eqnarray} && \int \frac{dk}{2\pi} \, \frac{e^{-ik(\eta-\eta^{\prime})}} {-k^2 + \left(2\pi n / a\right)^2 + M^2 - i\epsilon} \nonumber\\ && = \theta(\eta - \eta^{\prime}) \frac{i}{2\sqrt{\left(2\pi n / a\right)^2 + M^2}} + \theta(\eta^{\prime} - \eta) \frac{-i}{-2\sqrt{\left(2\pi n / a\right)^2 + M^2}}, \end{eqnarray} where $\epsilon>0$ and $\theta(x)$ is the step function. Here we define the new parameter, $\omega_n^2 \equiv \left(2\pi n / a\right)^2 + M^2$, and replace the integral into the following form: \begin{eqnarray} \int dM^2 \, \frac 1{2\sqrt{\left(2\pi n / a\right)^2 + M^2}} &=& \int d\omega_n \;=\; \omega_n. \end{eqnarray} Then we take the massless limit $M\to 0$ and perform the summation, \begin{eqnarray} \sum_{n=-\infty}^{+\infty} \omega_n &=& \frac{4\pi}{a}\sum_{n=1}^{\infty} n \;=\; \frac{4\pi}{a}\zeta(-1) \;=\; -\frac{\pi}{3a}. \end{eqnarray} At last we arrive at the same form of Eq. (\ref{PartitionFuncResult.eq}), \begin{eqnarray} \frac1{i}\ln Z \;=\; \frac 1{2\pi} \int_{-\infty}^{\infty}dt \; \frac 1{12} \frac 1{aD(t)}. \end{eqnarray} \newcommand{\PRL}[3] {{Phys. Rev. Lett.} {\bf #1}, #2 (#3)} \newcommand{\PR}[3] {{Phys. Rev.} {\bf #1}, #2 (#3)} \newcommand{\PRA}[3] {{Phys. Rev. A} {\bf #1}, #2 (#3)} \newcommand{\PRD}[3] {{Phys. Rev. D} {\bf #1}, #2 (#3)} \newcommand{\PL}[3] {{Phys. Lett.} {\bf #1}, #2 (#3)} \newcommand{\PLA}[3] {{Phys. Lett. A} {\bf #1}, #2 (#3)} \newcommand{\PLB}[3] {{Phys. Lett. B} {\bf #1}, #2 (#3)} \newcommand{\NuP}[3] {{Nucl. Phys.} {\bf #1}, #2 (#3)} \newcommand{\PTP}[3] {{Prog. Theor. Phys.} {\bf #1}, #2 (#3)} \newcommand{\Nature}[3] {{Nature (London)} {\bf #1}, #2 (#3)} \newcommand{\PKNAW}[3] {{Proc. K. Ned. Akad. Wet.} {\bf #1}, #2 (#3)} \newcommand{\Physica}[3]{{Physica (Utrecht)} {\bf #1}, #2 (#3)} \newcommand{\JMP}[3] {{J. Math. Phys.} {\bf #1}, #2 (#3)} \newcommand{\PRSLA}[3] {{Proc. R. Soc. London, Ser A} {\bf #1}, #2 (#3)} \newcommand{\AP}[3] {{Ann. Phys. (N.Y.)} {\bf #1}, #2 (#3)} \newcommand{\JPA}[3] {{J. Phys. A} {\bf #1}, #2 (#3)} \newcommand{\ZhETF}[3] {{Zh. \'{E}ksp. Teor. Fiz. Pis'ma. Red.} {\bf #1}, #2 (#3)} \newcommand{\JETP}[3] {{JETP Lett.} {\bf #1}, #2 (#3)} \newcommand{\CMP}[3] {{Commun. Math. Phys.} {\bf #1}, #2 (#3)}
\section{The $(e,\nu)$ Correlation as a Probe of Physics Beyond the Standard Model} \label{sec: intro} In the Standard Model, nuclear $\beta$ decay is mediated by the exchange of W bosons which have only vector and axial-vector couplings. However, extensions of the standard model, such as super-symmetric theories with more than one charged Higgs doublet, or leptoquarks, naturally predict scalar or tensor weak couplings~\cite{he:95}. A general effective Hamiltonian for allowed $\beta$ transitions that respects Lorentz invariance is~\cite{ja:57} \begin{eqnarray} H = (\bar{\psi_p} \gamma_\mu \psi_n) (C_V \bar{\psi_e} \gamma_\mu \psi_\nu + C_V' \bar{\psi_e} \gamma_\mu \gamma_5 \psi_\nu) \nonumber\\ +(\bar{\psi_p} \gamma_\mu \gamma_5 \psi_n) (C_A \bar{\psi_e} \gamma_\mu \gamma_5 \psi_\nu + C_A' \bar{\psi_e} \gamma_\mu \psi_\nu) \nonumber\\ +(\bar{\psi_p} \psi_n) (C_S \bar{\psi_e} \psi_\nu + C_S' \bar{\psi_e} \gamma_5 \psi_\nu) \nonumber\\ +\frac{1}{2}(\bar{\psi_p} \sigma_{\lambda\mu} \psi_n) (C_T \bar{\psi_e} \sigma_{\lambda\mu} \psi_\nu + C_T' \bar{\psi_e} \sigma_{\lambda\mu} \gamma_5 \psi_\nu) \nonumber\\ ~ + {\rm Hermitian~conj.} \label{hamilt_eq} \end{eqnarray} where a term proportional to $(\bar{\psi_p} \gamma_5 \psi_n)$ has been neglected because nucleons are non-relativistic. In the Standard Model, $C_V = C_V'$, $C_A = C_A'$, and $C_S = C_S' =C_T = C_T'= 0$. Jackson {\em et al.}~\cite{ja:57} computed the nuclear-$\beta$-decay rate from this Hamiltonian; for an unoriented initial state and summation over the lepton helicities \begin{eqnarray} dW = dW_0 (1 + a {{\bf p_e} \cdot {\bf p_\nu} \over E_e E_\nu} + b {m_e \over E_e} )~, \label{eq: rate} \end{eqnarray} where \begin{eqnarray} a \xi = |M_F|^2 (|C_V|^2+|C_V'|^2-|C_S|^2-|C_S'|^2) \nonumber\\ -\frac{1}{3} |M_{GT}|^2 (|C_A|^2+|C_A'|^2-|C_T|^2-|C_T'|^2) \label{eq: a}\\ b \xi =\pm2 \gamma {\rm Re}[|M_F|^2(C_V^*C_S+C_V'^*C_S') \nonumber\\ +|M_{GT}|^2(C_A^*C_T+C_A'^*C_T')] \label{eq: b}\\ {\rm and} \nonumber \\ \xi = |M_F|^2 (|C_V|^2+|C_V'|^2+|C_S|^2+|C_S'|^2) \nonumber\\ +|M_{GT}|^2 (|C_A|^2+|C_A'|^2+|C_T|^2+|C_T'|^2)~, \end{eqnarray} where $M_F$ and $M_{GT}$ are rank-0 and rank-1 nuclear matrix elements. We have simplified the expressions by neglecting the small Coulomb effects, and adopting the allowed approximation (ignoring curvature in the lepton wave functions and the recoil-order corrections). In what follows we shall assume $C_V=C_V^\prime$ and express the scalar couplings in terms of: \begin{eqnarray} \tilde{C_S}= C_S/C_V, ~~{\rm and}~~\tilde{C_S^\prime}= C_S^\prime/C_V. \end{eqnarray} Precise measurements of the $e$-$\nu$ correlation coefficient, $a$, and of the Fierz-interference term, $b$, can potentially yield information about physics beyond the standard model. Equations 2-5 are complicated and depend on nuclear physics information through $M_F$ and $M_{GT}$. However, for pure Fermi or pure GT transitions the expressions simplify and the $M_F$ and $M_{GT}$ factors cancel. In pure Fermi transitions, a positron-neutrino correlation coefficient $a < 1$ would immediately imply the presence of scalar currents; while in GT decays a value $a > -1/3$ would imply tensor currents. To measure the $e$-$\nu$ correlation one must determine the daughter's velocity, because it is out of the question to detect the low-energy neutrino. But, in general, measuring the daughter's recoil velocity is not a trivial task: the velocities are very small and the interaction of the daughter with the surrounding medium can jeopardize the measurement. Fortunately nature has provided us with cases where the daughter states are unbound to proton emission, so that the daughter's velocity can be determined via the `Doppler' broadening of the beta-delayed proton groups. Protons are preferred over other $\beta$-delayed radiations because they are emitted before the daughter nucleus has slowed appreciably and can be detected with high resolution and good efficiency. Neutrons, on the other hand, are hard to detect with high resolution and high efficiency simultaneously. Gamma rays are generally emitted too slowly to probe the daughter velocity before it has been reduced by interaction with the surrounding medium. This paper discusses superallowed transitions from proton-rich nuclei with proton-unbound daughter states. These transitions are strong, which enhances the signal-to-noise ratio, and the widths of daughter states are ideally suited for probing the $e$-$\nu$ correlation: \begin{itemize} \item the time scale for proton decay is short enough so that the decays occur before the daughter looses any appreciable velocity due to interactions with the medium (in the superallowed decay of $^{32}$Ar, the daughter travels $\leq 2 \times 10^{-2}$ \AA~before the proton is emitted). \item because proton decays in these cases violate isospin symmetry the time scale for proton decay is long enough so the widths are much narrower than the lepton-recoil broadening (in the superallowed decay of $^{32}$Ar, the daughter state has a width of $\approx 20$ eV while the lepton-recoil broadening has a full-width at half-maximum of $\approx 25$ keV). \end{itemize} \section{Limits on Scalar Weak Interactions from $^{32}$Ar $\beta^+$ Decay} \label{sec: ar32} We performed an experiment at ISOLDE that measured with high precision the energies of protons following the $0^+ \rightarrow 0^+$ $\beta^+$ decay of $^{32}$Ar. Except for this experiment and a previous ISOLDE study of $^{32}$Ar by Schardt and Riisager~\cite{sc:93}, there are no other precise determinations of $e$-$\nu$ correlations in pure Fermi transitions, which explains why, prior to our new result, limits on scalar couplings were rather poor in comparison to those on tensor currents~\cite{bo:84}. We here describe the salient features of our experiment and the extracted limits on scalar currents. More details on the $^{32}$Ar experiment can be found in Ref.~\cite{ad:99}. Critical challenges for this kind of experiment are: \begin{enumerate} \item obtaining an intense and pure source of radioactivity, \item eliminating proton-$\beta^+$ summing, which distorts the shape of the proton peak, \item and optimizing the energy resolution of the proton counter. \end{enumerate} ISOLDE solved problem 1 by producing very pure beams of both $^{32}$Ar and $^{33}$Ar from a CaO target and a plasma ion source, providing an average of $\approx 94$ $^{32}$Ar's/s and $\approx 3900$ $^{33}$Ar's/s on our catcher foil over the 9-day-long run. We solved problem 2 by immersing our detection system in a $3.5$ T magnetic field. The $^{32}$Ar beam from ISOLDE was stopped in a $\approx$23 $\mu$g/cm$^2$ C foil at 45 degrees to the beam. Our proton detectors were located at $\pm 90$ degrees with respect to the beam and at about 1.6 cm from the beam spot. In the 3.5 T field superallowed $\beta$'s had a maximum radius of $\approx 0.55$ cm, while protons had $\approx 7.14$ cm. We solved problem 3 by using cooled PIN-diode proton detectors ($\approx 0.9 \times 0.9$ cm) and temperature-controlled electronics to obtain a proton energy resolution of $\approx 4.5$ keV ($\approx 3.0$ keV electronic noise). This greatly enhanced our sensitivity to the $e$-$\nu$ correlation. \begin{figure}[ht] \hspace{-1.2cm} \hbox{\psfig{figure=fig1.ps,height=22cm}} \vspace{-14cm} \caption{Intrinsic shapes of the delayed proton group from $^{32}$Ar $0^+ \rightarrow 0^+$ decay for $a=+1,~b=0$ (unshaded curve) and $a=-1,~b=0$ (light-shaded curve).} \label{fig: MC} \end{figure} Fig.~\ref{fig: MC} shows Monte-Carlo predictions for the shape of the proton peak assuming the proton detector had infinitely good energy resolution. If one assumes $b=0$ in Eq.~\ref{eq: rate} one can extract $a$ by producing linear combinations of these two shapes, folding them with the detector response function (assumed to be two exponentials with adjustable tails and areas, convoluted with a Gaussian of adjustable width) and determining the values of the parameters that minimize $\chi ^2$. Fig.~\ref{fig: fit_32ar} shows our best fit to a subset of our data containing approximately 1/10 of the statistics. The very good energy resolution obtained in this study also allowed us to extract valuable spectroscopic information on $^{32}$Cl. The fit was performed using an R-matrix parametrization of the resonances. Reduced total widths, $\Gamma_p$, and ratios $\Gamma_{p1}/\Gamma_{p0}$ can be extracted for many of the resonances. \begin{figure}[ht] \hspace{-1.2cm} \hbox{\vspace{-10cm} \psfig{figure=fig2.ps,height=28cm,width=16cm}} \vspace{-16cm} \caption{R-matrix fit to the $^{32}$Ar delayed proton spectrum. This spectrum, contains roughly 1/10 of our data.} \label{fig: fit_32ar} \end{figure} For the general case when $b \ne 0$ one has to fold an additional distribution (taking into account the $m/E$ term in Eq.~\ref{eq: rate}). We produced a grid in the $\tilde{C_S},\tilde{C_S^\prime}$ space and, for each point, minimized $\chi^2$ with respect to the response function parameters. The resulting confidence regions are shown in Fig.~\ref{fig: cs_lim}. We found that our $\tilde{C}_S$, $\tilde{C}_S^{\prime}$ constraints are well reproduced by the single parameter \begin{eqnarray} \tilde{a} &\equiv& a/(1 + 0.1913 b)~ \nonumber \\ \label{eq: result} \end{eqnarray} where $a$ and $b$ are given in Eqs.~\ref{eq: a}, \ref{eq: b}. In other words, replacing $m/E$ by an appropriate average reproduces the regions of interest and allows us to quote our results in terms of $\tilde{a}$. Our experiment yields the constraint \begin{eqnarray} \tilde{a} = 0.9989 \pm 0.0052({\rm stat.}) \pm 0.0036({\rm syst.})~ {\rm 68\%~c.l.} \end{eqnarray} \begin{figure}[ht] \hspace{-0.5cm} \hbox{\psfig{figure=fig3.ps,width=16cm}} \vspace{-14cm} \caption{95\% conf. limits on $\tilde{C}_S$ and $\tilde{C}_S^{\prime}$, including statistical and systematic errors. Left panel: time-reversal-even couplings. The annulus is from this work, the diagonal band is the Fierz interference result of Ref.~\protect\cite{or:89}. Right panel: time-reversal-odd couplings. The circles are from this work and correspond to phases of $\tilde{C}_S$ and $\tilde{C}_S^{\prime}$ of $\pm 90^{\circ}$, $+45^{\circ}$ and $-45^{\circ}$. The shaded oval is the constraint with no assumptions about this phase. The diagonal band is from the $R$-coefficient in $^{19}$Ne decay \protect ~\cite{sc:83}. } \label{fig: cs_lim} \end{figure} The systematic error was evaluated by redoing the whole analysis under different conditions. We found $\partial \tilde{a}/\partial \Delta = -1.2\times 10^{-3}$~keV$^{-1}$ where $\Delta$ is the $\beta$-decay endpoint; and $\partial \tilde{a}/\partial Q_p=-0.9 \times 10^{-3}$~keV$^{-1}$, where $Q_p$ is the energy of the emitted proton. We measured $Q_p$ with an uncertainty, $\delta Q_p= \pm 1.2$~keV, by alternating between $\sim 2$ h $^{32}$Ar runs with 10-15 min $^{33}$Ar runs that gave us a continuous calibration of the energy scale. The mass of $^{32}$Ar has been determined only to within 50 keV~\cite{au:95}, which would impose a systematic error of $\approx 6$\% on our measurement. Fortunately, as shown in Table~\ref{tab: IMME32}, the masses of all other members of the $T=2$ isospin multiplet are known with high precision. \begin{table} \begin{flushleft} \caption{Comparison of the measured mass excesses of the lowest $T=2$ quintet in $A=32$ to predictions of the Isospin-Multiplet Mass Equation [$P(\chi^2, \nu)=0.73$].} \label{tab: IMME32} \begin{tabular}{lrrr} \hline \hline isobar & $T_3$ & $M_{\rm exp}$~(keV)$^a$ & $M_{{\rm IMME}}$ (keV) \\ \hline $^{32}$Si & $+2$ & $-24080.9 \pm 2.2$ & $-24081.9 \pm 1.4$ \\ $^{32}$P & $+1$ & $-19232.88 \pm 0.20$$^b$ & $-19232.9 \pm 0.2$ \\ $^{32}$S & $0$ & $-13970.98 \pm 0.41$$^c$ & $-13971.1 \pm 0.4$ \\ $^{32}$Cl & $-1$ & $-8296.9\pm 1.2$$^d$ & $-8296.6 \pm 1.1$ \\ $^{32}$Ar & $-2$ & $-2180 \pm 50$ & $-2209.3 \pm 3.2 $ \\ \hline \end{tabular} \\ $^a$unless noted otherwise, ground state masses are from Ref.~\protect\cite{au:95}.\\ $^b$$E_x=5072.44 \pm 0.06$ keV from Ref.~\protect\cite{en:90}.\\ $^c$$E_x=12045.0 \pm 0.4$ keV from Ref.~\protect\cite{an:85,wa:98}.\\ $^d$from delayed proton energy measured here and masses of Ref.~\protect\cite{au:95}.\\ \end{flushleft} \end{table} We use the Isospin-Multiplet Mass Equation~\cite{an:85}, $M(T_3)=a +b T_3 +c T_3^2$, to obtain \begin{equation} \Delta = 6087.3 \pm 2.2~{\rm keV}. \end{equation} \begin{figure}[ht] \hspace{-0.5cm} \hbox{\vspace{-6cm} \psfig{figure=fig4.ps,width=14cm}} \vspace{-5cm} \caption{Plot of residuals of IMME fit to the $T=2$ quintuplet in $A=32$. Top panel: mass excesses in MeV. The points show the measured values and the continuous line shows the IMME fit. Bottom panel: fit residuals in keV.} \label{fig: IMME32} \end{figure} As shown in Fig.~\ref{fig: IMME32} we obtain an excellent fit to the IMME with $P(\chi^2,\nu)=0.71$. On the other hand, modifying the IMME by adding a $dT_3^3$ term we obtain $\Delta = 6086.7 \pm 4.9$ keV and $d =0.25\pm0.47$ keV, but with a lower probability, $P(\chi^2,\nu)=0.52$, indicating that there is no empirical basis for adding a term to the IMME. We expect the IMME to work very well for the $A=32$ multiplet where the states are relatively well-bound and the Coulomb barriers relatively high. Small non-zero $d$ terms (never more significant than $3 \sigma$) have been observed in light nuclei where the $T_3=T$ members are much closer to being unbound (or even unbound) and the Coulomb barriers much lower. We here assume $\delta \Delta=\pm 2.2$~keV which combined with the uncertainty in $Q_p$ yields a {\em kinematic} systematic error $\delta \tilde{a} = \pm 0.0032$. Note that even when $d$ is allowed to vary freely the uncertainty is only about twice this value. We also checked the dependence of $\tilde{a}$ on the {\em fitting regions} of the proton spectra; a 28\% variation in the width of the region changed $\tilde{a}$ by less than $\pm 0.00055$. We examined the dependence of our results on the form of the detector response function by re-analysing the data with a single-tail response function; by re-analyzing the data assuming that a weak Gamow-Teller peak lay under the tail of the $^{32}$Ar superallowed peak; and by simultaneously fitting the $^{33}$Ar and $^{32}$Ar superallowed peaks using a common response function. From these tests we inferred a {\em line-shape} systematic error of $\delta\tilde{a} = \pm 0.0016$. \begin{figure}[ht] \hspace{-0.5cm} \vspace{2cm} \hbox{\psfig{figure=fig5.ps,width=16cm}} \vspace{-10cm} \caption{Comparison of our constraints on scalar couplings with previous work. The light shaded area represents constraints from neutron $\beta$ decay alone: {\em i.e.} from measurements of $a$, $A$, $B$, and $t_{1/2}$~\cite{pdg:98}. A slightly darker area shows how these constraints improve when combined with measurements of the polarization of $\beta$'s from $^{14}$O and $^{10}$C~\cite{ca:91}. The darker shaded area shows the result of adding to the previous the constraints on Fierz terms from $^{22}$Na~\cite{we:68} and the measurement of $a$ in $^6$He~\cite{jo:63}. The darkest shaded area shows constraints from our results. The narrow area looking like a line at $-45 ^\circ$ is from constraints on Fierz terms from $0^+ \rightarrow 0^+$ transitions.} \label{fig: comparison} \end{figure} Figure~\ref{fig: comparison} compares our results to previous constraints on scalar couplings. For scalar interactions with $\tilde{C}_S=-\tilde{C}_S^{\prime}$ so that $b=0$, our data yield the $1\sigma$ constraint $|\tilde{C}_S|^2 \leq 3.6 \times 10^{-3}$. The corresponding lower limit on the mass of scalar particles with gauge coupling strength is $M_S = |\tilde{C}_S|^{-1/2} M_W \geq 4.1 M_W$. We note that data from neutron $\beta$ decay by itself does not place stringent constraints on scalar couplings because the measurement on the correlation coefficient is not very accurate ($\delta a/a \approx 5$\%) and neutron decay is sensitive to both scalar and tensor couplings. Moreover, because of the larger value of $<m/E>$ ({\em i.e.} the lower endpoint energy), the circle generated by the equivalent of Eq.~\ref{eq: result} has a large radius which weakens the neutron constraints. Even when supplemented by other data on the GT Fierz interference {\em etc.}, the neutron constraints are not as tight as those from the present work and limits on Fierz interference in $0^+ \rightarrow 0^+$ transitions. \section{Isospin mixing Corrections to $0^+ \rightarrow 0^+$ ${\cal F}t$ values and $|V_{ud}|$} Taken at face value, the $V_{ud}$ matrix element extracted from the ${\cal F}t$ values of nine $0^+ \rightarrow 0^+$ $\beta$-decay transitions implies non-unitarity of the Kobayashi-Maskawa matrix~\cite{ha:98}. Because of the importance and unexpected nature of this conclusion, it is worth reexamining whether any systematic effect could affect the ${\cal F}t$ values. One possibility concerns the corrections for isospin-symmetry violation in the parent and daughter nuclear wave functions. These corrections are usually separated into `configuration mixing' and `nucleon overlap' parts; the latter being dominant ($\approx $ four times the former). Although several authors ~\cite{to:77,or:95,sa:96} have performed independent calculations that agree reasonably well it would be valuable to check these calculations on additional transitions in neighboring nuclei. The cases we present below provide a good opportunity to check the calculated corrections; the `nucleon overlap' corrections are enhanced over those in the nine standard cases because the nuclei lie farther from the valley of stability. \subsection{Isospin Mixing in the Fermi decay of $^{32}$Ar} The isospin-mixing correction in $^{32}$Ar was calculated by B.A. Brown~\cite{br:98} using the SKX Skyrme interaction~\cite{br:98c}. This calculation yields $2.0 \pm 0.4$\% where the uncertainty is based on previous comparisons of similar calculations to measurements. The large size of the correction is due to the looser binding of the $d \frac{3}{2}$ and $s \frac{1}{2}$ proton states compared to the neutron states. For comparison, the average correction for the nine standard cases is $\approx 0.41$\%, while the correction for the neighboring decay of $^{34}$Cl is $\approx 0.61$\%. In $^{32}$Ar decay there is no mixing with GT transitions so that one can check the isospin-mixing correction simply by determining the half-life of the parent, and the branching ratio and endpoint of the superallowed transition. One can thus extract the ${\cal F} t$ value and compare with the prediction. Furthermore, if some Fermi strength is diverted into narrow $J^{\pi}=0^+;~T=1$ levels in the $^{32}$Cl daughter, it may be possible to identify these transitions because they will have $a = +1$ instead of $a=-1/3$. We expect to determine the $^{32}$Ar half life to $\approx 0.2$\% from our ISOLDE experiment (our preliminary value is $t_{1/2}=100.74 \pm 0.18$ ms) and we hope to determine the branching ratio with high precision at an upcoming experiment at MSU~\cite{ko:99}. The extraction of the $^{32}$Ar mass from the IMME presented in section~\ref{sec: ar32} yields the endpoint to $\approx 2.2$ keV, which allows to calculate the phase space factor to $\approx 0.3$\%. The $^{32}$Ar mass could also be measured at ISOLTRAP~\cite{bo:98} but due to the short $^{32}$Ar half-life it is not clear that one could get enough intensity to determine the endpoint to within a few keV. All of the above indicates that one could extract the ${\cal F} t$ value to $\approx 0.41$\%. \subsection{Isospin Mixing in the Fermi decay of $^{33}$Ar} The theoretical isospin-mixing calculations can also be tested by studying the superallowed decays of the $A=4n+1$ nuclei. These are mixed Fermi/GT transitions so that one needs to determine the $B({\rm F})/B({\rm GT})$ ratios as well as the half-lives, branching ratios and energy releases of the superallowed transitions. The $B({\rm F})/B({\rm GT})$ ratio can be obtained from the positron-neutrino correlation. As pointed out in Section~\ref{sec: intro} the $e$-$\nu$ correlation from mixed Fermi-GT transitions is not as useful for extracting information on scalar or tensor currents because one needs to know the relative amounts of each component. However, one can take the existing limits on scalar and tensor currents from other experiments and use the $e$-$\nu$ correlation to extract the $B$(GT)/$B$(F) ratio. For simplicity, we here assume the scalar and tensor couplings to be zero. We will soon gather the necessary information for the case of $^{33}$Ar. We will obtain the half-life of $^{33}$Ar and the absolute branching ratio of the superallowed transition from our experiment at MSU~\cite{ko:99}. The endpoint can be extracted from the Isospin-Multiplet Mass Equation. \begin{table} \begin{flushleft} \caption{Comparison of the measured mass excesses of the lowest $T=3/2$ quintet in $A=33$ to predictions of the Isospin-Multiplet Mass Equation [$P(\chi^2, \nu)=0.51$].} \label{tab: IMME33} \begin{tabular}{lrrr} \hline \hline isobar & $T_3$ & $M_{\rm exp}$~(keV)$^a$ & $M_{{\rm IMME}}$ (keV) \\ \hline $^{33}$P & $+3/2$ & $-26337.7 \pm 1.1 $ & $-26337.7\pm 1.1$ \\ $^{33}$S & $+1/2$ & $-21106.14 \pm 0.41$$^b$ & $-21106.15\pm 0.41$ \\ $^{33}$Cl & $-1/2$ & $-15460.2 \pm 1.0 $$^c$ & $-15460.10 \pm 1.0$ \\ $^{33}$Ar & $-3/2$ & $-9380 \pm 30$ & $-9399.54 \pm 3.4 $ \\ \hline \end{tabular} \\ $^a$unless noted otherwise, ground state masses are from Ref.~\protect\cite{au:95}.\\ $^b$excitation energy from Ref.~\protect\cite{en:90} and masses of Ref.~\protect\cite{au:95}.\\ $^c$from $^{32}$S($p,p$) resonance energy\cite{en:90} and masses of Ref.~\protect\cite{au:95}.\\ \end{flushleft} \end{table} Table~\ref{tab: IMME33} shows the mass excesses for the $T=3/2$ quartet and the corresponding values of the IMME fit. Here we get: \begin{equation} \Delta(^{33}{\rm Ar})= 6060.6 \pm 2.6 ~{\rm keV.} \end{equation} In addition there are plans to determine the $^{33}$Ar mass to $\approx 3$ keV using ISOLTRAP\cite{bo:98} (the mass of the daughter level is known to $\approx 1$ keV). \begin{figure}[ht] \hspace{-1.2cm} \hbox{\vspace{-4cm} \psfig{figure=fig6.ps,height=28cm,width=16cm}} \vspace{-15cm} \caption{R-matrix fit to the $^{33}$Ar delayed proton spectrum. This spectrum, contains roughly 1/10 of our data.} \label{fig: fit_33ar} \end{figure} We obtain $a$ from the $^{33}$Ar delayed proton spectrum, a typical example of which is shown in Fig~\ref{fig: fit_33ar}. Our fit to the ISOLDE data yields \begin{equation} a(^{33}{\rm Ar})= 0.944 \pm 0.002({\rm stat.}) \pm 0.003({\rm syst.}), \label{eq: a_33_ours} \end{equation} which implies $B({\rm GT})/B({\rm F})=0.044 \pm 0.002$~\cite{note1}. Our result is in reasonably good agreement with the the shell-model calculation of ~\cite{br:85} which predicts $B({\rm GT})/B({\rm F}) = 0.055$, but in strong disagreement with the previous determination of Schardt and Riisager who obtained~\cite{sc:93} \begin{equation} a(^{33}{\rm Ar}) > 1.02 \pm 0.04 \\ \label{eq: a_33_sr} \end{equation} ($2 \sigma$ error bars), which can be translated into an upper limit $B({\rm GT})/B({\rm F}) < 0.015 $ which is $\approx 4$ of their $\sigma$'s from our value. We evaluated the systematic errors in $a$ following the same procedure we used for $^{32}$Ar. Using the endpoint deduced from the IMME and the masses of Table~\ref{tab: IMME33}, combined with the known $Q_p = 2276.5 \pm 1.0$ keV, and the derivatives $\partial \tilde{a}/\partial \Delta = -9.1 \times 10^{-4}$~keV$^{-1}$; and $\partial \tilde{a}/\partial Q_p=-8.5 \times 10^{-4}$~keV$^{-1}$, extracted by redoing the analysis with different values of $\Delta$ and $Q_p$, yields $\delta \tilde{a} = 0.0023$. Varying the width of the fitting region by 28\% yields variations of $\delta \tilde{a} \approx 0.0004$. Simultaneously fitting the $^{32}$Ar and $^{33}$Ar data with a common detector response function yields $\delta \tilde{a} \approx 0.002$. These three uncertainties were combined in quadratures to give the total systematic error shown in Eq.~\ref{eq: a_33_ours}. The discrepancy between our result in Eq.~\ref{eq: a_33_ours} and that of Schardt and Riisager in Eq.~\ref{eq: a_33_sr} is due primarily to differences in the analysis rather than disagreement of the data itself. Dieter Schardt kindly made the raw data of Ref.~\cite{sc:93} available to us and our analysis of their data gave a result essentially consistent with Eq.~\ref{eq: a_33_ours}. We now can show the potential value of measuring the ${\cal F} t$ value for the superallowed transition by imagining that the total strength $B({\rm F})+B({\rm GT})$ has been determined to $\approx 0.3$\%. Combining this with our positron-neutrino correlation measurement would yield $B({\rm F})$ to $\approx 0.5$\%. The predicted~\cite{br:98} isospin-mixing correction for $^{33}$Ar is $\approx 1.2$\%. For comparison, the correction in $^{34}$Cl is $\approx 0.6$\%. The fact that the correction is enhanced, as in $^{32}$Ar, makes it easier to measure. So these measurements could check whether the calculated corrections are accurate to within $\approx$50\%. It should be noted that larger discrepancies (corrections should be $\approx 0.7$\% as opposed to $\approx 0.4$\%~\cite{ha:98a}) would be needed to explain away the apparent non-unitarity of the Kobayashi-Maskawa matrix. \\ \\ \noindent {\bf Acknowledgments}\\ We thank D. Forkel-Wirth for help setting up our apparatus at CERN. This work was supported in part by the USA National Science Foundation and the Warren Foundation (at the University of Notre Dame) and by the Department of Energy (at the University of Washington). \newpage
\section{Introduction} Recent advances in string theories suggest that a special 11-dimension theory (dubbed as M theory) \cite{mtheory} may be the theory of everything. Impacts of M theory on our present world can be studied with compactification of the 11 dimensions down to our $3+1$ dimensions. The path of compactification is, however, not unique. In this multi-dimensional world, the standard model particles live on a brane ($3+1$ dim) while there are other fields, like gravity and super Yang-Mill fields, live in the bulk. The scale at which the extra dimensions are felt is unknown, anywhere from TeV to Planck scale. Recent studies \cite{gut} show that if this scale is of order TeV and there are gauge and fermion fields living in the bulk that correspond to the Kaluza-Klein excitations of the gauge and fermion fields of the SM, early unification of gauge couplings can be realized below or even much below the original GUT scale. This is possible because the extra matters in the bulk accelerate the RGE running of the gauge couplings, which then change from logarithmic evolution to power evolution. Supersymmetry model building is also an active area in the framework of extra dimensions \cite{msusy}. Apart from the above, radical ideas like TeV scale string theories were also proposed \cite{tevstring}. Inspired by string theories a simple but probably workable solution to the gauge hierarchy was recently proposed by Arkani-Hamed, Dimopoulos and Dvali (ADD) \cite{theory}. They assumed the space is $4+n$ dimensional, with the SM particles living on a brane. While the electromagnetic, strong, and weak forces are confined to this brane, gravity can propagate in the extra dimensions. To solve the gauge hierarchy problem they proposed the ``new'' Planck scale $M_S$ is of the order of TeV in this picture with the extra dimensions of a very large size $R$. The usual Planck scale $M_G=1/\sqrt{G_N} \sim 1.22 \times 10^{19}$ GeV is related to this effective Planck scale $M_S$ using the Gauss's law: \begin{equation} R^n \, M_S^{n+2} \sim M_G^2 \;. \end{equation} For $n=1$ it gives a large value for $R$, which is already ruled out by gravitational experiments. On the other hand, $n=2$ gives $R \alt 1$ mm, which is in the margin beyond the reach of present gravitational experiments. The graviton including its excitations in the extra dimensions can couple to the SM particles on the brane with an effective strength of $1/M_S$ (instead of $1/M_G$) after summing the effect of all excitations collectively, and thus the gravitation interaction becomes comparable in strength to weak interaction at TeV scale. Hence, it can give rise to a number of phenomenological activities testable at existing and future colliders \cite{wells,shrock,mira,han,joanne,tom1,id1,id2,bere,desh,grae,nath,ours,tom2,atwood,yuan,id3,id4}. So far, studies show that there are two categories of signals: direct and indirect. The indirect signal refers to exchanges of gravitons in the intermediate states, while direct refers to production or associated production of gravitons in the final state \cite{wells,mira,han,ours,atwood,yuan}. Indirect signals include fermion pair, gauge boson pair production, correction to precision variables, etc. \cite{wells,shrock,han,joanne,tom1,id1,id2,bere,desh,grae,nath,tom2,id3,id4}. There are also other astrophysical and cosmological signatures and constraints \cite{others}. Processes that only occur via loop diagrams in the SM are especially interesting if the low scale gravity allows tree-level interactions. In the SM, the lowest order photon-photon scattering can only take place via box diagrams of order $\alpha^2$ (on amplitude level) \cite{photon} and, therefore, is highly suppressed. Thus, photon-photon scattering opens an interesting door for any tree-level photon interactions. Even if such new interactions are much weaker than the electroweak strength, these tree-level diagrams are only of order $\alpha_{\rm new}$. It stands a good chance that these new interactions can beat the standard model. In the framework of ADD, photons can scatter via exchanges of spin-2 gravitons in $s$-, $t$-, and $u$-channels and the most important is that the coupling strength can be as large as the electroweak strength. In this work, we shall study the photon-photon scattering $\gamma\gamma \to \gamma\gamma$ and demonstrate that it provides a unique channel to identify the low scale gravity interactions. Other interesting processes of the same category is $\gamma\gamma \to \nu \bar \nu$ and the cross-channel, $\gamma \nu \to \gamma \nu$, both of which do not have any tree-level contributions in the SM \cite{duane0}. We shall not pursue these two further in this paper. Similarly, a pair of gluons can scatter into gluons or photons via exchanges of gravitons, the latter of which is our attention at hadron colliders. The lowest order $gg\to \gamma\gamma$ scattering occurs via a $s$-channel exchange of graviton in the low scale gravity model whereas it has to be via box diagrams in the SM. Thus, the new gluon scattering will give rise to anomalous diphoton production, in addition to the $q \bar q \to G \to \gamma\gamma$ channel, at hadron colliders. However, the tree-level SM $q \bar q \to \gamma \gamma$ presents a large irreducible background, not to mention the jet-fake background. This makes the diphoton production at hadron colliders not as attractive as in $\gamma\gamma$ and $e^+ e^-$ colliders as a probe to the low scale gravity model. For completeness we also study the diphoton production at $e^+ e^-$ colliders. The organization of the paper is as follows. In the next section, we compare the photon-photon scattering cross section between the SM and the low scale gravity. In Sec. III, we calculate diphoton production at the Tevatron and obtain the present limit on the cut-off scale $M_S$ using the diphoton data, and then estimate the sensitivity reach at the Run II. In Sec. IV, we repeat the same exercise at $e^+ e^-$ colliders and obtain the limits using the diphoton data from LEPII, and estimate the sensitivity reach at the future linear $e^+ e^-$ colliders. We shall then conclude in Sec. V. \section{Photon-photon Scattering} We concentrate on the spin-2 component of the Kaluza-Klein (KK) states, which are the excited modes of graviton in the extra dimensions. The spin-0 component has a coupling to the gauge boson proportional to the mass of the gauge boson in the unitary gauge, which means it has a zero coupling to photons. We follow the convention in Ref.\cite{han}. There are three contributing Feynman diagrams for the process $\gamma\gamma \to \gamma\gamma$ in the $s$-, $t$-, and $u$-channels. The amplitudes for $\gamma(p_1) \gamma(p_2) \to \gamma(k_1) \gamma(k_2)$ are given by: \begin{eqnarray} i{\cal M}_1 &=& - \frac{\kappa^2}{8} D(t)\, B^{\mu\nu,\mu'\nu'}(p_1-k_1)\, \epsilon^\rho(p_1) \, \epsilon^\sigma(p_2) \, \epsilon^\alpha(k_1) \, \epsilon^\beta(k_2) \, \nonumber \\ &&\times [ -p_1 \cdot k_1\, C_{\mu\nu,\rho\alpha} + D_{\mu\nu,\rho\alpha}(p_1,-k_1) ]\; [ -p_2 \cdot k_2\, C_{\mu'\nu',\sigma\beta} + D_{\mu'\nu',\sigma\beta} (p_2,-k_2) ]\;, \\ i{\cal M}_2 &=& i{\cal M}_1 (k_1 \leftrightarrow k_2 ) \;,\\ i{\cal M}_3 &=& - \frac{\kappa^2}{8} D(s)\, B^{\mu\nu,\mu'\nu'}(p_1+p_2)\, \epsilon^\rho(p_1) \, \epsilon^\sigma(p_2) \, \epsilon^\alpha(k_1) \, \epsilon^\beta(k_2) \, \nonumber \\ &&\times [ p_1 \cdot p_2\, C_{\mu\nu,\rho\sigma} + D_{\mu\nu,\rho\sigma}(p_1,p_2) ]\; [ k_1 \cdot k_2\, C_{\mu'\nu',\alpha\beta} + D_{\mu'\nu',\alpha\beta} (-k_1,-k_2) ]\;, \end{eqnarray} where $\kappa =\sqrt{16\pi G_N}$ and $B_{\mu\nu,\rho\sigma}(k)$, $C_{\mu\nu,\rho\sigma}$ and $D_{\mu\nu,\rho\sigma}(p_1,p_2)$ can be found in Ref. \cite{han}. The propagator factor $D(s) =\sum_k i/(s-m_k^2 +i\epsilon)$, where $k$ sums over all KK levels. After some tedious algebra the square of the amplitude, summed over final and averaged over the initial helicities, is surprisingly simple: \begin{equation} \overline{\sum} |{\cal M}|^2 = \frac{\kappa^4}{8} \; |D(s)|^2 \; (s^4 + t^4 + u^4 ) \;, \end{equation} where we have taken $M_S^2 \gg s, |t|, |u|$ and in this case the propagator factor $D(s)=D(|t|)=D(|u|)$ \cite{han}, which is given by \begin{equation} \kappa^2 |D(s)| = \frac{16 \pi}{M_S^4}\; \times {\cal F} \;, \end{equation} where the factor ${\cal F}$ is given by \begin{equation} \label{F} {\cal F} = \Biggr \{ \begin{array}{l} \log \left( \frac{M_S^2}{s} \right ) \;\; {\rm for}\;\; n=2 \;, \\ \frac{2}{n-2} \;\;\;\;\;\;\;\;\;\;\; {\rm for}\;\; n>2 \;. \end{array} \end{equation} The angular distribution is \begin{equation} \label{aa-cos} \frac{d\sigma (\gamma\gamma \to \gamma\gamma)}{d |\cos\theta| } = \frac{\pi s^3}{M_S^8}\, {\cal F}^2 \, \biggr [ 1 + \frac{1}{8}\,(1+ 6 \cos^2\theta + \cos^4 \theta ) \biggr ] \;, \end{equation} where $|\cos \theta|$ is from $0$ to 1. Since the cross section scales as $s^3/M_S^8$, which implies larger cross sections at higher $\sqrt{s}$. The SM background calculation is well known and we do not repeat the expressions here. We used the results in Ref. \cite{photon} with the form factors from Ref. \cite{duane}. The process is via box diagrams with all charged fermions and the $W$ boson in the loop. At the low energy, the fermion contribution dominates, but once $\sqrt{s}$ gets above a hundred GeV the $W$ contribution becomes more important and completely dominates at higher $\sqrt{s}$. We show the cross sections in Fig. \ref{fig-aa}(a). This SM cross section decreases gradually when $\sqrt{s}$ is above 500 GeV. In contrast, the low scale gravity interactions give a monotonically increasing cross section. For $n=2$ and $M_S=4$ TeV the cross-over is at about $\sqrt{s}=600$ GeV. We notice that the signal cross section does not decrease very rapidly with $n$, unlike the production of real gravitons \cite{mira,ours}. In Fig. \ref{fig-aa}(b), we show the angular distribution for the low scale gravity and for the SM. The signal has a relatively flat distribution, as can be easily deduced from Eq. (\ref{aa-cos}). The ratio of the cross section at $|\cos\theta|=0$ to that at $|\cos\theta|=1$ is only $9/16$. On the other hand, the SM background is very steep around $|\cos\theta|=1$, and that is why a cut of $|\cos\theta|< \cos 30^\circ$ is imposed to reduce the background. Monochromatic photon beam can be realized using the back-scatter laser technique \cite{telnov} by shining a laser beam onto an electron or positron beam. A linear $e^+ e^-$ collider can be converted into an almost monochromatic photon-photon collider, with a center-of-mass energy about 0.8 of the parent $e^+ e^-$ collider and with a luminosity the same order as the parent, i.e., as large as 50--100 fb$^{-1}$ per year. Since the cross section for the SM is of the order of 10 fb, so there should be enough events for doing a counting experiment. A 5--10 \% deviation from the SM prediction would be at a $1.1-3.2\sigma$ level. We use the 5\% or 10\% deviation from the SM as the criterion for sensitivity reach. The sensitivity reach at the $\gamma\gamma$ collider is shown in Fig. \ref{fig-aa-limit}. The reach on $M_S$ is about 5--8 (4.5--7.5) times of the center-of-mass energy of the collider for $n=2,4,6$ using the 5\% (10\%) deviation criterion. As we shall see later, the sensitivity reach at photon-photon colliders is better than at $e^+ e^-$ and much better than at hadron colliders. \section{Diphoton production at the Tevatron} Diphoton production has been an interesting subject for CDF and D0. It can provide constraints on the $qq\gamma\gamma$ type contact interactions, anomalous $\gamma\gamma\gamma$ and $Z\gamma\gamma$ couplings. In the context of the low scale gravity, diphotons can be produced via quark-antiquark and gluon-gluon annihilation into virtual gravitons and the associated KK states. The gluon-gluon annihilation is very similar to the photon-photon scattering described in the last section. The main background is the SM lowest order process: $q\bar q \to \gamma\gamma$. \footnote{Since the lowest order diphoton production $q\bar q\to \gamma \gamma$ is much larger than the box process: $gg \to \gamma\gamma$, we shall neglect the latter in considering the SM background.} There are two contributing subprocesses: \begin{eqnarray} \frac{d\sigma (q\bar q \to \gamma \gamma)}{d\cos\theta^*} &=& \frac{1}{96\pi \hat s}\; \Biggr [ 2 e^4 Q_q^4 \, \frac{1+\cos^2 \theta^*}{1-\cos^2\theta^*} + 2 \pi e^2 Q_q^2\, \frac{\hat s^2}{M_S^4} \,(1+\cos^2\theta^*) \, {\cal F} \nonumber \\ && + \frac{\pi^2}{2}\, \frac{\hat s^4}{M_S^8} \, (1-\cos^4\theta^*) \, {\cal F}^2 \Biggr ] \;, \label{qq} \end{eqnarray} \begin{equation} \frac{d\sigma (gg \to \gamma \gamma)}{d\cos\theta^*} = \frac{\pi}{512}\, \frac{\hat s^3}{M_S^8} \, (1+ 6 \cos^2\theta^* + \cos^4\theta^*) \; {\cal F}^2 \;, \end{equation} where the factor ${\cal F}$ is given in Eq. (\ref{F}), and the $\theta^*$ is the scattering angle in the center-of-mass frame and $\cos\theta^*$ is from $-1$ to 1. In $q\bar q \to \gamma \gamma$, the effect of graviton exchanges first occurs in the interference term, which only scales as $\hat s^2/M_S^4$, and potentially more important than the square term of $\hat s^4/M_S^8$ at $\hat s \ll M_S^2$. Both CDF and D0 \cite{cdf-d0} have preliminary data on diphoton production. We are going to use their data to constrain $M_S$. CDF has measured the invariant mass $M_{\gamma\gamma}$ spectrum in the region 50 GeV $<M_{\gamma\gamma}<$ 350 GeV. However, since the data is only preliminary and in graphical form only, we can only use the reported number of events in the region $M_{\gamma\gamma}>150$ GeV: 5 events are observed where $4.5\pm0.6$ are expected with an integrated luminosity of 100 pb$^{-1}$. This data, though without binning information, is sufficient to place a constraint on $M_S$, because the signal for the low-scale gravity does not appear as a peak in the $M_{\gamma\gamma}$ spectrum but, instead, a gradual enhancement from about $M_{\gamma\gamma}\approx 150$ GeV towards higher $M_{\gamma\gamma}$. We use the Poisson statistics to calculate the 95\% CL upper limit to the number of {\it signal} events $N_{95}$, \footnote{The number of signal events is $N_{95}$ or less with 95\% confidence.} using \begin{equation} 0.95=1-\epsilon=1- \frac{e^{-(n_B + N_{95})}\; \sum_{n=0}^{n_{\rm obs}} \, \frac{(n_B + N_{95})^n}{n !} } {e^{-n_B}\; \sum_{n=0}^{n_{\rm obs}} \, \frac{n_B^n}{n !} } \;, \end{equation} where $n_B=4.5$ is the expected number of background events and $n_{\rm obs} =5$ is the number of observed events. We obtain $N_{95}=6.61$. We then normalized our calculation to the expected number of events (=4.5) after imposing the same selection cuts as CDF. With this normalization we can then calculate $M_S$, which gives a signal of 6.61 events in excess of the SM prediction. We obtain the 95\% CL lower limit on $M_S$: \begin{eqnarray} \mbox{Tevatron Run I:} \qquad \qquad & & M_S > 0.91\; {\rm TeV}\;\; {\rm for}\;\; n=2 \;\; {\rm and} \nonumber \\ &&M_S > 0.87\; {\rm TeV}\;\; {\rm for}\;\; n=4 \nonumber \;. \end{eqnarray} For D0, however, the highest bin in the measured $M_{\gamma\gamma}$ spectrum is 80--112 GeV. At such a low value, it is difficult to see the effect of $(\hat s^2/M_S^4)$. Thus, we expect the limit that would be obtained from the D0 data is somewhat smaller than using the CDF data. Next, we estimate the sensitivity reach at the Run II of the Tevatron, assuming a luminosity of 2 fb$^{-1}$. The effect of the low scale gravity on the $M_{\gamma\gamma}$ spectrum is shown in Fig. \ref{fig-te}. It is easy to understand why the enhancement is more likely at the large $M_{\gamma\gamma}$. To estimate the sensitivity we divide the $M_{\gamma\gamma}$ spectrum into bins: a bin width of 100 GeV for bins in $200\;{\rm GeV} < M_{\gamma\gamma} < 500$ GeV, and for 500 GeV to 1000 GeV we combine it into one bin only. This is to make sure that each bin should have at least a few events in the SM: see the first row of Table \ref{table-te} (we also use a selection efficiency of 50\%.) For each bin we assume the SM prediction as the number of events that would be observed: $n^{\rm obs}$, and we calculate the number of events predicted by a $M_S$ and $n$: $n^{\rm th}$. We then calculate the $\chi^2$ for this bin and sum over all bins, using \begin{equation} \chi^2(M_S,n)= \sum_{i={\rm bins}} \biggr[ 2\left( n^{\rm th}_i - n^{\rm obs}_i \right ) + 2 n^{\rm obs}_i \, \ln \left( \frac{n^{\rm obs}_i}{n^{\rm th}_i} \right ) \biggr ]\;. \end{equation} The $\chi^2$ then gives a goodness of the fit for the value of $M_S$ and $n$. The larger the $\chi^2$ the smaller the probability that the corresponding value of $M_S$ and $n$ is a true representation for the data. To place a 95\%CL lower limit on $M_S$ a $\chi^2=9.49$ is needed for 4 degrees of freedom. The number of events in each bin for $n=2$ and $M_S=1.5-2$ TeV, and for $n=4$ and $M_S=1.4-2$ TeV with the corresponding $\chi^2$ are shown in Table \ref{table-te}. We obtain a limit of \begin{eqnarray} \mbox{Tevatron Run II:} \qquad \qquad & & M_S>1.72\;{\rm TeV}\;\; {\rm for} \;\; n=2 \;\; {\rm and} \nonumber \\ &&M_S>1.43\;{\rm TeV}\;\; {\rm for} \;\; n=4 \nonumber \;. \end{eqnarray} We verified that the binning is not important for the limit. We repeat the procedures using only one large bin from 200 to 1000 GeV, and the 95\% CL lower limit on $M_S$ becomes 1.73 (1.38) TeV for $n=2\;(4)$. \begin{table}[th] \caption{ \label{table-te} The number of events that would be observed in each bin of $M_{\gamma\gamma}$ for the SM and for the low scale gravity at the Tevatron with $\sqrt{s}=2$ TeV and a luminosity of 2 fb$^{-1}$. The $\chi^2$ is calculated assuming the SM prediction is what would be observed. The cuts imposed are: $|\eta_\gamma|<1$, $p_{T\gamma}>20$ GeV, and a selection efficiency of 0.5 is assumed. } \medskip \begin{tabular}{c|cccc|c} & \multicolumn{4}{c|}{bin} & \\ model & 200--300 GeV & 300--400 GeV & 400--500 GeV & 500--1000 GeV&$\chi^2$ \\ \hline \hline SM & 47.68 & 11.98 & 3.65 & 1.81 & - \\ \hline \hline $n=2$ & & & & & \\ $M_S=2.0$ TeV & 50.27 & 14.40 & 5.53 & 4.84 & 3.80 \\ $M_S=1.9$ TeV & 50.82 & 14.92 & 5.93 & 5.54 & 5.24 \\ $M_S=1.8$ TeV & 51.53 & 15.59 & 6.46 & 6.45 & 7.33 \\ $M_S=1.75$ TeV & 51.96 & 15.99 & 6.78 & 7.01 & 8.70 \\ $M_S=1.7$ TeV & 52.47 & 16.45 & 7.15 & 7.66 & 10.37 \\ $M_S=1.6$ TeV & 53.75 & 17.61 & 8.09 & 9.30 & 14.87 \\ $M_S=1.5$ TeV & 55.49 & 19.25 & 9.38 & 11.58 & 21.72 \\ \hline \hline $n=4$ & & & & & \\ $M_S=2.0$ TeV & 48.24 & 12.62 & 4.22 & 2.96 & 0.64 \\ $M_S=1.9$ TeV & 48.38 & 12.77 & 4.37 & 3.28 & 0.97 \\ $M_S=1.8$ TeV & 48.54 & 12.97 & 4.55 & 3.72 & 1.49 \\ $M_S=1.7$ TeV & 48.76 & 13.24 & 4.80 & 4.38 & 2.39 \\ $M_S=1.6$ TeV & 49.07 & 13.62 & 5.15 & 5.35 & 3.89 \\ $M_S=1.5$ TeV & 49.52 & 14.16 & 5.65 & 6.87 & 6.53 \\ $M_S=1.4$ TeV & 50.20 & 14.94 & 6.40 & 9.35 & 11.30 \\ \end{tabular} \end{table} \section{Diphoton production at $e^+ e^-$ colliders} We can use Eq. (\ref{qq}) with $Q_q=-1$ and multiply it by 3 to derive the expression for $e^+ e^- \to \gamma \gamma$: \begin{equation} \label{ee} \frac{d\sigma (e^+ e^- \to \gamma \gamma)}{dz} = \frac{2 \pi}{s}\; \left( \alpha^2 \frac{1+ z^2}{1- z^2} + \frac{\alpha}{4}\, \frac{s^2}{M_S^4} \, {\cal F}\, (1+z^2) + \frac{1}{64}\, \frac{s^4}{M_S^8} \, {\cal F}^2 \, (1-z^4) \right ) \end{equation} where $z=|\cos\theta|$ is the polar angle of the outgoing photon and $z$ ranges from 0 to 1. The four LEP collaborations have been measuring the diphoton production $e^+ e^- \to \gamma\gamma$ \cite{lep-aa} and using the data to constrain the deviation from QED and generic types of contact interactions of order $1/\Lambda_n$, $n=6,7,8$. Since these contact interaction parameters $1/\Lambda_n$ can be converted from the QED cutoff parameter $\Lambda_{\pm}$, we shall stick with the QED cutoff parameter in the following discussion. The possible deviation from QED is usually characterized by a cutoff parameter $\Lambda_\pm$ corresponding to a modified angular distribution: \begin{equation} \label{zz} \frac{d\sigma}{dz} = \frac{2\pi \alpha^2}{s}\; \frac{1+z^2}{1-z^2} \; \left( 1 \pm \frac{s^2}{2 \Lambda_\pm^4}\, (1 - z^2 ) \right )\;, \end{equation} where $z = |\cos\theta|$ and ranges from 0 to 1. Each collaboration measured the $\cos\theta$ distribution and obtained the 95\% CL limit on $\Lambda_\pm$ by varying $\eta=1/\Lambda_\pm^4$ and maximizing the likelihood function. Since each experiment has their own procedures, we adopt a simple approach that takes their limits on $\Lambda_\pm$ and converts them into limits on $M_S$. Note that in Eq. (\ref{ee}) the third term is suppressed relative to the second term, we can, therefore, just take the first and the second term, then it will look like Eq. (\ref{zz}). The QED cutoff parameter $\Lambda_+$ is related to $M_S$ by \begin{equation} \label{convert} \frac{M_S^4}{\cal F} = \frac{\Lambda_+^4}{2 \alpha} \;. \end{equation} The limits from each LEP experiment and the corresponding limits on $M_S$ are tabulated in Table \ref{table-ee}. Note that we used only $\Lambda_+$ to calculate $M_S$. The limits on $M_S$ is at most about 1.4 TeV for $n=2$ and about 1 TeV for $n=4$. The result for $n=2$ is enhanced because of the logarithmic factor in ${\cal F}$. Using the value of $M_S \sim 1$ TeV we can verify the ratio of the third term to the second term in Eq. (\ref{ee}) and the third term is only about 2\% of the second term. It justifies the approximation that we take only the first two terms of Eq. (\ref{ee}). So far, the treatment is rather simple. A better limit can be obtained by combining the data on $\eta = 1/\Lambda_\pm^4$ from each LEP experiment. However, since some of the data on $\eta$ are not given in detail, we can only combine those with a central value and an error. We have the following available: (i) OPAL (183 GeV): $\eta=(1.04 \pm 1.34)\times 10^{-10}\;{\rm GeV}^{-4}$, (ii) L3 (183 GeV): $\eta=(-0.59 \stackrel{\scriptstyle +1.19}{\scriptstyle -1.13}) \times 10^{-10}\;{\rm GeV}^{-4}$, (iii) L3 (161,172 GeV): $\eta=(-0.77 \stackrel{\scriptstyle +2.83}{\scriptstyle -2.58}) \times 10^{-10}\;{\rm GeV}^{-4}$, and (iv) DELPHI (183 GeV): $\eta=(-1.4 \pm 1.5)\times 10^{-10}\;{\rm GeV}^{-4}$. We combine these data and assuming they are all gaussian we obtain $\eta=(-0.31 \stackrel{\scriptstyle +0.74}{\scriptstyle -0.73})\times 10^{-10} \;{\rm GeV}^{-4}$, the error of which is given in $1\sigma$. From this $\eta$ the corresponding 95\% CL limit on $\Lambda_\pm$ are $\Lambda_+ > 298$ GeV and $\Lambda_- > 279$ GeV. We can see that the combined limit on $\Lambda_+$ is still not as good as the single limit from ALPEH (189 GeV) or OPAL (189 GeV). Once the data from each LEP experiment are given we can certainly improve the limit by combining them. Thus, for the present moment the best limit is from OPAL (189 GeV): $\Lambda_+ > 345$ GeV, which converts to $M_S> 1.38\; (0.98)$ TeV for $n=2\;(4)$. The behavior of the new gravity interactions at higher $\sqrt{s}$ can be easily deduced from Eq. (\ref{ee}). The new interaction gives rise to terms proportional to $s^2/M_S^4$ and $s^4/M_S^8$, which get substantial enhancement at large $\sqrt{s}$: see Fig. \ref{fig-ee}(a). The angular distribution also becomes flatter because in the SM the distribution scales as $(1+z^2)/(1-z^2)$ whereas the terms arising from the new gravity interactions scale as $(1+z^2)$ and $(1-z^4)$, respectively, as shown in Fig. \ref{fig-ee}(b). Here we also attempt to estimate the sensitivity reach on the cut-off scale $M_S$ at the future linear $e^+ e^-$ colliders. Since the cross section is of the order of 0.1 to 1 pb for $\sqrt{s}=0.5-2$ TeV, it corresponds to about $10^3 - 10^4$ events for a mere yearly luminosity of 10 fb$^{-1}$. Thus, a 5\% (10\%) deviation from the SM prediction corresponds to a level of $1.6\sigma - 5\sigma$ ($3.2\sigma - 10\sigma$). In Fig. \ref{fig-ee-limit}, we show the sensitivity reach on $M_S$ by requiring a 5\% or 10\% deviation from the SM prediction. The reach on $M_S$ is about 3.5--5.5 (3--4.5) times of the $\sqrt{s}$ of the collider for the 5\% (10\%) criterion. \begin{table}[t] \caption[]{\label{table-ee} The 95\% CL limits on the QED cutoff parameter $\Lambda_\pm$ from LEP experiments \cite{lep-aa} and the corresponding 95\% limits on $M_S$ obtained using Eq. (\ref{convert}). We show only the result of the highest energy of each experiment whichever available. } \medskip \begin{tabular}{lc|cc} &95\%CL limit on $\Lambda_+$ and $\Lambda_-$ & \multicolumn{2}{c}{95\% CL limit on $M_S$ (TeV)} \\ & & $n=2$ & $n=4$ \\ \hline OPAL ($\sqrt{s}=189$ GeV): & $\Lambda_+ >345$ GeV & 1.38 & 0.98 \\ & $\Lambda_- >278$ GeV & & \\ \hline DELPHI ($\sqrt{s}=183$ GeV): & $\Lambda_+ >253$ GeV & 0.97 & 0.72 \\ & $\Lambda_- >225$ GeV & & \\ \hline L3 ($\sqrt{s}=183$ GeV): & $\Lambda_+ >262$ GeV & 1.01 & 0.74 \\ & $\Lambda_- >245$ GeV & & \\ \hline ALEPH ($\sqrt{s}=189$ GeV): & $\Lambda_+ >332$ GeV & 1.32 & 0.94 \\ & $\Lambda_- >265$ GeV & & \end{tabular} \end{table} \section{Conclusions} Diphoton production at $\gamma\gamma$, $p \bar p$, and $e^+ e^-$ colliders provides useful channels to search for the presence of the low scale gravity interactions, which are the effects of allowing gravity to propagate in the extra dimensions. Photon-photon colliders are able to give the best sensitivity reach on the cut-off scale $M_S$ of the low scale gravity model among the three. This is because $\gamma\gamma \to \gamma\gamma$ can only occur via box diagrams in the SM while in $e^+ e^-$ and $p \bar p$ collisions the tree-level contributions from the SM dominates. In addition to the total cross section, the angular distribution also serves as a tool to distinguish between the SM and the new gravity interactions, as seen in Fig. ~\ref{fig-aa}(b) and Fig. \ref{fig-ee}(b). The present limit from the LEPII diphoton data is about $M_S >1.4\;(1)$ TeV for $n=2\;(4)$, and it is only $M_S > 0.9$ TeV from the CDF diphoton $M_{\gamma\gamma}$ data. The sensitivity reach in $\gamma\gamma$ collisions is about 5--8 times of $\sqrt{s_{\gamma\gamma}}$ while it is only 3.5--5.5 times of the $\sqrt{s}$ at $e^+ e^-$ collisions. At the Run II of the Tevatron, the reach is only about 1.7 (1.4) TeV for $n=2\;(4)$. Finally, we emphasize the diphoton production at photon-photon colliders could provide a unique probe to the collider signature for the model of low scale gravity. \section*{\bf Acknowledgments} This research was supported in part by the U.S.~Department of Energy under Grants No. DE-FG03-91ER40674 and by the Davis Institute for High Energy Physics.
\section{Introduction} Past studies of the polarization properties of the Crab Nebula pulsar have been limited to low-frequency radio and visible wavelengths. The pulsar's steep radio spectrum, and interference from the radio-bright Nebula make observations above 1\,GHz difficult with single dish antennas. Thus, interpretations of the pulsar's emission geometry have only been made from the properties of its polarized profiles at visible wavelengths. Single dish average profile measurements of the radio polarization are available at frequencies between 110 and 1664\,MHz (Manchester, Huguenin, \& Taylor 1972\nocite{mht72}; Manchester 1971a\nocite{man71a}). Prior to the work described in Moffett \& Hankins (1996, hereafter Paper I), only three components of the pulsar's average profile were known; a steep-spectrum precursor, which is approximately 100\% linearly polarized (\cite{man71a}), plus a main pulse (MP) and interpulse (IP) which are roughly 15 to 25\% linearly polarized. The position angle (PA) remains constant between all three components, with very little change of position angle across them. The only other major radio polarization observations that have been published are measurements of the time-variable rotation measure used to probe the magnetic fields of the Crab Nebula's filaments (\cite{rci+88}). No high radio-frequency ($\nu > 1.7$\,GHz) polarization information for the Crab pulsar has been available. The visible wavelength (\cite{sjd+88}) and newly acquired ultraviolet (\cite{sdb+96}) polarization profiles show similar linear polarized fractions as the radio, but unlike the radio profiles, they show large PA variations. At the peak positions of the MP and IP, the fraction of visible linear polarization is about the same as in the radio regime, $\approx 14$ to 17\%. In the region after the IP in phase, the percentage polarization rises to $\approx 47 \pm 10$\%, and the position angle rises above the IP angle and remains nearly constant across the total intensity minimum. Narayan \& Vivekanand (1982)\nocite{nv82} found that the visible wavelength polarization PA sweeps of the MP and IP suggest that emission comes from two opposite poles of a pulsar whose magnetic axis is nearly orthogonal to the rotational axis. But arguments from $\gamma$-ray emission theory have surfaced recently (Manchester 1995\nocite{man95}; Romani \& Yadigaroglu 1995\nocite{ry95}) that question this type of geometry by claiming that emission arises from a wide cone in the outer magnetosphere. So far, the study of radio polarization has not improved our knowledge of the emission and field geometry. The lack of position angle variation in low frequency radio profiles is difficult to explain in terms of the simple rotating vector model (\cite{rc69}). Following the serendipidous discovery of additional components in the Crab pulsar's profile in Paper I, a program of polarization observations was scheduled to study the high radio-frequency polarization characteristics of these new components, perhaps improving the interpretation of the polarization and emission location for the Crab pulsar's radio components. \section{Observations} Observations were conducted during several sessions from October 1995 to October 1996 at the Very Large Array (VLA) of NRAO. Between February 22 and April 18, 1996, the data acquisition system was modified to double the number of filterbank channels that are recorded. Using the phased VLA, the coherent sum of undetected right-hand and left-hand circular polarization ($R$ and $L$) from all antennas is mixed to 150\,MHz and then split into 14 independent frequency channels by a MkIII VLBI filter bank. The filter bank output is sent to the VLA's High Time Resolution Processor (HTRP), which consists of a set of 14 multiplying polarimeters. Channels of detected and smoothed $LL$, $RR$, $RL\cos{\theta}$, and $RL\sin{\theta}$, where $\theta$ is the phase offset between $R$ and $L$, are continuously sampled by 12-bit, analog-to-digital converters in a PC and recorded on disk at a time resolution of 256 $\mu$s. The detector time constants are set to optimize sampling of the dispersed time series across the channel bandwidths. The observations were scheduled so that short scans (typically 30 to 40 minutes apart) of an unresolved calibrator point source were made between pulsar scans to keep the VLA phased, and to record on- and off-source data for flux and polarization calibration of the pulsar data. Within the duration of an observing session, anywhere from five to seven sets of measured Stokes parameter fluxes were recorded from the phase and flux calibrator, 3C138, with enough parallactic angle coverage for instrumental polarization calibration. The flux density and position angle of 3C138 are regularly monitored by the University of Michigan Radio Astronomy Observatory. The position angle of this source remains the same over our frequency coverage, with a value of $\psi = -12^\circ$. The pulsar data were folded off-line at the pulsar's topocentric period using a timing model initially provided by Nice (1995). Consequent observations of the Crab pulsar using the Princeton/Dartmouth Mark III Pulsar Timing System (\cite{skn+92}) provided time-of-arrival (TOA) information reduced using the program TEMPO (\cite{tw89}), which yielded new timing solutions for folding at later epochs. Individual channel data were folded into two-minute average profiles of all four detected polarizations prior to calibration and dedispersion. The gain amplitudes, relating the received voltage in the data acquisition system to flux density in Janskys for the $LL$ and $RR$ detector signals were determined by observation of the phase calibrator source and blank sky. The gain amplitude of the cross polarizations, $RL\cos{\theta}$ and $RL\sin{\theta}$, were found directly from solutions for the circular polarization gains $G_{\rm L}$ and $G_{\rm R}$ from $LL$ and $RR$, by using $G_{\rm RL} = \sqrt{G_{\rm R} \cdot G_{\rm L}}$. For data collected from the lower side-band channels of the MkIII VLBI videoconverters, the sign of the measured Stokes U was inverted, thus removing a known $180^\circ$ phase shift caused by the image-rejecting mixers within the videoconverters. Polarization calibration was completed following procedures similar to those used by McKinnon (1992)\nocite{mck92}. In his paper, the polarization characteristics of the phased VLA approximate those of a single dish antenna with circular polarization feeds. An ideal antenna with orthogonal circular polarization receivers has no cross-coupling. However, imperfections in the reflectors and receiving systems of antennas tend to change the received radiation from purely linear polarized sources into elliptical polarization (\cite{ck69}). McKinnon's method involves measuring the time-dependent Stokes parameters of a polarization calibrator source with respect to the changing parallactic angle, and solving for time-dependent and independent instrumental corrections. McKinnon used the polarization from a point in a pulsar's profile to perform a self-calibration, but he could not determine the absolute position angle. We could have used the Crab pulsar at 1.4\,GHz to perform such a self-calibration, but at higher frequencies we were limited by low signal to noise ratios. Instead we used a phase calibrator of known polarization characteristics to solve for the instrumental corrections {\em and} absolute position angle at 1.4, 4.9, and 8.4\,GHz. \section{Results} Our results at 1.4\,GHz (Figure \ref{lband_fig}) are similar to published observations at 1.664\,GHz (Manchester 1971a\nocite{man71a}; Manchester 1971b\nocite{man71b}), but with the addition of three more components (see Paper I, Figure 2). These three are labeled LFC for ``low frequency component'', as it mainly appears at $\nu < 2$\,GHz, and HFC1 and HFC2 for ``high frequency component 1 and 2'', as they appear only at $\nu\geq 1.4$\,GHz. The MP and IP are both linearly polarized, 25\% and 15\% respectively, and their relative PAs are nearly the same. The LFC component is more than 40\% linearly polarized, and it has a PA offset $\approx 30^\circ$ from the MP, which sweeps down toward the MP. There is also low-level emission after the IP, coincident in phase with components HFC1 and HFC2 found in the total intensity profiles at 4.9\,GHz (Paper I). These components are $> 50\%$ polarized, and their position angles appear to be relatively the same, offset from the MP by $60^\circ$. The circular polarization undergoes a sense reversal centered on the MP with an amplitude 1 -- 2\% of the MP peak. We can exclude this as a cross-coupling signature, even though our uncertainties for fitting the instrumental parameters were several times higher for individual frequency channels. The linear polarization is not strong, nor does it sweep rapidly across the pulse at our time resolution. So a coupling of linear to circular power should not produce sense-reversing circular polarization. We can make no comparisons as no previous circular polarization observations of average profiles have been published. The similarity of circular polarization signatures on separate observation dates, and in individual channels improves our confidence that these signatures are real. \placefigure{lband_fig} The profiles at 4.9\,GHz (Figure \ref{cband_fig}) are new results. We have successfully confirmed the detection of the HFC components found in Paper I. The pulsar was visible for only three out of seven observing sessions; the profile was formed from about 3.7 hours of data. We attribute the non-detections to heavy scintillation, which affected observations at 4.9\,GHz and higher frequencies. The IP, HFC1, and HFC2 are highly polarized, 50\% to 100\%, while the MP seems to have the same polarized fraction as at 1.4\,GHz. HFC1 and HFC2 share a common range of PA, but it sweeps through them with different slopes. The most important feature to note is the IP. Its relative flux density has increased with respect to the MP and it has shifted earlier in phase by $\approx 10^\circ$ (see Paper I, Figure 2). \placefigure{cband_fig} At 8.4\,GHz (Figure \ref{xband_fig}), the profiles show some polarization, and also confirm the profile morphology found in earlier total intensity observations. The pulsar was visible for only one out of three observing sessions for a total of 2.3 hours, which we again attribute to scintillation. The IP seen in the profile is substantially wider than at lower frequency and the fractional polarization of the IP and HFCs is reduced, due in part to an incomplete instrumental correction of the position angles. As in Paper I, no evidence is found of the MP, whose predicted flux from a spectral index of $\alpha_{\rm MP} = -3$ (Section \ref{spix_sec}) is below the noise level of the profile recorded at this frequency. \placefigure{xband_fig} Observations made on April 18, 19 and 20, 1996, were conducted at several frequencies within the 1.4\,GHz receiver band, and evidence for Faraday rotation was found between the separate frequencies. A rotation measure, RM$ = -46.9$\,rad\,m$^{-2}$, was found after comparing the position angles of the major components, and its effects have been removed from all profiles reported here. Past measurements show the RM near $-43.0$\,rad\,m$^{-2}$ (\cite{rci+88}), but it is known to be variable on time scales of months, as the line of sight to the pulsar passes through the Crab Nebula's filaments. After removing Faraday rotation effects, the position angles of the MP, HFC1 and HFC2 are found to align at 1.4 and 4.9\,GHz, and the PAs of the IP, HFC1 and HFC2 align at 4.9 and 8.4\,GHz. But, the IP is found to have a position angle difference between 1.4 and 4.9\,GHz of $90^\circ$ (see Figure \ref{angle3freq_fig}). So the IP has a discontinuous change of positional phase, flux, and polarization between 1.4 and 4.9\,GHz. It obviously cannot be the same component at both frequencies. \placefigure{angle3freq_fig} \section{Analysis} The unique and confusing discoveries described in the previous section are the first successful, fully polarimetric observations of the Crab pulsar above the 1.4-GHz band. In the following sections, this pulsar's emission geometry is explored by comparing properties of the its polarization profile with known properties of other pulsars, and possible emission geometry models. \subsection{Multiple Components} The components of the Crab pulsar appear in six distinct positions in rotational phase at all observed radio frequencies. The distribution of components is difficult to explain in a low-altitude, dipolar or hollow-cone emission models (\cite{ran83a}; \cite{lm88}), mainly because of their number and wide separation. Up to 5 components have been seen from ``normal'' pulsars (PSR B1237+25 and B1857-26). The separation of profile components is usually restricted to a small range of pulse phase ($< 30^\circ$), corresponding to a cone of emission above one pole of the star. However, a few interpulsars exist, whose components can be attributed to emission from the observer's line of sight passing above both poles (orthogonal rotator), or from one pole (aligned rotator). The phase separation between the MP and IP, $\Delta\phi_{\rm MP-IP} \approx 140^\circ$, is too low to argue for a line of site crossing of both poles. When compared to the high energy emission (infrared to $\gamma$-ray), the morphology of the MP and IP implies that they arise from a wide conal beam, high above a single pole (\cite{man95}). With the wide beam picture in mind, one apparent symmetry in the distribution of the Crab's components can be seen if we draw a line through the midpoint between the MP and IP, and the midpoint between HFC1 and HFC2 (see Figure 3 of Paper I). The midpoints between the component pairs are separated by $\approx 170^\circ$ at 4.9\,GHz. So the Crab's components could arise from conal emission regions above both poles, one wider than the other. In fact, the HFCs do show promise as a conal pair. Rankin's (1993)\nocite{ran93a} empirical relations for inner and outer conal width (assuming the Crab to be an orthogonal rotator, $\alpha = 90^\circ$) yield: \begin{displaymath} \begin{array}{cclcl} \rho_{\rm inner} & = & 2 \times 4.33^{\circ}\,P^{-1/2} = 47.3^{\circ}\\ \rho_{\rm outer} & = & 2 \times 5.75^{\circ}\,P^{-1/2} = 62.8^{\circ} \end{array} \end{displaymath} The phase separation of the HFCs, $\Delta\phi_{\rm HFC1-HFC2} \approx 56^\circ$, is within Rankin's predicted values for inner and outer conal widths for a pulsar of the Crab's period. It is possible that the HFCs are generated at low altitudes, and the MP and IP are generated at higher altitudes where the emission beam is much wider. However, we note that interpreting the frequency-dependent properties of the IP and the HFCs with this geometric model is quite difficult. Another set of components, the LFC, precursor, and main pulse, form what may be a cone/core triplet. The LFC to MP separation is $\approx 45^\circ$, nearly what one expects for the inner conal width, and the precursor behaves much like a core component, with its high polarization and steep spectrum. But why the MP is so much brighter than the LFC requires explanation. It is interesting to note that one pulsar, B1055-52, has a similar distribution of components (precursor, main pulse, and a strong interpulse located $155^\circ$ away) at low frequency (\cite{mha+76}). And like the Crab, it also has pulsed high energy emission X-rays (\cite{of93}), pulsed $\gamma$-rays (\cite{fbb+93}), and has been recently detected as continuum source at visible wavelengths (\cite{mcb97}). \subsection{Radius to Frequency Mapping? \label{freq_dep_sec}} Using the main pulse as the fidicial point of the Crab's profile (Paper I), we found that the separations from MP to IP, and from MP to the HFCs are frequency-dependent. (see Figure \ref{phase_sep_fig}). From 1.4 to 4.7\,GHz, the IP jumps $\approx 10^\circ$ earlier in phase, while the HFCs appear to make a smooth linear transit in phase between 1.4 and 8.4\,GHz. This property is reminiscent of the smooth phase shift of conal components in radius-to-frequency mapping (Cordes 1978\nocite{cor78}; Rankin 1983b\nocite{ran83b}; Thorsett 1991\nocite{tho91}). The phase separation of conal components usually can be best fit by a power law function, $\Delta\phi \propto \nu^{\eta}$, where $\-1.1 \leq \eta \leq 0.0$. The phase separations from the MP to both HFC1 and HFC2 are best fit with $\eta = 1$ (fit parameters found in Figure \ref{phase_sep_fig}). The HFCs are both moving toward later rotational phase with increasing frequency, unlike conal components of other pulsars, whose phase separation decreases to a common fiducial point. Curiously, the HFCs would merge at a common point at the MP phase, if their phases are extrapolated to a frequency above 60\,GHz. \placefigure{phase_sep_fig} \subsection{Spectral Index\label{spix_sec}} The amplitude calibration method for these observations was based on gains transferred from a standard extragalactic continuum calibration source, whereas the flux densities in the profiles presented in Paper I were estimated using known radiometer characteristics. We used the integrated flux density under the major components, and computed spectral indices for the MP and IP; $\alpha_{\rm MP} = -3.0$ for the MP, and $\alpha_{\rm IP} = -4.1$ for the IP at $\nu \leq 1.4$\,GHz. Independent of the uncertainty of the flux density measurements, the relative spectral index differences between components were determined simply through ratios of their integrated flux densities using the following relation: \begin{displaymath} {S_{\rm C1}(\nu_1)/S_{\rm C1}(\nu_2) \over S_{\rm C2}(\nu_1)/S_{\rm C2}(\nu_2)} = {\nu_1 \over \nu_2}^{(\alpha_{\rm C1} - \alpha_{\rm C2})} \end{displaymath} where the fluxes, $S_{\nu}$, and spectral indices, $\alpha$, correspond to the components C1 and C2. The spectral index difference of components using these ratios yields a spectral index, $\alpha_{\rm PC} \simeq -5.0$, for the precursor. The spectral indices we have found for the three major components of the Crab pulsar profile agree with previous measurements by Manchester (1971a)\nocite{man71a}. In Figure \ref{flux_vs_freq_fig}, we plot the flux density spectrum of the MP and IP, and the two HFC components. Below 1.4\,GHz, the IP follows a power-law spectral index of approximately -4, but above 1.4\,GHz, the plot shows that the IP has a flat spectral index, as do the HFC components, though no power-law can be determined from the plot. Such a turn-up or flattening of pulsar spectra has been observed by Kramer {\it et al.} (1996) \nocite{kxj+96} in two other pulsars. They have suggested that a transition from coherent to incoherent emission would cause changes in the expected flux density. Sampling pulsar radiation at very high frequencies gives limits to the bandwidth of the coherent emission mechanism. A simple extrapolation of the Crab pulsar spectrum from radio to infrared wavelengths (Fig. 4-2, Manchester \& Taylor 1977\nocite{mt77}) implies that the flux must rise and the emission mechanism must change. So the change in spectral index lends support to our hypothesis that the low frequency IP and high frequency IP are two different components. We should note, that when compared to other pulsars, the spectral indices of the Crab pulsar's MP and IP are much steeper than the components of other pulsars ($-1.5 < \alpha < -3$). The Crab's mean spectral index, $\alpha_{\rm crab} = -3.1$, is also greater than the average spectral index, $\alpha = -1.5$, of most detected pulsars (\cite{lyl+95}). \placefigure{flux_vs_freq_fig} \subsection{Polarization Properties} The polarization position angle of the Crab changes across the full period, though not significantly within components. There are no sudden well-defined PA sweeps (`S' shaped sweeps) within components, as seen at optical wavelengths. However, we should note the radio components are much narrower, and some polarization information is smeared by dispersion and scattering. The lack of PA variation between close components implies that the observer's line of sight trajectory does not fall close to the magnetic poles, where the position angle of field lines varies quickly. The fraction of linear to total intensity of the MP and HFCs is nearly constant from 1.4 to 4.9\,GHz. But the IP becomes substantially more polarized (from 20\% to 100\%) between the two frequencies, as well as undergoing a $90^\circ$ PA shift. The spectral change in phase and PA could be due to a mechanism (birefringence) affecting the propagation of the two orthogonal modes (ordinary or O-mode, and extraordinary or X-mode) of linear polarized radiation within the pulsar's magnetosphere (\cite{ba86}). The ordinary mode waves are forced to travel along magnetic field lines, while the extraordinary mode waves are unaffected. A sudden change in plasma conditions could cause one of the modes to be beamed out of the line of sight. However, this process is sensitive to frequency, and any transitions we see should be continuous. The change of the phase and PA of the IP between 1.4 and 4.9\,GHz is rather abrupt, but this does not rule out birefringence effects, since we have not yet seen the IP at an intermediate frequency (\cite{mh96}). In general, the polarized fraction of other pulsars {\em decreases} with frequency, and the position angle gradient is independent of frequency (\cite{xkj+96}). It is generally believed that pulsar {\em depolarization} toward higher frequencies is due to the instantaneous superposition of emitted orthogonal or quasi-orthogonal polarization modes (\cite{scr+84a}). Our results seems to indicate that one polarization mode dominates the emission from the IP and HFCs for $\nu > 1.4$\,GHz. Following the standard rotating vector model (RVM), the polarization position angle traces the projected magnetic field of the pulsar if emission occurs along the open field lines. The position angle of the RVM is given by Manchester \& Taylor (1977)\nocite{mt77} as \begin{equation} \psi(\phi) = \psi_0 + {\rm tan}^{-1} \left[ {\sin{\alpha} \sin{(\phi - \phi_0)} \over \sin{\zeta}\cos{\alpha} - \cos{\zeta}\sin{\alpha}\cos{(\phi - \phi_0)}} \right] \, , \label{RVM_eqn} \end{equation} where $\psi_0$ is a position angle offset, $\phi_0$ is the pulse phase at which position angle variation is most rapid, $\alpha$ is the inclination angle from the rotation axis to the magnetic axis, and $\zeta$ is the angle between the rotation axis and the observer's line-of-sight. The observer's impact angle with the magnetic axis is just the difference $\beta = \zeta - \alpha$. It can also be determined from the maximum slope of position angle with phase by using \begin{equation} \left[{d\psi \over d\phi}\right]_{\rm max} = {\sin{\alpha} \over \sin{\beta}} \label{impact_eqn} \end{equation} This simple geometric construct is only useful if emission is located close to the polar cap, since the line of sight angle $\zeta$ in the model passes through the center of the star (which is not the case in reality). We have made rudimentary fits of Eq(\ref{RVM_eqn}) to our polarization profiles in an attempt to match polarization signatures to the low altitude dipole model. In Figure \ref{pafit_fig}, we plot the position angles of the Crab's major components at 1.4, 4.9 and 8.4\,GHz with the best fit to the RVM overplotted. The RVM does not fit well for the case where the PA of the IP at all frequencies is left at its 1.4-GHz value, so we have shifted the IP position angle at 1.4\,GHz by $90^\circ$ to match the that at higher frequency. From the fit, the angle found between the rotation and magnetic axes is $\alpha = 56.0^\circ$, with one pole projected near the IP, and the other near the LFC. With $\alpha$ fixed in Eq(\ref{impact_eqn}), the slope of the position angle with phase yields an impact angle, $\beta \simeq 51^\circ$, for the IP. The LFC has a smaller impact angle, $\beta \simeq -30^\circ$, and the maximum slope of the fitted curve occurs just ahead of the LFC. At this location, $\beta \simeq -18^\circ$. The impact angles for the components near the interpulse are much larger than those found from other pulsars (Rankin 1993b\nocite{ran93b}; Lyne \& Manchester 1988\nocite{lm88}). The fitted impact angles are much larger than the polar cap width expected for the Crab pulsar, given by Goldreich \& Julian (1969)\nocite{gj69} is $\rho_{\rm PC} = \left( \frac{2\pi r }{c P} \right)^{-1/2} = 4.56^\circ$, where $\rho_{\rm PC}$ is the width of the polar cap, $r$ is the height above the surface, $c$ is the speed of light, and $P$ is the pulsar's rotational period. So our solution to the RVM fit appears to find emission both close to (LFC) and well away from (IP, HFC1 and HFC2) expected low altitude dipole fields above the polar cap. \placefigure{pafit_fig} Our radio wavelength RVM fit does not agree with values found from fitting the visible wavelength PA sweeps (\cite{nv82}): $\alpha = 86^\circ$, $\beta_{\rm MP} = 9.6^\circ$, and $\beta_{\rm IP} = -18^\circ$. The large inclination angle and the small observer impact angles of this fit imply that both poles sweep by the observer. However, the visible wavelength fits were obtained by only fitting for the maximum sweep through each component individually, not by fitting the position angle over the whole pulsar period. A simple comparison of the visible polarization profiles (\cite{sjd+88}) and the high frequency radio profiles show a few similarities. First, the PA of the MP and IP at 1.4\,GHz matches the visible PA at the phase of the radio components. And though PAs of components at high radio frequency do not match the visible, the position angles and polarized fraction of both the visible and radio profiles increase in the region occupied by HFC1 and HFC2. \section{Emission Geometry} It is difficult to interpret the emission geometry from the profile morphology and polarization measurements we have acquired. There are six sites in rotational phase where pulsed emission occurs, and some evidence of radius to frequency mapping. The HFC components appear to be separated by a width comparable to the conal width expected of a pulsar of this period (Rankin 1993), as do the LFC and MP pair. The precursor may even be a core-type component between the LFC and MP. However, the sweep of position angle through these components is shallow, suggesting that radiation comes from far outside a low-altitude emission cone. So far, our interpretation has followed a simple emission geometry proposed by Smith (1986)\nocite{smi86}, which places the location of emission at both low and high altitudes. The MP and IP are generated in the outer magnetosphere, near the light cylinder, where the dipolar fields are swept back, and the rotational phase of components and their position angles is not the same as above the polar cap. Although no clear evidence of field sweep back has been found for pulsars, if the emission does originate at high altitudes, the swept-back dipole fields of the pulsar would allow the MP and IP to be formed from either the two sides of the same dipole cone above one pole, or from just the leading edges of dipolar fields above both poles of an orthogonal rotator (\cite{sjd+88}). The precursor and LFC are then generated close to the surface of the star above one pole. However, this simplistic model does little to interpret the HFC components, how the IP's properties change, or the nature of the polarization position angle. Another model, proposed by Romani and Yadigaroglu (1995)\nocite{ry95}, ties $\gamma$-ray emission of several pulsars to particle production in an outer magnetospheric gap. Through Monte Carlo simulations of particles in the gap, they have generated $\gamma$-ray profiles similar to the Crab and Vela pulsars, and have successfully generated a polarization position angle profile similar to that of the optical polarization of the Crab, by projecting the magnetic fields (or polarization of high energy photons) in the outer gap. Using this model, it is even possible to find a less powerful outer gap surface that could drive particle acceleration at rotational phases where HFC1 and HFC2 reside (\cite{rom96}). The processes by which radio radiation is generated in the outer magnetosphere are still unknown, though they must be similar to normal pulsar radio production to yield comparable radio power and spectra. Romani and Yadigaroglu (1995)\nocite{ry95} also claim that one should see low altitude emission alongside the outer magnetospheric emission if the orientation of the pulsar allows it. This is true for the Crab pulsar's precursor as well as the Vela pulsar's single radio component, which is offset in phase from its X-ray emission. One last piece of information that may aid in efforts to interpret the polarization, is evidence for the Crab pulsar's orientation on the sky. Using optical images from HST, \cite{hss+95} link certain features found at visible wavelengths with structures found in ROSAT X-ray images. The wisps, arcs, and jet-like features, which probably came from interactions of the Nebula with a pulsar wind, show a cylindrical symmetry, implying that the spin axis of the pulsar is at an angle of $110^\circ$ east of north, projected $\approx 30^\circ$ out of the plane of the sky to the southeast. If the geometry proposed by Hester {\it et al.} is tied to the true spin axis of the pulsar, then the angle of the spin axis to the observer is $\alpha = 90^\circ - 30^\circ = 60^\circ$, very close to our fitted value for the observer impact angle to the spin axis determined through RVM fits. \section{Conclusion} We have presented new polarimetric observations of the Crab pulsar at frequencies between 1.4 and 8.4\,GHz which are difficult to interpret under the classical polar cap model. There are more than the typical number of components seen in other pulsars, and they arise from all over the pulsar's rotational phase. The new pulse components (LFC, HFC1 and HFC2) found in Paper I all have high linear polarization. We re-confirmed the phase shift and spectral change of the IP between 1.4 and 4.9\,GHz, and found that the component also undergoes a $90^\circ$ relative position angle shift with respect to the other components! A good fit is made of the low altitude, rotating-vector model to the polarization position angle at high frequencies, but the line of sight impact angles to the magnetic axis are very large, implying that emission is arising from angles beyond the width of the low altitude polar cap region. It appears that the MP and IP do not arise from low altitude dipole emission. However, the LFC and HFC components show some properties inherent to conal emission (just as the precursor exhibits core-type emission). The Crab profile appears to be associated with a mixture of low and high altitude emission, with the IP being the greatest mystery and exception to the rule. \acknowledgments The authors wish to thank Phil Dooley and the LO/IF group at NRAO-Socorro for their maintenance of the HTRP system. DM acknowledges support from a NRAO pre-doctoral fellowship, and from NSF grants AST 93-15285 and AST 96-18408. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. This research has made use of data from the University of Michigan Radio Astronomy Observatory which is supported by the National Science Foundation and by funds from the University of Michigan. We also acknowledge the use of NASA's Astrophysics Data System Abstract Service (ADS), and the SIMBAD database, operated at CDS, Strasbourg, France.
\section{Introduction} This paper is about lower bounds for certain decision problems over $\mathbb C$. (See ~\cite{BCSS98} for the model of computation and for background). In particular, we will provide lower bounds for the complexity of deciding, given $x$, if $p^d(x)=0$ for some explicit polynomials $p^d$. \par A related problem is to give lower bounds for the evaluation of explicit polynomials. This has been an active subject of research since ~\cite{STRASSEN74}. See~\cite{BCS} for modern developments and for bibliographical remarks. More recent results appeared in~\cite{AHMMP} and~\cite{AM98}. \par Most of those bounds use the Ostrowsky model of computation ~(\cite{BCS} page 6): sum and multiplication by an algebraic constant are free, and the complexity of a computation for polynomial $f(x)$ is the number of non-scalar multiplications, i.e., of multiplications of two polynomials in the variable $x$. For instance, Horner rule for a degree $d$ polynomial requires $d$ non-scalar multiplications. \par All those bounds apply trivially to the complexity of evaluating polynomials by a `machine over $\mathbb C$' as defined in~\cite{BCSS98}, or to the (multiplicative-branching) complexity of a computation tree for evaluating the same polynomial. \par Little is known, however, about the application of those bounds to decision problems (Over $\mathbb C$, in the sense of ~\cite{BCSS98}, or by a decision tree as in~\cite{BCS}, Definition (4.19) page 115. In this definition, each node of a computation tree can perform one algebraic operation or comparison, and therefore a natural measure of complexity is the depth of the tree). \par In this paper, only decision problems of the form below will be considered: let $X \subseteq \mathbb N \times \mathbb C$, and let $X_d = \{ x \in \mathbb C : (d,x) \in X \}$. Typically, $d$ is the problem size and $\# X_d \le d$. One can think of $X$ as the disjoint union of the zero-set of a family of polynomials of degree $\le d$, where $d \in \mathbb N$. The two following forms of a decision problem are natural in this setting: \begin{problem} \label{pnu} For any fixed $d$, decide wether $x \in X_d$. \end{problem} \begin{problem} \label{pu} Decide wether $(d,x) \in X$ \end{problem} \medskip \par Problem~\ref{pnu} is non-uniform, in the sense that we allow for a different machine over $\mathbb C$ or a different decision tree to be used for each value of $d$. However, we want a bound on the running time or on the multiplicative complexity of the tree, as a function of $d$. \par Problem~\ref{pu} is uniform. It is harder than Problem~\ref{pnu}, in the sense that it cannot be solved by a decision tree, since $\# X_d$ can be arbitrarily large. It requires a machine over $\mathbb C$, that will eventually branch according to the value of $d$. \par Lower bounds for Problem~\ref{pnu} are also lower bounds for Problem~\ref{pu}. \par A trivial, topological lower bound for Problems~\ref{pnu} and~\ref{pu} when $\# X_d = d$ is $\log_2 d$. Sharper known bounds come from the `Canonical Path' argument, see ~\cite{BCSS98} section 2.5: Let $f$ be a univariate polynomial. The complexity of deciding $f(x)=0$ is bounded below by the minimum of the complexity of evaluating $g(x)$, where $g$ ranges over the non-zero multiples of $f$. \par If one assumes some property of $f$ that propagates to its multiple $g$, then one eventually obtains a sharper, non-trivial lower bounds. \par In Lemma ~\ref{lower1} below, we will give conditions on the roots of $f$ that will provide lower bounds for the evaluation of $g$. Essentially, we will require a subset of the roots to be rapidly growing. This will imply a rapid growth property for the coefficients of $g$. Then, the results of~\cite{AHMMP, AM98} imply a lower bound for the complexity of evaluating $g$. Thus we will be able to construct specific polynomials that are hard to decide in the {\em non-uniform} sense, viz. \begin{lowerbound} \label{low1} The set $X = \{ (d,x) \in \mathbb Z \times \mathbb C : x = 2^{2^{di}}, 0 \le i \le d$, cannot be solved in time polylog($d$) in the setting of Problem~\ref{pnu}. \end{lowerbound} \begin{lowerbound} \label{low2} The set $Y = \{ (d,x) \in \mathbb Z \times \mathbb C : p^{d}(x)=0 \}$, where $p^d(t) = \sum _{i=0}^d 2^{2^{d (d-i)}} t^i$, cannot be decided in time polylog($d$) in the setting of Problem~\ref{pnu}. \end{lowerbound} \par In a more classical computer-science language, we can define the input size of some $(d,x)$ as $\log d$. This means that the integer $d$ is represented in binary notation, while variable $x$ can contain an arbitrary complex number. In that case, `time polylog($d$) in the setting of Problem~\ref{pnu}' can be refrased as $\ppoly$. The lower bounds above become now: $X \not \in \ppoly$ and $Y \not \in \ppoly$. \par Non-uniform lower bounds ~\ref{low1} and ~\ref{low2} can be compared to the following easier, uniform lower bound: \begin{lowerbound}\label{low3} The set $Z = \{ (d,x) \in \mathbb Z \times \mathbb C : q^d(x)=0 \}$, where $q^d(t) = \sum _{i=0}^d 2^{2^i} t^i$, cannot be decided in time polylog($d$) in the setting of Problem~\ref{pu}. \end{lowerbound} \par This means that the set $Z$, where $d$ is represented in binary notation and $x$ is a complex number, does not belong to $\mathcal P$ over $\mathbb C$. \medskip \par This work was written while the author was visiting the Mathematical Sciences Research Institute, Berkeley, CA. Thanks to Pascal Koiran, Jos{\'e} Luis Monta\~na, Luis Pardo and Steve Smale for their suggestions and comments. \section{Background and notations} \begin{definition} Let $K \subset L$ be finite algebraic extensions of $\mathbb Q$. Let $\nu$ be a valuation in $M_K$. Then we extend the notation $\nu$ to $L$ by: \[ \nu(x) = \frac{ \sum_{\mu} n_{\mu} \mu(x) }{\deg[L:K]} \] where the sum ranges over all the valuations $\mu$ of $L$ that are `above' $\nu$, and where $n_{\mu}$ is the `local degree' of $L:K$. The local degree is defined as $n_{\mu} = \deg[ L_{\mu} : K_{\nu} ]$, where $K_{\nu}$ is the completion of $K$ under the metric induced by the absolute value $|.|_{\nu}$. \end{definition} Recall that for $x \in K$, $\deg[L:K] \nu(x) = \sum_{\mu} n_{\mu} \nu(x)$. The case $K = \mathbb Q$ is an immediate consequence of Corollary 2 of Theorem 1 in Chapter II, p. 39 of \cite{LANG86}. \begin{definition} Let $g$ be a polynomial with algebraic coefficients in some extension $K$ of $\mathbb Q$. Let $\nu$ be a valuation in $M_K$. The {\em Newton diagram} of $g$ at $\nu$ is the (lower) convex hull of the set $\{ (i, \nu(g_i)), i=0 \cdots d \}$. \end{definition} The basic property of Newton diagrams used here is the following. \begin{proposition}\label{prop1} Suppose that $\zeta_1, \cdots, \zeta_d$ are the roots of a univariate polynomial $g \in K[x]$. Let the roots of $g$ be ordered so that \[ \nu(\zeta_1) \ge \cdots \ge \nu(\zeta_d) \] and let the increasing sequence $i_j$ assume the values $0$, $d$ and all the values of $i$ where: \[ \nu(\zeta_i) > \nu(\zeta_{i+1}) \] \par Then the sharp corners of the Newton diagram are precisely the points of the form $(i_j, \nu(g_{i_j}))$ for all $j$. \par Moreover, the slope of the segment $[(i_{j-1}, \nu(g_{i_{j-1}})), (i_{j}, \nu(g_{i_{j}}))]$ is precisely $- \nu(\zeta_{i_{j}})$. \end{proposition} \begin{proof}[Proof of Proposition~\ref{prop1}] The proof uses the following property of valuations: $\nu (\sum x_i) \ge \min \nu (x_i)$. Furthermore, when that minimum is attained in only one $x_i$, we have equality. \par Let $i_{j-1} < k < i_j$. Writing \begin{eqnarray*} g_{i_{j-1}} &=& g_d \sigma_{d-i_{j-1}}(\zeta_1, \cdots, \zeta_d) \\ g_{k} &=& g_d \sigma_{d-k}(\zeta_1, \cdots, \zeta_d) \\ g_{i_{j}} &=& g_d \sigma_{d-i_{j}}(\zeta_1, \cdots, \zeta_d) \\ \end{eqnarray*} \par one can pass to the valuation by: \begin{eqnarray*} \nu(g_{i_{j-1}}) &=& \nu(g_d) + \nu(\zeta_{i_{j-1}+1}) + \cdots + \nu(\zeta_d) \\ \nu(g_{k}) &\ge& \nu(g_d) + \nu(\zeta_{k+1}) + \cdots + \nu(\zeta_d) \\ \nu(g_{i_{j}}) &=& \nu(g_d) + \nu(\zeta_{i_{j}+1}) + \cdots + \nu(\zeta_d) \\ \end{eqnarray*} \par Subtracting, one obtains: \begin{eqnarray*} \nu(g_{i_j}) - \nu(g_{i_{j-1}}) &=& -\nu(\zeta_{i_{j-1}+1}) - \cdots - \nu(\zeta_{i_{j}}) \\ &=& - (i_j - i_{j-1}) \nu(\zeta_{i_j}) \\ \nu(g_{i_j}) - \nu(g_{k}) &\le& -\nu(\zeta_{k+1}) - \cdots - \nu(\zeta_{i_{j}}) \\ &\le& - (i_j - k) \nu(\zeta_{i_j}) \\ \end{eqnarray*} \par This concludes the proof. \end{proof} \medskip \par \section{Uniform lower bounds} We can now prove Lower Bound~\ref{low3}. \begin{proof}[Proof of Lower bound~\ref{low3}] \centerline{\resizebox{10cm}{6cm}{\includegraphics{data1.eps}}} The Newton diagram of $q^{d}$ at 2 is $\{ (i, 2^i): 0 \le i \le d\}$. (This latest set is convex, since the points lie on the curve $y=2^x$ and this curve is convex). Therefore, there is a unique root $\zeta$ of $q^{d}$ that minimizes $\nu(\zeta)$. \par Since $q^{d}_{d-1} = (-\sum \zeta_i) q^{d}_d$, where the sum ranges over all the roots, we have: \[ \nu_2 (q^{d}_{d-1}) = \nu_2 (q^{d}_{d}) + \min \nu_2(\zeta_i) = \nu_2 (q^{d}_{d}) + \nu_2(\zeta) \] Replacing by the actual values of the coefficients, one gets: \begin{equation}\label{eq1} \nu_2(\zeta) = -2^{d-1} \end{equation} \medskip \par Now, suppose that there is a machine $M$ that decides $q^{d}(t) = 0$ in time polylog($d$). One can assume without loss of generality that this machine has no constant but $0$ and $1$. Let its running time be bounded by $T=a (\log d)^b$. \par Let us fix $d > 2+T^2$. We will derive a contradiction. \par Let $g$ be the polynomial defining the canonical path (recall that $d$ is fixed now, so this is the path followed by generic $t \in \mathbb C$). It can be computed in time $\le T^2$, so we have the following bounds: \begin{eqnarray*} \deg g &\le 2^{T^2} \\ 0\le \nu_2 (g_p) &\le 2^{T^2} \end{eqnarray*} \par Since $\zeta$ is also a root of $g$, there are coefficients $g_i$ and $g_j$, $i \ne j$, such that: \begin{equation}\label{eq2} (j-i) \nu_2(\zeta) = \nu_2(g_i) - \nu_2 (g_j) \end{equation} \par Thus, $|\nu_2(\zeta)| \le |\nu_2(g_i)|+|\nu_2(g_i)|$. This implies: \[ |\nu_2(\zeta)| \le {2^{1+T^2}} < 2^{d-1} \] Replacing by equation~\ref{eq1}, one obtains $2^{d-1}<2^{d-1}$, a contradiction. \end{proof} \section{Non-uniform lower bounds} \begin{lemma}\label{lower1} Let $g=g(t)$ be a degree $D$ polynomial with algebraic coefficients. Let $\nu$ be a (non-archimedian) valuation of $K=\mathbb Q[g_0, \cdots, g_D]$. Let $\xi_1, \cdots \xi_D$ be the roots of $g$, and assume they are ordered in such way that: \[ \nu(\xi_1) \ge \cdots \ge \nu(\xi_D) \] Suppose that there is a subsequence $\zeta_j = \xi_{i_j+1}$, $j=1 \cdots d$, such that the following holds: \begin{enumerate} \item $\nu(\zeta_d) \ge 1$ \item $\nu(\zeta_{j}) \ge 2 (i_{j+1} - i_{j}) \ \nu(\zeta_{j+1})$, for $0 \le j \le d-1$. \end{enumerate} \medskip \par Then $g$ cannot be evaluated in less than \[ L \ge \sqrt{ \frac{d}{28 \log_2 D+1} } \] multiplications. \end{lemma} \begin{proof}[Proof of Lemma~\ref{lower1}] We can assume without loss of generality that the ordering of the $\xi_i$ satisfies: \[ \cdots \xi_{i_j} < \xi_{i_j+1} = \zeta_j \le \xi_{i_j +2} \cdots \] For $j \in \{1, \cdots , d-1 \}$ we have: \[ \nu ( g_{i_j} ) - \nu (g_{i_{j+1}} ) = \nu(\xi_{i_j+1}) + \cdots + \nu(\xi_{i_{j+1}}) \] Hence, using $\nu(\xi_{i_{j+1}}) > \nu(\zeta_d) \ge 1$: \[ \nu(\zeta_j) \le \nu ( g_{i_j} ) - \nu (g_{i_{j+1}} ) \le (i_{j+1} - i_{j}) \nu (\zeta_j) \] By the same argument, for $j \in \{0, \cdots , d-2\}$: \[ \nu(\zeta_{j+1}) \le \nu ( g_{i_{j+1}} ) - \nu (g_{i_{j+2}} ) \le (i_{j+2} - i_{j+1}) \nu (\zeta_{j+1}) \] \par Hence, \[ \frac{ \nu ( g_{i_j} ) - \nu (g_{i_{j+1}} ) } { \nu ( g_{i_{j+1}} ) - \nu (g_{i_{j+2}} ) } \ge \frac {\nu(\zeta_{j})} {(i_{j+1} - i_{j}) \nu (\zeta_{j+1})} \ge 2 \] \par Set $G_j = \nu(g_{i_j})$ for $j=0, \cdots, d-1$. We know that the $G_j$ are such that $|G_{j+1} - G_j| < \frac{1}{2} |G_j - G_{j-1}|$. Hence \[ \# \{ \sum s_j G_j, s_j \in \{0;1\} \} = 2^d \] \par Hence: \[ \# \{ \nu (\prod_{s \in S} g_{s}) , S \subset \{0, \cdots, D\} \} \ge 2^d \] \par and hence \[ \mu(g) = \# \{ \sum_{S \subset \{0, \cdots, D\}} \theta_S \prod_{s \in S} g_{s} , \theta_S \in \{0;1\} \} \ge 2^{2^d} \] \par By Lemma~1 in~\cite{AHMMP} or by Lemma 4 in~\cite{AM98}, \[ \mu(g) \le 2^{ (D+1)^{28 L^2} } \] and hence, taking logs: \[ (D+1)^{28 L^2} \ge 2^d \] Taking logs again: \[ 28 L^2 \ge \frac{d}{\log_2 D+1} \] and hence: \[ L \ge \sqrt{ \frac{d}{28 \log_2 D+1} } \] \end{proof} \par Note: Lemma~1 in~\cite{AHMMP} is slightly more general than Lemma~4 in~\cite{AM98}. However, using Lemma~4 in~\cite{AM98} it is possible to replace all the appearances of the number 28 in the statement and proof of Lemma~\ref{lower1} above by the number 21. \par \begin{proof}[Proof of Lower Bound~\ref{low2}] We see from its Netwon diagram that the polynomial $p$ has distinct roots $\zeta_1, \cdots \zeta_d$ with: \[ \nu_2 ( \zeta_i ) = 2^{d (d-i+1)} - 2^{d (d-i)} = 2^{d (d-i)} (2^d - 1) \] \par So we have $\nu_2 (\zeta_d) = 2^d - 1 > 1$, and \begin{equation}\label{rationu} \nu_2(\zeta_i) / \nu_2(\zeta_{i+1}) = 2^d \end{equation} \par Assume that there are $a, b$ such that for each $d$, there is a machine $M$ over $\mathbb C$ deciding $p(t)=0$ in time $T= a (\log d)^b$. Its generic path is defined by a polynomial $g(t)$ of degree $\le 2^T$. \par Let us fix $d > 28 (T+1) T^2$. In particular $d \ge T+1$. We are in the conditions of Lemma~\ref{lower1}, where $D=2^T$. From that Lemma, it follows that \[ T \ge \sqrt{ \frac{d}{28 \log_2 2^T +1} } \ge \sqrt{ \frac{d}{28 (T+1)} } \] Hence, \[ 28 T^2 (T+1) \ge d \] contradicting our choice of $d$. \end{proof} Equation ~(\ref{rationu}) holds trivially in the proof of Lower bound~\ref{low1}. The rest of the proof is verbatim the same. \bibliographystyle{plain}
\section{Introduction} Although numerical methods for the integration of ordinary differential equations (ODEs) based on the use of the matrix exponential have long history, the subject has acquired new relevance recently with two developments. The first, which is irrelevant to the theme of this paper, is the introduction of Krylov subspace techniques and their application to large stiff systems of differential equations \cite{hochbruck98eif}. The other development is motivated by the philosophy of \textit{geometric integration\/} and its purpose is to recover under discretization important qualitative and geometric features of the underlying dynamical system. Examples of such methods can be found {\em inter alia\/} in \cite{casas96ffa,crouch93nio}. An important technique in geometric integration is the use of Lie-group actions, which lend themselves to the design of very effective time-stepping methods for ODEs evolving on homogeneous manifolds. Such methods have been recently studied in \cite{munthe-kaas97hor} and \cite{engo98otc}. Methods based on the use of the classical Magnus and Fer expansions for integrating ODEs on Lie-groups can be brought into this formalism \cite{iserles97loi,iserles99ots,zanna97car}. All such methods require a repeated evaluation of a matrix exponential, often of large matrices. Inasmuch as typically one can expect the replacement of the exact exponential by a suitable approximant (a rational function, say, a Krylov subspace approximant or a Schur factorization), the context of Lie-group methods imposes a crucial extra requirement. The approximant in question, applied to an arbitrary element of the Lie algebra $\GG{g}$, must produce an outcome in the Lie group $G$, otherwise the whole purpose of the calculation, dicretizing within $G$, will be null and void. This can be done is {\em some,\/} but by no means, all Lie algebras of interest and we refer the reader to \cite{celledoni98atm} for a more substantive discussion of this issue. Let $G$ be a finite-dimensional Lie group. For all practical purposes, we may assume that $G$ is a subgroup of the {\em general linear group\/} $\CC{GL}(n)$, the set of all nonsingular $n\times n$ matrices. We denote by $\GG{g}$ the Lie algebra corresponding to $G$, observing that it is a subalgebra of $\GG{gl}(n)$, the Lie algebra of all $n\times n$ matrices. Our concern in this paper is with differential equations that evolve on a manifold ${\cal M}$ subject to the action of $\CC{G}$. For simplicity we can assume that ${\cal M}$ coincides with $\CC{G}$ and the action is of $\CC{G}$ on itself. The numerical solution of such differential equations can be obtained considering the {\em pull-back\/} on $\GG{g}$, by means of the exponential map, of the vectorfield defining the equation. We can compute the corresponding flow by a Lie-algebra discretization method and recover the approximation of the original problem via exponentiation. Given an integration method of order $p$, we consider order-$p$ approximants $F(tB)$ for $\exp (tB)$, where $B\in\GG{g}$ and $t\geq0$. We require that $F(tB)\in \CC{G}$, whence it is easy to prove that important qualitative features of the original equation and the order of the discretization are retained. In \cite{celledoni98atm} we have introduced low-rank splitting methods for the construction of the approximant $F$, as the first attempt to provide a comprehensive treatment of this issue. Although the constraint $F(tB)\in \CC{G}$ represents remarkable advantage in many applications, such as problems in which the conservation of invariants is at issue in numerical modelling (volume conservation in meteorology, invariance under rotations in the theory of mechanical systems and in robotics), it should not be interpreted as the sole purpose of our analysis. Our methods are relevant also for the approximation of $\exp(tB)$ in the more general setting $B\in \GG{gl}(n)$. Suppose in fact that $B\in \GG{gl}(n)$ and we want to approximate $\exp(tB)$. It is always possible to write $B$ as a sum of a matrix $B_s\in\GG{sl}(n)$ (the {\em special linear algebra\/} of $n\times n$ matrices with zero trace) and a diagonal matrix $B_d$ whose nonzero entries are equal to $\delta=\CC{tr}\,(B)/n$. Then $B_s=B-B_d$ and $[B_s ,B_d ]=0$ so that $\exp(tB)=\exp(tB_s)\exp(t\delta)$. This fact is a particular case of what is known in Lie theory as the Levi decomposition \cite{humphreys72itl,varadarajan84lgl}. Using this decomposition of the matrix $B$, if necessary in tandem with some scaling and squaring technique, the approximation of $\exp(tB)$ can be always reduced to the approximation of $\exp(tB_s)$ with $B_s \in \GG{sl}(n)$. As long as we can assure that our approximation of $\exp(tB_s)$ resides in $\CC{SL}(n)$, the outcome is an approximant $F(tB)$ of $\exp(tB)$ that shares with the exact exponential the feature that $\det F(tB)=\exp (\CC{tr}\, B)$. It is possible to prove that, given a splitting $B=\sum_{i=1}^{k}B_i$, the function \begin{displaymath} \CC{e}^{\frac12 tB_1}\CC{e}^{\frac12 tB_2}\cdots \CC{e}^{\frac12 tB_{k-1}} \CC{e}^{tB_k}\CC{e}^{\frac12 tB_{k-1}}\cdots \CC{e}^{\frac12 tB_2} \CC{e}^{\frac12 tB_1}, \end{displaymath} known as the generalized {\em Strang splitting,\/} approximates $\exp(tB)$ to order 2. As long as $B_1,B_2,\ldots,B_k\in\GG{g}$, it follows at once from the definition of a Lie group that the approximant resides in $G$. Moreover, $2k-1$ is the least number of exponentials that render such a splitting into a second-order approximant \cite{celledoni98atm}. The Strang splitting is time reversible, hence it follows readily from classical theory that the order can be raised from 2 to 4 by composing three Strang splittings with different time steps \cite{yoshida90coh}. In that case we need to evaluate $3k$ exponentials and multiply $6k$ matrices. In the case of low-rank splittings which have been considered by \citeasnoun{celledoni98atm} this results in the following count of flops: $4n^3$ for order $2$, $12n^3$ for order $4$. In this paper we present composition methods in which the number of exponentials $k$ equals the dimension $d$ of the Lie algebra. Our construction allows us to increase the order of the approximation without increasing the number of exponentials to evaluate and multiply together. Letting $\{V_1,\dots, V_d\}$ be a basis of $\GG{g}$, we write $F(tB)$ as a composition of exponentials of the type $\exp(\alpha_i(t)V_i)$, where each $\alpha_i(t)$ for $i=1,\dots ,d$ is a scalar function. In general $d=\O{ n^2}$, however, with an appropriate choice of the basis elements, the computation of each exponential $\exp(\alpha_i(t)V_i)$ requires $\O{n}$ flops, while the formation of their product adds just $2n^3+\O{n^2}$ flops. The challenging part of the computation is the construction of the functions $\alpha_1(t) ,\dots ,\alpha_d(t)$, and the cost of their calculation depends on the desired order of the approximation. Naive complexity analysis might have indicated that the total cost is growing exponentially in $d$ as the order increases. Yet, the cost remains relatively modest for small orders and the method lends itself very well to the exploitation of sparsity in the matrix $B$. In the sequel we show how this approach can be turned into an efficient numerical method and we obtain algorithms of order up to $4$ with a cost of $\O{n^3}$ for dense matrices. Our approach can be interpreted as representing the solution using {\em canonical coordinates of the second kind,\/} an approach that has been pioneered by \citeasnoun{owren98imb} in the context of general Lie-group methods. Having said this, the more restrictive framework of exponential approximants possesses a very great deal of special structure. This can be exploited so as to produce efficient and competitive algorithms that approximate $\exp(tB)$, $B\in\GG{g}$, in the Lie group $G$. \section{The technique of coordinates of second kind for the approximation of the exponential matrix} Let $\CC{G}$ be a Lie group and $\GG{g}$ its corresponding $d$-dimensional Lie algebra. We choose a basis $\{V_1,\dots, V_d\}$ of $\GG{g}$, whence every element $Y\in\CC{G}$ sufficiently close to the identity can be represented in a unique fashion as \begin{displaymath} Y=\exp(\gamma_1V_1) \exp(\gamma_2V_2)\cdots \exp(\gamma_dV_d), \end{displaymath} where $\exp : \GG{g}\rightarrow \CC{G}$ is the exponential map. This representation is known as representation in canonical coordinates of the second kind \cite{varadarajan84lgl}. This representation is global in the case of solvable Lie algebras. We restrict ourselves to the case $\GG{g}\subseteq\GG{gl}(n)$, $\CC{G}\subseteq\CC{GL}(n)$, when $\exp$ is the usual matrix exponential. Given $B\in\GG{g}$, we can represent it in a unique fashion as \begin{displaymath} B=\sum_{i=1}^{d}\beta_i V_i. \end{displaymath} It is possible then to write $\exp(tB)$ in the form \begin{displaymath} U(t)=\exp(tB)=\exp(g_1(t)V_1) \exp(g_2(t)V_2)\cdots \exp(g_d(t)V_d). \end{displaymath} Letting $\MM{g}=[g_1,\dots,g_d]^{\CC{T}}$, $\MM{\beta}=[\beta_1,\dots,\beta_d]^{\CC{T}},$ it can be proved that the vector function $\MM{g}$ obeys a differential equation of the form \begin{displaymath} \frac{d\MM{g}}{dt}=\MM{f}(\MM{\beta},\MM{g}), \qquad \MM{g}(0)=0, \end{displaymath} where $\MM{f}$ is a suitable function of $\MM{\beta}$ and $\MM{g}$, for sufficiently small $t$ \cite{wei63ogr}. Given a solvable Lie algebra $\GG{g}$ \citeasnoun{wei63ogr} prove results on the global representation of $U$. However, an explicit form of $\MM{f}$ is known only for very simple examples of low-dimensional Lie algebras. In this paper we seek polynomials ${\alpha}_1\approx g_1,\dots,{\alpha}_d\approx g_d$ of a suitable degree so that \begin{displaymath} \exp(tB)\approx \exp({\alpha}_1(t)V_1) \exp({\alpha}_3(t)V_2) \cdots \exp({\alpha}_d(t)V_d). \end{displaymath} Differentiation yields \begin{equation} \label{eq:2.1} \sum_{i=1}^{d}\beta_i V_i=\sum_{i=1}^{d}g_i'(t)\prod_{j=1}^{i-1}\CC{e}^{g_jV_j}V_i \prod_{j=i-1}^{1}\CC{e}^{-g_jV_j}. \end{equation} Evaluating this expression at the origin gives the first-order condition \begin{equation} \label{eq:2.2} g'_i(0)=\beta_i,\qquad i=1,2,\ldots,d. \end{equation} Further differentiations of \R{eq:2.1} lead to higher-order conditions. Let us define the functions \begin{equation} \label{eq:2.3} P_i(\MM{g})=\exp(\CC{ad}_{g_1V_1})\circ \cdots\circ\exp(\CC{ad}_{g_{i-1}V_{i-1}})(V_i),\qquad i=1,2,\ldots,d, \end{equation} where the {\em adjoint operator\/} $\CC{ad}_{x}:\GG{g}\rightarrow \GG{g}$ is defined as $\CC{ad}_{x}(y)=[x,y]$ for any $x,y \in\GG{g}$, $[x,y]=xy-yx$ being the matrix commutator. Note that $P_i(\MM{g}(0))=V_i$, $i=1,2,\ldots,d$. Moreover, the right-hand side of \R{eq:2.1} can be written in the simplified form The function \begin{displaymath} {\cal T}(\MM{g})=\sum_{i=1}^{d}g_i'(t)P_i(g). \end{displaymath} Since the derivatives of the left hand side of \R{eq:2.1} vanish, the conditions for order $p\geq 1$, can be obtained by solving the equations \begin{equation} \label{eq:2.4} \left.\frac{\CC{d}^r}{\CC{d}t^r}{\cal T}(\MM{g})\right|_{\,t=0}=0, \qquad r=1,2,\ldots , p-1, \quad p\geq 1, \end{equation} where \begin{equation} \label{eq:2.5} \frac{\CC{d}^r}{\CC{d}t^r}{\cal T}(\MM{g})=\sum_{i=1}^{d}\sum_{k=1}^{r}{r\choose k} \frac{\CC{d}^{r-k+1}g_i} {\CC{d}t^{r-k+1}}\frac{\CC{d}^{k}P_i}{\CC{d}t^{k}}. \end{equation} In particular, \begin{Eqnarray*} \frac{\CC{d}}{\CC{d}t}{\cal T}(\MM{g})&=&\sum_{i=1}^{d} \left(g_i''P_i +g_i'\frac{d}{dt}P_i\right),\\ \frac{\CC{d}^2}{\CC{d}t^2}{\cal T}(\MM{g})&=&\sum_{i=1}^{d} \left( g_i'''P_i+2g_i''\frac{\CC{d}}{\CC{d}t}P_i+g_i'\frac{\CC{d}^2} {\CC{d}t^2}P_i\right) ,\\ \frac{\CC{d}^3}{\CC{d}t^3}{\cal T}(\MM{g})&=&\sum_{i=1}^{d} \left( g_i^{IV}P_i+3g_i'''\frac{\CC{d}}{\CC{d}t}P_i+3g_i'' \frac{\CC{d}^2}{\CC{d}t^2}P_i+g_i'\frac{\CC{d}^3}{\CC{d}t^3}P_i\right). \end{Eqnarray*} Solving \R{eq:2.4} for $r=1$ results in the values of $g_i''(0)$ for $i=1,2,\ldots ,d$ that allow us to construct an order-$2$ approximant. Substituting such values in \R{eq:2.4} for $r=2$ yields $g_i'''(0)$ for $i=1,2,\ldots ,d$ and consequently an approximant of order $3$. Similar procedure can be used to construct recursively approximants of arbitrarily high order. The main part of the computation is the evaluation of the $k$-th derivative of $P_i(\MM{g})$ at $t=0$. Expanding the exponentials in \R{eq:2.3} we obtain \begin{displaymath} P_i(\MM{g})=\prod_{k=1}^{i-1} (I+\CC{ad}_{g_kV_k}+\Frac12 \CC{ad}^2_{g_kV_k}+\Frac16 \CC{ad}^3_{g_kV_k}+\ldots )(V_i) \end{displaymath} and, after further algebra, \begin{Eqnarray*} P_i(\MM{g})&=&\left\{I+\sum_{k=1}^{i-1}\CC{ad}_{g_kV_k}+ \sum_{k=2}^{i-1} \sum_{l=1}^{k-1}\CC{ad}_{g_lV_l}\CC{ad}_{g_kV_k}\right.\\ &&\mbox{}+\Frac12\sum_{k=1}^{i-1}\CC{ad}^2_{g_kV_k}+ \sum_{k=3}^{i-1}\sum_{l=2}^{k-1}\sum_{j=1}^{l-1}\CC{ad}_{g_jV_j} \CC{ad}_{g_lV_l}\CC{ad}_{g_kV_k}\\ &&\left.\mbox{}+\Frac12\sum_{k=2}^{i-1}\sum_{l=1}^{k-1} \left(\CC{ad}_{g_lV_l}\CC{ad}^2_{g_kV_k}+\CC{ad}^2_{g_lV_l} \CC{ad}_{g_kV_k}\right)+\Frac16\sum_{k=1}^{i-1}\CC{ad}^3_{g_k V_k}+\dots \right\}\left(V_i\right). \end{Eqnarray*} Similarly to \cite{owren97rkm}, we write $P_i(\MM{g})$ in the form \begin{displaymath} P_i(\MM{g})=I+\sum_{r=1}^{\infty}\sum_{j_1=1}^{i-1} \sum_{j_2=j_1}^{i-1}\cdots \sum_{j_r=j_{r-1}}^{i-1} \frac{1}{\MM{j}!}g_{j_1}\dots g_{j_r}\CC{ad}_{V_{j_1}}\circ\cdots\circ\CC{ad}_{V_{j_r}}(V_i). \end{displaymath} Here $\MM{j}=(j_1,\dots j_r)$ is a multi-index of integer elements with $1\leq j_r\le i-1$ and $\MM{j}!:=q_1!q_2!\dots q_{i-1}!$ where $q_k$ is the number of occurrences of $k$ in $(j_1,j_2\ldots j_r)$. A general expression for the $k$-th derivative of $P_i$ is given as follows: since $g_i(0)=0$, we may let $f_i(t)=g_i(t)/t$, $i=1,2,\ldots,d$. We can then rewrite $P_i$ in the form \begin{displaymath} P_i(\MM{g})= I+\sum_{r=1}^{\infty}t^r\sum_{j_1=1}^{i-1}\sum_{j_2=j_1}^{i-1}\cdots \sum_{j_r=j_{r-1}}^{i-1}\frac{1}{\MM{j}!}f_{j_1}\cdots f_{j_r}\CC{ad}_{V_{j_1}}\circ\cdots\circ\CC{ad}_{V_{j_r}}(V_i). \end{displaymath} By following the construction in\cite{owren97rkm} we obtain \begin{equation} \label{eq:2.6} \begin{meqn} \displaystyle \left. \frac{\CC{d}^kP_i}{\CC{d}t^k}\right|_{t=0}&=& \displaystyle \sum_{r=1}^{k} \sum_{\delta_1+\dots+\delta_r=k} \frac{k!}{\prod_{\nu =1}^{\mu}(\delta_{\nu}-1)!}\sum_{1\le j_1\le\dots\le j_{\mu}\le i-1}\frac{1}{\MM{j}!}\\[18pt] &&\left.\displaystyle \mbox{}\times f_{j_1}^{(\delta_1-1)}\cdots f_{j_{\mu}}^{(\delta_{\mu}-1)}\right|_{t=0} \CC{ad}_{V_{j_1}}\circ\cdots\circ \CC{ad}_{V_{j_{\mu}}}(V_i). \end{meqn} \end{equation} Substituting \R{eq:2.6} in \R{eq:2.4} and \R{eq:2.5} we obtain the conditions for arbitrary order $p$. In particular we obtain the following formulae for the derivatives of $P_i(\MM{g})$ at $t=0$, \begin{Eqnarray*} \left.\frac{\CC{d}P_i}{\CC{d}t}\right|_{t=0}&=&\sum_{k=1}^{i-1} \CC{ad}_{V_k}\left(V_i\right)g_k'(0),\\ \left. \frac{\CC{d}^2P_i}{\CC{d}t^2}\right|_{t=0}&=&\sum_{k=1}^{i-1} \left(\sum_{l=1}^{k-1}2\CC{ad}_{V_l}\CC{ad}_{V_k}(V_i)g_k'(0) g_l'(0)+\CC{ad}^2_{V_k}(V_i)[g_k'(0)]^2+\CC{ad}_{V_k}(V_i)g_k''(0)\right),\\ \left. \frac{\CC{d}^3P_i}{\CC{d}t^3}\right|_{t=0}&=& 6\sum_{k=3}^{i-1}\sum_{l=2}^{k-1}\sum_{j=1}^{l-1}\CC{ad}_{V_j} \CC{ad}_{V_l}\CC{ad}_{V_k}(V_i)g_j'(0)g_l'(0)g_k'(0)\\ &&\mbox{}+3\sum_{k=2}^{i-1}\sum_{l=1}^{k-1}\left(\CC{ad}_{V_l} \CC{ad}^2_{V_k}(V_i)g_l'(0)[g_k'(0)]^2+\CC{ad}^2_{V_l}\CC{ad}_{V_k} (V_i) [g_l'(0)]^2g_k'(0)\right)\\ &&\mbox{}+\sum_{k=1}^{i-1}\CC{ad}_{V_k}^3(V_i)[g_k'(0)]^3\\ &&\mbox{}+3\sum_{k=2}^{i-1}\sum_{l=1}^{k-1}\CC{ad}_{V_l} \CC{ad}_{V_k} (V_i)\left(g_k''(0)g_l'(0)+g_k'(0)g_l''(0)\right)\\ &&\mbox{}+3\sum_{k=1}^{i-1}\CC{ad}^2_{V_k}(V_i)g_k''(0)g_k'(0)\\ &&\mbox{}+\sum_{k=1}^{i-1}\CC{ad}_{V_k}(V_i)g_k'''(0). \end{Eqnarray*} Substitution readily produces order conditions. Specifically, \begin{equation} \label{eq:2.7} \sum_{i=1}^{d} g_i''(0)V_i=-\sum_{i=1}^{d}g_i'(0)\sum_{k=1}^{i-1}\CC{ad}_{V_k}(V_i)g_k'(0), \end{equation} are conditions for order $2$, while \begin{Eqnarray} \sum_{i=1}^{d} g_i'''(0)V_i&=& - \sum_{i=1}^{d}\left\{ 2g_i''(0)\sum_{k=1}^{i-1}\CC{ad}_{V_k}(V_i)g_k'(0)\right. \nonumber\\ &&\mbox{}+g_i'(0)\sum_{k=1}^{i-1}\left[2\sum_{l=1}^{k-1} \CC{ad}_{V_l}\CC{ad}_{V_k}(V_i)g_k'(0)g_l'(0)\right.\label{eq:2.8}\\ &&\left.\left.\mbox{}+\CC{ad}^2_{V_k}(V_i)(g_k'(0))^2+\CC{ad}_{V_k} (V_i)g_k''(0)\right]\right\},\nonumber \end{Eqnarray} are the order-3 conditions. Finally, conditions for order 4 are \begin{Eqnarray} \sum_{i=1}^{d} g_i^{\CC{IV}}(0)V_i&=&- \sum_{i=1}^{d}\left\{ 3g_i'''(0)\sum_{k=1}^{i-1}\CC{ad}_{V_k}(V_i)g_k'(0)+\right. \nonumber\\ &&\mbox{}+3g_i''(0)\left[\sum_{k=1}^{i-1}\left(\sum_{l=1}^{k-1} 2\CC{ad}_{V_l}\CC{ad}_{V_k}(V_i)g_k'(0)g_l'(0)\right.\right.\nonumber\\ &&\left.\left.\mbox{}+ \CC{ad}^2_{V_k}(V_i)(g_k'(0))^2+\CC{ad}_{V_k} (V_i)g_k''(0)\right)\right]\nonumber\\ &&\mbox{}+g_i'(0)\left[6\sum_{k=3}^{i-1}\sum_{l=2}^{k-1}\sum_{j=1}^{l-1} \CC{ad}_{V_j}\CC{ad}_{V_l}\CC{ad}_{V_k}(V_i)g_j'(0)g_l'(0)g_k'(0)\right. \nonumber\\ &&\mbox{}+3\sum_{k=2}^{i-1}\sum_{l=1}^{k-1}\left(\CC{ad}_{V_l} \CC{ad}^2_{V_k}(V_i)g_l'(0)(g_k'(0))^2+\CC{ad}^2_{V_l}\CC{ad}_{V_k} (V_i)(g_l'(0))^2g_k'(0)\right)\label{eq:2.9}\\ &&\mbox{}+\sum_{k=1}^{i-1}\CC{ad}_{V_k}^3(V_i)(g_k'(0))^3\nonumber\\ &&\mbox{}+3\sum_{k=2}^{i-1}\sum_{l=1}^{k-1}\CC{ad}_{V_l}\CC{ad}_{V_k} (V_i)\left(g_k''(0)g_l'(0)+g_k'(0)g_l''(0)\right)\nonumber\\ &&\mbox{}+3\sum_{k=1}^{i-1}\CC{ad}^2_{V_k}(V_i)g_k''(0)g_k'(0) \nonumber\\ &&\mbox{}+\left.\left. \sum_{k=1}^{i-1}\CC{ad}_{V_k}(V_i)g_k'''(0) \right]\right\}. \nonumber \end{Eqnarray} In Figure \ref{fig:1} we have plotted along the $y$ axis the $2$-norm of the error of the approximation of $\exp(tB)$ with the second-kind coordinates (SKC) methods of order ranging from $1$ to $4$. The values of the error are plotted against time, (along the $x$-axis), to logarithmic scale for matrices of $\GG{sl}(5)$. The methods have been implemented using the standard basis defined in section 3. \begin{figure}[tb] \begin{center} \leavevmode \psfig{file=skc5.eps,width=8cm} \caption{Error in the approximation of the exponential with WN technique.} \label{fig:1} \end{center} \end{figure} The computation of $g''(0)$, $g'''(0)$ and $g^{\CC{IV}}(0)$ is obtained directly implementing the formulas \R{eq:2.7}, \R{eq:2.8},\R{eq:2.9} respectively. This implementation does not depend on the choice of the particular basis of $\GG{sl}(5)$, but the number of commutators that must be computed with this approach is $\O{d^p}$ for $p=2,3,4$. Even if we assume that the $V_i$s are very sparse matrices and that the cost of computing each commutator is $\O{1}$ operations, the total cost exceeds $\O{n^{2p}}$ flops for $p=2,3,4$ where $n$ is the dimension of the matrix. Such expense is not acceptable for a competitive method of approximation of $\exp(tB)$. Fortunately, it can be decreased a very great deal by an appropriate choice of the basis $\{V_1,V_2,\ldots,V_d\}$. This is the theme of the next section. \setcounter{equation}{0} \section{Choosing a basis} The choice of the right basis and sparse representation of commutators are critical to the implementation of the SKC methods. Recalling the order conditions \R{eq:2.7}--\R{eq:2.9}, our aim is to choose a basis so that terms of the form $\CC{ad}_{V_{i_1}}\CC{ad}_{V_{i_2}} \cdots \CC{ad}_{V_{i_s}} V_j$ can be represented in the most economical manner. We recall that, given the basis $\{V_1,V_2,\ldots,V_d\}$ of a $d$-dimensional Lie algebra $\GG{g}$, the {\em structure constants\/} are the numbers $c_{k,l}^i$, $k,l,i=1,2,\ldots,d$, such that \begin{displaymath} [V_k,V_l]=\sum_{i=1}^d c_{k,l}^i V_i \end{displaymath} \cite{humphreys72itl}. Let \begin{displaymath} B=\sum_{k=1}^d \beta_k V_k. \end{displaymath} Then an order-1 condition is always \begin{equation} \label{eq:3.1} g_k'(0)=\beta_k,\qquad k=1,2,\ldots,d. \end{equation} To obtain the order-2 condition we substitute \R{eq:3.1} in \R{eq:2.7} and express commutators in terms of structure constants, \begin{Eqnarray*} \sum_{k=1}^d g_k''(0)V_k&=&-\sum_{l=1}^d \beta_l \sum_{j=1}^{l-1} [V_j,V_l] \beta_j=-\sum_{l=1}^d \beta_l \sum_{j=1}^{l-1} \sum_{k=1}^d c_{j,l}^k V_k\\ &=&-\sum_{k=1}^d \left( \sum_{l=1}^d \sum_{j=1}^{l-1} \beta_l c_{j,l}^k \beta_j \right) V_k. \end{Eqnarray*} Since $c_{j,l}^k=-c_{l,j}^k$, we thus deduce that \begin{equation} \label{eq:3.2} g_k''(0)=\sum_{l=1}^d \sum_{j=1}^{l-1} \beta_l c_{l,j}^k \beta_j,\qquad k=1,2,\ldots,d. \end{equation} Likewise, substituting in \R{eq:2.8}, \begin{Eqnarray*} \sum_{k=1}^d g_k'''(0)V_k&=&-\sum_{i=1}^d \left\{ 2g_i''(0)\sum_{l=1}^{i-1} [V_l,V_i]\beta_l +\beta_i \sum_{l=1}^{i-1} \left(2\sum_{j=1}^{l-1} [V_j,[V_l,V_i]] \beta_l\beta_j \right.\right.\\ &&\left.\left.\mbox{}+[V_l,[V_l,V_i]]\beta_l^2 +[V_l,V_i]g_l''(0)\right) \right\}. \end{Eqnarray*} Note that \begin{displaymath} [V_j,[V_l,V_i]]=\sum_{s=1}^d c_{l,i}^s [V_j,V_s] =\sum_{k=1}^d \sum_{s=1}^d c_{l,i}^s c_{j,s}^k V_k. \end{displaymath} Therefore \begin{Eqnarray*} \sum_{k=1}^d g_k'''(0)V_k&=&-2\sum_{i=1}^d g_i''(0) \sum_{l=1}^{i-1} \sum_{k=1}^d c_{l,i}^k \beta_l V_k-2\sum_{i=1}^d \beta_i \sum_{l=1}^{i-1} \sum_{j=1}^{l-1} \sum_{k=1}^d \sum_{s=1}^d c_{l,i}^s c_{j,s}^k \beta_l \beta_j V_k\\ &&\mbox{}-\sum_{i=1}^d \beta_i \sum_{l=1}^{i-1} \sum_{k=1}^d \sum_{s=1}^d c_{l,i}^s c_{l,s}^k \beta_l^2 V_k-\sum_{i=1}^d \sum_{l=1}^{i-1} \beta_i \sum_{k=1}^d c_{l,i}^k g_l''(0)V_k \end{Eqnarray*} and we deduce that \begin{Eqnarray*} g_k'''(0)&=&\sum_{i=1}^d \sum_{l=1}^{i-1} c_{i,l}^k [2g_i''(0)\beta_l +\beta_i g_l''(0)]+ 2\sum_{i=1}^d \sum_{l=1}^{i-1} \sum_{j=1}^{l-1} \sum_{s=1}^d c_{i,l}^s c_{j,s}^k \beta_i\beta_l \beta_j\\ &&\mbox{}-\sum_{i=1}^d\sum_{i=1}^d \sum_{l=1}^{i-1} \sum_{s=1}^d c_{i,l}^s c_{l,s}^k \beta_i \beta_l^2,\qquad k=1,2,\ldots,d. \end{Eqnarray*} Bearing in mind that for order $p$ we require \begin{displaymath} \alpha_k(t)=\sum_{r=1}^p \frac{1}{r!} g_k^{(r)}(0)t^r,\qquad k=1,2,\ldots,d, \end{displaymath} we observe that the sheer volume of calculations required for the evaluation of the functions $\alpha_1,\alpha_2,\ldots,\alpha_d$ is prohibitive for, say, order 3, unless most of the structure constants vanish. Fortunately, bases of finite-dimensional Lie algebras which are `sparse' (in the sense that a very high proportion of structure constants vanish) are known. They are associated with {\em root space decompositions\/} of Lie algebras \cite{humphreys72itl} and, in the case of semisimple algebras, are known as {\em Chevalley bases\/} \cite{carter95llg}. Wishing to avoid too much Lie-algebraic terminology in a numerical analysis paper, we reserve our exposition to just three examples which are the most important in a range applications. \vspace{8pt} \noindent {\bf The orthogonal group} Let $\GG{g}=\GG{so}(n)$, the Lie algebra of $n\times n$ skew-symmetric matrices. It corresponds to two important Lie groups: the orthogonal group $\CC{O}(n)$ of $n\times n$ orthogonal matrices and its subgroup, the special orthogonal group $\CC{SO}(n)$ of matrices with unit determinant. Its dimension is $d=\frac12n(n-1)$. We let \begin{displaymath} F_{i,j}=\MM{e}_i{\MM{e}_j}^{\CC{T}}-\MM{e}_j{\MM{e}_i}^{\CC{T}}, \qquad i=1,2,\ldots,n, \quad j=i+1,i+2,\ldots ,n, \end{displaymath} where $\MM{e}_i$ is the $i$-th canonical vector of $\BB{R}^n$. In other words, $F_{i,j}$ is a matrix whose $(i,j)$-th element is $1$, the $(j,i)$-th element equals $-1$ and zero otherwise. We can trivially expand each $B\in \GG{so}(n)$ as $B=\sum_{i=1}^{n}\sum_{j=i+1}^nb_{i,j}F_{i,j}$. $U(t):=\exp(tF_{i,j})$ is simply an {\em Euler rotation\/} in the $(i,j)$ plane: it is identity matrix, except that \begin{displaymath} \left[ \begin{array}{cc} U_{i,i} & U_{i,j}\\ U_{j,i} & U_{j,j} \end{array}\right]=\left[ \begin{array}{rr} \cos(tF_{i,j}) & \sin(tF_{i,j})\\ -\sin(tF_{i,j}) & \cos(tF_{i,j}) \end{array}\right]. \end{displaymath} Noting that \begin{displaymath} [F_{i,j} , F_{l,k} ] = \begin{case} -F_{j,k}, & i=l,\; j\neq k,\\ -F_{i,l}, & i\neq l,\; j=k, \\ F_{i,k}, & i\neq k,\; j=l, \\ F_{j,l}, & i=k,\; j\neq l, \\ O, & \CC{otherwise,} \end{case} \end{displaymath} the order conditions are simplified as follows, \begin{Eqnarray*} p\ge 1:\qquad g'_{i,j}(0)&=&b_{i,j},\\ p\ge 2:\qquad g''_{i,j}(0)&=& \sum_{s=j+1}^nb_{j,s}b_{i,s}-\sum_{r=i+1}^{j-1}b_{r,j}b_{i,r}\\ &&\mbox{}+\sum_{r=1}^{i-1}b_{r,j}b_{r,i},\qquad 1\leq i<j\leq n, \end{Eqnarray*} and similarly for higher-order terms. Thus, the cost of computing the coefficients for the second-order method is just $\frac12(n-2)(n-1)n \approx\frac12 n^3$ flops. In comparison, a naive computation of \R{eq:3.2}, without exploiting sparsity of structure constants, requires $\frac18 (n^2-n-2)(n-1)^2n^2\approx \frac18 n^6$ flops. A more classical composition method for $B\in\GG{so}(n)$ is the \textit{Strang splitting\/} which we can write in the form \begin{displaymath} \CC{e}^{tb_{1,2}F_{1,2}/2}\cdots \CC{e}^{tb_{n-2,n}F_{n-2,n}/2} \CC{e}^{tb_{n-1,n}F{n-1,n}} \CC{e}^{tb_{n-2,n}F_{n-2,n}/2}\cdots \CC{e}^{tb_{1,2}F_{1,2}/2} \end{displaymath} \cite{celledoni98atm}. It gives a second-order approximant to $\exp(tB)$ whose calculation requires $\approx 4n^3$ flops, in comparison with $\approx 3n^3$ for the second-order CSK method. We note that, in the specific case of $\GG{so}(n)$, diagonal Pad\'e approximants to the exponential provide an alternative to our method, since they map the algebra to $\CC{O}(n)$. Having said this, for dense matrices $B$ the cost of evaluating the second-order approximant $(I-\frac12 tB)^{-1}(I+\frac12 tB)$ with, say, LU factorization is $\O{n^3}$, comparative with our method. \vspace{8pt} \noindent {\bf The special linear group} Let $\GG{g}=\GG{sl}(n)$, the set of $n\times n$ matrices with zero trace, whence $d=n^2-1$. We split the algebra in the first instance into diagonal and off-diagonal parts: in the terminology of Lie algebras, the subspace spanned by the diagonal elements is a {\em Cartan subalgebra\/} \cite{carter95llg} or {\em maximal toral algebra\/} \cite{humphreys72itl} of $\GG{sl}(n)$. Specifically, our basis is \begin{displaymath} \{E_{i,j}\,:\, i,j=1,2,\ldots,n\; i\neq j\}\cup \{D_i\,:\, i=1,2,\ldots,n-1\}. \end{displaymath} where \begin{Eqnarray*} E_{i,j}&=&\MM{e}_i\MM{e_j}^{\CC{T}}, \qquad i,j=1,2,\ldots , n,\quad i\neq j,\\ D_i&=&\MM{e}_i\MM{e_i}^{\CC{T}}-\MM{e}_{i+1}\MM{e}_{i+1}^{\CC{T}}, \qquad i=1,2,\ldots,n-1. \end{Eqnarray*} The exponentials of $E_{i,j}$ and $D_i$ are trivial, \begin{displaymath} \CC{e}^{tE_{i,j}}=I+tE_{i,j},\qquad \CC{e}^{tD_i}=\CC{e}^t D_i. \end{displaymath} We order the elements by taking first $E_{i,j}$, $i\neq j$, in lexicographic order, followed by $D_1,D_2,\ldots,D_{n-1}$. The commutator table is \begin{Eqnarray*} [E_{i,j},E_{r,s}]&=& \begin{case} E_{i,s}, & i\neq s,\; j=r,\\ -E_{r,j}, & i=s,\;j\neq r,\\ \sum_{l=s}^{r-1} D_l, & i=s<j=r,\\ -\sum_{l=r}^{s-1} D_l, & i=s>j=r,\\ O, & \mbox{otherwise,} \end{case}\qquad \begin{array}{l} i,j,r,s=1,2,\ldots,n,\\ i\neq j,\quad r\neq s, \end{array} \\{} [E_{i,j},D_r]&=& \begin{case} -E_{r,j}, & i=r,\; j\neq r+1,\\ -E_{i,r+1}, & i\neq r,\; j=r+1,\\ -2E_{r,r+1}, & i=r,\; j=r+1,\\ E_{r+1,j}, & i=r+1,\; j\neq r,\\ E_{i,r}, & i\neq r+1,\; j=r,\\ 2E_{r+1,r}, & i=r+1,\; j=r,\\ O, & \mbox{otherwise,} \end{case}\qquad \begin{array}{l} i,j=1,\ldots,n,\quad i\neq j,\\ r=1,\ldots,n-1, \end{array}\\{} [D_i,D_j]&=&O,\qquad i,j=1,2,\ldots,n-1. \end{Eqnarray*} In general, for a $d$-dimensional Lie algebra there are $(d-1)d^2$ structure constants. In the case of $\GG{sl}(n)$ this means that up to $\approx n^6$ structure constants may be nonzero. Yet, using the above basis results in just $2(n-1)n^2+4(n-2)(n-1)+\frac23 (n^2-1)n\approx \frac73 n^3$ nonzero structure constants and substantial saving in the implementation of the SKC technique. Letting \begin{Eqnarray*} [E_{i,j},E_{r,s}]&=&\sum_{(k,l)} c_{(i,j),(k,l)}^{(k,l)} E_{k,l} + \sum_k c_{(i,j),(r,s)}^{(k)} D_k,\\{} [E_{i,j},D_r]&=&\sum_{(k,l)}c_{(i,j),r}E_{k,l} \end{Eqnarray*} (note that $[D_r,E_{i,j}]=-[E_{i,j},D_r]$ and $[D_i,D_j]=O$) we thus have \begin{Eqnarray*} c_{(i,j),(r,s)}^{(k,l)}&=& \begin{case} +1, & k=i,\; l=s,\; r=j,\; s\neq i,\\ -1, & k=r,\; l=j,\; r\neq j,\; s=i,\\ 0, & \mbox{otherwise}, \end{case}\\ c_{(i,j),(r,s)}^{k}&=& \begin{case} +1, & i=s<j=r,\; k=s,s+1,\ldots,r-1,\\ -1, & i=s>j=r,\; k=r,r+1,\ldots,s-1,\\ 0, & \mbox{otherwise}, \end{case}\\ c_{(i,j),r}^{(k,l)}&=& \begin{case} +1, & k=i=r+1,\;l=j,\;j\neq r \mbox{\ or\ } k=i,\; l=j=r,\; i\neq r+1,\\ -1, & k=i=r,\;l=j,\;j\neq r+1 \mbox{\ or\ } k=i,\; l=j=r+1,\; i\neq r,\\ +2, & k=i=r+1,\; l=j=r,\\ -2, & k=i=r,\; l=j=r+1,\\ 0, & \mbox{otherwise}, \end{case}\\ c_{(i,j),r}^{k}&=&c_{r,s}^{(k,l)}=c_{r,s}^k=0. \end{Eqnarray*} Letting \begin{displaymath} B=\sum_{k\neq l} \beta_{k,l} E_{k,l}+\sum_{k} \gamma_k D_k, \end{displaymath} and ordering the pairs $(k,l)$, $k\neq l$, in lexicographic order, we thus have \begin{Eqnarray*} g_{k,l}'(0)&=&\beta_{k,l},\\ g_k'(0)&=&\gamma_k,\\ g_{k,l}''(0)&=&\sum_{(i,j)\succ(r,s)} \beta_{i,j} c_{(i,j),(r,s)}^{(k,l)} \beta_{r,s} +\sum_{(i,j),r} \beta_{i,j} c_{(i,j),r}^{k} \gamma_r\\ &=&\sum_{i=1}^{k-1} \beta_{k,i}\beta_{i,l}-\sum_{i=k+1}^n \beta_{k,i}\beta_{i,l} +\beta_{k,l}(\gamma_{k-1}+\gamma_l -\gamma_k -\gamma_{l-1}),\\ g_k''(0)&=&\sum_{(i,j)\succ(r,s)} \beta_{i,j} c_{(i,j),(r,s)}^{k} \beta_{r,s} =-\sum_{i=1}^k \sum_{j=k+1}^n \beta_{i,j}\beta_{j,i}, \end{Eqnarray*} where $\gamma_0=\gamma_n=0$. \vspace{8pt} \noindent{\bf The Lorenz group} This is the 6-dimensional group $\CC{SO}(3,1)$ of $4\times4$ matrices $A$ such that $AJA^{\CC{T}}=J$, where $J=\CC{diag}(1,1,1,-1)$ \cite{carter95llg}. It has important applications in special relativity theory. he corresponding {\em Lorenz algebra\/} $\GG{so}(3,1)$ consists of all matrices $B$ such that $BJ+JB^{\CC{T}}=O$. It is easy to verify that each element of $\GG{so}(3,1)$ can be written in the form \begin{displaymath} B=\left[ \begin{array}{rrrr} 0 & b_1 & b_2 & b_3\\ -b_1 & 0 & b_4 & b_5\\ -b_2 & -b_4 & 0 & b_6\\ b_3 & b_5 & b_6 & 0 \end{array}\right],\qquad b_1,b_2,\ldots,b_6\in\BB{R}. \end{displaymath} Choosing the basis \begin{Eqnarray*} &&\left\{\left[ \begin{array}{rrrr} 0 & 1 & 0 & 0\\ -1 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 \end{array}\right],\;\, \left[ \begin{array}{rrrr} 0 & 0 & 1 & 0\\ 0 & 0 & 0 & 0\\ -1 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 \end{array}\right],\;\, \left[ \begin{array}{rrrr} 0 & 0 & 0 & 0\\ 0 & 0 & 1 & 0\\ 0 & -1 & 0 & 0\\ 0 & 0 & 0 & 0 \end{array}\right],\;\, \left[ \begin{array}{rrrr} 0 & 0 & 0 & 1\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 1 & 0 & 0 & 0 \end{array}\right],\;\, \left[ \begin{array}{rrrr} 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 1\\ 0 & 0 & 0 & 0\\ 0 & 1 & 0 & 0 \end{array}\right],\;\, \left[ \begin{array}{rrrr} 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 1\\ 0 & 0 & 1 & 0 \end{array}\right]\right\}, \end{Eqnarray*} we obtain the commutator table \begin{displaymath} \begin{array}{llllll} \hspace*{-12.5pt}&[V_1,V_2]=-V_3, &[V_1,V_3]=V_2, &[V_1,V_4]=-V_5, &[V_1,V_5]=V_4, &[V_1,V_6]=O,\\{} \hspace*{-12.5pt}&[V_2,V_1]=V_3, &[V_2,V_3]=-V_1, &[V_2,V_4]=-V_6, &[V_2,V_5]=O, &[V_2,V_6]=V_4,\\{} \hspace*{-12.5pt}&[V_3,V_1]=-V_2, &[V_3,V_2]=V_1, &[V_3,V_4]=O, &[V_3,V_5]=-V_6, &[V_3,V_6]=V_5,\\{} \hspace*{-12.5pt}&[V_4,V_1]=V_5, &[V_4,V_2]=V_6, &[V_4,V_3]=O, &[V_4,V_5]=V_1, &[V_4,V_6]=V_2,\\{} \hspace*{-12.5pt}&[V_5,V_1]=-V_4, &[V_5,V_2]=O, &[V_5,V_3]=V_6, &[V_5,V_4]=-V_1, &[V_5,V_6]=V_3,\\{} \hspace*{-12.5pt}&[V_6,V_1]=O, &[V_6,V_2]=-V_4, &[V_6,V_3]=-V_5, &[V_6,V_4]=-V_2, &[V_6,V_5]=-V_3. \end{array} \end{displaymath} Thus, out of 180 structure constants, just 24 are nonzero -- and they all equal $\pm1$. After brief claculation, we drive for example the polynomials $\alpha_k$ that yield an order-2 CSK approximant, \begin{Eqnarray*} \alpha_1(t)&=&\beta_1 t+\Frac12(\beta_2\beta_3-\beta_4\beta_5)t^2,\\ \alpha_2(t)&=&\beta_2 t-\Frac12(\beta_1\beta_3+\beta_4\beta_6)t^2,\\ \alpha_3(t)&=&\beta_3 t+\Frac12(\beta_1\beta_2-\beta_5\beta_6)t^2,\\ \alpha_4(t)&=&\beta_4 t-\Frac12(\beta_1\beta_5+\beta_2\beta_6)t^2,\\ \alpha_5(t)&=&\beta_5 t+\Frac12(\beta_1\beta_4-\beta_3\beta_6)t^2,\\ \alpha_6(t)&=&\beta_6 t+\Frac12(\beta_3\beta_5-\beta_2\beta_4)t^2, \end{Eqnarray*} where $B=\sum_{k=1}^6 \beta_k V_k$. \setcounter{equation}{0} \section{Time symmetry} An approximant $F(tB)\approx\exp(tB)$ is said to be {\em time symmetric\/} if $F(tB)F(-tB)=I$, $t\geq0$. Time symmetric approximants are important for a number of reasons, not least being that they lend themselves to the {\em Yo\v{s}ida technique,\/} which allows their order to be increased \cite{yoshida90coh}. The techniques of the last section are not time symmetric. Here we describe their modification, which results in a time-symmetric approximant. Me mention in passing that it is possible to envisage two distinct techniques to obtain high-order algorithms based on canonical coordinates of the second kind. The first, implicit in the work of the previous section, consists of evaluating the numbers $g_k^{(l)}(0)$ for $l=1,2,\ldots,p$, where $p$ is the order of the method. The alternative, the subject matter of the present section, consists in combining a second-order or a fourth-order approximant across a number of steps to obtain a higher-order method. Given the splitting \begin{displaymath} B=\sum_{l=1}^s C_l, \end{displaymath} it is well known that the \emph{Strang splitting\/} The approximation \begin{equation} \label{eq:4.1} F(tB)=\CC{e}^{tC_1/2}\cdots \CC{e}^{tC_{s-1}/2}e^{tC_s} \CC{e}^{tC_{s-1}/2}\cdots \CC{e}^{tC_1/2} \end{equation} is of order $2$ and time symmetric. Note that, as a consequence of time symmetry, for sufficiently small $t\geq0$ we can represent $F(tB)=\CC{e}^{{\cal F}(t)}$ where the matrix function ${\cal F}(t)$ is odd. It is precisely this feature that allows the application of the Yo\v{s}ida technique. The clear reason for \R{eq:4.1} being time symmetric is that it is palindromic in the alphabet $\{C_1,C_2,\ldots,C_s\}$. This provides a clue how to modify techniques based on canonical coordinates of the second kind so as to render them time symmetric. Given a basis $\{V_1,V_2,\ldots,V_d\}$ of the Lie algebra $\GG{g}$, we approximate $\CC{e}^{tB}$ by the product \begin{equation} \label{eq:4.2} \exp[\alpha_1(t)V_1]\cdots \exp[\alpha_{d-1}(t)V_{d-1}] \exp[\alpha_d(t)V_d] \exp[\alpha_{d-1}(t)V_{d-1}]\cdots \exp[\alpha_{1}(t)V_1], \end{equation} where $\alpha_1,\alpha_2,\ldots,\alpha_d$ are {\em odd\/} polynomials. Taking $\alpha_l=\frac12\beta_l t$, $l=1,2,\ldots,d-1$ and $\alpha_d=\beta_d t$ yeilds the second-order Strang splitting. In the sequel we seek higher-order methods of this kind. Using the Baker--Campbell--Hausdorff (BCH) formula it is possible to express the product of exponentials at the right hand side of \R{eq:4.2} as a single exponential \cite[ p. 141]{varadarajan84lgl}. Due to the symmetric arrangement of the exponentials in \R{eq:4.2}, the BCH formula is an expansion in odd powers of $t$. If this expansion converges, which is always the case for sufficently small $t$, it makes sense to write the equation \begin{equation} \label{eq:4.3} tB=2\sum_{i=1}^{d-1}\alpha_i(t)V_i+\alpha_d(t)V_d+ \sum_{k=1}^{\infty}Q^{2k}(\MM{\alpha}). \end{equation} Here we denote by $Q^{2k}(\MM{\alpha})$ the terms of order $\O{t^{2k+1}}$ in the BCH formula applied to \R{eq:4.3}. Moreover, we let $\alpha_{i}^{2k}(t)$ be the polynomial obtained by truncating the expansion of $\alpha_i(t)$ after the first $k$ terms, and we denote the remainder by $r_i^{2k}(t)$. In other words, \begin{displaymath} \alpha_i(t)=\alpha_i^{2k}(t)+r_i^{2k}(t),\qquad r_i^{2k}(t)=\O{t^{2k+1}},\qquad i=1,2,\ldots,d. \end{displaymath} From \R{eq:4.3} we deduce \begin{displaymath} 2\sum_{i=1}^{d-1}\alpha_i^{2(k)}(t)V_i+\alpha_d^{2(k)}(t)V_d=tB- \sum_{r=1}^{k-1}Q^{2r}(\MM{\alpha})+\O{t^{2k+1}}. \end{displaymath} Noting that \begin{displaymath} Q^{2r}(\MM{\alpha})=Q^{2r}(\MM{\alpha}^{2(k-1)}+\MM{r}^{2(k-1)}) =Q^{2r}(\MM{\alpha}^{2(k-1)})+\O{t^{2k+r}}, \end{displaymath} we obtain \begin{equation} \label{eq:4.4} 2\sum_{i=1}^{d-1}\alpha_i^{2(k)}(t)V_i+\alpha_d^{2(k)}(t)V_d =tB-\sum_{r=1}^{k-1}Q^{2r}(\MM{\alpha}^{2(k-1)})+\O{t^{2k+1}}. \end{equation} Dropping the $\O{t^{2k+1}}$ terms in \R{eq:4.4}, it is possible to compute $\MM{\alpha}^{2k}$ from $\MM{\alpha}^{2(k-1)}$. This gives a procedure to derive a sequence of successively increaing-order approximants of $\exp(tB)$. It is easy to see that the approximants \begin{displaymath} F^{2k}(tB)=\exp(\alpha_1^{2k}(t)V_1)\cdots \exp(\alpha_d^{2k}(t)V_d)\cdots \exp(\alpha_{1}^{2k}(t)V_1), \end{displaymath} of $\exp(tB)$ are such that $F^{2k}(tB)F^{2k}(-tB)=I$, hence time symmetry, the reason being the symmetric arrangements of the exponentials in $F^{2k}(tB)$ and the odd-power expansion of the functions $\alpha_i^{2k}$. The BCH and symmetric BCH formulae for $k$-terms have an exceedingly complicated expansion, which can be obtained recursively. In what follows we will make use just of the term $Q^{2}(\alpha)$, demonstrating how it is possible to compute it explicitely for particular choices of the basis. In the remainder of this section we consider the implementation of time-symmetric CSK methods. We split $B$ as before and commence by considering the Strang splitting \R{eq:4.1} except that, to simplify notation, we arrange the terms in reverse ordering, \begin{equation} \label{eq:4.5} \CC{e}^{tC_s/2}\cdots \CC{e}^{tC_2/2}\CC{e}^{tC_1}\CC{e}^{tC_2/2}\cdots \CC{e}^{tC_s/2}. \end{equation} \begin{lemma} The term $Q^2$ of the BCH formula applied to \R{eq:4.5} is \begin{equation} \label{eq:4.6} Q^2= \frac{t^3}{12}\sum_{l=2}^s [C_1+\cdots+C_{l-1} +\Frac12 C_l,[C_1+\cdots+C_{l-1},C_l]]. \end{equation} \end{lemma} \begin{proof} See the appendix. \end{proof} Let us next consider the case $\GG{g}=\GG{so}(n)$, choosing the same sparse basis as in Section~2. Therefore, according to \R{eq:4.6}, we have \begin{equation} \label{eq:4.7} \begin{meqn} Q^2&=&\displaystyle \frac{t^3}{12}\sum_{i=1}^{n-1}\sum_{j=i+1}^n [b_{1,2}F_{1,2}+ \ldots + b_{i,j-1}F_{i,j-1},[b_{1,2}F_{1,2}+\ldots+b_{i,j-1} F_{i,j-1},\\ &&\mbox{}\displaystyle b_{i,j}F_{i,j}]]+ \frac{1}{24}\sum_{i=1}^{n-1}\sum_{j=i+1}^n b_{i,j} [F_{i,j},[b_{1,2}F_{1,2}+\dots+b_{i,j-1}F_{i,j-1},b_{i,j}F_{i,j}]]. \end{meqn} \end{equation} We compute separately each part of this sum. Exploiting the commutator table of our basis we have \begin{Eqnarray*} &&[b_{1,2}F_{1,2}+\ldots+b_{i,j-1}F_{i,j-1},b_{i,j}F_{i,j}]\\ &=& b_{i,j}\left(-\sum_{s=i+1}^{j-1}b_{i,s}F_{s,j}-\sum_{r=1}^{i-1} b_{r,j}F_{r,i}+\sum_{t=1}^{i-1}b_{t,i}F_{t,j}\right), \end{Eqnarray*} and noting that $b_{i,s}=-b_{s,i}$, we deduce that \begin{equation} \label{eq:4.8} [b_{1,2}F_{1,2}+\dots+b_{i,j-1}F_{i,j-1},b_{i,j}F_{i,j}]=b_{i,j}\left( \sum_{\stackrel{\scriptstyle t=1}{t\neq i}}^{j-1}b_{t,i}F_{t,j}-\sum_{r=1}^{i-1}b_{r,j}F_{r,i}\right). \end{equation} Commuting the right-hand side with $F_{i,j}$ gives \begin{equation} \label{eq:4.9} \left[F_{i,j},\sum_{\stackrel{\scriptstyle t=1}{t\neq i}}^{j-1}b_{t,i}F_{t,j}-\sum_{r=1}^{i-1} b_{r,j}F_{r,i}\right]= \sum_{\stackrel{\scriptstyle t=1}{t\neq i}}^{j-1} b_{t,i}F_{t,i}+\sum_{r=1}^{i-1}b_{r,j}F_{r,j}. \end{equation} Let $\MM{b}_1,\MM{b}_2,\ldots,\MM{b}_n$ be the columns of $B$ and denote \begin{displaymath} \MM{b}_l^{s}=\sum_{k=1}^{s-1}b_{k,l}\MM{e}_k,\qquad k=1,2,\ldots,n. \end{displaymath} Then \R{eq:4.8} yields \begin{displaymath} [b_{1,2}F_{1,2}+\dots+b_{i,j-1}F_{i,j-1},b_{i,j}F_{i,j}]=b_{i,j}( \MM{b}_{i}^{j}\MM{e}_j^{\CC{T}}-\MM{e}_j{\MM{b}_{i}^{j}}^{\CC{T}}) -b_{i,j}(\MM{b}_{j}^{i}\MM{e}_i^{\CC{T}} -\MM{e}_i{\MM{b}_{j}^{i}}^{\CC{T}}), \end{displaymath} while \R{eq:4.9} gives \begin{Eqnarray*} &&[F_{i,j}, [b_{1,2}F_{1,2}+\cdots +b_{i,j-1}F_{i,j-1},b_{i,j}F_{i,j}]]\\ &=& b_{i,j}(\MM{b}_{i}^{j}\MM{e}_i^{\CC{T}}-\MM{e}_i{\MM{b}_{i}^{j}} ^{\CC{T}})-b_{i,j}(\MM{b}_{j}^{i}\MM{e}_j^{\CC{T}}-\MM{e}_j {\MM{b}_{j}^{i}}^{\CC{T}}). \end{Eqnarray*} Multiplying the latter by $b_{i,j}$ and summing in $i$ and $j$ we can evaluate \R{eq:4.7} in $n^3$ operations. Note that we count separately multiplications and additions, for example, we assume that the cost of Euclidean inner product of two vectors of length $n$ is $2n$ operations. We now assemble together our results to calculate \R{eq:4.7}. We proceed by splitting the sum $b_{1,2}F_{1,2}+\cdots+b_{i,j-1}F_{i,j-1}$ in three parts, whereby \begin{Eqnarray*} &&[b_{1,2}F_{1,2}+\cdots+b_{i,j-1}F_{i,j-1},\MM{b}_{i}^{j} \MM{e}_j^{\CC{T}}-\MM{e}_j{\MM{b}_{i}^{j}}^{\CC{T}}-(\MM{b}_{j}^{i} \MM{e}_i^{\CC{T}} -\MM{e}_i{\MM{b}_{j}^{i}}^{\CC{T}})]\\ &=&\sum_{l=1}^{i}\sum_{k=1}^{l-1}[(\MM{b}_{l}^{l}\MM{e}_l^{\CC{T}} -\MM{e}_l{\MM{b}_{l}^{l}}^{\CC{T}}),\MM{b}_{i}^{j}\MM{e}_j^{\CC{T}} -\MM{e}_j{\MM{b}_{i}^{j}}^{\CC{T}}-(\MM{b}_{j}^{i}\MM{e}_i^{\CC{T}} -\MM{e}_i{\MM{b}_{j}^{i}}^{\CC{T}})]\\ &&\mbox{}+\sum_{l=i+1}^{j-1}\sum_{k=1}^{i}[(\MM{b}_{l}^{i+1}\MM{e}_l^{\CC{T}} -\MM{e}_l{\MM{b}_{l}^{i+1}}^{\CC{T}}), \MM{b}_{i}^{j}\MM{e}_j^{\CC{T}} -\MM{e}_j{\MM{b}_{i}^{j}}^{\CC{T}} -(\MM{b}_{j}^{i}\MM{e}_i^{\CC{T}}-\MM{e}_i{\MM{b}_{j}^{i}}^{\CC{T}})]\\ &&\mbox{}+\sum_{l=j}^{m}\sum_{k=1}^{i-1}[(\MM{b}_{l}^{i}\MM{e}_l^{\CC{T}} -\MM{e}_l{\MM{b}_{l}^{i}}^{\CC{T}}), \MM{b}_{i}^{j}\MM{e}_j^{\CC{T}} -\MM{e}_j{\MM{b}_{i}^{j}}^{\CC{T}}-(\MM{b}_{j}^{i}\MM{e}_i^{\CC{T}} -\MM{e}_i{\MM{b}_{j}^{i}}^{\CC{T}})]. \end{Eqnarray*} Finally, \begin{Eqnarray*} &&[b_{1,2}F_{1,2}+\cdots+b_{i,j-1}F_{i,j-1},\MM{b}_{i}^{j} \MM{e}_j^{\CC{T}}-\MM{e}_j{\MM{b}_{i}^{j}}^{\CC{T}}-(\MM{b}_{j}^{i} \MM{e}_i^{\CC{T}} -\MM{e}_i{\MM{b}_{j}^{i}}^{\CC{T}})]\\ &=&-{\MM{b}_i^i}^{\CC{T}}\MM{b}_i^jF_{i,j}+\MM{b}_i^i {\MM{b}_j^i}^{\CC{T}}-\MM{b}_j^i{\MM{b}_i^i}^{\CC{T}}+ {\MM{b}_j^{i}}^{\CC{T}}\MM{b}_j^iF_{j,i}-(\MM{b}_j^i {\MM{b}_i^j}^{\CC{T}}-\MM{b}_i^j{\MM{b}_j^i}^{\CC{T}})\\ &&\mbox{}\sum_{l=1}^{i-1}b_{l,i}(\MM{b}_l^l\MM{e}_j^{\CC{T}}- \MM{e}_j{\MM{b}_l^l}^{\CC{T}})-{\MM{b}_l^l}^{\CC{T}}\MM{b}_i^jF_{l,j} -b_{l,j}(\MM{b}_l^l\MM{e}_i^{\CC{T}}-\MM{e}_i{\MM{b}_l^l}^{\CC{T}}) +{\MM{b}_l^l}^{\CC{T}}\MM{b}_j^iF_{l,i}\\ &&\mbox{}\sum_{l=i+1}^{j-1}b_{l,i}(\MM{b}_l^{i+1}\MM{e}_j^{\CC{T}} -\MM{e}_j{\MM{b}_l^{i+1}}^{\CC{T}})-{\MM{b}_l^{i+1}}^{\CC{T}} \MM{b}_i^{j}F_{l,j}+b_{i,l}(\MM{b}_{j}^{i}\MM{e}_l^{\CC{T}}- \MM{e}_l{\MM{b}_{j}^{i}}^{\CC{T}})+{\MM{b}_l^{i+1}}^{\CC{T}}\MM{b}_j^i F_{l,i}\\ &&\mbox{}+\sum_{l=j+1}^{n}{\MM{b}_l^{i}}^{\CC{T}}\MM{b}_j^iF_{l,i} -{\MM{b}_l^i}^{\CC{T}}\MM{b}_i^jF_{l,j}. \end{Eqnarray*} We analyse the computational costs of the previous formula, summing over $i$ and $j$ and showing that \R{eq:4.7} can be computed in $\O{n^3}$ operations. Note that, since $\sum_{l=i+1}^{j-1}b_{i,l}\MM{e}_l=\MM{b}_i^i-\MM{b}_i^j$, we have \begin{displaymath} \MM{b}_i^i{\MM{b}_j^i}^{\CC{T}}-\MM{b}_j^i{\MM{b}_i^i}^{\CC{T}}+ \sum_{l=i+1}^{j-1}b_{i,l}(\MM{b}_{j}^{i}\MM{e}_l^{\CC{T}}- \MM{e}_l{\MM{b}_{j}^{i}}^{\CC{T}})-(\MM{b}_j^i{\MM{b}_i^j}^{\CC{T}} -\MM{b}_i^j{\MM{b}_j^i}^{\CC{T}})=-2(\MM{b}_j^i{\MM{b}_i^j}^{\CC{T}} -\MM{b}_i^j{\MM{b}_j^i}^{\CC{T}}). \end{displaymath} It is more convenient to write the previous expression in the form \begin{Eqnarray*} &&-2\sum_{i=1}^{n-1}\sum_{j=i+1}^n b_{i,j}(\MM{b}_j^i {\MM{b}_i^j}^{\CC{T}}-\MM{b}_i^j{\MM{b}_j^i}^{\CC{T}})\\ &=&-\sum_{i=1}^n\left( \sum_{j=i+1}^{n}2b_{i,j}\MM{b}_j^i\right) {\MM{b}_i^i}^{\CC{T}}-{\MM{b}_i^i}\left(\sum_{j=i+1}^{n}2b_{i,j} \MM{b}_j^{i}\right)^{\!\!\CC{T}}\\ &&\mbox{}-\sum_{i=1}^{n-1}\sum_{j=i+1}^n2b_{i,j} \left(\MM{b}_j^i(\MM{b}_i^j-\MM{b}_i^i)^{\CC{T}} -(\MM{b}_i^j-\MM{b}_i^i){\MM{b}_j^i}^{\CC{T}}\right). \end{Eqnarray*} The first part of this sum is computed in about $\frac23 n^3$ operations and the second part, exploiting the equality \begin{displaymath} \sum_{i=1}^{n-1}\sum_{j=i+1}^n2b_{i,j}\MM{b}_j^i(\MM{b}_i^j -\MM{b}_i^i)^{\CC{T}}=2\sum_{i=1}^{n-1}\sum_{k=n-1}^{i+2}b_{i,k} \left(\sum_{l=n}^{k+1}b_{i,l}\MM{b}_l^i\right)\MM{e}_k^{\CC{T}}, \end{displaymath} can also be computed in $\frac23 n^3$ operations. Adding terms of the type $\alpha F_{l,j}$ and $\beta F_{l,i}$ leads to \begin{displaymath} -\sum_{l=1}^i{\MM{b}_l^l}^{\CC{T}}\MM{b}_i^jF_{l,j}- \sum_{l=i+1}^{j-1}{\MM{b}_l^{i+1}}^{\CC{T}}\MM{b}_i^{j}F_{l,j} -\sum_{l=j+1}^{n}{\MM{b}_l^i}^{\CC{T}}\MM{b}_i^jF_{l,j} =-L_i^i\MM{b}_j^i{\MM{e}_j}^{\CC{T}}-{\MM{e}_j}(-L_i^i\MM{b}_j^j)^{\CC{T}}, \end{displaymath} and \begin{displaymath} \sum_{l=1}^{i-1}{\MM{b}_l^l}^{\CC{T}}\MM{b}_j^iF_{l,i} +\sum_{l=i+1}^{j-1}{\MM{b}_l^{i+1}}^{\CC{T}}\MM{b}_j^{i}F_{l,i} +\sum_{l=j+1}^{n}{\MM{b}_l^i}^{\CC{T}}\MM{b}_j^iF_{l,i} =L_i^j\MM{b}_i^j{\MM{e}_i}^{\CC{T}}-{\MM{e}_i}(L_i^j\MM{b}_i^j)^{\CC{T}}, \end{displaymath} where the matrix $L_i$ is the lower triangular part of $b_{1,2}F_{1,2}+\ldots+b_{i-1,n}F_{i-1,n}$ and we denote by $L_{i}^{s}$, $s=i,j$, the matrix $L_i$ with zeros along its $s$-th row. Summing up with respect to $i$ and $j$, we obtain \begin{Eqnarray*} \sum_{i=1}^{n-1}\sum_{j=i+1}^nb_{i,j}(-L_i^i\MM{b}_j^i){\MM{e}_i}^{\CC{T}} &=&\sum_{i=1}^{n-1}(-L_i^i)\left(\sum_{j=i+1}^nb_{i,j}\MM{b}_j^i\right) {\MM{e}_i}^{\CC{T}}\\ \sum_{i=1}^{n-1}\sum_{j=i+1}^nb_{i,j}L_i^j\MM{b}_i^j{\MM{e}_j}^{\CC{T}} &=&\sum_{j=2}^{n}\left(\sum_{i=1}^{j-1}b_{i,j}\MM{c}_i\right) {\MM{e}_j}^{\CC{T}} \end{Eqnarray*} where we have used the notation $\MM{c}_i:=L_i\MM{b}_i^i$ for $i=1,\dots,n-1.$ The cost of computing the first sum is $\frac43n^3$, while the cost of computing the second is $2n^3$ operations. Finally the terms \begin{displaymath} \sum_{i=1}^{n-1}\sum_{j=i+1}^n\sum_{l=1}^ib_{i,j}b_{l,i} (\MM{b}_l^l\MM{e}_j^{\CC{T}}-\MM{e}_j{\MM{b}_l^l}^{\CC{T}})= \sum_{l=1}^{n-1}\MM{b}_l^l\left(\sum_{j=l+2}^n c_{j,l}\MM{e}_j^{\CC{T}}\right)-\left(\sum_{j=l+2}^n c_{j,l}\MM{e}_j\right){\MM{b}_l^l}^{\CC{T}} \end{displaymath} with $c_{j,l}=\sum_{i=l}^{j-1}b_{i,j}b_{l,i}$, \begin{displaymath} \sum_{i=1}^{n-1}\sum_{j=i+1}^n\sum_{l=1}^ib_{i,j}b_{l,j} (\MM{b}_l\MM{e}_i^{\CC{T}}-\MM{e}_i\MM{b}_l^{\CC{T}})= \sum_{l=1}^{n-1}\MM{b}_l^l\left(\sum_{i=l+1}^{n-1} d_{i,l}\MM{e}_i^{\CC{T}}\right)-\left(\sum_{i=l+1}^{n-1} d_{i,l}\MM{e}_i\right){\MM{b}_l^l}^{\CC{T}} \end{displaymath} with $d_{i,l}=\sum_{j=i+1}^{n}b_{i,j}b_{l,j}$, and \begin{Eqnarray*} &&\sum_{i=1}^{n-1}\sum_{j=i+1}^n\sum_{l=i+1}^{j-1}b_{i,j}b_{l,i} (\MM{b}_l^{i+1}\MM{e}_j^{\CC{T}}-\MM{e}_j{\MM{b}_l^{i+1}}^{\CC{T}})\\ &=&\sum_{i=1}^{n-1}\sum_{l=i+1}^{n-1}b_{l,i}\left(\MM{b}_l^{i+1} {(\MM{b}_i^{l+1}-\MM{b}_i)}^{\CC{T}}-(\MM{b}_i^{l+1}-\MM{b}_i) {\MM{b}_l^{i+1}}^{\CC{T}}\right); \end{Eqnarray*} can be computed in about $\frac23n^3$, $\frac23n^3$ and $\frac12n^3$ operations respectively. Collecting the contributions of all the terms in the sum we obtain a total count of $7\frac12n^3$ operations. At the present time it is not clear that this method of computation of $Q^2$ in the $\GG{so}(n)$ case is optimal form the point of view of complexity theory. We did not try any other ordering of the basis elements and it is not at all certain that different orderings could give better constants in front of the term $n^3$. Given that the construction of the (second-order) Strang splitting carries a cost of $4n^3$ operations, the total flop count for constructing a symmetric fourth order SKC approximation of an exponential in $\GG{so}(n)$ by our algorithm is $11\frac12n^3$. This is marginally better than obtaining an order-4 approximation by the Yo\u{s}ida technique from three Strang splittings which, as pointed out in \cite{celledoni98atm}, requires $12n^3$ flops. \setcounter{equation}{0} \section{Sparse matrices} In a naive formulation, the method of canonical coordinates of the second kind is considerably too expensive for practical computation. This, however, can be alleviated by the use of a sufficiently `sparse' basis of the underlying Lie algebra $\GG{g}$. As explained in Section~2, choosing a basis so that an overwhelming majority of structure constants vanish renders the algorithm strikingly more effective. It is important to emphasize that this has nothing to do with the structure of the matrix $B\in\GG{g}$, which need not be sparse. Yet, in most practical computations (in particular when $n$ is large) one can expect $B$ to be sparse and structured. Good algorithms should be able to exploit this phenomenon. In the case of SKC methods we identify two mechanisms that allow us to exploit sparsity. Although this aspect of our methods is still a matter for active investigation, the interim results are substantive enough to warrant publication. For simplicity, we describe the first mechanism just in the case of a tridiagonal $B\in\GG{so}(n)$, hence \begin{displaymath} B=\left[ \begin{array}{ccccc} 0 & \beta_1 & 0 & \cdots & 0\\ -\beta_1 & 0 & \ddots & \ddots & \vdots\\ 0 & \ddots & \ddots & \ddots & 0\\ \vdots & \ddots & \ddots & 0 & \beta_{n-1}\\ 0 & \cdots & 0 & -\beta_{n-1} & 0 \end{array}\right]=\sum_{k=1}^{n-1}\beta_k F_{k,k+1}, \end{displaymath} where the matrices $F_{k,l}=\MM{e}_k\MM{e}_l^{\CC{T}}-\MM{e}_l\MM{e}_k ^{\CC{T}}$ have been introduced in Section~2. Since \begin{displaymath} b_{k,l}= \begin{case} \beta_k, & l=k+1,\\ 0, & \mbox{otherwise,} \end{case} \end{displaymath} it is easy to substitute in the general formulae for the order-2 method: \begin{displaymath} g_{i,j}'(0)= \begin{case} \beta_i, & \!\!\!\! j=i+1,\\ 0, & \!\!\!\! \mbox{otherwise,} \end{case}\quad g_{i,j}''(0)= \begin{case} \beta_{i-1}\beta_i, & \!\!\!\! j=i-2,\\ 0, & \!\!\!\! \mbox{otherwise,} \end{case} \qquad 1\leq i<j\leq n, \end{displaymath} Arranging the elements of the basis in lexicographic order, we thus obtain the second-order approximant \begin{displaymath} \CC{e}^{\beta_{n-1}t F_{n-1,n}} \CC{e}^{\beta_{n-2}t F_{n-2,n-1}} \CC{e}^{\frac12\beta_{n-1}\beta_{n-1}t^2 F_{n-2,n}} \cdots \CC{e}^{\beta_2 tF_{2,3}} \CC{e}^{\frac12 \beta_2\beta_3 t^2F_{2,4}} \CC{e}^{\beta_1 tF_{1,2}} \CC{e}^{\frac12 \beta_1\beta_2 t^2F_{1,3}}. \end{displaymath} In other words, the cost of the approximation is just $\O{n}$ flops. Similar situation pertains to \begin{displaymath} B=\left[ \begin{array}{ccccc} \gamma_1 & \eta_1 & 0 & \cdots & 0\\ \mu_1 & \gamma_2 & \ddots & & \vdots\\ 0 & \ddots & \ddots & \ddots & 0\\ \vdots & & \ddots & \gamma_{n-1} & \eta_{n-1}\\ 0 & \cdots & 0 & \mu_{n-1} & \gamma_n \end{array}\right]\in \GG{sl}(n). \end{displaymath} Choosing the same basis and terminology as in Section 2 we can readily ascertain that \begin{Eqnarray*} g_{k,k-2}''(0)&=&\mu_{k-2}\mu_{k-1}\qquad k=3,4,\ldots,n\\ g_{k,k-1}''(0)&=&-(\gamma_{k-2}-2\gamma_{k-1}+\gamma_k)\mu_{k-1}, \qquad k=3,4,\ldots,n,\\ g_{k,k+1}''(0)&=&(\gamma_{k-1}-2\gamma_k+\gamma_{k+1})\eta_k,\qquad k=2,3,\ldots, n-1,\\ g_{k,k+2}''(0)&=&-\eta_k\eta_{k+1},\qquad k=1,2,\ldots,n-2,\\ g_{k,l}''(0)&=&0,\qquad |k-l|\geq 3 \end{Eqnarray*} and $g_k''(0)=-\eta_k\mu_k$, $k=1,2,\ldots,n-1$. Thus, a second-order approximant to a tridiagonal $B\in\GG{sl}(n)$ is itself quindiagonal and its computation requires just $\O{n}$ flops. Higher-order approximants and matrices with greater bandwidth lend themselves to similar treatment, although the savings are less striking. In a sense, the situation is parallel to that of approximating $\exp(tB)$ by a rational approximant, when savings accrue from sparse matrix-inversion methods, except that in our case the result is assured to belong to the right Lie group. Another observation which is highly pertinent to the approximation of exponentials of sparse matrices has been made in \cite{iserles99hli}. Suppose that $B$ is a banded matrix of bandwidth $s\geq 3$. In general, $F(t)=\exp(tB)$ is a dense matrix. Yet, as is easy to illustrate by computer experiments, $F(t)$ is very near to a banded matrix. Specifically, given $\varepsilon>0$, there exists $r=r(t,\varepsilon)\geq s$ such that all the elements of $F(t)$ outside a band of width $r$ are less than $\varepsilon$ in magnitude. Moreover, tight upper bounds on $r$ can be derived with relative ease. The idea thus is to set to zero all the elements outside bandwidth $r$. The outcome is a banded approximant to the exponential. Moreover, with an appropriate choice of basis elements, this means that the functions $\alpha_i$ are {\em set to zero\/} for elements that possess terms exclusively outside the band. Consequently corresponding exponentials equal identity and need not be included in the product. Thus, the cost scales with the size $r$ of the bandwidth. Similar phenomenon has been already encountered in the context of $\GG{so}(n)$ and $\GG{sl}(n)$, when our choice of basis and order has implied a banded structure of the exponential. The present mechanism is different, even if the net outcome is similar. \setcounter{equation}{0} \section{Numerical experiments} Our numerical experiments are organized as follows. We fist consider a test on random matrices in $\GG{so}(50)$, illustrating the performance of methods based on the use of second kind coordinates techniques for full and sparse matrices. The third and last example is the solution of a third-order ODE using {\rm Runge--Kutta/Munthe-Kaas\/} (RK/MK) methods described in \cite{munthe-kaas97hor}. We use the {\tt Matlab} toolbox {\tt DiffMan\/} for the integration of ODEs on manifolds, comparing the usual implementation of RK/MK methods, whereby the the exponential is approximated to machine accuracy, with a version of the methods obtained using the time-symmetric fourth-order approximation from Section~4. All experiments have been performed in {\tt Matlab} and we have computed the error while comparing the results with the built-in function {\tt expm} which calculates the exponential to nearly machine accuracy. We evaluated the the error computing $\|\CC{e}^{-tB}F(tB)-I\|_{\rm F}$ where $F(tB)$ is the SKC approximation of $\exp(tB)$ and $\|\cdot\|_{\rm F}$ denotes the Frobenius norm. The matrices have been generated randomly using the {\tt Matlab} function {\tt rand} and scaling the Frobenius norm so that $\|B\|_{\rm F}=1$. We approximate $\exp(tB)$ with a single step of the methods for different values of $t$, ($t=1/2^k$ and $k=1,\ldots ,5$). In both the first two figures the norm of the error is plotted (along the $y$-axis) to a logarithmic scale with respect to $t$. Figure~\ref{fig:2} reports the results of our first test, where we have considered a full matrix in $\GG{so}(50)$. In the plots the error norm is indicated with the symbols `$*$' (SKC, time symmetric, order 4) and `$\circ$' (Strang splitting, order 2). \begin{figure}[Htbp] \begin{center} \leavevmode \psfig{file=exwn1.eps,width=8cm} \caption{Error versus time in the $\GG{so}(50)$ (full case).} \label{fig:2} \end{center} \end{figure} In the next example, illustrated in Figure~\ref{fig:3}, the same methods have been applied to a sparse matrix in $\GG{so}(50)$, with four non-zero diagonals (i.e., bandwidth 5). In both the examples the methods give the correct order. In the second case, however, the count of flops is drastically reduced. We counted the number of flops using the {\tt Matlab} function {\tt flops}. In the first case the cost for constructing $Q^2$ amounts to $9.62n^3$ while in the second we counted $0.95n^3$ flops. As it is easy to understand, the described implementation of the methods allows to take advantage immediately of the sparsity structure of the matrix $B$, working directly on the nonzeros entries of $B$ {\em \`a la\/} Section~4. \begin{figure}[Htbp] \begin{center} \leavevmode \psfig{file=exwn2.eps,width=8cm} \caption{Error versus time in the $\GG{so}(50)$ sparse case.} \label{fig:3} \end{center} \end{figure} The last example is concerned with the use of the techniques described in this paper in substituting the exponentials computed to machine accuracy in the integration methods of \cite{munthe-kaas97hor}. The experiments have been performed using the {\tt Matlab} toolbox {\tt DiffMan.\/} We use a RK/MK method of order four. The example is a problem whose solution is the soliton originating in the {\em Korteweg--de Vries\/} (KdV) equation. It is a third-order ODE obtained performing a symmetry reduction on the KdV equation. The resulting ODE can be written as a three-dimensional system, \begin{displaymath} y'=\left[ \begin{array}{ccc} 0&1 &0\\ 0&0&1\\ -9y(2)&3&0\\ \end{array} \right]y \end{displaymath} with $y(0)=[1,0,-1.5]^{\CC{T}}$ and $t\in [0,5]$. The solution of the ODE $f=y_1(t)$ can be easily derived explicitely and it is $f(t)=\alpha\CC{sech}\,(t\beta)$, $\alpha=1$, $\beta=1/2\sqrt(3).$ In Figure \ref{fig:4} we plot the analytic solution (solid line) on a grid of $161$ points. The dotted line is the numerical solution obtained with the {\tt Matlab} routine {\tt ode45} with absolute and relative tolerance $1.0e-4$. The method produced this solution in $69$ steps, and it was implemented with step-size control procedure. The dashed-dotted line is the numerical solution obtained with the RK/MK method using SKC symmetric tecnhniques for the approximation of the exponential, with fixed step-size $h$. \begin{figure}[Htbp] \begin{center} \leavevmode \psfig{file=fexwn2.eps,width=8cm} \caption{The soliton originating in the KdV equation.} \label{fig:4} \end{center} \end{figure} In Figure \ref{fig:5} we plot the error (along the $y$-axis) with respect to the numerical solution obtained with the {\tt Matlab} routine {\tt ode45} to a logarithmic scale, versus the stepsize $h=1/2^k$ and $k=1,\ldots ,5$ for the cases of the implementation of RK/MK with the {\tt expm} function of {\tt Matlab} (marked with $+$) and approximating $\exp$ to order four with a SKC technique ($\circ$). The line marked with $\ast$ representes the error of the numerical solution given by the RK/MK method implemented with SKC technique for the approximation of the exponential, measured with respect to the numerical solution obtained by the same method with the use of the exact exponential ({\tt expm} routine of {\tt Matlab}). \begin{figure}[Htbp] \begin{center} \leavevmode \psfig{file=fexwn.eps,width=8cm} \caption{RK/MK: global error at $t=5$ with {\tt expm} and SKC} \label{fig:5} \end{center} \end{figure} It is interesting to note in this case that substituting the exact exponential with suitable fourth-order approximant does not lead to a significant deterioration in the quality of the RK/MK method and the overall error does not change much. Note that in the present case the primary variable is a vector, rather than a matrix. In general, if the underlying ODE can be written in a vector form, i.e.\ as an action of a Lie group on $\BB{R}^n$, we need to approximate $\exp(tB)\MM{v}$, where $\MM{v}\in\BB{R}^n$, rather than the matrix $\exp(tB)$. This leads to obvious savings in the SKC techniques, similarly, say, to the approach of rational functions. In particular, the cost of composing exponentials is $\O{n^2}$, rather than $\O{n^3}$, operations. \section*{Acknowledgments} The authors are grateful to Brynjulf Owren for many fruitful discussions, to Per Christian Moan for bringing the reference \cite{wei63ogr} to their attention, to the Numerical Analysis group of DAMTP Cambridge, and to the Geometric Integration members during the fall semester 1998 at MSRI Berkeley. Research at MSRI is supported in part by NSF grant DMS-9701755. \bibliographystyle{agsm}
\section{Introduction} The unidentified gamma-ray point sources concentrated along the Galactic plane present a long outstanding puzzle (Swanenburg et al. 1981). Their positions, timing, and energy spectra should provide a means for their identification. However, the gamma-ray error boxes are sufficiently large that multiple optical and x-ray candidates are usually found for Galactic sources. The only convincing identifications have been made through detection of a pulsed gamma-ray signal at a period known from radio or x-ray observations (Bignami \& Hermsen 1983, Halpern \& Holt 1992, Thompson et al. 1994, Ramanamurthy et al. 1995). The known gamma-ray pulsars have spectra significantly harder than those of most of the unidentified sources. Recently, Merck et al. (1996) studied the energy spectra of the unidentified Galactic gamma-ray sources and identified only eight sources with energy spectra sufficiently hard to be pulsar candidates. Here, we report on x-ray and optical observations in the field of one of the unidentified Galactic gamma-ray sources with a hard gamma-ray spectrum, 2EG~J0635+0521. In \S 2, we report the discovery of an x-ray point source with an unusually hard spectrum within the error box of 2EG~J0635+0521. In \S \S 3 and 4, we describe the properties of an optical counterpart of the x-ray source. We conclude, in \S 5, with comments on identification of the x-ray and gamma-ray sources and on the possible nature of the source. \begin{figure}[tb] \epsscale{0.9} \plotone{f1.ps} \figcaption{X-ray image of SAX~J0635+0533 for the 5--12~keV band. Shown are the position of the optical counterpart (cross), and the 95\% confidence error ellipse for 2EJ~0635+0521 (dashed line).} \label{fig_image} \end{figure} \begin{figure}[tb] \epsscale{0.7} \plotone{f2.ps} \figcaption{X-ray spectrum of SAX~J0635+0533. Shown are data from the LECS (1.0-4.0~keV), MECS (1.8--10~keV), and PDS (12--60~keV). The solid line is a fit of a powerlaw of photon index of 1.5 with absorption.} \label{fig_xray_spectrum} \end{figure} \section{X-ray observations} Using the {\it Satellite Italiano per Astronomia X} (BeppoSAX), we observed a field centered on a region of unusually high x-ray hardness and brightness found in an x-ray survey of the Monoceros supernova remnant (Leahy, Naranan, \& Singh 1986) and located within the error box of the EGRET source 2EG~J0635+0521 (Thompson et al. 1995). The observation was carried out from 23 Oct 1997 3:20:26 UT to 24 Oct 1997 0:11:52 UT. The field of view of the BeppoSAX narrow field instruments covered approximately half of the error box of 2EG~J0635+0521. The x-ray images from the LECS and the MECS 2 and 3 detectors reveal a single point source located at RA = $\rm 06^{h} 35^{m} 17^{s}.4$, DEC = $+05^{\circ} 33' 20''.9$ (J2000). There is a 1' systematic uncertainty in the position reconstruction. We denote the source as SAX~J0635+0533. This source lies within the error box of 2EG~J0635+0521 (see Figure~1) and near the peak of the gamma-ray emission as reconstructed by Jaffe et al. (1997). The image is consistent with that of a single point source, see Fig.~1. We find no evidence for diffuse hard x-ray emission. The x-ray spectrum of the source is quite hard with a spectral index near 1.5 and significant emission detected out to 40~keV. Shown in Figure~2 are data from the LECS, the combined MECS 2 and 3, and the PDS. We used an extraction radius of 2' for the MECS and 4' for the LECS, the standard background files, and response matrixes calculated for the source position 3' off-axis. The x-ray spectrum from all the instruments can be fitted with a simple powerlaw model with absorption. This fit is acceptable with $\chi^2/{\rm DoF} = 97/81$. The fitted column density is $(2.0 \pm 0.3) \times 10^{22} \rm \, cm^{-2}$ and the powerlaw index is $1.50 \pm 0.08$. The unabsorbed flux from the source is $1.2 \times 10^{-11} \rm \, erg \, cm^{-2} \, s^{-1}$ in the 2-10~keV band. We also fitted an exponentially cutoff powerlaw which gave $\chi^2/{\rm DoF} = 96/80$. The exponential cutoff does not improve the fit. The 90\% confidence lower bound on the cutoff energy is 37~keV. We also fitted a Raymond-Smith model spectrum which gave a worse fit $\chi^2/{\rm DoF} = 122/83$ and an unphysically high temperature, $kT > 50 \rm \, keV$. We searched for pulsations in the x-ray data over a period range from 0.030~s to 1000~s. We found no significant pulsed signal. From the period with the highest single trial significance, we place an upper limit on the pulsed fraction of 26\%. The continuum of the timing power spectrum is consistent with the Poisson level for frequencies above 0.002~Hz. This field was previously viewed with the Einstein IPC (region 3 in Leahy et al. 1986). The source was not detected with Einstein because, we now know, it fell on a strongback arm. This field was also observed for 4371~s by the {\it ROSAT\/} PSPC on 1992 October~9. The position of SAX~J0635+0533 falls $34^{\prime}$ off axis in this image, where the point-spread function is significantly degraded. Nevertheless, there may be a very weak source present at this position. Folding the spectrum of SAX~J0635+0533 through the {\it ROSAT\/} response, we find that the predicted number of counts is roughly consistent with the number observed. There is no strong evidence for variability of the source. \begin{figure}[tb] \epsscale{0.7} \plotone{f3.ps} \figcaption{Finding chart for SAX~J0635+0533 from the Digitized Sky Survey. Field is $4^{\prime}\times 4^{\prime}$. North is up, east is left. The X-ray error circle has a radius of $1^{\prime}$. The position of the optical counterpart is (J2000) RA = $06^{\rm h}35^{\rm m}18^{\rm s}\!.29$, Dec = $+05^{\circ}33^{\prime}6^{\prime\prime}\!.3$.} \label{fig_finding_chart} \end{figure} \begin{figure*}[tb] \epsscale{1.3} \plotone{f4.ps} \figcaption{Composite optical spectrum of SAX~J0635+0533 obtained on the KPNO 2.1m telescope on 1998 January 29 and 30. In the middle panel, the flux scale refers to the lower trace. The upper trace is the same spectrum multiplied by a factor of 6.} \label{fig_optical_spectrum} \end{figure*} \begin{figure}[tb] \epsscale{0.7} \plotone{f5.ps} \figcaption{Keck HIRES spectrum of selected emission lines, H$\beta$ and He~I~$\lambda$5876, normalized to the continuum and smoothed with a 14~km~s$^{-1}$ running boxcar filter.} \label{fig_keck_spectrum} \end{figure} \section{Optical observations} In order to identify the optical counterpart of SAX~J0635+0533, we obtained optical spectra of seven stars within and around the $1^{\prime}$ radius error circle on 1997 Dec. 10, using the 2.4~m Hiltner telescope of the MDM Observatory. The spectra covered the wavelength range 4500--7500~\AA\ at 4~\AA\ resolution. The brightest star within the error circle (Figure~3) is the only one with broad emission lines and other spectral features that are characteristic of a reddened Be star. Be stars are often hard X-ray sources because of emission from an unseen compact companion such as a neutron star. No other star in the error box has optical spectroscopic properties resembling those of known hard X-ray sources. On this basis we identify the Be star with SAX~J0635+0533. This star is listed in the USNO A1.0 astrometric catalog of the Palomar Observatory Sky Survey (Monet et al. 1996), at position (J2000) RA = $06^{\rm h}35^{\rm m}18^{\rm s}\!.29$, Dec = $+05^{\circ}33^{\prime}6^{\prime\prime}\!.3$. The B1950 coordinates are RA = $06^{\rm h}32^{\rm m}38^{\rm s}\!.11$, Dec = $+05^{\circ}35^{\prime}34^{\prime\prime}\!.3$. The same catalog gives approximate magnitudes of $B = 12.7$ and $R = 12.7$. There is no 20~cm radio source present at this position in the NVSS catalog (Condon et al. 1998). Additional optical spectra were obtained with the Goldcam spectrograph on the KPNO 2.1~m telescope on 1998 January 28, 29, and 30. On the first two nights the spectra covered the wavelength range 3867--7516~\AA\ with a resolution of 5--6~\AA, while on the third night the spectra covered the range 5548--8613~\AA\ with a resolution of 4~\AA. Three sequences of 300~s exposures were taken on each night (at the beginning middle, and end of the night), yielding a total of 39 spectra. Lastly, we obtained one more spectrum on the Hiltner telescope on 1998 April 7. All of the spectra obtained are consistent with each other in continuum flux and emission-line properties. The mean of several of the KPNO spectra is shown in Figure~4, and their emission-line properties are listed in Table~1. The quoted line widths have been corrected for instrumental resolution. The strong H$\alpha$ emission line has a narrow core and very broad wings. It can be fitted well by the sum of a broad and a narrow Gaussian. The narrow component has an equivalent width (EW) of 27~\AA\ and a full-width at half maximum (FWHM) of 330~km~s$^{-1}$, while the broad component has EW = 8~\AA\ and FWHM = 1300~km~s$^{-1}$. The wings of H$\alpha$ extend to a full width at zero intensity (FWZI) of $2100\pm200$~km~s$^{-1}$. A 300~s exposure on SAX~J0635+0533 was obtained by Goeff Marcy using the HIRES instrument (Vogt et al. 1994) on the Keck~I telescope, on 1998 January 26. The resolution is 2~km~s$^{-1}$. Figure~5 shows the profiles of H$\beta$ and He~I~$\lambda$5876, after normalizing them by the local continuum level and smoothing with a 14~km~s$^{-1}$ running boxcar filter. The emission lines are double peaked, with peak separation of $\approx 200$~km~s$^{-1}$ and FWZI $\approx 500$~km~s$^{-1}$. Thus it appears that the narrow component identified in the moderate-resolution spectra as having FWHM $\approx 330$~km~s$^{-1}$ is actually resolved into a disk-like profile, similar to what is seen in the Be star X-ray transient A0535+26 by Clark et al. (1998), and attributed by them to a circumstellar disk around the Be star. The individual KPNO spectra were used to search for radial velocity variations of the H$\alpha$ line over the three nights, but no significant variations were found. The scatter in the measured velocities appears to be random, and defines an upper limit to the real radial velocity variations on this time scale of 30~km~s$^{-1}$. \section{Mean Spectrum, Reddening, and Distance} The {\it observed} Johnson magnitudes (Johnson \& Morgan 1951) measured from the mean spectrum of Figure~4 are $B=13.81$, $V=12.83$, $R=11.98$. Absorption in many diffuse interstellar bands (DIBs) is apparent in Figure~4, and the equivalent widths of these can be used to estimate reddening and distance. We measured four of the best-studied interstellar absorption features; their EWs are listed in Table~1. In addition to the EWs, the depth of the $\lambda 4430$ feature was also measured to be 13\% below the continuum (this quantity is a reddening indicator in addition to the EW). There are several calibrations of the DIBs in the literature. The most extensive work is that of Herbig (1975) who studied the correlation of the EWs of many DIBs with reddening. Wampler (1963, 1966) noted that the calibration of the $\lambda 4430$ feature as a reddening indicator depends on Galactic longitude. More recently, T\"ug \& Schmidt-Kaler (1981) also investigated this issue and came up with their own calibration for the depth of the $\lambda 4430$ feature. We estimated the extinction $E(B-V)$ to SAX~J0635+0533 using each of these calibrations, and adopted a mean value of $E(B-V)=1.2 \pm 0.2$, where the uncertainty corresponds to the dispersion among different calibrations. Assuming that $A_{\rm V} = 3.2\, E(B-V)$, we find $A_{\rm V} = 3.8 \pm 0.6$. The Mon~OB2 association is a cluster of massive stars at a distance of 1620~pc (Garmany \& Stencel 1992), and within $1^{\circ}$ of SAX~J0635+0533. The H~I column densities to 11 stars in this cluster have been measured with IUE using Lyman $\alpha$ absorption and are in the range $1.4-3.8 \times 10^{21} \rm \, cm^{-2}$ (Diplas \& Savage 1984). The estimated $E(B-V)$ to SAX~J0635+0533 corresponds to $N_{\rm H} = 6.0 \times 10^{21}$~cm$^{-2}$ using standard conversions (Savage \& Mathis 1979). Thus we can be confident that SAX~J0635+0533 is no closer than the 1620~pc distance to Mon OB2. The mean column density to the 11 stars is $1.93 \times 10^{21} \rm \, cm^{-2}$. Assuming that $A_{\rm V} = 1$ for $N(H~I) = 1.6 \times 10^{21} \rm \, cm^{-2}$, this corresponds to $A_{\rm V} = 0.8$~mag/kpc. If we assume that 0.8 mag/kpc is a typical value along the line of sight, the distance to SAX~J0635+533 is probably not further than 5~kpc. The total Galactic 21~cm column density in this direction is $7.0 \times 10^{21}$~cm$^{-2}$ (Stark et al.\ 1992), which is larger than the column density estimated for SAX~J0635+0533 from interstellar absorption. Thus, we regard the estimated extinction from the DIBs as compatible with the 21~cm H~I column, while the larger X-ray measured value, $N_{\rm H} = (2.0 \pm 0.3) \times 10^{22}$~cm$^{-2}$, implies that there must be some circumstellar gas that is responsible for the extra X-ray absorption. We note that the emission-line ratio of H$\alpha$/H$\beta$ can be used with recombination theory to derive an upper limit to $E(B-V)$ of 1.4, and that this upper limit is also well below the X-ray measured column. After correction for reddening, the Johnson magnitudes estimated from the spectrum become $B=8.74$, $V=8.99$, and $R=9.14$. Such colors are characteristic of an early B type star. For a distance range of 2.5--5~kpc, the absolute visual magnitude would be in the range $-3.1$ to $-4.5$, and the stellar classification in the range B2V to B1III. \section{Discussion} Assuming 200 Be/X-ray binaries within 5~kpc (Rappaport \& van den Heuvel 1982) distributed uniformly in Galactic longitude and in a latitude distribution with an rms width of $2.2^{\circ}$, we estimate there is a 4\% chance of finding a spurious overlap of a Be/X-ray binary in an error box the size of that for 2EG~J0635+0521. This probability is reasonably low. In addition, both Be/X-ray binaries and the low-latitude EGRET sources are preferentially found near OB associations (Kaaret \& Cottam 1996). This may suggest that some of the unidentified EGRET sources are, in fact, associated with Be/X-ray binaries. While identification of SAX~J0635+0533 and its optical counterpart with 2EG~J0635+0521 must remain tentative awaiting an improved gamma-ray position or detection of periodicity in both x-rays and gamma-rays, the positional coincidence and the hard spectrum of the x-ray source are suggestive and here we entertain some speculations based on the possible identification. Assuming the association of the x-ray source SAX~J0635+0533 with the optical counterpart discussed above and with 2EG~J0635+0521, allows us to find the luminosity of the source based on the optical distance estimate. The 2--10 keV X-ray luminosity would be $9 - 35 \times 10^{33} \rm \, ergs \, s^{-1}$ and the gamma-ray luminosity for 0.1--5~GeV would be $1.3 - 5 \times 10^{35} \rm \, ergs \, s^{-1}$, both assuming isotropic emission. The x-ray to gamma-ray flux ratio is $0.07$. If the identification is correct, then the source would be a gamma-ray emitting x-ray binary. It is possible that the binary contains a neutron star and that both the x-ray and gamma-ray emission are magnetospheric pulsar emission. If the binary is wide, as is tentatively suggested by our non-detection of radial velocity variations, then the pulsar emission may be be relatively uninfluenced by the presence of the companion star. There are several pieces of evidence in favor of pulsar emission. The ratio of x-ray to gamma-ray flux is consistent with those of ``Vela-like'' pulsars. The power-law x-ray spectrum with photon index near 1.5 is similar to that of most isolated gamma-ray pulsars above 2~keV (Wang et al. 1998). The gamma-ray flux is constant over 14 different observations (McLaughlin et al. 1996), as would be expected for a pulsar. And finally, as noted above, 2EG~J0635+0521 was identified by Merck et al. (1996) as one of only eight unidentified Galactic gamma-ray sources with energy spectra sufficiently hard to be pulsar candidates. Alternatively, either the x-ray emission, gamma-ray emission, or both may arise from interaction of the energetic particle wind from the pulsar with the wind or the radiation from the companion. Such mechanisms have been suggested for Be/millisecond radio pulsar binary PSR~J1259-63 and the variable radio Be/X-ray binary system LSI~61$^{\circ}$~303. In PSR~J1259-63, emission up to 200~keV is thought to arise from a shock interaction of the energetic particle wind from a pulsar with the wind from the Be-star (Grove et al. 1995). The x-ray photon index, 1.5--1.9 (Kaspi et al. 1995), is similar to that we find for SAX~J0635+0533. However, PSR~J1259-63 was not detected by EGRET (Tavani et al. 1996) and the x-ray emission is highly variable. It is possible that the shock acceleration mechanism could operate to higher energies if the shock region is less compact than that of PSR~J1259-63 at periastron (Tavani \& Arons 1997). The gamma-ray source 2CG~135+01 has long had a tentative association with the unusually variable radio source and Be/X-ray binary LSI~61$^{\circ}$~303 (Gregory \& Taylor 1978; for recent work see Strickman et al. 1998). It has been suggested (Maraschi \& Treves 1981) that the source contains a young pulsar and that the gamma-rays arise from inverse-Compton scattering of the optical photons from the B-star in the relativistic wind from the pulsar at the boundary between the pulsar wind and the stellar wind. However, 2CG~135+01 is a variable gamma-ray source (Tavani et al. 1998). If 2EG~J0635+0521 is similar to 2CG~135+01, then the constancy of its gamma-ray flux (McLaughlin et al. 1996) would suggest a low eccentricity orbit. Finally, it may be that the emission is powered not by rotation of a neutron star, but rather by accretion. The recently reported detection of GeV pulsations from Cen~X-3 with a frequency equal to that detected contemporaneously in x-rays (Vestrand, Sreekumar, \& Mori 1997), albeit not with an outstandingly high significance, suggests the possibility that 2EG~J0635+0521 could be an accretion powered system. The mechanism for gamma-ray production could be shock acceleration or a magnetospheric process. If SAX~J0635+0533 is similar to Cen~X-3, the x-ray emission would be from accretion and should be pulsed. However, the lower bound we obtained for the exponential cutoff energy of the x-ray spectrum is significantly higher than the cutoffs typically found in high-mass x-ray binaries (White, Swank, \& Holt 1983) and would be more suggestive of either a low magnetic field neutron star or a black hole. Identification of SAX~J0635+0533 with the gamma-ray source 2EG~J0635+0521 would argue against interpreting the gamma-ray emission as due to cosmic-ray production via shock acceleration in the supernova remnant (Esposito et al. 1996; Jaffe et al. 1997). We note that we detect no diffuse hard x-ray emission in the BeppoSAX data. Possible ways to strength the identification of the x-ray/optical source and the gamma-ray source would be to detect a periodicity in the x-ray, optical, or radio which could then be detected in gamma-rays or to extend the x-ray spectrum and join it to the gamma-ray spectrum. Finally, if the gamma-ray emission extends to higher energies, an improved gamma-ray position may be obtained from a ground-based TeV telescope such as the Whipple. \acknowledgments We thank Fabrizio Fiore and the Sax Data Center for assistance with the SAX data. We thank Arlin Crotts for obtaining identification spectra at MDM Observatory, and Geoff Marcy for obtaining the Keck HIRES spectrum. P.~K. acknowledges support from NASA grant NAG5-7389. M.~E. acknowledges support from Hubble fellowship grant HF-01068.01-94A from Space Telescope Science Institute, which is operated for NASA by the Association of Universities for Research in Astronomy, Inc., under contract NAS~5-26255.
\section{\bf Introduction } With Superkamiokande in operation and SNO in preparation the study of solar neutrinos has entered a decisive stage, when neutrino oscillations can be directly discovered. At present we know almost certainly that something happens to neutrinos either inside the Sun or on the way between the Sun and Earth. This knowledge has been mainly provided by solar neurino experiments and by helioseismic observations. \\*[1mm] 1. {\em Helioseismic data} confirm the Standard Solar Models (SSMs) with precision sufficient for reliable prediction of neutrino fluxes. The seismic data (in agreement within a fraction of percent with the SSM ) are valid down to radial distance $0.05 R_{\odot}$, where production of $B$- and $Be$ - neutrinos has maximum, while the other neutrinos are mostly produced at larger distances. At smaller radial distances, where production of neutrinos falls down due to decreasing of volume, the seismic data still exist, though with worse precision. Acoustic frequencies comprise the set of seismic data. Nothing in physics is measured with greater precision than frequencies. This is why seismic measurements give the super-precise data on density and sound speed inside the Sun. Within fraction of percent seismically measured density and\\ ---------------------\\ {\em $^*$ Plenary talk at 19th Texas Symposium, 1998} \newpage \noindent sound speed are different from the SSM predictions (especially at distance $0.7 R_{\odot}$) and this difference is statisticlally significant. It might imply that some physical processes are not included in SSM's, and they are of great interest for physics of the Sun. But not for solar neutrinos! This statistically significant difference, {\em e.g.} in measured and predicted sound speed, produces negligible difference in neutrino fluxes, which is out of interest for any present (and most probably for any future) solar-neutrino experiment. Almost for half century we thought that solar neutrinos with their tremendous penetrating power give us the best way to look inside the Sun. We see now that seismic observations do it with higher precision, while solar neutrinos give the unique information about neutrino properties.\\*[1mm] 2. {\em Nuclear cross-sections} is now the dominat source of uncertainties in the calculated solar-neutrino fluxes. The impressive progress exists here too. In the LUNA experiment at Gran Sasso the cross-section of one of the most intriguing reaction,$^3He+^3\!\!He \to ^4\!\!He+2p$, was measured at energy corresponding to maximum of the Gamow peak in the Sun. The famous speculations about solving or ameliorating the SNP due to increase of this cross-section at very low energy, have been now honorably buried. In the nearest future most of cross-sections relevant to SNP will be measured in the LUNA experiment at very low energy. There is also considerable progress in calculations of cross-sections and screening of nuclear reactions in the solar plasma. A rather exceptional case is cross-section of Hep reaction $p+^3\!\!He \to ^4\!\!He+e^++\nu$, in which neutrinos with the highest energies are produced. Uncertainties in calculation of this cross-section are very large. \\*[1mm] 3. {\em Solar Neutrino Problem (SNP)} is a deficit of neutrino fluxes ( as compared to the SSM prediction) detected in all solar-neutrino experiments (Homestake, SAGE, GALLEX, Kamiokande and Superkamiokande). This deficit is described by factor $\sim 3$ for Homestake and by factor $\sim 2$ for all other experiments.\\*[1mm] 4. {\em Astrophysical solution} to SNP is strongly disfavoured by combination of any two solar-neutrino experiments, {\em e.g.} the boron and chlorine experiments (Superkamiokande and Homestake) or the gallium and boron experiments (GALLEX/SAGE and Superkamiokande). The ratio of $Be$ to $B$ neutrino fluxes, extracted from each pair of experiments mentioned above, is negative (or too small). This is the essence of failure of astrophysical solution. The arbitrary variation of temperature and unknown cross-sections do not solve a problem of $Be/B$ ratio.\\*[1mm] {\em Solar neutrino experiments have a status of disappearance oscillation experiment}.\\*[1mm] But solar-neutrino oscillations are not proved yet. In this paper I will discuss the status of different oscillation solutions to SNP. \section{Status of Astrophysical Solution to SNP} The global rates of four solar-neutrino experiments \cite{Inoue,GALLEX,SAGE,Hom}, as reported up to 1999, are listed in Table 1 and compared with calculations of Bahcall and Pinsonneault 1998 \cite{BP98}. \begin{table}[t] \caption{ The solar-neutrino data of 1998 compared with the SSM prediction , Bahcall and Pinnsoneault 1998 \cite{BP98}. The data of Superkamiokande are given in units $10^6~cm^{-2}s^{-1}$.} \vspace{3mm} \center{\begin{tabular}{||c|c|c|c||} \hline & DATA & SSM [5] & DATA/SSM \\ \hline SUPERK\ & & & \\ &$2.42 \pm 0.08$ &5.15 &$0.47\pm 0.02$\\ GALLEX\ & & & \\ (SNU) &$77.5\pm 7.7$ &129 &$0.60\pm0.06$\\ SAGE\ & & & \\ (SNU) &$66.6\pm 8.0$ &129 &$0.52\pm0.06$\\ HOMEST\ & & & \\ (SNU) &$2.56\pm 0.23 $ &7.7 &$0.33\pm0.03$ \\ \hline \end{tabular}} \end{table} The deficit of detected neutrino fluxes seen in the last column of Table 1 is impossible to explain by astrophysics or/and nuclear physics. This conclusion is based on the following.\\*[1mm] \noindent (i) Compatibility of the boron (Superkamiokande) and chlorine (Homestake) signals or boron and gallium (GALLEX/SAGE) signals results in unphysically small ratio $\Phi_{\nu}(Be)/\Phi_{\nu}(B)$ \cite{BaBe}-\cite{HaLa98}. The best fit value of this ratio is negative. The statement above is model-independent.\\*[1mm] (ii) The arbitrary variation of unknown nuclear cross-sections and the central temperature cannot bring $\Phi_{\nu}(Be)/\Phi_{\nu}(B)$ ratio in agreement with observations \cite{lasthope}.\\*[1mm] (iii) Seismic observations of the density and sound speed radial profiles confirm SSM at distances down to $0.05 R_{\odot}$ with accuracy better than fraction of percent \cite{Dz}-\cite{Tu-Ch} and at the center with accuracy better than $4\%$ for sound speed \cite{ferrara1}. \\*[1mm] (iv) As minimum, SSM is a good approximation to realistic model of the Sun. In this case there must be a track, when changing the parameters of SSM and/or introducing the new physical phenomena, one arrives from SSM neutrino fluxes to the observed ones. Such a track does not exist \cite{HaLa98}.\\*[1mm] We shall analyze the last item at some details. But two comments are in order now. There are two other, more model-dependent arguments against astrophysical solution to SNP. If one takes three major components of neutrino fluxes ($pp$, $Be$ and $B$) as independent and positive, and the CNO neutrino flux (which gives much smaller contribution) - according to SSM, then arbitrary variation of those three fluxes do not give acceptable fit to the observational data at 99.99$\%$CL \cite{BKS}. The deficit of $B$ neutrinos seen in the Superkamiokande data (Table 1) is another problem for astrophysical solution. Some time ago many people thought that with extreme and correlated uncertainties in $pBe$-cross-section and in the central temperature $T_c$ this discrepancy can be eliminated. Now the situation looks like follows. In the helioseismically constrained solar models (HCSM) \cite{HCSM} the central temperature $T_c= 1.58 \cdot 10^7~K$ within maximum uncertainty $\Delta T_c/T_c = 1.4 \%$. Taking $T_c$ 1.4$\%$ lower and $S_{17}$ 40$\%$ lower we obtain the minimum $B$-neutrino flux $3.0 \cdot 10^6~cm^{-2}s^{-1}$, {\em i.e.} 7.4$\sigma$ higher than the measured one. \vspace{-18mm} \begin{figure}[htb] \epsfxsize=9.5truecm \centerline{\epsffile{figure1.eps}} \vspace{-40mm} \caption{\em Neutrino fluxes allowed by arbitrary $^3\!He$ mixing accompanied by independent variations of temperature, $S_{34}$ and $S_{17}$. The solid lines limit the allowed region in case $^3\!He$ radial profile is the same as in SSM, and the dashed lines limit the region when both, temperature and $^3\!He$ profiles are varying. Some trajectories from the SSM allowed regions are shown for illustration. The best fit is given by point 11, separated by more than $6 \sigma$ from experimentally allowed region (shown as ''$3 \sigma$ exp.'' in the figure).} \end{figure} The most recent attempt \cite{Gough,Schatz,CuHa} to reconcile the astrophysical solution with measured neutrino fluxes involves an old idea of $^3\!He$-mixing in the solar core. In SSM's $^3\!He$-abundance is very low in the $Be,B$-neutrino production zone. $^3\!He$ is accumulated at much larger distance $r \sim 0.3R_{\odot}$. It is assumed that due to some process (it could be gravity wave induced diffusion \cite{Schatz} or non-linear instability \cite{Gough}) the ''fresh'' $^3\!He$ is brought into solar core. It could happen as the short repeating episodes. Then neutrinoless channel in nuclear reactions, $^3\!He + ^3\!He \to ^4\!\!He + 2p$, is enhanced and the central temperature $T_c$ decreases too. A general analysis of astrophysical solution, which includes arbitrary $^3\!He$-mixing has been recently performed in ref.(\cite{BFL99}). The $^3\!He$-mixing was assumed not to be accompanied by mixing of other elements. Additionally all other relevant parameters in the neutrino production zone were varying within wide range: $T_c$ - within $\pm 5\%$, $S_{17}$ - within $\pm 40\%$ and $S_{34}$ - in the range $(-20\% +40\%)$. The temperature and $^3\!He$ radial profiles were also varying. The results are presented in Fig.1 as allowed regions between two limiting curves, thin solid ones (''temp.'') and two broken ones (''$^3\!He$ profile''). The best fit is at least $6 \sigma$ away from observationally allowed region. It can be interpreted as well that there is no allowed track from the SSM's region (''SSM'') to the observationally allowed region (see Fig.1). A trajectory with variation of temperature $T_c$ is shown by thick solid line (''temp.'') \section{Oscillation Solutions} Due to oscillations, electron neutrino emitted from the Sun can be found at the Earth as neutrino with another flavor: muon, tau or sterile neutrino. These neutrinos either do not give a signal in the detector ({\em e.g.} muon neutrinos in gallium or chlorine detectors) or interact weaker due to NC ({\em e.g.} muon- or tau-neutrinos in Superkamiokande). I will not discuss in this review sterile neutrinos. Atmospheric neutrino oscillations imply that $\nu_{\mu}$ and $\nu_{\tau}$ neutrinos are maximally mixed. In this case solar $\nu_e$ neutrino oscillates with equal probability to each of those neutrinos. The probability to find emitted electron neutrino in the same flavor state in the detector $P_{\nu_e \to \nu_e}$ is called survival probability or (less precisely) suppression factor. In general case survival probability depends on energy, $P_{\nu_e \to \nu_e}(E)$, {\em i.e.} solar neutrinos are suppressed in energy-dependent way and actually this property allows to solve SNP with help of oscillations. Four oscillation solutions are currently discussed in the literature (see Table 2.) \vspace{-3mm} \begin{table}[htb] \caption{ Oscillation Solutions to SNP.} \center{\begin{tabular}{|c|c|c|} \hline & $\sin^2 2\theta$ & $\Delta m^2$ ($eV^2$)\\ & best fit & best fit \\ \hline MSW $\;\;\;\;$\ & & \\ $\;\;\;$ SMA & $ 6.0\cdot 10^{-3}$ & $5.4\cdot 10^{-6}$\\ $\;\;\;$ LMA & $ 0.76$ & $1.8\cdot 10^{-5}$\\ $\;\;\;$ LOW & $ 0.96$ & $7.9\cdot 10^{-8}$\\ VO $\;\;\;\;\;\;\;$\ & & \\ $\;\;\;$Just-so & $0.75$ & $8.0\cdot 10^{-11}$\\ $\;\;$EIS & $\sim 1$ & $ 10^{-9} - 10^{-3}$\\ RSFP $\;\;\;$ & & \\ & small & $ 10^{-8} - 10^{-7}$\\ \hline \end{tabular}} \end{table} \par \vspace*{-0.5cm} 1. {\em MSW solution} \cite{MSW}.\\ MSW effect in the Sun is a resonance conversion of $\nu_e$ into $\nu_{\mu}$ or $\nu_{\tau}$. For neutrino energies at interest it occurs in the narrow layer, $\Delta R \sim 0.01 R_{\odot}$, at the distance $R \sim 0.1 R_{\odot}$ from the center of the Sun. There are three MSW solutions to SNP , which explain the global rates in all four solar-neutrino experiments: Small Mixing Angle (SMA) MSW, Large Mixing Angle (LMA) MSW and LOW solution with low probability (it appears only at $99\%CL$). The best fits of these solutions to the rates are reported in Table 2. 2. {\em Vacuum Oscillations (VO)}\\ The concept of vacuum oscillations was first put forward by B.Pontecorvo \cite{Pont} (for a review see \cite {BiPe}). The survival probability for $\nu_e$ neutrino with energy $E$ at distance $r$ is given by \begin{equation} P_{\nu_e\to\nu_e}= 1 - \sin^2 2\theta \sin^2 \left( \frac{\Delta m^2} {4E} r \right), \label{eq:srv} \end{equation} where $l_v= 4\pi E/\Delta m^2$ is the vacuum oscillation length. At $\Delta m^2 = 8\cdot 10^{-11}~eV^2$ (the best fit) the oscillation length of neutrino with energy $E_{\nu} \sim 3~MeV$ is $\l_{\nu} \sim 1\cdot 10^{13}~cm$, {\em i.e.} of order of distance between the Sun and Earth. That is why this VO solution is called {\em just-so}. Since observational data need large suppression of neutrino flux, by factor $\sim 2$, $\sin^2 2\theta \sim 1$ is needed: see Eq.(\ref{eq:srv}). Thus just-so VO solution must be large mixing angle solution. The best fit values are given in Table 2. 3. {\em EIS VO solution}\\ VO with Energy Independent Suppression (EIS) occurs when $\Delta m^2 \gg 10^{-10}~eV^2$. In this case the oscillation length is much smaller than distance between the Sun and Earth. Oscillatory function in Eq.(\ref{eq:srv}) is averaged to factor 1/2 and hence suppression factor is energy independent and equals to $1-\frac{1}{2}\sin^2 2\theta$. Since this suppression should be of order 0.5, $\;\; \sin^2 2\theta \sim 1$ is needed. On the other hand one must assume $\Delta m^2 \ll 10^{-3}~eV^2$ because of non-observation of $\nu_e$ oscillation in the atmospheric neutrinos. The energy independent suppression is excluded by observed rates at $99.8\% CL$ \cite{BKS}. However, the Homestake data give the dominant contribution to this conclusion. If these data are arbitrarily excluded from analysis, EIS VO survives. I will not give more attention to discussion of EIS VO solution. Further details a reader can find in references \cite{Perk,FoVo,Comf,Gold}. 4. {\em RSFP solution}\\ The Resonant Spin-Flavor Precession (RSFP) describes two physical effects working simultaneously: the spin-flavor precession, when neutrino spin (coupled to magnetic moment) precesses around magnetic field, changing simultaneously neutrino flavor, and the resonant, density-dependent effect, which produces difference in potential energy of neutrinos with different flavors (similar to the MSW effect). This complex transition occurs in the external magnetic field due to presence of non-diagonal (transition) neutrino magnetic moments. The RSFP was first recognized in ref.'s \cite{Akh,Lim}. For excellent review see \cite{Akh-rev}. This theory had a predecessor. The precession of neutrino magnetic moment around magnetic field converts left-chiral electron neutrino $\nu_{eL}$ into sterile right component $\nu_{eR}$, suppressing thus $\nu_e$-flux\cite{VVO}. However, the suppression effect in this case is energy-independent and thus contradicts to the observed solar-neutrino rates. In ref's \cite{VVO,BaFi} the matter effect was included and in \cite{SchVa} spin-flavor precession was discovered. The observed rates in solar-neutrino experiments can be explained only by the RSFP, because only this type of precession give the energy-dependent suppression factor. Majorana neutrino can have only transition magnetic moment. RSFP induces the transition $\nu_{eL}$ to $\bar{\nu}_{\mu R}$, i.e. electron neutrino to muon antineutrino, which can scatter off the electron due to NC. The survival probability is similar to that of SMA MSW (see Fig.3 from \cite{Nun}). Neutrino mixing is not needed directly for RSFP effect, but it is needed indirectly to provide the transition magnetic moment of the Majorana neutrino. To be a solution to SNP, RSFP needs a transition magnetic moment $\mu \sim 10^{-11} \mu_B$, magnetic field in the resonance layer of the Sun, $B \sim 20 - 100~kG$ and $\Delta m^2$ in the range $10^{-8} - 10^{-7}~eV^2$ (see Section 6). \section{Signatures of Oscillation Solutions} A common signature of most neutrino oscillation solutions is distortion of $B$-neutrino spectrum. The survival probabilities for SMA MSW, LMA MSW and just-so VO are shown in Fig.2. The survival probabilities for RSFP are similar to SMA MSW and shown in Fig.3. One can see there that LMA MSW (and LOW too) predicts small distortion of $B$-neutrino spectrum spectrum in the region of observation $5- 15~MeV$. For EIS VO the distortion is absent. The strongest spectrum distortion one can expect for SMA MSW and just-so VO. However, spectrum of recoil electrons are distorted weaker than that of neutrinos, because of cross-section and averaging over energy bins in observations ({\em e.g.} see Fig.8). {\em The absence of distortion of neutrino or recoil-electron spectra is not a general argument against neutrino oscillations}. {\em Anomalous NC/CC ratio} is another common signature of neutrino oscillations which can be observed in SNO. The NC events will be seen there by detection of neutrons produced in $\nu+D \to p+n+\nu$ reaction. Oscillation $\nu_e \to \nu_{\mu} (\nu_{\tau})$ does not change NC interaction but changes CC \newpage \begin{figure}[t] \begin{center} \psfig{bbllx= 7pt, bblly=50pt, bburx=550pt, bbury=720pt, file=figure2.ps, height=8.5cm , clip=} \end{center} \vspace{-15mm} \caption{\em Electron neutrino survival probabilities.} \end{figure} \vspace*{-10mm} \samepage \vspace*{-10mm} \begin{figure}[h] \epsfxsize=7.6truecm \centerline{\epsffile{figure3.eps}} \vspace{-10mm} \caption{\em RSFP survival probabilities from \cite{Nun} compared with SMA MSW (best fit) survival probability, shown by thick solid line. The four other curves correspond to different radial profiles of magnetic field.} \vspace*{-60mm} \end{figure} \vspace*{-10mm} \newpage \noindent interaction and thus the ratio of NC/CC rate. In case of oscillation to sterile neutrino the NC/CC ratio is not changed. Therefore, {\em the normal ratio NC/CC is not a general argument against neutrino oscillations}. {\em MSW solutions} have very distinct signatures. They are day/night effect, zenith angle dependence of solar-neutrino flux and difference in day/night neutrino spectra. All these effects are caused by MSW matter effect in the Earth. Another related effect is seasonal variation of neutrino flux caused by longer nights in winters. This effect is smaller than integrated day/night effect. For recent calculations see \cite{Valle} The signature of {\em just-so VO} is anomalous seasonal variation of neutrino flux \cite{Pom}. Due to ellipticity of the Earth's orbit, the distance between the Sun and Earth changes with time, causing $7\%$ variation of the flux due to $r^{-2}$ effect (''geometrical'' seasonal variation). Due to just-so VO the flux of $\nu_e$ neutrinos changes additionally due to dependence of survival probability (\ref{eq:srv}) on distance. As follows from Eq.(\ref{eq:srv}) neutrinos with different $E$ have different phases and it weakens the observed effect, when averaged over interval $\Delta E$. In case of monochromatic $Be$ neutrinos VO seasonal time variations are strongest For the detailed calculations of anomalous seasonal variations see \cite{KrPe} and for the recent calculations \cite{GlKe,MiSm,GeRo,Sm,Fo,MaPe}. For {\em EIS VO solution} the anomalous seasonal variations are absent, because the oscillation length is too small. It results in a signature, which can be observed by BOREXINO: $Be$-neutrino flux is suppressed by a factor $\sim 2$, but does not show anomalous seasonal variations. \samepage {\em RSFP} has two signatures. As a result of RSFP electron neutrinos oscillate inside the Sun into muon/tau antineutrinos. Due to vacuum oscillations on the way to the Earth these neutrinos oscillate to electron antineutrinos. The latter oscillations are suppressed by mixing angle, which is small in case of RSFP. However, even small fluxes of electron antineutrino can be reliably detected ({\em e.g.} by KamLand \cite{ASuz}). \samepage The second signature is prediction of 11-year periodicity for $Be$-neutrino flux \cite{Akh-rev}.\\ \newpage \noindent RSFP occurs in the resonant layers, which are located at different distances for neutrinos of different energies: for $pp$ and $B$ neutrinos the resonant layers are located near the solar center and at the periphery, respectively, where magnetic field is weak and RSFP too. The resonant layer for $Be$ neutrinos is located at intermediate distance, where magnetic field is large and RSFP is strongest. 11-year variations of neutrino flux is caused by periodic variation of toroidal magnetic field at the bottom of convective zone. The magnetic activity of the sun exhibits quasi-periodic time variations with the mean period $11~yr$. This periodicity is thought to be originated due to toroidal field, generated in so-called {\em overshoot layer} by dynamo mechanism and located near the bottom of convective zone. Theoretically, magnetic field there can reach $100~kG$. This field rises through convective zone to the surface of the sun. The $11~yr$ periodicity should be observed most effectively by neutrino detectors sensitive to $Be$-neutrinos: Homestake, GALLEX and BOREXINO. In particular (see Fig.3) when toroidal magnetic field disappears (due to change of magnetic polarity) survival probability increases from $\sim 0.1$ to $\sim 1$. Since $B$-neutrino flux is also suppressed by factor $\sim 0.4 - 0.6$, $11~yr$ variations should be seen in the combined Kamiokande and Superkamiokande data. \section{708-day Superkamiokande Data} After 708 days of solar-neutrino observations Superkamiokande has not found direct evidences for neutrino oscillations. There are only some indications to the distortion of the spectrum of recoil electrons, which will be discussed in this Section. \samepage {\em The spectrum of the recoil electrons} (708 d) is shown in Fig.4 \cite{Inoue,Tots} as the ratio to (undistorted) spectrum calculated in BP98 SSM \cite{BP98}. The spectrum is suppressed by overall factor 0.47, but there is no distortion of the spectrum, except the high energy excess at $E_e \geq 13~MeV$. In principle, this excess can be a result of low statistics or small systematic errors at the end of the boron neutrino spectrum. For example, due to very steep end of the electron spectrum, even small systematic \newpage \noindent error in electron energy (e.g. due to calibration) could enhance the number of events in the highest energy bins. Another possible explanation of this excess~\cite{BaKr} is that the Hep neutrino flux might be significantly larger (about a factor 10--20) than the SSM prediction. The Hep flux depends on solar properties, such as $^3$He abundance and the temperature, and on $S_{13}$, the zero-energy astrophysical $S$-factor of the \mbox{$p+{}^3He \rightarrow {}^4He + e^+ + \nu$} reaction. Both SSM based~\cite{BaKr} and model-independent~\cite{BDFR} approaches give a robust prediction for the ratio $\Phi_{\nu}(Hep)/S_{13}$. Therefore, this scenario implies a cross-section larger by a factor 10--20 than the present calculations. Such a large correction to the calculation does not seem likely, though is not excluded. The signature of Hep neutrinos, the presence of electrons above the maximum boron neutrino energy, can be tested by the SNO experiment. The observed excess is difficult to explain by neutrino oscillations. The oscillation parameters $\sin^2 2\theta$ and $\Delta m^2$, which correspond to the allowed global rates in four neutrino experiments, result in the recoil electron spectra in bad agreement with the excess. The spectra for the best fit MSW solutions (LMA shown by short-dash lines and SMA -- by long-dash lines) are displayed in Fig.4 (calculations by K.Inoue). {\em Night/Day excess} after 201.6, 504 and 708 days of observations, is found to be small, consistent with zero within $1.7 \sigma$: \vspace{-3mm} $$ \frac{N-D}{N+D}\times 100 =- 0.4 \pm 3.1 \pm 1.7 \;\;\;\;\; (201.6 d) $$ \vspace{-3mm} $$ \frac{N-D}{N+D}\times 100 =+ 2.3 \pm 2.0 \pm 1.4 \;\;\;\;\; (504 d) $$ \vspace{-3mm} $$ \frac{N-D}{N+D}\times 100 =+ 2.9 \pm 1.7 \pm 0.39\;\;\;\;\; (708 d) $$ The observed (708 d) excess has the right sign, but is statistically insignificant ($1.7\sigma$). Note that systematic error is much smaller than observed effect. Statistics, if increased by factor 5, can make the effect statistically significant. For clear discussion see \cite{Lisi}. {\em Absence of day/night effect does not rule out MSW solution.} The zenith-angle dependence is not seen in Superkamiokande data. \newpage The Superkamiokande data for day/night effect, zenith-angle dependence and recoil-electron energy spectrum (especially absence of distortion in most part of the spectrum) have the great restriction power for the oscillation solutions. \vspace*{-5mm} \begin{figure}[h] \begin{center} \psfig{bbllx= 55pt, bblly=175pt, bburx=540pt, bbury=655pt, file=figure4.ps, height=7.5cm , clip=} \end{center} \vspace{-13mm} \caption{\em Energy spectrum of recoil electrons from 708d of Superkamiokande data \cite{Inoue,Tots}. Plotted is ratio of the observed spectrum to one predicted by BP SSM \cite{BP98}. The long- and short- dash lines show SMA MSW and LMA MSW spectra, respectively, for best-fit rates (calculations by K.Inoue).} \end{figure} \vspace*{-10mm} \section{Status of Oscillation Solutions} I will discuss here the status of MSW, VO and RSFP solutions in the light of 708d Superkamiokande data.\\*[1mm] {\bf 6.1 MSW solutions}\\*[1mm] The regions in oscillation parameter space allowed by global rates do not explain the high energy excess in the recoil-electron spectrum \cite{Suz} (see Fig.4). MSW solutions need an alternative explanation of this excess, {\em e.g.} by Hep neutrinos (see Section 5). The status of MSW solutions is determined by {\em combined} restrictions due to global rates, day/night effect, zenith-angle flux dependence and the energy spectrum (under \newpage \noindent assumption of arbitrary $S_{13}$). The result of such analysis is expressed in goodness of the total fit $\chi^2$. \vspace{-13mm} \begin{figure}[h] \epsfxsize=11.5truecm \centerline{\epsffile{figure5.eps}} \vspace{-15mm} \caption{\em Status of MSW oscillation solutions after 504 days of Superkamiokande data \cite{BKS}.} \end{figure} \vspace{-5mm} The data of Superkamiokande for 504 days disfavoured LMA MSW solution \cite{BKS}. In Fig.5 the upper panel shows the regions allowed (at $99\%$ CL) by the global rates. The middle panel includes additionally restrictions due to spectrum and the low one includes three restrictions (rates, spectrum and zenith-angle dependence). All regions are shown at $99\%$ CL. One can see how LMA and LOW solutions disappear. In Fig.6 the 708d data of Superkamiokande are shown as allowed by rates (upper panel), the regions excluded by day/night effect (middle panel) and excluded by spectrum (low panel). Note, that in the middle panel the regions above and below the central one are allowed at $68\%$ CL. \newpage \begin{figure}[ht] \begin{center} \includegraphics[bbllx=30pt,bblly=154pt,bburx=535pt,bbury=660pt, height=50mm,width=65mm]{figure6.ps} \end{center} \vspace{-2mm} \begin{center} \includegraphics[bbllx=30pt,bblly=154pt,bburx=535pt,bbury=660pt, height=50mm,width=65mm]{figure7.ps} \end{center} \vspace{-2mm} \begin{center} \includegraphics[bbllx=30pt,bblly=154pt,bburx=535pt,bbury=660pt, height=50mm,width=65mm]{figure8.ps} \end{center} \vspace{-12mm} \caption{\em Status of MSW solutions after 708 days of Superkamiokande data (courtesy of Y.Totsuka)} \vspace{-20mm} \end{figure} \vspace{-40mm} \newpage \noindent The visual inspection shows that all three MSW solutions are allowed. However, the quantitative analysis of 708d data presented by K.Inoue \cite{Inoue} shows that the SMA MSW solution is not acceptable at $90\%$ CL if day/night effect and spectrum (with free Hep flux) are included in the analysis simultaneously. LMA MSW solution is more favourable. In conclusion, exclusion of any MSW solution looks unstable and statistics-dependent. I mean that conclusions are changing too drastically with accumulation of data and with method of analysis (inclusion or not day and night spectra, inclusion Hep flux as a free parameter {\em etc}). Inclusion of too many data together may be misleading if the data are partially inconsistent. Finally, I would like to remind a reader that the data of Superkamiokande are still preliminary, and conclude that it is premature to speak of exclusion of any MSW solutions.\\*[1mm] {\bf 6.2 VO Solutions}\\*[1mm] If high energy excess in the spectrum is due to Hep neutrinos, just-so VO solution fits the rates and the spectrum. In case the excess is due to oscillations, the regions in oscillation parameter space allowed by the rates (Fig.7, upper panel) are excluded by energy spectrum (low panel). The spectrum is well fitted by vacuum oscillations with $\Delta m^2=4.2\cdot 10^{-10}~eV^2$ and $sin^2 2\theta=0.93$ \cite{Suz}, but this point is located outside the regions allowed by the rates (Fig.7), {\em i.e.} it does not represent the SNP solution. The status of this point has been further analysed in \cite{BFL98}. To explain both the excess and the rates it was assumed that boron neutrino flux is 15--20\% smaller than the SSM prediction, and that the chlorine signal is about 30\% larger than the Homestake observation. This assumed $3.4\sigma$ increase of the chlorine signal could have a combined statistical and systematic origin. In practice, the SSM boron neutrino flux and the Homestake signal were rescaled with help of parameters $f_B$ and $f_{Cl}$, as $\Phi_B= f_B \Phi_B^{SSM}$ and $R_{Cl} = 2.56 f_{Cl}$~SNU, where 2.56~SNU is the Homestake signal. \newpage For each pair $f_B$ and $f_{Cl}$ there were found the parameters ($\Delta m^2, \sin^22\theta$), that explain the observed rates, and $B$-neutrino spectrum was calculated for these parameter values. \samepage \begin{figure}[h] \begin{center} \includegraphics[bbllx= 15pt, bblly=150pt, bburx=540pt, bbury=660pt, height=60mm,width=75mm]{figure9.ps} \end{center} \begin{center} \includegraphics[bbllx= 15pt, bblly=150pt, bburx=530pt, bbury=660pt, height=60mm,width=75mm]{figure10.ps} \end{center} \vspace{-5mm} \caption{\em Just-so VO solution: regions allowed by rates (upper panel) and excluded by spectrum (low panel)-- courtesy of Y.Totsuka.} \vspace{-28.1mm} \end{figure} \vspace{-40mm} \newpage In particular, for $f_B=0.8$ and $f_{Cl}=1.3$ the oscillation parameters ($\Delta m^2=4.2 \cdot 10^{-10}$~eV$^2, \sin^22\theta=0.93$) give a good fit to all rates ($\chi^2$/d.o.f. = 3.0/3) and to the spectrum with the excess (see Fig.8). More generally, the oscillation parameters give rates in agreement with the experiments at the $2\sigma$ level when $0.77 \leq f_B \leq 0.83$ and $1.3 \leq f_{Cl} \leq 1.55$. These VO solutions will be referred to as HEE VO (with HEE for High Energy Excess) to distinguish them from ordinary just-so VO solutions. \vspace{-8mm} \begin{figure}[h] \begin{center} \psfig{bbllx= 100pt, bblly=270pt, bburx=510pt, bbury=660pt, file=figure11.ps, height=7cm , clip=} \end{center} \vspace{-13mm} \caption{\em Ratio of the vacuum oscillation spectra to the SSM spectrum. The solid curve corresponds to the HEE VO solution with $\Delta m^2=4.2\cdot 10^{-10}$~eV$^2$ and $\sin^2 2\theta=0.93$. The dashed and dotted curves correspond to the VO solutions of Refs.~\cite{HaLa98} and \cite{BKS}, respectively.} \end{figure} \vspace{-8mm} \noindent The anomalous seasonal variations are rather unusual in the HEE VO solution. They are described by time dependence of survival probability for the electron neutrino: $P_{\nu_e \to \nu_e}(t)$. In particular, for $Be$-neutrinos with energy $E=0.862$~MeV it equals to \begin{equation} P(t)= 1-\sin^2 2\theta \sin^2\left( \frac{\Delta m^2 a}{4E}\, (1+e\cos \frac{2\pi t}{T} ) \, \right) \label{seas} \end{equation} where $a=1.496\cdot 10^{13}$~cm is the semi-major axis, $e=0.01675$ is the eccentricity of the Earth's orbit, and $T=1$~yr is the orbital period. As seen in Fig.9, the case of the HEE VO (solid curve) is dramatically different from the just-so VO case: there are two maxima and minima during one year and the survival probability oscillates between $1-\sin^2 2\theta \approx 0.14$ and 1. The explanation is obvious: the HEE VO solution has a large $\Delta m^2$, which results in a phase $\Delta m^2 a/(4E) \approx 93$, large enough to produce two full harmonics during one year, when the phase changes by about 3\% due to the factor $(1+e\cos 2\pi t/T)$. The HEE VO solution predicts (see Fig.9) that $Be$ electron neutrinos should arrive almost unsuppressed during about four months a year! According to the SSM, beryllium neutrinos contribute 34.4~SNU out of the total gallium signal of 129~SNU. Therefore, the strong $^7$Be neutrino oscillation predicted by the HEE VO solution also implies an appreciable variation of total gallium signal. In Fig.9 the dotted curve shows this variation for the HEE VO solution, which can be compared with the weaker variation in the just-so VO solution (dashed-dotted curve). \vspace{-10mm} \begin{figure}[hb] \begin{center} \psfig{bbllx= 100pt, bblly=270pt, bburx=560pt, bbury=660pt, file=figure12.ps, height=6cm,width=7.5cm} \end{center} \vspace{-12mm} \caption{\em Anomalous seasonal variations of the beryllium neutrino flux and gallium signal for the VO and HEE VO solutions. The survival probability $P_{\nu_e\to \nu_e}$ for Be neutrinos is given for the HEE VO (solid curve) and for just-so VO (dashed curve) as function of time ($T$ is an orbital period). The dotted (dash-dotted) curve shows the time variation of gallium signal in SNU for the HEE VO and for just-so VO~\cite{BKS} solutions.} \end{figure} \vspace{-40mm} \newpage In Figs. 10 - 15 the predictions of HEE VO solution for seasonal time variations are compared with observations of GALLEX, SAGE, Homestake and Superkamiokande. While the agreement of each of the observational data with the HEE VO solution might appear accidental and not sta- \begin{figure}[h] \begin{center} \includegraphics[bbllx=105pt,bblly=260pt,bburx=510pt,bbury=660pt, height=45mm,width=75mm]{figure13.ps} \end{center} \vspace{-10mm} \caption{\em Seasonal variations predicted by the HEE VO in comparison with the GALLEX data \cite{GALLEX}. The fit with the HEE VO has $\chi^2$/d.o.f.=0.87/4, while a time independent fit gives $\chi^2$/d.o.f.=1.36/5.} \vspace{-5mm} \begin{center} \includegraphics[bbllx=105pt,bblly=260pt,bburx=510pt,bbury=660pt, height=45mm,width=75mm]{figure14.ps} \end{center} \vspace{-10mm} \caption{\em Seasonal variations predicted by the HEE VO in comparison with the SAGE preliminary data \cite{SAGE}. The fit with the HEE VO has $\chi^2$/d.o.f.=8.9/5, while a time independent fit gives $\chi^2$/d.o.f.=3.8/5.} \end{figure} \vspace{-40mm} \newpage \noindent tistically significant, the combined agreement between this model and experiment looks like indication in favour of the HEE VO solution. \begin{figure}[ht] \begin{center} \includegraphics[bbllx=105pt,bblly=260pt,bburx=510pt,bbury=660pt, height=43mm,width=75mm]{figure15.ps} \end{center} \vspace{-10mm} \caption{\em Seasonal variations predicted by the HEE VO in comparison with the Homestake data \cite{Hom}. The fit with the HEE VO gives $\chi^2$/d.o.f.=1.4/5, while the fit with constant (no-oscillation) gives $\chi^2$/d.o.f=3.1/5.} \begin{center} \includegraphics[bbllx=100pt,bblly=260pt,bburx=510pt,bbury=654pt, height=43mm,width=75mm]{figure16.ps} \end{center} \vspace{-8mm} \caption{\em Seasonal variations predicted by the HEE VO for $E_e > 10~MeV$ in comparison with the Superkamiokande data at the same energies. The fit with the HEE VO gives $\chi^2$/d.o.f.=2.7/7, while the one with the geometrical effect only gives $\chi^2$/d.o.f.=2.3/7.} \end{figure} \vspace{-40mm} \newpage The anomalous seasonal variation of $\nu_{Be}$- flux predicted by the HEE VO will be reliably tested by BOREXINO and LENS. \vspace{-10mm} \begin{figure}[h] \begin{center} \psfig{bbllx= 90pt, bblly=260pt, bburx=530pt, bbury=670pt, file=figure17.ps, height=5.5cm, width=7.5cm} \end{center} \vspace{-10mm} \caption{\em The same as in Fig.13 for $E_e > 11.5~MeV$. The fit with the HEE VO gives $\chi^2$/d.o.f.=2.7/7, while the one with the geometrical effect only gives $\chi^2$/d.o.f.=2.3/7.} \vspace{-10mm} \end{figure} \vspace{2mm} {\bf 6.3 RSFP solution}\\*[1mm] As was recently demonstrated \cite{Nun}, the RSFP solution can successfully explain the rates (see also \cite{Babu,Petc} for early calculations ) and high-energy excess in the Superkamiokande spectrum. This solution has more free parameters to fit the data. For the Majorana neutrino they are: $\Delta m^2$, transition magnetic moment $\mu_{\nu}$, scale of toroidal magnetic field in the convective zone, $B$, and radial profile for magnetic field, $B(r)$, in the wide range of distances. The mixing angle is an arbitrary parameter in the RSFP solution which determines the magnetic moment, but it must be small enough, $sin 2\theta < 0.25$ \cite{Akh-rev}. In Fig.15 the calculated recoil-electron spectra (for four magnetic radial profiles) are compared with the 504d Superkamiokande data. The agreement is reasonably good, though from deflections of the individual points one can guess that $\chi^2$ is not very small. In Fig.16 the regions explaining the rates in four solar-neutrino experiments are shown in parameter space $\Delta m^2$ and the mean magnetic field $<B>$ for two magnetic radial \newpage \noindent profiles \cite{Nun}. One can see there the allowed range of parameters. One of the signatures of the RSFP solution, the time-variation of the neutrino signal, is probably testable now. There are two widely discussed effects: 11 year periodicity and (June+December)/(March+Sept) ratio of fluxes. \samepage There are some indications to 11yr periodicity in the Homestake signal, especially when correlations with various solar phenomena (sun spot number, green coronal line, surface magnetic field {\em etc}) are included. My personal opinion is that correlation is allowed as an argument, if the the time variability of the signal is established. Such a lesson was taught us by a bitter experience with Cyg X-3, when correlation with X-ray variability was used as a proof of high-energy gamma-ray signal. Meanwhile, the Homestake signal is perfectly compatible \cite{Lis} with constant flux ($\chi^2/d.o.f. = 0.6$). The data of all other detectors are also consistent with time-independent flux. \vspace{-10mm} \begin{figure}[htb] \epsfxsize=7.5truecm \centerline{\epsffile{figure19.eps}} \vspace{-10mm} \caption{\em RSFP spectra for recoil electrons for four different magnetic radial profiles, compared with 504d Superkamiokande data.} \vspace{-10mm} \end{figure} The suppression of neutrino flux in the RSFP model disappears when magnetic field vanishes. It happens in two cases: when polarity of magnetic field in the Sun changes and when neutrino flux arrives, propagating in the plane of solar equator ( June 5 and December 5). This effect is strongest for $Be$-neutrinos (see Fig.3). \newpage The GALLEX data do not show an excess of the rate in June and December or the deficit in March and September. Using three month intervals centered at June 5 and December 5 (''high'' rates) and at March 5 and September 5 (''low'' rates) the GALLEX collaboration has obtained as a mean rate $78.5 \pm 12~SNU$ for 20 ''high'' runs and $90 \pm 12~SNU$ for 19 ''low'' runs, {\em i.e.} within limited statistics the wrong-sign effect (T.Kirsten, private communication). In Figs. 13--14 one can see a similar wrong-sign effect in Superkamiokande data. On the other hand there are no accurate model calculations of this effect in the RSFP models, to compare with the data above. \vspace{-15mm} \begin{figure}[ht] \epsfxsize=7truecm \centerline{\epsffile{figure18.eps}} \vspace{-15mm} \caption{\em The RSFP regions allowed by rates for two magnetic field radial profiles \cite{Nun}}. \end{figure} \vspace{-13mm} \section{\bf Future Experiments} {\em SNO}, e.g. \cite{McD}, is 1 kt heavy water detector, which will start to operate this year. In contrast to Superkamiokande, electron neutrinos will be \newpage \noindent directly seen here in the CC current reaction $\nu_e +D \to p+p+e^{-}$. Thanks to the large cross-section, event rate is close to that of Superkamiokande. Neutrino energy is given by electron energy and mass difference of $D$ and $H$. The SNO data can be helpful in detection of Hep neutrinos above the end of $B$-neutrino spectrum. The NC reactions $\nu_x+D \to \nu_x+ p + n$, seen by presence of neutrons result in anomalous NC/CC ratio in case of oscillation of $\nu_e$ to active neutrino component. This is a signature of neutrino oscillation. SNO is more sensitive than Superkamiokande to the day/night effect, which is a ``smoking gun'' of MSW effect. This is because $\nu_e$ neutrinos are directly measured in CC reaction. Detection of day/night effect in Superkamiokande, SNO and probably ICARUS is the last hope for this effect, because all other planned now detectors are not sensitive to it. {\em ICARUS} \cite{ICARUS} is a liquid argon detector. Detection of solar neutrinos is based on CC-reaction $\nu_e+^{40}\!\!Ar \to ^{40}\!\!K + e^{-}$ and $\nu e$ scattering. With excellent energy resolution (about $5\%$) and low threshold of electron detection (about $5~MeV$), ICARUS has great potential for super-precise measurement of electron spectrum and flux of Hep neutrinos. {\em KamLand} \cite{ASuz} is 1kt liquid scintillator detector for $\bar{\nu}_e$ neutrinos, based on the Reines reaction $\bar{\nu}_e +p \to e^+ + n$. Solar $\bar{\nu}_e$ can be detected, though without clear signature of solar origin ({\em e.g.} no directionality). Detection of $\bar{\nu}_e$ neutrinos with $E_{\nu} \sim 3~MeV$ from a nuclear reactor at distance $L \sim 100~km$ can test the oscillations with $\Delta m^2 \sim E_{\nu}/L \sim 3\cdot 10^{-6}~eV^2$, {\em i.e.} close to that of LMA MSW solution. {\em HELLAZ} \cite{HELLAZ} is a low temperature and high pressure hellium detector, registering neutrinos in $\nu e$ scattering. The recoil electron energy and scattering angle are measured with high precision and thus the energy of neutrino is known with comparable accuracy. This detector is designed for $pp$ neutrinos, but probably $Be$ neutrinos can be registered too. {\em BOREXINO} and {LENS} are two low-energy neutrino detectors, complementary in physical interpretation of the results. BOREXINO will start to operate in the beginning of the next millennium. LENS is a new proposal to the Gran Sasso Laboratory based on the recent idea put forward by R.S.Raghavan \cite{Ragh}. BOREXINO at Gran Sasso is 300t liquid scintillator detector for registering $Be$-neutrinos due to $\nu e$ scattering. It measures CC+NC signal from $\nu_e$ neutrinos together with NC signal from $\nu_{\mu}$ and $\nu_{\tau}$, in case of oscillation to active components. LENS is a liquid scintillator detector loaded by Yb or Gd nuclei. Neutrinos are detected due to reactions $\nu_e + ^{176}\!Yb \to ^{176}\!Lu^* + e^-$ (or $\nu_e + ^{160}\!Gd \to ^{160}\!Tb^* + e^-$) with a threshold 244 keV (300 keV). The prompt signal from electron is accompanied by a delayed signal from photon or conversion electron from an excited nucleus. This strongly reduces the background. LENS will detect $pp$ and $Be \; \nu_e$-neutrinos. \samepage The combination of the BOREXINO and LENS data will provide us with the following physical information. (i) With $pp$- and $Be$- neutrino fluxes measured separately, the whole neutrino spectroscopy of the Sun will be completed. The suppression of neutrino fluxes at different energies will be explicitly known. (ii) With $Be \; \nu_e$-neutrino flux known from LENS, BOREXINO will give a signal from $\nu_{\mu}$ and $\nu_{\tau}$, in case of $\nu_e$ oscillation to active neutrinos. Thus, the combination of both experiments have a status of {\em appearance oscillation experiment}. In case of $\nu_e$ oscillation to sterile neutrino, BOREXINO should not show the signal in excess of that predicted by LENS. (iii) Just-so VO and HEE VO solutions predict strong seasonal variation of $Be$ neutrino flux. The BOREXINO/LENS observations will confirm or reject these models. EIS VO model predicts absence of anomalous seasonal variation accompanied by suppression (by factor $\sim 2$) of both $pp$ and $Be \; \nu_e$ neutrinos. This is also can be tested by the combined BOREXINO and LENS data. (iv) Both detectors can observe 11yr periodicity in $Be$-neutrino flux and measure (June+Dec)/(March+Sept) ratio, which are signatures of the RSFP solution. \vspace{-15mm} \section{\bf Conclusions} Solar Neutrino Problem (SNP) is deficit of neutrino fluxes as compared to the SSM predictions detected in all solar-neutrino experiments. The astrophysical (including nuclear physics) solution to SNP is excluded or strongly disfavoured. SNP has a status of disappearance oscillation experiment. No direct signature of oscillations has been found yet. Currently there are six oscillation solutions to SNP: SMA MSW, LMA MSW, LOW MSW, Just-so VO, EIS VO, and RSFP. Two of them are disfavoured: EIS VO (vacuum oscillations with energy-independent suppression) is excluded by observed rates at $99.8 \%$ CL and can survive if the Homestake result is excluded from analysis; LOW MSW is seen only at $\geq 99\%$ CL. Distortion of $B$-neutrino spectrum (as compared with the SSM specrum) is a common signature of oscillation solutions (which is absent only in EIS VO and weak in the LMA MSW).The ratio of the observed electron spectrum (708d of Superkamiokande data) to that of predicted by the SSM model is flat at $5.5~MeV \leq E_e \leq 13~MeV$ and has an excess at $E_e \geq 13~MeV$. This excess cannot be explained by the MSW solutions and by those just-so VO solutions, which explain the rates. It is not excluded that this excess is due to Hep neutrinos or small systematic experimental error. Day/Night effect and zenith-angle dependence of neutrino flux is a signature of MSW solutions. After 708d of Superkamiokande observations this effect (in percent) is $2.9 \pm 1.7 \pm 0.30$, i.e. consistent with zero at $1.7 \sigma$. Statistics, if increased by factor 5, might make this effect statistically significant. SNO, in the operation soon, is more sensitive than Superkamiokande to day/night effect (due to CC events). There are still chances that day/night effect will be discovered in this round of observations. If not, the future detectors planned at present (BOREXINO, LENS and HELLAZ), will not also be able to see it. The HEE VO solution with $\Delta m^2=4.2\cdot 10^{-10}~eV^2$ and $\sin^2 2\theta=0.93$ explains the spectrum with high energy excess and the rates, if $B$-neutrino flux is assumed to be $15 - 20 \%$ smaller than in SSM and if the chlorine signal is about $30\%$ larger than in the Homestake observations. This solution predicts high amplitude semi-annual time variation of $Be$-neutrino flux, that can be reliably observed by BOREXINO. Another oscillation solution which explains all rates and Superkamiokande spectrum (including high energy excess) is the RSFP model. An open problem for this model is prediction of 11yr (or 22yr) variations and (June+Dec)/(March+Sept) excess, that are not observed. Future low-energy neutrino detectors, BOREXINO and LENS, are very sensitive to VO solutions and they will either confirm or reject them. \\[1mm] {\bf Acknowledgments}\\[1mm] \samepage I am grateful to Gianni Fiorentini and Marcello Lissia for enjoyable permanent collaboration and discussions. I am very much indebted to Yoji Totsuka and Kunio Inoue who provided me with the Superkamiokande data in the form of ps-files. I have learned much about Superkamiokande data from discussions with Kunio Inoue. I would like to thank Plamen Krastev for preparing a compilation of figures from Ref.\cite{BKS} and for discussions. Sandro Bettini and Till Kirsten are thanked for discussions and useful remarks. \samepage I am honoured and grateful to the organizers of 19th Texas Symposium for inviting me for a plenary talk. I appreciate very much their efforts to the excellent organization of the conference. \samepage
\section{Introduction} The structure and morphology of the inner Milky Way are difficult to determine due both to dust obscuration and to our edge-on view. The canonical picture of the Milky Way as an axisymmetric spiral galaxy was enshrined in the models of Schmidt (1965), Bahcall \& Soneira (1980), Ostriker \& Caldwell (1983), Kent (1992), and others. However, the suggestion by de Vaucouleurs (1964) that the Galaxy is barred has been supported by many recent studies (cf.\ the reviews of Blitz {\it et al.\ }\ 1993 and Kuijken 1996). What was once thought of as the bulge now seems to be, at least in part, a thickened bar. Lines of evidence for a bar include: the infrared surface brightness distribution (Blitz \& Spergel 1991; Dwek {\it et al.\ }\ 1995), the distribution of Mira variables (Whitelock \& Catchpole 1992), IRAS point sources (Weinberg 1992, Nikolaev \& Weinberg 1997), the magnitude offset of bulge stars at positive and negative longitudes (Stanek 1995; Stanek {\it et al.\ }\ 1997), OH/IR stars (Sevenster 1995), and the gas motions near the Galactic center (\eg\ Liszt \& Burton 1980; Binney {\it et al.\ }\ 1991). Several groups have used infrared photometry, especially from the COBE/DIRBE data, to deduce the density distribution in the Galactic bar (\eg\ Blitz \& Spergel 1991; Dwek {\it et al.\ }\ 1995; Binney, Gerhard \& Spergel 1997). It has long been known, from both 21~cm and mm observations of gaseous emission lines, that the kinematics of gas toward the Galactic center ($|l| \lesssim 10\hbox{$^{\circ}$}$) are inconsistent with purely circular motions (\eg\ Rougoor \& Oort 1960, Kerr \& Westerhout 1965, Oort 1977). Figure \ref{fig-lisztdata} shows the \hi\ longitude--velocity ($\ell-V$) diagram constructed from the data of Liszt \& Burton (1980; see also Burton \& Liszt 1983). This diagram shows the distribution of \hi\ radial velocities at galactic longitudes $13\hbox{$^{\circ}$} > \ell > -11\hbox{$^{\circ}$}$. Most gas is approaching at negative longitudes and receding at positive, which is the general sense of rotation of the Milky Way, but there is significant emission from gas moving in the opposite sense on both sides; such gas is inconsistent with simple circular orbits and is said to have ``forbidden velocities.'' Forbidden velocities in excess of 100 \hbox{$\rm {km}~\rm s^{-1}$}\ are observed throughout the range $-6\hbox{$^{\circ}$} < \ell < 6\hbox{$^{\circ}$}$. A variety of explanations for the non-circular motions have been proposed including explosive outflows (cf.\ Oort 1977), spiral density waves (\eg\ Scoville, Solomon \& Jefferts 1974), and barlike perturbations. If the non-circular motions do result from gas flow in a non-axisymmetric potential, observation and detailed modeling of the gas kinematics should provide strong constraints on the mass distribution in the inner Galaxy. In fact, flow patterns in barred galaxy models have already been shown to provide qualitative fits to the observations (\eg\ Peters 1975; Liszt \& Burton 1980; van Albada 1985b; Mulder \& Liem 1986; Binney {\it et al.\ }\ 1991). Features in diagrams such as Figure \ref{fig-lisztdata} contain information about the distribution of gas in space and velocity within the disk of the Galaxy. But because we cannot determine the distance to individual parcels of gas, there is no unique way to invert the observed \lv diagram to determine the two-dimensional distribution of gas in the Galaxy; the projection into $\ell$ and $V$ space is highly degenerate. Even if such a deprojection were available, we still could not use the flow pattern to deduce the galactic gravitational potential directly, since the gas is also subject to pressure forces and its motion is governed by the non-linear equations of fluid dynamics. Thus the data need to be interpreted by comparison with models. Binney {\it et al.\ }\ (1991) compare stellar orbits in a barred model with the CO and \hi\ \lv diagrams, which offers some insight, but omits the effects of the strong shocks expected in gas flows in a bar. Subsequently, several numerical methods have been employed to construct improved models for the gas. Jenkins \& Binney (1994) used sticky particles, Englmaier \& Gerhard (1998) used smoothed particle hydrodynamics (SPH), while Fux (1997,1999) combined SPH and $N$-body techniques to attempt to build a fully self-consistent model of the inner Milky Way. Fux (1999) has compared the gas kinematics in such a model to ``arm'' features in the CO and \hi\ \lv diagrams to constrain the properties of the bar; his approach is complementary to ours, concentrating on high-density regions of the \lv diagram. Most modeling efforts have been devoted to observations of the dense molecular gas while comparatively little attention has been devoted to the \hi\ data. Here we focus on the \lv diagram for the \hi, which is less affected by two principal limitations of the molecular data: the \lv diagram for the \hi\ is both more symmetric and more complete than the corresponding CO plots. In particular, CO (Dame {\it et al.\ }\ 1987; Bally {\it et al.\ }\ 1988) is not detected where \hi\ emission is present in some significant regions of the \lv plane; for example, between $\ell = 0\hbox{$^{\circ}$}$ and $-6\hbox{$^{\circ}$}$, the \hi\ emission extends to $\sim -270~\hbox{$\rm {km}~\rm s^{-1}$}$ while the CO emission extends to $\sim -220~\hbox{$\rm {km}~\rm s^{-1}$}$ only (Figure 4 of Dame {\it et al.\ }). More importantly, \hi\ emission extends to higher forbidden velocities over a wider angular range in comparison with that observed in CO. We attempt to place constraints on the properties of the Galactic bar by comparing the \hi\ \lv diagram with similar plots synthesized from many fluid-dynamical models in various potentials. The full gas velocity field allows us to determine which regions of the Galaxy are responsible for prominent features of the \lv diagram. Our goal is not to identify a unique model, but rather to infer properties of the inner Galaxy that appear to be required by the data. We conclude that the Galaxy must have a strong bar that rotates fairly quickly and has a central density high enough to produce an inner Lindblad resonance. The bar must have a semi-major axis $a \gtrsim 3$~kpc, and be viewed obliquely, with the bar major axis between 30\hbox{$^{\circ}$}\ and 40\hbox{$^{\circ}$}\ to the Sun--Galactic Center line. \section{The Galactic longitude--velocity diagram} \label{sec-lvdiag} \subsection{Observational data} We use the \hi\ observations of the inner Galaxy by Burton \& Liszt (1978, 1983; and Liszt \& Burton 1980), which produced the \lv diagram shown in Figure \ref{fig-lisztdata}. These data have uniform coverage of the longitude range $\ell = -11\hbox{$^{\circ}$}$ to +13\hbox{$^{\circ}$}, with spatial resolution $\sim 0.5\hbox{$^{\circ}$}$, well matched to the resolution of our simulations, and good velocity resolution ($2.75~\hbox{$\rm {km}~\rm s^{-1}$}$) and sensitivity. H.\ Liszt kindly provided the data in electronic form. The spectra are taken on an 0.5\hbox{$^{\circ}$}\ grid in \el\ and $b$; because we are comparing to 2-D simulations, we summed the data along the $b$ axis. We also smoothed in $V$ with a Gaussian of $\sigma = 5.5~\hbox{$\rm {km}~\rm s^{-1}$}$. A high-velocity \hi\ cloud at $\ell=8\hbox{$^{\circ}$}, b=-4\hbox{$^{\circ}$}$ and $V=-210~\hbox{$\rm {km}~\rm s^{-1}$}$ (``Shane's feature,'' Saraber \& Shane 1974), was excluded from the dataset. \placefigure{fig-lisztdata} \begin{figure*}[t] \psfig{figure=lisztdata.ps,width=7.0truein,angle=-90} \caption{The longitude--velocity diagram of \hi\ in the inner Galaxy from the data of Liszt \& Burton (1980). The lowest contour is 0.125\degk\ in antenna temperature summed over $b$, or 1.25 \lvunit\ of atomic gas (H+He), and the contours increase geometrically by a factor of 2. The forbidden quadrants are positive $V$ at negative \el, and negative $V$ at positive \el. The band of emission at $|V| \lesssim 100~\hbox{$\rm {km}~\rm s^{-1}$}$ is foreground from the disk. The filled circles are those data points on the extreme-velocity contour that we try to reproduce in the simulations. \label{fig-lisztdata} } \end{figure*} Plots of individual latitude slices (Burton \& Liszt 1978; Liszt \& Burton 1980) show that the velocity ``peaks'' in Figure \ref{fig-lisztdata} are prominent at latitudes near $b = 0\hbox{$^{\circ}$}$. The broad band of emission (sometimes called the ``main maximum'') at $-100 < V < 100\ \hbox{$\rm {km}~\rm s^{-1}$}$ at all longitudes is present over the entire latitude range observed by Liszt \& Burton ($-6\hbox{$^{\circ}$} < b < 6\hbox{$^{\circ}$}$). Since the half-thickness of the gas layer is approximately 250 pc inside the Solar radius to 4 kpc radius, and the thickness may be only 100 kpc inside 4 kpc (Mihalas \& Binney 1981; Jackson \& Kellman 1974), the band of emission is presumably from disk gas that is relatively close by. The velocity extent of the band is large for the velocity dispersion of the gas as derived by Gunn, Knapp \& Tremaine (1979), even given the 1000:1 density contrast. It is presumably attributable to line-of-sight integration over substantial bulk motions in the disk such as spiral arm streaming motions (Burton \& Liszt 1983). Foreground gas is also responsible for 21 cm absorption against the central continuum source at $\ell=0\hbox{$^{\circ}$}$, $b=0\hbox{$^{\circ}$}$. This absorption appears at negative velocity (Burton \& Liszt 1978, 1993) and is visible in the summed data in some of the intermediate contours in Figure~\ref{fig-lisztdata}, although it is not conspicuous in the extreme contour. The absorption at negative velocities implies that the negative-velocity gas at $\ell=0\hbox{$^{\circ}$}$, $b=0\hbox{$^{\circ}$}$ is between the Sun and the Galactic Center, while the positive velocity gas at that position is behind the Center (Burton \& Liszt 1978). The filled circles in Figure \ref{fig-lisztdata} mark the points of the observed extreme-velocity contour (EVC) we will use for comparison to the simulations. Because we are interested in the motions of the gas in the inner Galaxy, we do not use that portion of the EVC that appears to be substantially influenced by foreground disk gas, but we retain the data point at $\ell=0\hbox{$^{\circ}$}$ since the extreme contour there is not much affected by absorption. The \lv diagram is not perfectly two-fold rotationally symmetric in many respects. Here we simply note that the shapes of the velocity peaks in the EVC differ: that at positive \el\ lies at 3\hbox{$^{\circ}$}\ while the most negative velocity is at $\ell =-4\hbox{$^{\circ}$}$, although the magnitudes are similar. More detailed plots of the \hi\ \lv diagram reveal other non-symmetric features in the interior of the diagram, including the well-known ``3-kpc expanding arm'' (\eg\ Peters 1975; Burton \& Liszt 1983, 1993), which is marginally visible in Figure~\ref{fig-lisztdata} at $-3\hbox{$^{\circ}$} > \ell > -9\hbox{$^{\circ}$}$ near $-100~\hbox{$\rm {km}~\rm s^{-1}$}$. Some investigators (\eg\ Kerr 1967; Liszt \& Burton 1980) have also presented evidence that the \hi\ gas distribution in the inner Galaxy is tilted out of the Galactic plane. By summing the data over $b$, we have suppressed this aspect, which would be difficult to address in any case since our models are two-dimensional. \subsection{Interpreting the extreme-velocity contour} One must make assumptions in order to extract information on the structure of the Galaxy from the \lv diagram. The simplest approach is to assume that the Galaxy is axisymmetric and the gas moves on circular orbits. With this assumption (and others noted below), the \lv diagram can be used to determine the rotation curve of the Galaxy interior to the Solar circle by the tangent-point method (cf.\ Gunn {\it et al.\ }\ 1979, Mihalas \& Binney 1981). The critical feature of the \lv diagram in this method is the extreme-velocity contour (EVC), which is the outer contour of the gas distribution in longitude--velocity space; it is the highest absolute radial velocity observed along the line of sight at each \el. In the tangent-point method, the EVC in the upper left and lower right quadrants only is used; gas at forbidden velocities is ignored. The extreme observed velocity needs to be corrected for instrumental resolution and the velocity dispersion of the gas, which is assumed to have a uniform value (Gunn {\it et al.\ }\ 1979), to find the terminal velocity at each longitude, $v_t(\ell)$ (see Section \ref{sec-evclevel} below). With the further assumptions that some \hi\ gas exists at every tangent point and that the circular angular frequency, $\Omega(R)$, decreases monotonically from the center, $v_t(\ell)$ yields the Galactic rotation curve $\Theta(R)$ directly through the equation $\Theta(R_0~|\rm{sin}~\ell |) = |v_t(\ell)| + \Theta_0~|\rm{sin}~\ell|$. As the correction term for the circular velocity of the LSR is small at longitudes near $0\hbox{$^{\circ}$}$, the EVC on the maximum side (positive $V$ at positive \el, negative $V$ at negative \el) is approximately the rotation curve under the axisymmetric assumption. For circular orbits, on one side of the Galactic center all the gas should be coming towards the Sun, and on the other side it should be going away. Hence, the EVC on the non-maximum side, in the upper right and lower left quadrants of the \lv diagram, should be featureless and close to 0 \hbox{$\rm {km}~\rm s^{-1}$}\ (as long as the circular frequency at $R_0$ is less than the circular frequency in the inner Galaxy, which is true for any reasonable rotation curve). The velocity dispersion of the gas and bulk motions in the disk will push the EVC beyond 0 \hbox{$\rm {km}~\rm s^{-1}$}, but apart from these effects the non-maximum EVC should not tell us much. \placefigure{fig-axisymlv} \begin{figure*}[t] \psfig{figure=axisymlv.ps,width=7.0truein,angle=-90} \caption{The longitude-velocity diagram of gas in an axisymmetric model. There is no gas in the forbidden quadrants, beyond the velocity dispersion-induced spread. The lowest contour is 1.7 \lvunit\ of total gas. The contours increase geometrically by a factor of 2. The filled circles are the EVC from the data in Figure \ref{fig-lisztdata}. Simply rearranging some of the mass of this model into a bar yields a model that fits these data points quite well. \label{fig-axisymlv} } \end{figure*} Figure \ref{fig-axisymlv} shows an \lv diagram for a model with gas all on circular orbits. The rotation curve that gives rise to this \lv diagram is plotted in Figure \ref{fig-axirotcurv}. The contrast with Figure \ref{fig-lisztdata} is instructive. Gas at forbidden velocities in the Milky Way is clearly inconsistent with a simple circular flow pattern. \placefigure{fig-axirotcurv} \begin{figure*}[t] \psfig{figure=rotcurv.ps,width=7.0truein} \caption{The rotation curve for the axisymmetric model. This model is the axially symmetrized mass distribution of our Model 1. \label{fig-axirotcurv} } \end{figure*} The EVC is still a useful probe of the Galactic mass distribution even when the gas is {\it not\/} on circular orbits, provided that the observed tracer is ubiquitous in the disk and that the non-circular motions are caused by streaming in a non-axisymmetric potential, as first proposed by de Vaucouleurs (1964). As long as the observations are sensitive enough to pick up the tracer in regions of low density, the EVC depends almost solely on the velocity field, and variations in the fraction of gas mass in a given tracer phase are much less important. Here we discount the alternative possibility that non-circular motions arise from explosions or other violent events near the Galactic Center (cf.\ Oort 1977). Neutral hydrogen is ubiquitous in the Galactic disk and is readily detectable through its 21 cm emission. It is clearly more widespread than CO in the inner Galaxy, since there are no ``holes'' in the \hi\ \lv diagram (Figure \ref{fig-lisztdata}) in contrast with that for the CO (\eg Figure 4 of Dame {\it et al.\ }\ 1987, and Figure 4 of Bally {\it et al.\ }\ 1988). Additionally, as noted earlier, the negative velocity peak of the CO \lv diagram reaches only to $-220~\hbox{$\rm {km}~\rm s^{-1}$}$ at $\ell=-2\hbox{$^{\circ}$}$ while that peak reaches $-270~\hbox{$\rm {km}~\rm s^{-1}$}$ at $\ell=-4\hbox{$^{\circ}$}$ in the \hi\ \lv diagram, and the forbidden emission extends further in \hi\ than in CO, especially for negative velocities at $0\hbox{$^{\circ}$} < \ell < 5\hbox{$^{\circ}$}$. The interior of an \lv diagram for CO shows much substructure with strong density contrasts, whereas that for \hi\ exhibits only mild variations (Figure~\ref{fig-lisztdata}). Interior features, in both molecular and atomic gas, provide extra information to constrain models; \eg\ Fux (1999) attempts to match them to an SPH gas flow in a model of the Galaxy. The additional substructure in molecular emission, which traces gas of higher density, is probably caused by variations both in the atomic fraction and in molecular emissivity (\eg\ temperature). Such variations, even if they are well understood, would be very hard to model, however. The EVC of the \hi\ \lv diagram, on the other hand, is insensitive to density variations. All successful models of the inner Milky Way should therefore match it provided only that there is {\it some\/} atomic gas everywhere in the flow. The smoothness of the EVC in Figure~\ref{fig-lisztdata} gives us grounds to hope that this requirement is fulfilled. \section{Simulations of the gas flow} \label{sec-simul} We use a two-dimensional grid-based gas dynamical code to simulate the gas flow in models for the galactic potential. The code was originally written by G.\ D.\ van Albada to model gas flow in barred galaxy potentials (van Albada 1985a, 1985b) and kindly provided by E.\ Athana\-ssoula. She used it (Athana\-ssoula 1992b) to study gas flow patterns in various barred potentials. \subsection{The fluid code} The code is an second-order, flux-splitting Eulerian grid code for an isothermal gas in an imposed gravitational potential representing the stellar component and halo of the Galaxy. We neglect the self-gravity of the gas in order to reduce computational requirements. We justify this omission on the grounds that the gas surface density is considerably less than that of the stellar bulge and disk, especially in the inner regions of the Galaxy with which we are primarily concerned (see Section~\ref{sec-propmodel} below). Our grid has 200 by 400 cells, each 50 pc square, and we enforce a 180\hbox{$^{\circ}$}\ rotation symmetry, so that the grid is effectively 400 by 400. The grid is fixed with respect to the barred potential, and both rotate at a steady pattern speed; the bar is aligned at 45\hbox{$^{\circ}$}\ to the grid axes. The time step is variable, chosen automatically via a Courant condition, and is generally approximately 0.1 Myr. The sound speed of the gas is taken to be 8 \hbox{$\rm {km}~\rm s^{-1}$}\ (cf.\ Gunn {\it et al.\ }\ 1979), corresponding to a temperature of $\sim 10^4~\degk$. Varying the sound speed within reasonable limits of a few \hbox{$\rm {km}~\rm s^{-1}$}\ does not materially affect the derived gas flow. By its nature, the code approximates the interstellar medium as an Eulerian fluid, smooth on scales of the grid cell size. Without some idealization it is hopeless to simulate the extremely complex dynamics of the multiphase ISM, which has structure on all observed scales and a vast assortment of energy inputs and outputs. Some authors (Jenkins \& Binney 1994; Combes 1996) have suggested that smooth-fluid models using the Euler equations, such as grid codes and smooth-particle hydrodynamics, are not appropriate for the clumpy ISM, and have advocated various sticky-particle methods. Sticky-particle codes may be well suited to simulating the dynamics of the self-gravitating molecular cloud component, which Jenkins \& Binney implicitly probed by comparing to CO observations. However, the \hi\ in the neutral ISM is much less clumpy; it is not clear that the neutral ISM is made up of discrete clouds, especially over scales of $\gtrsim 50$~pc, the grid scale we use. Essentially, applying the Euler equations to the ISM simply asserts that the ISM has a pressure or sound speed defined in a coarse-grained sense, over scales greater than the code's resolution. Englmaier \& Gerhard (1997) used an SPH code to simulate flow in one of the model potentials that Athana\-ssoula (1992b) used with the Eulerian grid code. For equivalent input parameters, Englmaier \& Gerhard obtained results very similar to Athana\-ssoula's, which reassures us that the simulations are not dependent on the fluid-dynamical algorithm.\footnote{Englmaier \& Gerhard found that increasing the sound speed of the gas to 20--25 \hbox{$\rm {km}~\rm s^{-1}$}\ changed the flow pattern. However, such a large value implies an unreasonably high temperature for the ISM, and is inconsistent with the value found by Gunn {\it et al.\ }\ (1979).} A limitation of particle codes is their inability to represent large density contrasts. By design, spatially adaptive particle codes resolve structure well in high density regions, but the finite number of particles precludes adequate representation of the fluid properties in very low density regions. Grid codes, on the other hand, cannot resolve spatial structure below a few grid cells, but can handle nearly any density contrast with no increase in overhead, and represent low and high density regions equally. In a case such as the gas in the Milky Way bar, where the geometry and scales of interest are largely fixed by the stellar potential, spatial adaptivity is less essential and grid codes are generally more efficient. The grid's advantage in density contrast is especially important since the gas in low density regions will prove crucial to match the observed emission in the forbidden quadrants of the \lv diagram, as discussed further in Section \ref{subsec-invertlv}. \subsection{Simulation procedure} \label{subsec-simproc} We begin each simulation in a quasi-equilibrium state, with the mass of the bar redistributed in an axisymmetric configuration, the gas on circular orbits, and a uniform gas surface density of 5 \hbox{$\rm M_{\odot}~{\rm pc^{-2}}$}. We turn on the bar by linear interpolation between the initial axisymmetric state and its fully barred shape, reaching its final state in 0.1 Gyr. The bar growth time is approximately equal to the orbital period at a radius of 3~kpc. Different choices for the growth time and initial density do not particularly affect the results, save that the final gas density distribution scales overall proportionally to the constant chosen for the initial density. We continue the simulation to 0.2 Gyr to allow the gas flow to ``settle'' after the bar has grown, and to 0.3 Gyr to verify that the flow has stabilized. The gas response can never reach a completely steady state, because the gas inside co-rotation continuously loses energy in shocks and flows toward the center.\footnote{The gas build up in the center can be significant if the code is run for many rotation periods, \eg\ several Gyr. This effect can be lessened by the use of a ``gas-recycling'' provision in the code. However, we found that gas recycling caused long-period oscillations in the flow with the fine grid used here, probably because it redistributes energy over the grid (G.\ van Albada, private communication). The oscillations do not occur on coarser grids, such as those used by Athana\-ssoula (1992b), presumably due to the higher numerical diffusivity. Since we are not interested in the long-term evolution of the flow, we avoid this numerical problem by turning gas recycling off.} Gas continues to accumulate in the center, but there is very little change in the gas velocity field from 0.2 to 0.3 Gyr. We use the gas density and velocity fields at 0.2 Gyr to construct \lv diagrams as would be seen by an observer in the plane of the simulation. The observer is placed 8.5 kpc from the Galactic center and in the LSR, moving with a velocity of $\Theta_0 = 220~\hbox{$\rm {km}~\rm s^{-1}$}$ toward $\ell = 90\hbox{$^{\circ}$}$, and at a given viewing angle -- the angle between the bar major axis and the Sun-Galactic Center line. (The effect of a different LSR motion is discussed below in Section~\ref{sec-lsrmotion}). The viewing angle is varied to find the best value, as detailed below in Section~\ref{sec-bestfit}. For each cell in the simulation grid, we calculate the longitude of the cell and the angle it subtends, and the radial velocity of the gas in the cell. The gas density in the cell and its distance from the Sun determine the observed brightness. The brightness distribution is convolved and sampled in longitude to model the angular beamwidth of the telescope and the $0.5\hbox{$^{\circ}$}$ sampling of the observed positions, and convolved in velocity to include the effects of the sound speed of the gas ($c_s = 8~\hbox{$\rm {km}~\rm s^{-1}$}$) and the velocity resolution of the observations (smoothed with a Gaussian of $\sigma = 5.5~\hbox{$\rm {km}~\rm s^{-1}$}$). \subsection{Model gravitational potentials} Our models for the gravitational potential are similar to those used by Athana\-ssoula (1992a,b). They have three components: an ellipsoidal bar, a centrally concentrated bulge, and an extended component to represent both the disk and halo. We model the bar as a prolate Ferrers $n=1$ ellipsoid with semimajor axis $a$ and semiminor axis $b$. The bar density is given by \begin{equation} \rho(x,y,z) = \left\{ \begin{array}{ll} \rho_{0,{\rm bar}} \, (1 - u^2) & {\rm if}~u^2 < 1, \\ 0 & {\rm if}~u^2 > 1, \end{array} \right. \end{equation} where \begin{equation} u^2 = \frac{x^2}{a^2} + \frac{y^2}{b^2} + \frac{z^2}{b^2}. \end{equation} This model for the bar is convenient because its gravitational field is analytic (Binney \& Tremaine 1987), but it is a crude model for the real bar (\eg\ Dwek {\it et al.\ }\ 1995). We compensate for one of its principal weaknesses by adding a bulge component. Ferrers bars are not very centrally concentrated; the bulge component allows us to increase the central concentration and to adjust its strength relative to the bar. The bulge is a modified Hubble profile sphere with core radius $r_c$ and density given by \begin{equation} \rho(r) = \rho_{0,{\rm bul}} \, \left[1 + \left( \frac{r}{r_c} \right)^2 \right]^{-3/2}. \end{equation} The ``bulge'' component can be viewed as effectively part of the bar; our treatment of the two as separate analytical components does not imply that we regard them as distinct, either photometrically or kinematically. We use $M_{\rm bul}$ to refer to the bulge mass within 1 kpc of the Galactic center, since this is most analogous to the central concentration of the bar; the {\it total\/} mass of a modified Hubble profile sphere diverges at large radii. The extended component has the potential \begin{equation} \Phi(R) = \Phi_0 \ln\left( 1 + \sqrt{1 + (R/R_c)^2} \right), \end{equation} where $R_c$ is scale length. If all the mass that gives rise to this potential were to reside in the disk, it would have the surface density of a Rybicki disk (given by Zang 1976 and derived independently by Hunter, Ball \& Gottesman 1984): \begin{equation} \Sigma(R) = \Sigma_0 \frac{R_c}{\sqrt{R_c^2 + R^2}}, \end{equation} with $\Phi_0 = 2\pi G\Sigma_0R_c$. The rotation curve of this potential becomes asymptotically flat at large radius, making it suitable for modeling the contribution both of the axisymmetric part of the stellar disk and of the dark matter halo. As the simulation is two-dimensional, it is insensitive to the three-dimensional forms of these density distributions; any distribution that yielded similar forces in the plane could be substituted. Thus, mass can be traded off between the axisymmetric components; for example, it is unimportant that the density of the Hubble bulge falls off slowly, since the small additional contribution to the rotation curve (cf.\ Figure~\ref{fig-axirotcurv}) could be absorbed into the rotation curve of the disk or halo. The total potential is specified by seven parameters: a central density and scale length for each of the bulge and ``disk,'' and a central density and two axis lengths for the bar. Our only constraint is that the rotation curve should be roughly flat outside $R_0$, with a circular velocity from 200--220 \hbox{$\rm {km}~\rm s^{-1}$}\ at 8.5 kpc. An eighth parameter, the Lagrange or corotation radius $R_L$, is required to fully specify a model; choosing $R_L$ is equivalent to specifying a pattern speed for the bar. The gas flow pattern is determined by the adopted potential, but the \lv diagram further depends on the viewing angle \philsr\ between the Sun--Galactic center line and the major axis of the bar. We varied the parameters by trial and error and examined the \lv diagrams after each run to learn the effects of changes in bar size, bar mass, bulge mass, Lagrange radius and so on. Our goal was to find a model or models that matched the observations reasonably well, rather than systematically to explore the parameter space, which is impractical given the large number of parameters. We did run some series to explore the effect of varying a parameter, most notably, varying the Lagrange radius while holding all other parameters constant. In all, we ran 51 models; their parameters are given in Table 1. The table is sorted by the goodness of fit as measured by the RMS deviation in velocity between model and data (discussed further in Section~\ref{sec-bestfit}). The best fit viewing angle and the goodness of fit are tabulated in the last two columns of Table 1. The models are numbered best to worst; the number, naturally, does not correspond to the order in which the models were run, since we improved the models by learning from past results -- Model 1 was actually the 46th model run. \section{Comparison to observations} \label{sec-compar} We compared the outer envelope -- the extreme-velocity contour -- of the synthesized \lv diagrams to that of the data. The observed EVC used is the contour of 0.125~\degk~degrees of antenna temperature summed over $b$, or 1.25~\lvunit\ of atomic gas (H and He), using the calibration given by Liszt \& Burton (1980). The data points used are shown by the filled circles in Figure \ref{fig-lisztdata}. As discussed in Section~\ref{sec-lvdiag}, those portions of the EVC that show signs of contamination from foreground disk emission are excluded from comparisons to models. \subsection{The EVC contour level} \label{sec-evclevel} The position of the observed extreme-velocity contour is determined by the actual terminal velocity envelope of the gas, extended by the velocity broadening due to the gas sound speed and the instrumental resolution. Since the flux level at which the EVC can be observed is also limited by the noise in the observations, the EVC is not an intrinsic property of the Galaxy, but also depends on the observational parameters. In the tangent-point method, the observed EVC must be corrected to yield the terminal velocity envelope. In practice, it is conventional to assume that (1) the difference between the terminal velocity $v_t(\el)$ and the EVC is some constant $\Delta V$, and (2) $\Delta V$ can be determined by observations near $\ell \simeq \pm 90\hbox{$^{\circ}$}$, where the actual terminal velocity is expected to be zero (cf.\ Gunn {\it et al.\ }\ 1979). As the data we are using do not cover $\ell \simeq \pm 90\hbox{$^{\circ}$}$, we cannot make use of this method to derive $\Delta V$. In order to compare the observations and simulations, we have constructed simulated \lv diagrams which take into account the velocity dispersion of the gas and the instrumental resolution. But the absolute level at which to place the EVC in the simulated \lv diagram is not constrained, since we do not know $\Delta V$ for the observations. Fortunately, both simulated and observed \lv diagrams have fairly sharp edges, in the sense that the flux falls off rapidly with increasing $|V|$ -- see Figures \ref{fig-lisztdata} and \ref{fig-bestlv}. The lowest contours simply trace the falloff profile of the velocity dispersion and instrumental resolution. We compared simulated \lv diagrams to that observed, examining the fall off at the edges of the distribution, to set the level for the EVC in the simulated \lv diagram. Placing the EVC at $\sim 1.7$ \lvunit of total simulation gas produced a reasonably good match but EVC levels of 1.25 -- 2.5~\lvunit\ were almost equally acceptable. Because the \lv diagrams do have sharp edges, changing the flux level of the comparison EVC, even by a factor of 2, does not have a strong effect. We ran comparisons of the entire series of models at EVC contour levels from 0.625 -- 5.0~\lvunit\ and verified that the small changes caused by different choices for the EVC level produce only minor changes in the rank ordering of models, and do not affect our conclusions. We note that comparing the simulation EVC at 1.7~\lvunit\ of total gas to the observed \hi\ EVC at 1.25~\lvunit\ of atomic gas could be interpreted to mean that the gas is 75\% atomic; levels of 1.25 -- 2.5~\lvunit\ would imply atomic fractions of 100\% -- 50\%. However, the comparison is not reliable for this purpose. The edges of the EVC largely represent gas in low density regions, and the molecular fraction is undoubtedly higher in high density regions -- CO emission does not generally extend to the velocities of the \hi\ EVC (Dame {\it et al.\ } 1987). Additionally, the inferred fraction would be changed if the value of the initial gas surface density used in the simulations were changed. It is, however, comforting that the inferred atomic fraction is close to but less than 1. (The actual atomic mass fraction in the inner Galaxy is perhaps 50\%; cf.\ Bronfman {\it et al.\ }\ 1988; Bloemen {\it et al.\ }\ 1986.) \subsection{Best fit and viewing angle} \label{sec-bestfit} To rank the models by the quality of their fit to the data, we compute the root-mean-square deviation in velocity between the location of the simulated EVC and the observed data points. The RMS velocity deviation is not an ``error'' in a statistical sense; it serves as a figure-of-merit for ranking the models. The RMS places a relatively high weight on large deviations, which penalizes gross differences between model and data more than large numbers of small differences. For a given model, the position of the observer with respect to the bar must be specified to construct an \lv diagram. We define the ``viewing angle'' \philsr\ to be the angle between the bar major axis and the Galactic Center--to--Sun line, so that 0\hbox{$^{\circ}$}\ is an end-on bar, 90\hbox{$^{\circ}$}\ is side-on, and values between 0\hbox{$^{\circ}$}\ and 90\hbox{$^{\circ}$}\ put the near end of the bar in the first Galactic quadrant ($0\hbox{$^{\circ}$} < \ell < 90\hbox{$^{\circ}$}$). We determined the best-fit viewing angle for each model iteratively by synthesizing \lv diagrams and computing the RMS deviation at viewing angle intervals of 10\hbox{$^{\circ}$}, 4\hbox{$^{\circ}$}, and 1\hbox{$^{\circ}$}, successively, narrowing the search interval at each step. The best-fit viewing angle for each model and the corresponding RMS velocity deviation are tabulated in Table 1, sorted by the goodness of fit. The viewing angle given is for the best fit between 0\hbox{$^{\circ}$}\ and 90\hbox{$^{\circ}$}; these are the realistic models since many lines of evidence place the near end of the bar in this quadrant. For the few models that have a better fit outside this quadrant, that result is given in the table footnotes. \section{Results: I. The best model} \label{sec-results1} \subsection{Properties of the model} \label{sec-propmodel} Our primary result is that we have found a model which reproduces the outer contour of the \lv diagram fairly well. This model is model 1 in Table 1; a number of the models that are runners-up are closely related to it. Model 1 has a bar with semimajor axis 3.6 kpc and Lagrangian radius 5.0 kpc, corresponding to a pattern speed of 41.9 \hbox{$\rm{km}~\rm{s}^{-1}~\rm{kpc}^{-1}$}. The best-fit \lv diagram is shown in Figure \ref{fig-bestlv} and the RMS velocity deviation is 16.54~\hbox{$\rm {km}~\rm s^{-1}$}. The minimum in RMS deviation is well localized at a viewing angle of 34\hbox{$^{\circ}$}\ to the bar major axis, although changes of a few degrees ($< 5\hbox{$^{\circ}$}$) are possible without greatly worsening the fit. The localization in RMS deviation is similar for all of the better models. The effects of changes in the viewing angle are discussed further in Section \ref{sec-results2}. \placefigure{fig-bestlv} \begin{figure*}[ht] \psfig{figure=bestlv.ps,width=7.0truein,angle=-90} \caption{The \lv diagram for the best-fitting model, Model 1, as viewed at 34\hbox{$^{\circ}$}. The lowest contour is 1.7 \lvunit\ of total gas and the contours increase geometrically by a factor of 2. The filled circles are the observed data points. The RMS velocity deviation between model and data is 16.54~\hbox{$\rm {km}~\rm s^{-1}$}. \label{fig-bestlv} } \end{figure*} Figure \ref{fig-surfmass} shows the surface density distribution of the combined bar, bulge, and disk+halo components, as projected along the $z$ axis of the Galaxy. The figure shows the central 8 kpc by 8 kpc region of the model, in what is essentially a face-on view -- although the contours are in surface density of mass, not light. This plot demonstrates the influence of the bulge component, which makes the central concentration of the bar much higher than that of a Ferrers bar in isolation. It is also clear that the full surface density distribution is less elongated than the bar component alone, with an axis ratio of about 3:1 as compared to the bar component's axis ratio of 4:1. \placefigure{fig-surfmass} \begin{figure*}[ht] \psfig{figure=surfmass.ps,width=7.0truein,angle=-90} \caption{The face-on surface density of the mass distribution for Model 1, as projected along the $z$-axis, and shown for the innermost 8 kpc by 8 kpc of the Galaxy. The lowest contour is at 400~\hbox{$\rm M_{\odot}~{\rm pc^{-2}}$}, and the contours increase geometrically by a factor of $\sqrt{2}$. \label{fig-surfmass} } \end{figure*} When the mass density is integrated over $-100 < z < 100$~pc, the lower estimate for the thickness of the gas layer, the resulting distribution is similar to that of Figure \ref{fig-surfmass} but with mass surface density lower by a factor of about four. Within this range of $z$, at the bar end the mass surface density is 140~\hbox{$\rm M_{\odot}~{\rm pc^{-2}}$}; in the central region at $R<0.5$~kpc the mean mass surface density is 2900~\hbox{$\rm M_{\odot}~{\rm pc^{-2}}$}. A comparison of these mass surface densities with the gas surface density suggests our neglect of gas self-gravity is justified. \placefigure{fig-gasdensell} \begin{figure*}[htb] \psfig{figure=gasdensell.ps,width=7.0truein} \caption{The face-on gas density in the innermost 8 kpc by 8 kpc of Model 1, as viewed from above the plane of the simulation. The grayscale runs from 0 to 15 \hbox{$\rm M_{\odot}~{\rm pc^{-2}}$}\ of total gas, with dark representing higher density. The bar is at a 45\hbox{$^{\circ}$}\ angle to the grid, running from lower left to upper right. The ellipse represents the points at which the bar density drops to zero, with semimajor axis 3.6 kpc. The sense of the bar rotation is clockwise. The inclined line from top to bottom of the figure represents the Sun--Galactic Center line for the best-fit viewing angle of 34\hbox{$^{\circ}$}, with the Sun above the top of the figure. \label{fig-gasdensell} } \end{figure*} The gas density in the innermost 8 kpc square of this model is shown in Figure \ref{fig-gasdensell}, also in a face-on view. The long, straight high density features in the bar are shocks, with transverse velocity jumps of $\sim 200~\hbox{$\rm {km}~\rm s^{-1}$}$, extending out to 2.9 kpc from the galactic center. They are parallel but offset from each other; near the center the straight shocks join onto an oval or nuclear ring of high density gas that is also the location of shocks. The semi-major axis of this oval is 0.5 kpc. The gas surface density in the shocks is 5--20~\hbox{$\rm M_{\odot}~{\rm pc^{-2}}$}; within $R<0.5$~kpc the mean gas surface density is 130~\hbox{$\rm M_{\odot}~{\rm pc^{-2}}$}. The straight, offset shocks and the inner oval are characteristic of gas flow in strongly barred potentials with an inner Lindblad resonance (Athana\-ssoula 1992b). Dust lanes with morphologies similar to the high density gas in Figure \ref{fig-gasdensell} are observed in many barred galaxies. The dust lanes are presumably caused by the high gas density at the shock (Prendergast 1962, unpublished; see also van Albada \& Sanders 1982; Prendergast 1983; Athanassoula 1992b). Spectroscopy of barred galaxies shows sharp velocity jumps at the location of the dust lane (\eg\ Pence \& Blackman 1984; Lindblad {\it et al.\ }\ 1996; Regan, Vogel \& Teuben 1997; Weiner {\it et al.\ }\ 1999). Figure \ref{fig-gasdensell} also shows that there is little gas in the lens region of the galaxy, inside 3~kpc; barred galaxies often show a central hole in the gas distribution swept clear by the angular momentum transport of the bar (\eg\ NGC 1300, England 1989; NGC 1398, Moore \& Gottesman 1995; and NGC 4123, Weiner {\it et al.\ }\ 1999). Outside 3~kpc, the gaseous disk is relatively quiescent; the bar does not drive a large response in the outer disk. The disk does not exhibit spiral patterns outside the bar radius; spirals in the outer disk could be driven by spirals in the stellar disk and/or the self-gravity of the gas, which we have neglected in order to concentrate on the inner Galaxy. \placefigure{fig-streamlines} \begin{figure*}[ht] \psfig{figure=streamlines.ps,width=7.0truein,angle=-90} \caption{The gas velocity field in the innermost 8 kpc by 8 kpc of Model 1. The bar is at a 45\hbox{$^{\circ}$}\ angle to the grid as before. For clarity, only every fourth cell in the grid is plotted. \label{fig-streamlines} } \end{figure*} The gas velocity field as seen in a non-rotating frame, in the inner $8 \hbox{ kpc} \times 8$ kpc region, is shown in Figure \ref{fig-streamlines}. For clarity, we have plotted only every fourth cell. The velocity changes abruptly at the shocks along the density peaks. Essentially, gas in the bar moves up to the shock at relatively high velocity and hits the shock, dissipating energy. The post-shock gas then streams back down the bar, gaining velocity quickly as it moves away from the shock and falls down the potential well. Gas streamlines in the bar are elongated along the bar, in the manner of the $x_1$ family of stellar orbits in bars, but are clearly not symmetric about the major axis of the bar, unlike the $x_1$ orbits. The shocks are located along the leading edge of the streamlines and are approximately parallel to the bar; the major axis of the elongated streamlines is rotated approximately $5\hbox{$^{\circ}$}$ ahead (toward the leading side) of the bar major axis. This angle, which we will refer to as the ``lead angle,'' is closely related to the pattern speed of the bar, to be discussed further in Section~\ref{sec-patspeed}. Near the center of the bar, the major axes of the streamlines change, so that the streamlines are elongated across the bar more than along it, similar to the $x_2$ family of stellar orbits present in bars with inner Lindblad resonances (Athana\-ssoula 1992a,b). The central oval of high gas density corresponds to this family of streamlines. Again, the streamlines are rotated by an oblique angle with respect to the bar, unlike the $x_2$ stellar orbits, which are perpendicular to the bar major axis. \subsection{Inverting the projection into \lv space} \label{subsec-invertlv} The plot of the gas streamlines offers some understanding of the features in the \lv diagram, but the effect of projection into \lv space is much clearer in Figure \ref{fig-velcontour}. This figure plots the radial velocity observed in Model 1 as a function of position on the grid, i.e.\ over the plane of the Galaxy, showing the radial velocity {\it before\/} it is projected into the \lv diagram. \placefigure{fig-velcontour} \begin{figure*}[ht] \psfig{figure=velcontour.ps,width=7.0truein} \caption{A contour plot over the inner $8 \hbox{ kpc } \times 8$ kpc region of the radial velocity at which each grid cell in Model 1 is observed when the viewing angle is 34\hbox{$^{\circ}$}. The contours are from --250 \hbox{$\rm {km}~\rm s^{-1}$}\ to 250 \hbox{$\rm {km}~\rm s^{-1}$}\ at intervals of 50 \hbox{$\rm {km}~\rm s^{-1}$}; contours below zero are dashed. The bar is at a 45\hbox{$^{\circ}$}\ angle to the grid, as before. The inclined line from top to bottom is the Sun--Galactic Center line ($\ell = 0\hbox{$^{\circ}$}$), with the Sun above the figure. \label{fig-velcontour} } \end{figure*} The gas at forbidden velocities moves toward the Sun at $\ell > 0\hbox{$^{\circ}$}$, the side where most of the gas is moving away, and vice versa at $\ell < 0\hbox{$^{\circ}$}$. It is clear from Figure \ref{fig-streamlines} and Figure \ref{fig-velcontour} that forbidden velocities belong to low-density gas approaching the shocks. The preshock region with forbidden velocities extends all the way out to the shock tip at 2.9 kpc. However, the magnitude of the forbidden radial velocities in the preshock region falls below 100~\hbox{$\rm {km}~\rm s^{-1}$}\ at about 1.5 kpc from the Galactic center, which corresponds roughly to $\ell = \pm 6\hbox{$^{\circ}$}$ for the viewing angle of 34\hbox{$^{\circ}$}. Past this point, emission at forbidden velocities is obscured in the Milky Way by the band of emission from foreground gas. Identifying the emission in the forbidden quadrants with the low-density preshock gas may explain why the forbidden emission is much more extensive in \hi\ than in CO, while the peaks in the \hi\ \lv diagram, which come from higher density regions, are present in the CO \lv diagram. The identification of forbidden velocities with the preshock gas also illuminates some of the difficulty Jenkins \& Binney (1994) had matching their sticky-particle models to the data. The \lv diagrams they presented have very little emission in the forbidden quadrants. However, their maps of gas density in the plane of the simulation show that the apparent lack of emission is because there are very few particles in the preshock regions at any given time, as discussed in Section \ref{sec-simul}. The high and narrow peaks in the EVC at $\ell \sim +3\hbox{$^{\circ}$} \hbox{ and } -4\hbox{$^{\circ}$}$, or $\sim 0.6$~kpc projected distance from the Galactic center, have no counterparts in the equivalent axisymmetric model (Figure~\ref{fig-axisymlv}). The origin of these peaks can also be understood from Figure \ref{fig-streamlines}; the elongation of the orbits caused by the strong ellipticity of the gravitational potential results in high gas velocities roughly parallel to the bar major axis. The observed high radial velocities arise from the gas on elongated orbits just as it passes the oval of high-density gas (Figure \ref{fig-velcontour}). The EVC declines rapidly beyond the peak because the gas at larger radii does not fall as deeply into the bar's potential well, and is on less elongated orbits. Many authors (e.g.\ Gunn {\it et al.\ }\ 1979; Gerhard \& Vietri 1986; Liszt 1992; Burton \& Liszt 1993) have noted that the peaks in the EVC and the rapid decline imply an unusual rotation curve if the gas is assumed to move on circular orbits; the inferred rotation curve also shows a sharp rise and rapid decline. These features are more naturally explained by gas flow in a triaxial potential (\eg\ Gerhard \& Vietri 1986, Burton \& Liszt 1993). Simulations such as model 1 show that not only the EVC peaks, but also the forbidden emission, are accounted for by gas flows in a strong bar. As Burton \& Liszt emphasize, comparisons with a derived rotation curve instead of the full \lv diagram both embody incorrect assumptions about the inner Galaxy and discard valuable data from the forbidden quadrants of the \lv diagram. Figure \ref{fig-velcontour} can also be used to determine the location within the plane of the Galaxy of a feature in the \lv diagram, or an object whose longitude and radial velocity are known but whose distance is uncertain. For example, the 3-kpc expanding arm goes approximately through the points $(\el,V)$ = ($-10\hbox{$^{\circ}$}$, $-100~\hbox{$\rm {km}~\rm s^{-1}$}$), ($-5\hbox{$^{\circ}$}$, $-75~\hbox{$\rm {km}~\rm s^{-1}$}$), (0\hbox{$^{\circ}$}, $-50~\hbox{$\rm {km}~\rm s^{-1}$}$), (+2.5\hbox{$^{\circ}$}, $-35~\hbox{$\rm {km}~\rm s^{-1}$}$) (Liszt \& Burton 1980). Locating these points on Figure \ref{fig-velcontour} shows that they lie approximately on a arc centered on the Galactic center and of $\sim 2.5$~kpc radius, suggesting that the 3-kpc arm could be a spiral arm at about that radius with a small pitch angle, and that its motion is consistent with the overall Galactic velocity field, removing the need for large anomalous expansion velocities. In fact, an arm at approximately the right position is visible in Figure~\ref{fig-gasdensell}. We note that even though the simulation is bisymmetric, the synthesised \lv diagram has some asymmetry because one end of the bar is closer to the Sun than the other. The observed \lv diagram is somewhat more asymmetric than the model, however. We cannot rule out the possibility that the observed asymmetry is due to actual asymmetries in the gas distribution or the shape of the Galaxy. However, the asymmetries in \hi\ are considerably smaller than those in the CO \lv diagram (Dame {\it et al.\ }\ 1987; Bally {\it et al.\ }\ 1988). The most obvious deviation of this model from the observations is that it is not as strongly peaked at positive \el\ as the data, although this is essentially due to the asymmetry in the peaks of the data, since the model compromises by slightly overestimating the peak at negative \el. The model also produces a strong diagonal feature in the interior of the \lv diagram from about (+3\hbox{$^{\circ}$}, +100~\hbox{$\rm {km}~\rm s^{-1}$}) to (--3\hbox{$^{\circ}$}, --100~\hbox{$\rm {km}~\rm s^{-1}$}) which is not present in Figure~\ref{fig-lisztdata}. This feature is caused by the nuclear ring of high density gas, which is not seen in the observations both because the gas is probably in the $H_2$ phase (cf.\ Rubin, Kenney \& Young 1997) and because those parts of this feature with $|V|<100\;$km~s$^{-1}$ are obscured by the ``main maximum'' of foreground from disk gas. The gas density and velocity fields in our model 1 are consistent with those observed in external barred galaxies. In particular, the straight shock regions with high gas densities can be identified with the straight dust lanes along the bar seen in many barred galaxies, which are generally thought to be the locations of shocks (Prendergast 1962, unpublished; see also \eg\ Prendergast 1983; van Albada \& Sanders 1982; Athana\-ssoula 1992b). Model 1 provides the best fit among our models, but it is by no means a unique solution to the problem of reproducing the \lv diagram. A slightly different choice of parameters could conceivably do better, and it is almost certain that some potential with components other than the particular analytic forms we used could improve on Model 1. However, the Galaxy is likely to resemble Model 1 in certain major respects, such as viewing angle, bar size, possession of an ILR, and high pattern speed. These conclusions are partly drawn from our experience with other, less well-fitting models, which we now discuss. \section{Results: II. Other models} \label{sec-results2} In this section we describe other models to illustrate the influence of variations in some of the major parameters. This exercise allows us to infer the properties that a successful model is likely to possess in order to reproduce the observations. A natural question to ask is whether the adverse consequences of changing one parameter can be compensated for by changes to other parameters. In general, the effects of the parameters are sufficiently interlinked that attempting to compensate by making one change has other unintended consequences. Given the number of parameters, it is impractical to test for all possible compensatory changes, but we do not believe that large variations in important parameters can be compensated away. Although Model 1 was one of the last models to be run, it is a close variant of Model 3, which we had tried much earlier (our 24th run); we tried a number of variations to improve Model 3 before actually succeeding. Our experience makes it seem unlikely that some other radically different model could fit equally well or better, but we cannot rule out the possibility. \subsection{Changes in viewing angle} \label{sec-viewangle} Changing the angle from which Model 1 is viewed is not properly a different model for the potential, but can drastically change the resulting \lv diagram. Figure \ref{fig-viewangle} illustrates the systematic changes that occur when Model 1 is viewed at angles $-10$\hbox{$^{\circ}$}, $-5$\hbox{$^{\circ}$}, $+5$\hbox{$^{\circ}$}, and $+10$\hbox{$^{\circ}$}\ from the optimum value of $\philsr = 34\hbox{$^{\circ}$}$. Viewing a model more nearly end-on than optimum, as in Figure \ref{fig-viewangle}(a) and (b), produces both higher peaks in the EVC and steeper declines from the peaks. It also reduces the extent of the gas in the forbidden quadrants, relative to the height of the peaks. Once again, reference to Figure \ref{fig-streamlines} reveals the reasons for these changes. \placefigure{fig-viewangle} \begin{figure*}[htb] \psfig{figure=viewangle.ps,width=7.0truein,angle=-90} \caption{Longitude-velocity diagrams for Model 1 when it is viewed at angles other than the optimal viewing angle of $\philsr = 34\hbox{$^{\circ}$}$. The lowest contour is 1.7 \lvunit\ of total gas and the contours increase geometrically by a factor of 4. \label{fig-viewangle}} (a) $\philsr = 24\hbox{$^{\circ}$}$, more end-on; RMS velocity deviation = 31.09~\hbox{$\rm {km}~\rm s^{-1}$}.\\ (b) $\philsr = 29\hbox{$^{\circ}$}$, RMS = 22.10~\hbox{$\rm {km}~\rm s^{-1}$}.\\ (c) $\philsr = 39\hbox{$^{\circ}$}$, more side-on; RMS = 22.28~\hbox{$\rm {km}~\rm s^{-1}$}.\\ (d) $\philsr = 44\hbox{$^{\circ}$}$, RMS = 36.94~\hbox{$\rm {km}~\rm s^{-1}$}. \end{figure*} In a more end-on view of the bar, the elongated orbits that produce the velocity peaks are projected more onto the line of sight of the observer, making the peaks higher. Counter to what might be expected, the peaks do not move significantly closer together in a more end-on view because the streamlines in this part of the flow are curved, and the region contributing to the peaks rotates somewhat. The curve of the streamlines is caused by the presence of an ILR, because the $x_2$ orbit family forces the elongated streamlines in the inner region of the bar away from the center. The more end-on view also means that the region with highly to moderately elongated orbits subtends a smaller angle, and so the fall-off with increasing $|\el|$ is more rapid. The relative deficiency of gas in the forbidden quadrants occurs because the shocks, and the preshock regions responsible for the forbidden emission, subtend a smaller angle when the bar is viewed more end-on. In the more end-on view, the projected components of the velocities of the preshock gas are larger, which compensates somewhat, but the slope of the decline in the EVC from the peaks into the forbidden quadrants is steeper. Clearly, the peaks could be lowered by reducing the central density of the model, but the more end-on view would then yield too little emission in the forbidden quadrants. The effects of a more side-on view of the bar, as seen in Figure \ref{fig-viewangle}(c) and (d), are essentially exactly the opposite. The velocity peaks drop and their slope is gentler. The extent of the gas in the forbidden quadrants increases, but the lower projected velocities give a gentler slope to the EVC. Gross variations in viewing angle, to the point where, for example, a model is viewed fully side-on at $\philsr \sim 90\hbox{$^{\circ}$}$, can produce \lv diagrams that deviate somewhat from these rules of thumb. For example, some models such as numbers 6, 9, 12, and 18 can produce high velocity peaks at side-on viewing angles because the innermost streamlines derived from $x_2$ orbits (approximately perpendicular to the bar) are viewed end-on. These models are of little practical interest, since a number of other lines of evidence rule out such large viewing angles -- for example, a grossly side-on view cannot produce the magnitude offset between bulge stars at positive and negative longitudes, as shown by Stanek {\it et al.\ }\ (1997). (We note that models 6, 9, 12, and 18 are all slow bars in which $R_L \geq 2.4a$; see below.) \subsection{Motion of the LSR} \label{sec-lsrmotion} The \lv diagrams were constructed by assuming that the LSR is moving with a circular (tangential) velocity $\Theta_0 = 220$~\hbox{$\rm {km}~\rm s^{-1}$}\ relative to the Galactic Center, with no radial motion. We tested the effect of assuming a different velocity of the LSR relative to the Galactic Center. A radial motion of --5 to +10~\hbox{$\rm {km}~\rm s^{-1}$}, positive outward, can be accommodated; values outside this range significantly worsen the models' fit to the data. The best values of the radial motion are between 0 and +5~\hbox{$\rm {km}~\rm s^{-1}$}. The fits are not sensitive to reasonable variations of the circular speed, since the data are near $\ell = 0\hbox{$^{\circ}$}$; values of $\Theta_0$ from 160 to 240 \hbox{$\rm {km}~\rm s^{-1}$}\ were tested and yielded acceptable fits. Varying the LSR motion has a minimal effect on the relative ranking of the models. The non-circular motion predicted by the models for gas at the solar position is small. For Model 1, the gas at the solar position has a tangential velocity of 211~\hbox{$\rm {km}~\rm s^{-1}$}, and a radial motion of --0.7~\hbox{$\rm {km}~\rm s^{-1}$}\ (inward). Model 1 has an outer Lindblad resonance (OLR) near the solar position, but the gas is on an essentially circular orbit. The OLR could have observable effects on the kinematics of stars in the solar neighborhood, in either mean velocity or dispersion. The nature of the effects is not simple to predict (cf.\ Kalnajs 1992, Kuijken \& Tremaine 1992, Weinberg 1994); moreover Dehnen's (1998) analysis of Hipparcos data shows that the velocity structure of nearby stars is quite complicated. \subsection{Varying the pattern speed} \label{sec-patspeed} We created a sequence of models including Model 1 to test the effect of varying the Lagrange radius or, equivalently, the pattern speed of the bar. The sequence in Lagrange radius $R_L = 4$, 5, 6, 7, and 8 kpc yielded models 2, 1, 5, 8, and 4 respectively. This sequence includes most of the best-fitting models (Model 3 is closely related).\footnote{ The groups of models \{33, 26, 22, 6\}, \{23, 12, 9\}, \{19, 13\}, and \{21, 18\} also comprise sets where only the Lagrange radius is changed -- the effects are similar.} Figure \ref{fig-radlagmosaic} shows face-on views of the gas density in this sequence of models, like that of Figure \ref{fig-gasdensell} for Model 1; Figure \ref{fig-radlaglv} shows \lv plots for Models 2, 5, 8, and 4, to be compared with Figure \ref{fig-bestlv}. \begin{figure*}[htb] \psfig{figure=radlaglv.ps,width=7.0truein,angle=-90} \caption{Longitude-velocity diagrams for a sequence of models with the same parameters as Model 1 except for the Lagrange radius. Model 1 (Figure \ref{fig-bestlv}) is between Models 2 (panel a) and 5 (panel b) in the sequence of increasing $R_L$. The lowest contour is 1.7 \lvunit\ of total gas and the contours increase geometrically by a factor of 4. \label{fig-radlaglv}} (a) Model 2, $R_L$ = 4.0 kpc, bar viewing angle $31\hbox{$^{\circ}$}$. \\ (b) Model 5, $R_L$ = 6.0 kpc, bar viewing angle $20\hbox{$^{\circ}$}$. \\ (c) Model 8, $R_L$ = 7.0 kpc, bar viewing angle $15\hbox{$^{\circ}$}$. \\ (d) Model 4, $R_L$ = 8.0 kpc, bar viewing angle $8\hbox{$^{\circ}$}$. \end{figure*} The streamlines in Model 1 are not symmetric about the bar major axis; in fact the major axis of the streamlines is rotated by about 5\hbox{$^{\circ}$}\ with respect to it, the ``lead angle'' referred to in Section~\ref{sec-results1}. \placefigure{fig-radlagmosaic} \begin{figure*}[htb] \psfig{figure=radlagmosaic.ps,width=7.0truein} \caption{The gas density in the inner 8 by 8 kpc of the grid for a sequence of five models with different Lagrange radii and all other parameters fixed. The bar is at 45\hbox{$^{\circ}$}\ to the grid in each model; the ellipse indicates the outer edge of the bar, at which the bar density drops to zero. The slanted line indicates the Sun--Galactic Center line for the best-fit viewing angle. The bar semimajor axis is $a = 3.6$~kpc. \label{fig-radlagmosaic}} (a) Model 2, $R_L$ = 4.0 kpc. Shock lead angle $\approx 0\hbox{$^{\circ}$}$, bar viewing angle $31\hbox{$^{\circ}$}$. \\ (b) Model 1, $R_L$ = 5.0 kpc. Shock lead angle $\approx 5\hbox{$^{\circ}$}$, bar viewing angle $34\hbox{$^{\circ}$}$. \\ (c) Model 5, $R_L$ = 6.0 kpc. Shock lead angle $\approx 10\hbox{$^{\circ}$}$, bar viewing angle $20\hbox{$^{\circ}$}$. \\ (d) Model 8, $R_L$ = 7.0 kpc. Shock lead angle $\approx 15\hbox{$^{\circ}$}$, bar viewing angle $15\hbox{$^{\circ}$}$. \\ (e) Model 4, $R_L$ = 8.0 kpc. Shock lead angle $\approx 25\hbox{$^{\circ}$}$, bar viewing angle $8\hbox{$^{\circ}$}$. \end{figure*} Figure \ref{fig-radlagmosaic} shows that the lead angle increases with the Lagrange radius, as far as 25\hbox{$^{\circ}$}\ for the slowest bar. The somewhat surprising result that several models with grossly different Lagrange radii and lead angles all appear to fit the \lv data reasonably well arises because the models simply compensate by moving the best-fit viewing angle synchronously with the changes in the lead angle. The best-fit viewing angle stays roughly constant with respect to the {\it shocks}, which means that it also changes in a clockwise sense with respect to the bar, causing \philsr\ to decrease. Thus changes in viewing angle are strongly coupled to the angle the gas streamlines make with the bar. The systematic change in the location of the shocks has a relatively simple explanation. As the Lagrange radius is increased, the bar pattern speed slows (for $R_L = 4.0$, 5.0, 6.0, 7.0, 8.0, $\Omega_p = 54.2$, 41.9, 34.9, 30.2, 26.6 \hbox{$\rm{km}~\rm{s}^{-1}~\rm{kpc}^{-1}$}\ respectively). Inside the Lagrange or corotation radius, gas overtakes the gravitational potential well of the bar; the shocks are caused as the gas climbs out of the well, slows down, and piles up (Prendergast 1983). Although the shape of the gas streamlines is dependent on the full gas-dynamics, the magnitude of the velocity for gas at a given radius is roughly set by the gravitational acceleration from the mass interior to it, which is the same in all five models. In a frame co-rotating with the bar, if the bar is slower, the gas is moving faster as it overtakes the bar, so it climbs farther out of the potential well before the shock pile-up occurs. Therefore, in slower pattern speed models, the shocks are farther ahead of the bar, in the sense of more positive lead angle. The increased speed of the gas relative to the potential also increases the strength of the shocks. The behavior of the shocks rules out slow bars, if we demand that the Milky Way bar should resemble bars in other galaxies. In external galaxies, the prominent dust lanes frequently seen along the bar run along the ``leading'' sides of the bar; the morphology of these dust lanes and exemplary galaxies are discussed by Athana\-ssoula (1992b). Strong bars generally have straight dust lanes while weaker bars sometimes have curved dust lanes; in both cases, the dust lanes are generally parallel to the bar, as in the shocks of Model 1, or angled slightly in the sense of smaller lead angle. These dust lanes are identified with the high-density shocks, such as those in Figure \ref{fig-gasdensell}, as discussed above. We know of no barred galaxies that have dust lanes with a lead angle of more than a few degrees; Athana\-ssoula (1992b) argued that therefore strong bars rotate quickly. Merrifield \& Kuijken (1995) have also showed that the bar in NGC 936 rotates quickly, via a completely independent method. The position of the shocks in models 4 and 8 ($R_L = 7.0$ and 8.0 kpc), and in all other slow bar models we have run, is grossly inconsistent with what we know about barred galaxies. We reject these models for this reason, even though some of them formally fit the \lv diagram well. \subsection{Bar strength and shape} \label{sec-barshape} The streamline plot of Figure \ref{fig-streamlines} and the contours of observed velocity shown in Figure \ref{fig-velcontour} suggest that the bar has to be strong and fairly elongated. Only a massive bar can produce the large non-circular motions needed to put gas at the forbidden velocities observed in the \lv diagram. If the gas streamlines are less elongated than those seen in Figure~\ref{fig-streamlines}, the regions with forbidden velocities are smaller and subtend a smaller range of Galactic longitude. \placefigure{fig-weakbars} \begin{figure*}[htb] \psfig{figure=weakbars.ps,width=7.0truein,angle=-90} \caption{Longitude-velocity diagrams for several models with smaller bars of lower mass than that of Model 1. These models have $M_{bar}$ = 80\% that of Model 1, and $M_{bul}$ between 75\% and 125\% that of Model 1. The lowest contour is 1.7 \lvunit\ of total gas and the contours increase geometrically by a factor of 4. There is a deficiency of gas in the forbidden quadrants in these models. \label{fig-weakbars}} (a) Model 10, viewing angle 27\hbox{$^{\circ}$}. \\ (b) Model 11, viewing angle 35\hbox{$^{\circ}$}. \\ (c) Model 13, viewing angle 19\hbox{$^{\circ}$}. \\ (d) Model 14, viewing angle 38\hbox{$^{\circ}$}. \end{figure*} In our best model, the bar component has a mass of $M_{\rm bar} = 9.8~\times~10^9$~\msun, and the mass of the bulge component (within 1 kpc radius) is $M_{\rm bul} = 5.4~\times~10^9$~\msun. The effect of a weaker bar on the \lv plots is shown in Figure~\ref{fig-weakbars}. Models 10, 11, 13, and 14 show a significant deficit of gas in the forbidden quadrants, notably at $\ell \sim -5\hbox{$^{\circ}$}$. These models have smaller bars than Model 1 with lower $M_{bar}$ (even though the bar density $\rho_{0,bar}$ is somewhat higher). Models 6, 9, and 12, which also have less massive bars than Model 1, do somewhat better at producing material in the forbidden quadrants, but only because the weaker forcing potential is partly compensated for by the stronger shocks that occur in a slow-rotating bar, as noted above. However, models 6, 9, and 12, like all the other slow-bar models, have shocks in an implausible position and are not viable models for the Galaxy. The bar must also be strong in the sense of having a large axis ratio. The formal axis ratio of the Ferrers bar in Model 1 is 4:1, although the actual axis ratio of the total mass distribution, when the bulge and disk are included, is closer to 3:1 (cf.\ Figure ~\ref{fig-surfmass}). \placefigure{fig-fatbars} \begin{figure*}[htb] \psfig{figure=fatbars.ps,width=7.0truein,angle=-90} \caption{Longitude-velocity diagrams for several models with bars that have lower axis ratios than Model 1. The lowest contour is 1.7 \lvunit\ of total gas and the contours increase geometrically by a factor of 4. These models have EVCs whose peaks are not as sharp or as high as those observed. \label{fig-fatbars}} (a) Model 29, $a$ = 3.6 kpc, $b$ = 1.2 kpc, $a/b$ = 3.0, viewing angle 27\hbox{$^{\circ}$}. \\ (b) Model 37, $a$ = 4.0 kpc, $b$ = 1.3 kpc, $a/b$ = 3.1, viewing angle 55\hbox{$^{\circ}$}. \\ (c) Model 39, $a$ = 2.5 kpc, $b$ = 1.0 kpc, $a/b$ = 2.5, viewing angle 35\hbox{$^{\circ}$}. \\ (d) Model 46, $a$ = 3.0 kpc, $b$ = 1.3 kpc, $a/b$ = 2.3, viewing angle 32\hbox{$^{\circ}$}. \end{figure*} Models with smaller axis ratios generally do not reproduce the data well. Figure~\ref{fig-fatbars} shows \lv diagrams for several models whose bar components have axis ratios smaller than that of Model 1, with $a:b$ from 3.1:1 to 2.3:1. The axis ratios of the total mass distributions are fatter still. Although these models have different bar lengths, their appearance in \lv diagrams is similar: they produce EVCs that are gently sloped, not sharply peaked as seen in the observations. In particular, the decline of the EVC away from the peaks is fairly sharp in the observations, but much too gentle in the models with low axis ratio bars. Even an extremely centrally concentrated model but wide-barred potential such as Model 46, which has a small dense bulge, does not produce sharp peaks. More massive bars -- longer, more dense, or both -- do not successfully produce sharper peaks or better models: the most massive bars are models 45, 42, 37, 27, 43 and 44 ($M_{bar}$ = 41.2, 27.2, 22.7, 17.5, 15.9, and 15.9 $\times 10^9 \msun$, respectively). As discussed above, the peaks in the \lv diagram are produced by the strongly non-circular motions inside the bar while the steep decline in the EVC as $|\ell|$ increases further is linked to the weakening of the non-circular motions as the quadrupole field decays quickly with Galactocentric distance. An axisymmetric model with an unusual mass distribution could be made to produce this behavior but could not, of course, give rise to forbidden velocities. A strong and elongated bar is favored to produce both forbidden velocities and the narrow peaks in the EVC. \subsection{The presence of an inner Lindblad resonance} \label{sec-ilr} As already noted, the peaks of the \lv diagram arise from orbits just outside the oval of high density gas in the center, where the streamlines rotate to be highly angled to the bar rather than closely aligned with it. This rotation of the streamlines is related to the presence of an inner Lindblad resonance (ILR) (Athana\-ssoula 1992a,b). Bars with an ILR have a family of stellar orbits near the center that are elongated perpendicular to the bar rather than along it, and the rotated streamlines are related to these orbits. Bars without an ILR have only streamlines elongated along the bar; these streamlines would yield peak gas velocities as they pass the center (Athana\-ssoula 1992b). \placefigure{fig-noilrbars} \begin{figure*}[htb] \psfig{figure=noilrbars.ps,width=7.0truein,angle=-90} \caption{Longitude-velocity diagrams for several models which are less centrally concentrated than Model 1. The lowest contour is 1.7 \lvunit\ of total gas and the contours increase geometrically by a factor of 4. These models exhibit gently changing EVCs lacking the observed peaks. \label{fig-noilrbars}} (a) Model 27, $M_{bul}/M_{bar} = 0.019$, viewing angle 20\hbox{$^{\circ}$}. \\ (b) Model 50, $M_{bul}/M_{bar} = 0.13$, viewing angle 20\hbox{$^{\circ}$}. \\ (c) Model 48, $M_{bul}/M_{bar} = 0.16$, viewing angle 23\hbox{$^{\circ}$}. \\ (d) Model 42, $M_{bul}/M_{bar} = 0.17$, viewing angle 55\hbox{$^{\circ}$}. \end{figure*} An inner Lindblad resonance forces the elongated orbits away from the center, causing the highest bar-induced streaming velocities to occur some distance out, and producing sharply defined peaks in the \lv diagram that are several degrees apart. If the bar did not have an ILR, the EVC peaks are not necessarily as sharply defined, nor can they be separated by several degrees in longitude, as is observed. Model 27 is the least centrally concentrated of all our models, and its best-fit \lv diagram has EVCs without dominant peaks; the positive velocity EVC is nearly flat from $\ell=-3\hbox{$^{\circ}$}$ to $\ell=10\hbox{$^{\circ}$}$. The least centrally concentrated potentials are models 27, 51, 45, 50, 48, and 42 ($M_{bul} / M_{bar}$ = 0.019, 0.096, 0.11, 0.13, 0.16, and 0.17 respectively). Figure~\ref{fig-noilrbars} shows \lv diagrams for four of these weakly-concentrated potentials, which have EVCs with weak or gentle peaks. The observed strength and separation of the peaks in the \lv diagram suggests that the Galactic bar must have an ILR. The central mass concentration, represented in our model by the ``bulge'' component, is responsible for the ILR, and is also necessary to cause the sharply rising peaks in the EVC. Our adopted modified Hubble profile for the central mass component has a uniform density core, whereas the luminosity density in the Milky Way rises all the way to the center as the $\sim -1.8$ power of the radius (Becklin \& Neugebauer 1968). The finite resolution of the grid code vitiates attempts to simulate the effects of a central cusp; strong gradients in the angular velocity on scales below a few grid cells cannot be accurately represented. However, the small core radius in our best model, $r_c=0.2$ kpc (four simulation grid cells), is well inside the ILR feature at $R \sim 0.4$~kpc. The existence of the ILR implied by the EVC peaks requires only a concentrated mass within that radius, so our conclusion is little affected by the details of the density profile. \section{Discussion} We have shown that gas flow in a barred model of the Galaxy can fit many of the observed features of the \hi\ \lv diagram, most notably the emission in the forbidden quadrants and the sharp peaks in velocity. Our best fit model was arrived at through adjusting the free parameters by trial and error. Although the model has been tuned, the number of parameters is relatively small for a model of the Galactic potential. Furthermore, the complexity of the dynamics governing the gas response to the potential makes constructing a reasonably good model a non-trivial pursuit. This model does show that it is possible, and that the reservations of Jenkins \& Binney (1994) regarding the ability of simple gas-dynamical models to reproduce the data are perhaps too pessimistic. Our preferred model has a bar semi-major axis of 3.6 kpc. The bar component itself has an axis ratio of 4:1, although the ``bulge'' in our model should also be considered as part of the bar, and the axis ratio of the bar+bulge is somewhat fatter, approximately 3:1. The bar in this model rotates quickly, with a Lagrange radius of 5.0 kpc (bar pattern speed 42 \hbox{$\rm{km}~\rm{s}^{-1}~\rm{kpc}^{-1}$}) and the bar major axis is inclined at 34\hbox{$^{\circ}$}\ to our line of sight. Our model 1 differs in a number of important respects from the models of the Milky Way bar proposed by Binney {\it et al.\ }\ (1991), and further developed by Jenkins \& Binney (1994) and by Englmaier \& Gerhard (1998). These authors favor a considerably smaller bar and a higher pattern speed, placing corotation at $R\sim 3.5$~kpc, because they employ a cusped $x_1$ orbit to give the narrow peaks at $\ell \simeq \pm 3\hbox{$^{\circ}$}$. The $x_2$ orbit family, which is much less extensive in their models than in ours, gives a smaller peak very close to $\ell=0$. While their models were developed to interpret the CO \lv diagram, they fail to account for the large and extensive forbidden velocities seen in \hi. The strength and size of the bar are required by forbidden velocities in excess of 100 \hbox{$\rm {km}~\rm s^{-1}$}\ extending as far as $\el = \pm 6\hbox{$^{\circ}$}$. Were the true viewing angle much less than our preferred 34\hbox{$^{\circ}$}, as favored in some studies, the bar would have to be considerably longer to produce the observed forbidden velocities. Our constraint on viewing angle is not independent of the bar pattern speed, however, since slower bars give better fits when viewed at smaller angles. Models with a Lagrange radius of $R_L$ = 4.0 to 6.0 kpc (bar pattern speed $\Omega_p$ = 54 to 35 \hbox{$\rm{km}~\rm{s}^{-1}~\rm{kpc}^{-1}$}), i.e.\ fast-rotating bars, are favored; models with higher $R_L$ (lower $\Omega_p$) have shock patterns in the gas that differ drastically from those observed in other barred galaxies. It is unlikely that the viewing angle could be forced below 25\hbox{$^{\circ}$}\ to the bar major axis. We interpret the narrow velocity peaks at $\ell \simeq \pm 3\hbox{$^{\circ}$}$ as the signature of gas streaming along the bar past a nuclear ring in the Milky Way which lies close to the location of the inner Lindblad resonance. If the bar is strong, the high speed of these streams does not require an unusual radial mass profile -- the flow patterns shown in Figures \ref{fig-axisymlv} \& \ref{fig-bestlv} arise from two mass distributions that both, when azimuthally averaged, give the circular velocity curve shown in Figure~\ref{fig-axirotcurv}. The mass distribution in the inner Galaxy does have to be sufficiently concentrated for an ILR to be present, however; if this were not the case, the peaks would lie much closer to $\ell=0$. The location of the peaks at $\ell \simeq \pm 3\hbox{$^{\circ}$}$ requires the semi-major axis of the nuclear ring to be $\sim 400$ pc -- on the small end of the distribution of nuclear rings seen in other barred galaxies (Buta \& Crocker 1993). As nuclear rings in external galaxies are generally highly gas rich (Helfer \& Blitz 1995; Sofue 1996; Rubin, Kenney \& Young 1997), it is no surprise that the associated velocity peaks in the Milky Way stand out in CO as well as \hi. We note that the rotation curve of our preferred model, shown in Figure \ref{fig-axirotcurv}, indicates that the bulge and bar components together dominate the rotation curve in the inner few kpc of the Galaxy. We cannot isolate the contribution of the dark halo component, since our analytical model lumps the dark halo and the axisymmetric part of the disk together. However, since the Galaxy {\it does} have a disk, it is clear that the dark halo cannot be very dominant in this model. Although this potential is not a unique model of the Galaxy, as discussed above in Section \ref{sec-results1}, we believe that any model that fits the \lv diagram will have to have non-axisymmetric motions as strong as those in Model 1 and, hence, a bulge+bar which dominates the rotation curve in the inner part of the Galaxy. Englmaier \& Gerhard (1998) modeled the gas flow in the inner Galaxy, using models derived from COBE photometry. They found that the luminous matter must dominate over dark matter inside the solar circle, in order to match the terminal velocity curve in the non-forbidden quadrants. We do not claim that because our model 1 gives a reasonable fit, the mass distribution in the inner Galaxy must necessarily be very close to the analytic form we have assumed. The real mass distribution in the inner Galaxy is undoubtedly more complex than our simple analytical model. A different form of mass distribution will yield somewhat different results for the best-fitting model parameters. However, we believe that the real mass distribution will resemble Model 1 in its chief details: the strength and size of the bar, presence of an ILR, and viewing angle which is not too close to end-on. We have not attempted to satisfy the many other constraints on the shape of the inner Galaxy, such as COBE photometry, simultaneously. The model is broadly consistent with some results, such as the bar viewing angle determined by the IRAS point sources (Weinberg 1992), the magnitude offset of red clump stars (Stanek {\it et al.\ }\ 1997), and the distribution of OH/IR stars (Sevenster {\it et al.\ } 1999). Fux (1999) has compared the appearance of arm features produced in a self-consistent model with features in the CO and \hi\ \lv diagrams; his preferred model has a bar of similar length, with an ILR, and which rotates quickly, but the preferred viewing angle is somewhat smaller, 25\hbox{$^{\circ}$}, and the bar is fatter. Fux's comparison of models to data emphasizes high-density gas, while ours probes mostly low-density gas, which may be responsible for some of the differences. The viewing angles in Fux's best model and in ours are both incompatible with models which invoke a fairly end-on bar to account for the high microlensing optical depth towards the Galactic Bulge (Zhao \& Mao 1996; see also Fux 1997). The \lv diagrams synthesized from fluid models are sensitive to the details of the potential and the viewing angle, and the comparison with the data is unaffected by extinction. For these reasons we believe that the technique has great power to discriminate among candidate models of the inner Milky Way. We may eventually hope to identify a model of the Galactic bar that satisfies photometric constraints and fits both the CO and \hi\ kinematic data. \acknowledgments We are grateful to Dick van Albada and Lia Athana\-ssoula for providing us with the gas dynamics code and for helpful comments and advice on its use, to Harvey Liszt for providing the data of Liszt \& Burton (1980) in electronic form, and to an anonymous referee for a thoughtful report. This work was supported by NSF grant AST 96/17088 and NASA LTSA grant NAG 5-6037. BJW acknowledges support from a Carnegie postdoctoral fellowship.
\section*{Introduction} The determination of the nuclear incompressibility $K$ is still a matter of debate, despite a remarkable number of works on the subject$^{1)}$. In the present contribution, we present self-consistent calculations of the nuclear collective modes associated with a compression and expansion of the nuclear volume, namely the isoscalar giant monopole and dipole resonances (ISGMR and ISGDR, respectively). In fact, we share the point of view that the most reliable way to extract information on $K$ is to perform that kind of calculations, having as the only phenomenological input a given effective nucleon-nucleon interaction, and choose the value of $K$ corresponding to the force which can reproduce the experimental properties of the compression modes in finite nuclei. The ISGMR, or ``breathing mode'', is excited by the operator $\sum_{i=1}^A r_i^2$ and it has been identified in many isotopes along the chart of nuclei already two decades ago. However, this systematics has never allowed an unambigous determination of $K$$^{1)}$. This was one of the motivations for the recent experimental program undertaken at the Texas A\&~M Cyclotron Institute, which has allowed the extraction of experimental data for the ISGMR of better quality as compared to the past, by means of the analysis of the results of inelastic scattering of 240 MeV $\alpha$-particles. We refer to other contributions in these proceedings for reports on these experimental data$^{2)}$. Monopole strength functions turn out to be quite fragmented for nuclei lighter than $^{90}$Zr. For nuclei like $^{208}$Pb, $^{144}$Sm, $^{116}$Sn and $^{90}$Zr, however, one is able to identify a single peak which, together with a high-energy extended tail, exhausts essentially all the monopole Energy Weighted Sum Rule (EWSR). These medium-heavy nuclei are, therefore, those suited for the extraction of information about the nuclear incompressibility and we concentrate ourselves on them in the present work. The ISGDR is excited by the operator $\sum_i r_i^3 Y_{10}$ and corresponds to a compression of the nucleus along a definite direction, so that it has been called sometimes the ``squeezing mode''. Although some first indication about the energy location of this resonance dates back to the beginning of the eighties, a more clear indication about its strength distribution in $^{208}$Pb has been reported only recently$^{3)}$. Measurements have been done also for other nuclei, namely $^{90}$Zr, $^{116}$Sn and $^{144}$Sm (as in the case of the giant monopole resonance). There is some expectation that the study of this mode can help to shed some light on the problem of nuclear incompressibility. Actually, at first sight this compressional mode seems to provide us with a new problem. A simple assumption like the scaling model (illustrated for the present purposes in Ref.$^{4)}$) would lead to two different values of the finite-nucleus incompressibility $K_A$ if applied to the ISGMR and the ISGDR with the input of their experimental energies. The hydrodynamical model gives two results which are closer$^{4)}$ but which still make us wonder about the validity of methods based on extracting $K_A$ and extrapolating it to large values of $A$, for the determination of $K$. This points again to the necessity of reliable microscopic calculations of the compressional modes, in order to reproduce the experimental data and extract the value of $K$ from the properties of the force which is used. Our calculations are performed within the framework of self-consistent Hartree-Fock (HF) plus Random-Phase Approximation (RPA). We use effective forces of Skyrme type$^{5-7)}$ and we look at their predictions for the properties of ISGMR and ISGDR. The parametrizations we employ span a large range of values for $K$ (from 200 MeV to about 350 MeV). In particular, we focus mainly on two original aspects: firstly, we look at the effects of pairing correlations in open-shell nuclei; secondly, we study if the picture obtained at mean-field level is altered by the inclusion of the coupling of the giant resonances to more complicated nuclear configurations. This inclusion is necessary if one wishes to understand theoretically all the contributions to the resonance width and may shift the resonance centroid. About the first aspect, it is well known that pairing correlations are important in general to explain the properties of ground states and low-lying excited states in open-shell nuclei. Since we wish to see how these correlations affect in particular the compressional modes, we take them into account by extending the HF-RPA approach to a quasi-particle RPA (QRPA) on top of a HF-BCS calculation. About the second aspect, we recall that if we start from a description of the giant resonance as a superposition of one particle-one hole (1p-1h) excitations, in their damping process we must take care of the coupling with states of 2p-2h character. They are in fact known to play a major role and give rise to the spreading width $\Gamma^\downarrow$ of the giant resonance which is usually a quite large fraction of the total width. Within mean field theories, only the width associated with the resonance fragmentation (Landau width) and the escape width $\Gamma^\uparrow$ are included (the latter, provided that 1p-1h configurations with the particle in the continuum are considered). In the past, we have developed a theory in which all the contributions to the total width of giant resonances are consistently treated and we have obtained satisfactory results when applying it to a number of cases. In particular, we will recall what has been obtained$^{8)}$ for the case of the ISGMR in $^{208}$Pb. We also report about a new calculation for the ISGDR in the same nucleus. \section*{Formalism: a brief survey} For all nuclei we consider, we solve the HF equations on a radial mesh and, in the case of the open-shell isotopes, we solve HF-BCS equations. A constant pairing gap $\Delta$ is introduced (for neutrons in the case of $^{116}$Sn and for protons in the case of $^{90}$Zr and $^{144}$Sm), and at each HF iteration the quasi-particle energies, the occupation factors and the densities to be input at the next iteration are determined accordingly. $\Delta$ is obtained from the binding energies of the neighboring nuclei$^{9)}$. The states included in the solution of the HF-BCS equations are those below a cutoff energy given by $\lambda_{HF}+8.3$ MeV ($\lambda_{HF}$ being the HF Fermi energy), in analogy with the procedure of Ref.$^{10)}$. Using the above self-consistent mean fields we work out the RPA or QRPA equations (respectively on top of HF or HF-BCS), in their matrix form. Discrete positive energy states are obtained by diagonalizing the mean field on a harmonic oscillator basis and they are used to build the 1p-1h (or 2 quasi-particles) basis coupled to $J^\pi$=0$^+$ or 1$^-$. The dimension of this basis is chosen in such a way that more than 95\% (typically 97-99\%) of the appropriate EWSR is exhausted in the RPA or QRPA calculation. More details, especially on the way the QRPA equations are implemented, will be given in Ref.$^{11)}$. As mentioned in the previous section, in the case of $^{208}$Pb we perform also calculations that go beyond this simple discrete RPA. This is done along the formalism described in Ref.$^{12)}$, which is recalled here only very briefly. We label by $Q_1$ the space of discrete 1p-1h configurations in which the RPA equations are solved. To account for the escape width $\Gamma^\uparrow$ and spreading width $\Gamma^\downarrow$ of the giant resonances, we build two other orthogonal subspaces $P$ and $Q_2$. The space $P$ is made of particle-hole configurations where the particle is in an unbound state orthogonal to all the discrete single-particle levels; the space $Q_2$ is built with the configurations which are known to play a major role in the damping process of giant resonances: these configurations are 1p-1h states coupled to a collective vibration. Using the projection operator formalism one can easily find that the effects of coupling the subspaces $P$ and $Q_2$ to $Q_1$ are described by the following effective Hamiltonian acting in the $Q_1$ space: \begin{eqnarray} {\cal H} (E) \equiv Q_1 H Q_1 & + & W^\uparrow(E) + W^\downarrow(E) \nonumber \\ = Q_1 H Q_1 & + & Q_1 H P {\textstyle 1 \over \textstyle E - PHP + i\epsilon} P H Q_1 \nonumber \\ \ & + & Q_1 H Q_2 {\textstyle 1 \over \textstyle E - Q_2 H Q_2 + i\epsilon} Q_2 H Q_1, \nonumber \label{H_eff}\end{eqnarray} where $E$ is the excitation energy. For each value of $E$ the RPA equations corresponding to this effective, complex Hamiltonian ${\cal H} (E)$ are solved and the resulting sets of eigenstates enable us to calculate all relevant quantities, in particular the strength function associated with a given operator. To evaluate the matrix elements of $W^\downarrow$, we calculate the collective phonons with the same effective interaction used for the giant resonance we are studying (within RPA), and we couple these phonons with the 1p-1h components of the giant resonance by using their energies and transition densities. \section*{Results for the isoscalar monopole resonance} As recalled in the introduction, Youngblood {\it et al.}$^{2)}$ have recently measured the ISGMR strength distribution with fairly good precision, in the nuclei $^{90}$Zr, $^{116}$Sn, $^{144}$Sm and $^{208}$Pb. In their work, they also compare the experimental centroid energies with the calculations of Blaizot {\it et al.}$^{13)}$ performed by using RPA and employing the finite-range Gogny effective interaction: a value of the nuclear incompressibility $K$ = 231 MeV is deduced. In the following, we denote as centroid energy the ratio $E_0\equiv m_1/m_0$ ($m_0$ and $m_1$ being the non-energy-weighted and energy-weighted sum rules, respectively). If we try to compare the experimental values with calculations done at the same RPA level but using the zero-range Skyrme effective interactions, we can infer a different conclusion with respect to the value of $K$. Among the Skyrme type interactions, the parametrization which gives probably the best account of the experimental centroid energies in the nuclei studied by the authors of Ref.$^{2)}$, is the SGII force$^{6)}$. The results are shown in Table 1. The force SGII is characterized by a value of the nuclear incompressibility $K$ = 215 MeV: since it reproduces very well the ISGMR centroid energy in $^{208}$Pb, and it slightly overestimates those in the other isotopes, one would conclude that $K$ is of the order of or slightly less than 215 MeV. This conclusion is inferred by means of simple RPA. It is of course legitimate to wonder if calculations beyond this simple approximation could lead to different values of the nuclear incompressibility. We first consider the effect of pairing correlations. In the case of $^{116}$Sn, the centroid energy of 17.18 MeV obtained with the SGII force in RPA, becomes 17.19 MeV if one turns to QRPA. A very small shift is found also when other forces are used (for instance, with the recently proposed SLy4 force$^{7)}$, one obtains 17.51 MeV and 17.59 MeV for RPA and QRPA, respectively) and when other nuclei are considered. In general, although we know that pairing correlations play a crucial role not only to explain the ground-state of open-shell nuclei but also their low-lying excited states, it appears that they do not affect so much the giant resonances like the ISGMR (or ISGDR, anticipating results of the next section) which lie at relatively high excitation energy compared to the pairing gap $\Delta$. Civitarese {\it et al.}$^{14)}$ found also small shifts (of the order of 100-150 keV) for the ISGMR and ISGQR when pairing correlations are taken into account: this shift is larger than that obtained in the present work, but it is the result of a different (non self-consistent) model. The present conclusion for the nuclear incompressibility is therefore similar to that obtained by Hamamoto {\it et al.}$^{15)}$, since they find that the Skyrme interaction which provides the best results for the ISGMR is the SKM$^*$ parametrization and this is very similar to SGII (the associated nuclear incompressibility being 217 MeV). Our study, however, is done in a more general framework since we have analyzed also the role of pairing correlations. If we finally consider the results of calculations beyond mean field$^{8)}$ (which include not only the continuum coupling but also the coupling with the 2p-2h type states) performed for $^{208}$Pb we find that it is also possible to reproduce rather well the total width of the ISGMR, which is around 3 MeV. This width is actually in large part a consequence of fragmentation (or Landau damping): at least three states share, at the level of RPA, the resonance strength, but continuum as well as 2p-2h couplings are able to give to each peak the correct width so that the overall lineshape coincides with the experimental findings. We stress that the coupling with the 2p-2h type states is also responsible for a downward shift of the ISGMR centroid and peak energies, which is of the order of 0.5 MeV. One may argue that this affects the extraction of the value of $K$ from theoretical calculations. Actually, since the value of $K$ associated with a given force is obtained by a calculation of nuclear matter at the mean field level, it is legitimate to draw conclusions about $K$ from the comparison with the experiment of the ISGMR results for finite nuclei obtained again at the mean field level. But the fact that a given force is able to account for the ISGMR linewidth enforces our confidence about its reliability. And it would be of course legitimate either, to compare the centroid energies obtained after 2p-2h coupling with experiment provided the value of $K$ associated with the force is calculated by including the same couplings at the nuclear matter level. No such calculations in nuclear matter have been done so far, to our knowledge. \begin{table}[here] \caption{Experimental and theoretical values of the centroid energies $E_0\equiv m_1/m_0$ for the ISGMR and ISGDR. The theoretical values are obtained with the Skyrme-RPA approach, using the SGII interaction. All values are in MeV.} \vspace{0.3cm} \begin{tabular}{|c|r|r|r|r|} \hline & \multicolumn{2}{|c|}{ISGMR} & \multicolumn{2}{|c|}{ISGDR} \\ \cline{2-5} & Exp. & Theory & Exp. & Theory \\ \hline $^{90}$Zr & 17.9 & 19.1 & 26.2 & 27.1 \\ $^{116}$Sn & 16.0 & 17.2 & 23.0 & 26.3 \\ $^{144}$Sm & 15.3 & 16.2 & 24.2 & 25.4 \\ $^{208}$Pb & 14.2 & 14.1 & 20.3 & 24.1 \\ \hline \end{tabular} \label{table1} \end{table} \vspace{0.2cm} \section*{Results for the isoscalar dipole resonance} A peculiar feature of calculations of this giant resonance is the appearance of a spurious state in the calculated spectrum. When diagonalizing the RPA matrix on a 1$^-$ basis, we expect to see among all states the spurious state at zero energy corresponding to the center-of-mass motion and we expect as well that it exhausts the whole strength associated with the operator $\sum_i r_i Y_{10}$. Due to a lack of complete self-consistency (some part of the residual interaction, like the two-body spin-orbit and Coulomb forces, are usually neglected in the RPA because their effect should be rather small) and to numerical inaccuracies, this is not the case. The spurious state comes out in practice at finite energy and its wave function does not overlap completely with that of the exact center-of-mass motion: as a consequence, the remaining RPA eigenstates are not exactly orthogonal to the true spurious state, and their spurious component must be projected out. This is not difficult if RPA is done in the discrete p-h space. In any case, it can be shown$^{16)}$ that this projection procedure is equivalent to replacing the $\sum_i r_i^3 Y_{10}$ operator with $\sum_i (r_i^3-\eta r_i) Y_{10}$, $\eta$ being ${5\over 3}\langle r^2 \rangle$. Once this projection is done, we find that a substantial amount of $r^3 Y_{10}$ strength still remains in the 10 - 15 MeV region, in addition to the strength in the 20 - 25 MeV region. The data do not show any strength in the lower region, i.e., experimental centroid energies correspond to the energy region above 15 MeV. Therefore, to have a meaningful comparison with experiment we will refer from now on to theoretical centroid energies calculated in the interval 15 - 40 MeV. In Table 1 we show the ISGDR centroid energies obtained with the SGII force. Especially in the case of $^{208}$Pb and $^{116}$Sn, it can be noticed that RPA calculations tend to overestimate the value of the centroid energy, the discrepancy being less severe in the other two cases. One may wonder if this is a special feature of SGII, although this force has been said to behave rather well for the monopole case. Fig. 1 shows that this is not the case: the centroid energies obtained in RPA with a number of different Skyrme parametrizations are plotted as a function of their incompressibility $K$, and it can be noticed that all forces systematically overestimate the experimental values of the centroid energies. Gogny interactions have also been used to study the ISGDR in this and other nuclei$^{17)}$, but they also predict too large centroid energies. The same can be said about relativistic models like relativistic RPA$^{18)}$ or time-dependent relativistic mean field$^{19)}$. We may conclude that the case of the ISGDR in $^{208}$Pb is a kind of exception among the giant resonances studied within the self-consistent HF-RPA approach, as usually one never finds such large discrepancies between theory and experiment ($\sim$ 4-5 MeV). We finally address the question whether the coupling with more complicated configurations, which has been seen to be responsible of a downward shift of the resonance energy, can diminish this discrepancy in the case of $^{208}$Pb. We have done for the ISGDR a calculation of the type described above in the monopole case. The resulting strength function, which includes continuum and 2p-2h coupling, is depicted in Fig. 2. One can see that the total width of the resonance is quite large, and the theoretical value of about 6 MeV compares well with the experimental result which is about 7 MeV$^{20)}$. Although the resonance lineshape is accounted for by theory, the downward shift with respect to the RPA result is only about 1 MeV. The fundamental problem why the ISGDR energy in $^{208}$Pb cannot be reproduced by theoretical models still remains. \section*{Conclusion} In this paper, we have considered the isoscalar monopole and dipole resonances in a number of nuclei and we have tried to reproduce their properties by means of HF-RPA, or QRPA on top of HF-BCS, or more sophisticated approach which takes care of the continuum properly and of the coupling with nuclear configurations which are more complicated than the simple 1p-1h. In general, we have found that the effect of pairing correlations is quite small as these resonances lie at high energy with respect to the pairing gap $\Delta$. Concerning the RPA results, the situation looks different in the case of monopole and dipole. In the former case, the Skyrme-type force SGII is able to reproduce well the centroid energy in $^{208}$Pb, and it slightly overestimates this energy in other medium-heavy nuclei which have been measured accurately in recent experiments. This would allow us to extract a value of the nuclear incompressibility around 215 MeV. In the case of the ISGDR, however, the same Skyrme force overpredicts this centroid energy in $^{208}$Pb by about 4 MeV. Other parametrizations of Skyrme type cannot do better, and the problem is not solved if one turns to Gogny interactions or to relativistic models. Therefore, although in other nuclei this discrepancy between theory and experiment can be less than in the case of $^{208}$Pb, we can say that the ``squeezing mode'', which could be taken as a further probe of the nuclear incompressibility besides the well-known ``breathing'' monopole oscillation, is challenging us with a new problem. Calculations beyond mean field do not change substantially our conclusions about the centroid energies. However, we stress that these calculations are necessary for a proper account of the giant resonances lineshape, and in fact in our case they have been able to reproduce the total width of the ISGMR in $^{208}$Pb, and also of the ISGDR although its centroid is overestimated. We would like to thank Umesh Garg for stimulating discussions and for communicating experimental data prior to publication, and also Jean-Paul Blaizot for useful discussions. \begin{figure}[h \caption{RPA centroid energies of ISGDR calculated in $^{208}$Pb with various Skyrme interactions, as a function of compression modulus $K$. } \end{figure} \begin{figure}[h \caption{Calculated ISGDR strength distribution of operator $(r^3 - \eta r) Y_{10}$ including damping effects of particle-hole-plus-phonon coupling (see text). The nucleus is $^{208}$Pb. } \end{figure} \section*{References} \re 1) J.\ P.\ Blaizot: Phys.\ Rep.\ {\bf 64}, 171 (1980); J.\ M.\ Pearson: Phys.\ Lett.\ {\bf B 271}, 12 (1991); S.\ Shlomo and D.\ H.\ Youngblood: Phys.\ Rev.\ {\bf C 47}, 529 (1993). \re 2) D.\ H.\ Youngblood, H.\ L.\ Clark and Y.- W.\ Lui: Phys.\ Rev.\ Lett.\ {\bf 82}, 691 (1999). \re 3) B.\ Davis {\it et al.}: Phys.\ Rev.\ Lett.\ {\bf 79}, 609 (1997). \re 4) S.\ Stringari: Phys.\ Lett.\ {\bf B106}, 232 (1982). \re 5) M.\ Beiner, H.\ Flocard, N.\ Van Giai and Ph.\ Quentin: Nucl.\ Phys.\ {\bf A238}, 29 (1975). \re 6) N.\ Van Giai and H.\ Sagawa: Phys.\ Lett.\ {\bf B106}, 379 (1981). \re 7) E.\ Chabanat, P.\ Bonche, P.\ Haensel, J.\ Meyer and R.\ Schaeffer: Nucl.\ Phys.\ {\bf A635}, 231 (1998). \re 8) G.\ Col\`o, P.\ F.\ Bortignon, N.\ Van Giai, A.\ Bracco and R.\ A.\ Broglia: Phys.\ Lett.\ {\bf B276}, 279 (1992). \re 9) A.\ Bohr and B.\ M.\ Mottelson, Nuclear Structure, vol. I, W.A. Benjamin 1969, Eqs. (2.92) and (2.93). \re 10) N.\ Tajima, S.\ Takahara and N.\ Onishi: Nucl.\ Phys.\ {\bf A603}, 23 (1996). \re 11) E.\ Khan, N.\ Van Giai and G.\ Col\`o: to be published. \re 12) G.\ Col\`o, N.\ Van Giai, P.\ F.\ Bortignon and R.\ A.\ Broglia: Phys.\ Rev.\ {\bf C50}, 1496 (1994). \re 13) J.\ P.\ Blaizot, J.\ F.\ Berger, J.\ Decharg\'e and M.\ Girod: Nucl.\ Phys.\ {\bf A591}, 435 (1995). \re 14) O.\ Civitarese, A.\ G.\ Dumrauf, M.\ Reboiro, P.\ Ring and M.\ M.\ Sharma: Phys.\ Rev.\ {\bf C43}, 2622 (1991). \re 15) I.\ Hamamoto, H.\ Sagawa and X.\ Z.\ Zhang: Phys.\ Rev.\ {\bf C56}, 3121 (1997). \re 16) N.\ Van Giai and H.\ Sagawa: Nucl.\ Phys.\ {\bf A371}, 1 (1981). \re 17) J.\ Decharg\'e and L.\ \v{S}ips: Nucl.\ Phys.\ {\bf A407}, 1 (1983). \re 18) N.\ Van Giai and Z.\ Y.\ Ma: these proceedings. \re 19) D.\ Vretenar {\it et al.}: these proceedings. \re 20) U.\ Garg {\it et al.}: Proc. Topical Conference on Giant Resonances, Varenna ( Nucl.\ Phys.\ {\bf A}, to be published; H.\ L.\ Clark {\it et al.}: ibid.\ ; U.\ Garg: private communication. \end{document}
\section{Introduction} Before the success of Weinberg-Salam model, the weak interaction was characterized by the dimensional Fermi's coupling constant, $G_{F}= (300 Gev)^{-2}$, far below the electro-weak scale. But later it turns out that the dimensional coupling constant is the low energy effective coupling which is determined by the dimensionless electro-weak coupling constant and the vacuum expectation value of Higgs scalar field through the spontaneous symmetry breaking. The weakness of the weak interaction is originated from the large vacuum expectation value of Higgs field \cite{Weinberg}. From this lesson, it is suspected that gravity may be also characterized by a dimensionless coupling constant $\xi$ with the gravitational constant $G_{N}$ given by the inverse square of the vacuum expectation value of a scalar field. The weakness of the gravity can be associated with a symmetry breaking at a very high energy scale. It has been independently proposed by Zee \cite{Zee}, Smolin \cite{Smolin}, and Adler \cite{Adler} that the Einstein-Hilbert action can be replaced by the induced gravity action \begin{equation} S = \int d^{4}x \sqrt{-g}\{\frac{1}{2}\xi \phi^{2} R + \frac{1}{2} \partial_{\mu}\phi \partial^{\mu}\phi - V(\phi)\}, \label{action} \end{equation} where the coupling constant $\xi$ is dimensionless. The potential $V(\phi)$ is assumed to have its minimum value at $\phi = \sigma$, then the above action is reduced to the well known Einstein-Hilbert action with gravitational constant $G_{N}=\frac{1}{8\pi\xi\sigma^{2}}$. In the analogy of the $SU(2) \times U(1)$ symmetry of the electro-weak interactions, we can consider a symmetry which may be broken through a spontaneous symmetry breaking in the gravitational interactions. The most attractive candidate symmetry is the Weyl's conformal symmetry which rejects the Einstein-Hilbert action, but admits the induced gravity action Eq.(\ref{action}) with a specific conformal coupling, $\xi = \frac{1}{6}$. In Riemann space, the conformal coupling is unique with $\xi=\frac{1}{6}$. However, introducing the vector torsion, an extended conformal coupling is possible in induced gravity \cite{yoon95a,yoon95b} because the vector torsion plays the role of a conformal gauge field in Riemann-Cartan space \cite{Smolin,Nieh,yoon88}. It is found that induced gravity at conformal coupling should have conformal invariance for consistency \cite{yoon95b,yoon97}. We investigate the conformal coupling in induced gravity with a gradient torsion. \section{Conformal Couplings in Induced Gravity} The induced gravity action Eq.(\ref{action}) is invariant under the conformal transformation, \begin{equation} g_{\mu\nu}'(x) = exp(2\rho) g_{\mu\nu}(x), ~~~ \phi'(x) = exp(-\rho) \phi(x), \end{equation} at the conformal coupling $\xi = \frac{1}{6}$ for a conformally invariant scalar potential. In Riemann-Cartan space, an extension of conformal coupling with the torsion in induced gravity is possible. It is found that the minimal extension to Riemann-Cartan space is sufficient for our purpose. The conformal transformation of the affine connections $\Gamma^{\gamma}_{~\beta \alpha}$ is determined from the invariance of the tetrad postulation, \begin{equation} D_{\alpha} e^{i}_{\beta}\equiv\partial_{\alpha}e^{i}_{\beta} + \omega^{i}_{j\alpha} e^{j}_{\beta}- \Gamma^{\gamma}_{~\beta \alpha}e^{i}_{\gamma} = 0, \label{cor} \end{equation} under the following tetrads $e^{i}_{\alpha}$ transformations; \begin{equation} (e^{i}_{\alpha})'= exp(\rho) e^{i}_{\alpha}. \label{cor1} \end{equation} The spin connections $\omega^{i}_{j\alpha}$ are conformally invariant like other gauge fields. The affine connections and the torsions which are the antisymmetric components of the affine connections transform as follows; \begin{equation} (\Gamma^{\gamma}_{~\beta \alpha})' =\Gamma^{\gamma}_{~\beta \alpha} + \delta^{\gamma}_{~\beta} \partial_{\alpha} \rho, ~~~ (T^{\gamma}_{~\beta \alpha})'= T^{\gamma}_{~\beta \alpha}+ \delta^{\gamma}_{\beta}\partial_{\alpha}\rho - \delta^{\gamma}_{\alpha}\partial_{\beta}\rho. \label{tor1} \end{equation} Therefore, the trace of the torsion $T^{\gamma}_{~\gamma \alpha}$ effectively plays the role of a conformal gauge field. In general, the torsion can be decomposed into three components; \begin{equation} T^{\alpha}_{~\beta\gamma}= \Sigma^{\alpha}_{~\beta \gamma} + A^{\alpha}_{~\beta \gamma} - \delta^{\alpha}_{\gamma}S_{\beta} + \delta^{\alpha}_{\beta}S_{\gamma}, \label{tor2} \end{equation} where $\Sigma_{[\alpha\beta\gamma]} \equiv 0, ~~ \Sigma^{\alpha}_{~\alpha\gamma} \equiv 0, ~~ A_{\alpha\beta\gamma} \equiv T_{[\alpha\beta \gamma]}$. The traceless part of torsion $C^{\alpha}_{~\beta \gamma} \equiv \Sigma^{\alpha}_{~\beta\gamma} + A^{\alpha}_{~\beta\gamma}$ is conformally invariant. \begin{equation} (S_{\alpha})' = S_{\alpha} + \partial_{\alpha}\rho, ~~~ (C^{\alpha}_{~\beta \gamma})'= C^{\alpha}_{~\beta \gamma} . \label{vec} \end{equation} Because the minimal extension to Riemann-Cartan space is sufficient for our purpose, we impose the conformally invariant torsionless condition; \begin{equation} C^{\alpha}_{~\beta \gamma} \equiv 0. \label{zero} \end{equation} This condition is the conformally invariant extension of the torsionless condition in Riemann space, $ T^{\alpha}_{~\beta \gamma}\equiv 0$. For this minimally extended Riemann-Cartan space, the affine connection can be written in terms of $g_{\mu\nu}$ and $S_{\alpha}$; \begin{equation} \Gamma^{\alpha}_{~\beta\gamma}= \{^{\alpha}_{\beta\gamma}\}+ S^{\alpha} g_{\beta\gamma} - S_{\beta}\delta^{\alpha}_{\gamma}. \label{cristo} \end{equation} Introducing the conformally covariant derivative $D_{\alpha}$ for scalar field $\phi$, \begin{equation} D_{\alpha}\phi \equiv \partial_{\alpha}\phi + S_{\alpha}\phi, \end{equation} we have an extended conformal coupling of induced gravity up to total derivatives as follow; \begin{equation} I =\int d^{4}x \sqrt{-g} \{\frac{\xi}{2}R(\Gamma)\phi^{2} + \frac{1}{2}D_{\alpha}\phi D^{\alpha}\phi - \frac{1}{4}H_{\alpha\beta}H^{\alpha\beta} - V(\phi)\}, \label{gen} \end{equation} where we have excluded the curvature square terms. Now, the coupling $\xi$ is a dimensionless arbitrary constant. Using Eq.(\ref{cristo}) we can rewrite this action in terms of Riemann curvature scalar $R(\{\})$; \[ I=\int d^{4}x \sqrt{-g} \{\frac{\xi}{2}R(\{\})\phi^{2} + \frac{1}{2}\partial_{\alpha}\phi\partial^{\alpha}\phi + (1-6\xi) S^{\alpha}(\partial_{\alpha}\phi)\phi \] \begin{equation} + \frac{1}{2}(1-6\xi)S_{\alpha}S^{\alpha}\phi^{2} - \frac{1}{4}H_{\alpha \beta}H^{\alpha \beta} - V(\phi)\}. \label{iaction} \end{equation} The more general form of induced gravity action can be considered \cite{buch1,buch2}, but we restrict the couplings to be conformal. In the limit of $\xi \rightarrow \frac{1}{6}$, this extended conformal coupling is reduced to the ordinary conformal coupling in Riemann space decoupled from the vector torsion. \section{Dilaton Gravity from Induced Gravity with Gradient Torsion} Analyzing the equations of motion for the action Eq.(\ref{iaction}) with an effective potential $V_{eff}(\phi;S_{\alpha},g_{\beta\gamma})$ which depends on metric and torsion in general, we obtain the following two equations of motion and a constraint for the scalar potential; \begin{equation} \nabla_{\mu}H^{\mu\nu}= -(1-6\xi)\{(\partial^{\nu}\phi)\phi+S^{\nu}\phi^{2}\} +\frac{\partial V_{eff}(\phi;S_{\alpha},g_{\beta\gamma})}{\partial S_{\nu}}~, \label{box1} \end{equation} \[ \xi\phi^{2}G_{\mu\nu}= (H_{\mu\alpha}H_{\nu}^{~\alpha} -\frac{1}{4}g_{\mu\nu}H_{\alpha\beta}H^{\alpha\beta}) - (\partial_{\mu}\phi\partial_{\nu}\phi -\frac{1}{2} g_{\mu\nu}\partial_{\alpha}\phi\partial^{\alpha}\phi) - (1-6\xi)\phi^{2}(S_{\mu}S_{\nu}-\frac{1}{2}g_{\mu\nu}S_{\alpha}S^{\alpha}) \] \[ - (1-6\xi)(S_{\mu}\phi\partial_{\nu}\phi+S_{\nu}\phi\partial_{\mu}\phi - g_{\mu\nu}S^{\alpha}\phi\partial_{\alpha}\phi) + \xi\{\nabla_{\mu}(\phi\partial_{\nu}\phi)+\nabla_{\nu}(\phi\partial_{\mu}\phi) -g_{\mu\nu}\Box\phi^{2}\} \] \begin{equation} - g_{\mu\nu}V_{eff}(\phi;S_{\alpha},g_{\beta\gamma}) + 2\frac{\partial V_{eff}(\phi;S_{\alpha},g_{\beta\gamma})} {\partial g^{\mu\nu}}~, \label{long} \end{equation} \begin{equation} 4V_{eff}(\phi;S_{\alpha},g_{\beta\gamma}) -\phi\frac{\partial V_{eff}(\phi;S_{\alpha},g_{\beta\gamma})}{\partial\phi} =2\frac{\partial V_{eff}(\phi;S_{\alpha},g_{\beta\gamma})} {\partial g^{\mu\nu}}g^{\mu\nu} +\nabla_{\nu}\frac{\partial V_{eff}(\phi;S_{\alpha},g_{\beta\gamma})} {\partial S_{\nu}}~, \label{combix} \end{equation} where all covariant derivatives are in Riemann space with the Christoffel connections \cite{yoon95b,yoon97}. The constraint Eq.(\ref{combix}) requires that the metric independent bare potential should be quartic in the scalar field, $V_{o}(\phi)=\frac{\lambda}{4!}\phi^{4}$, and the deviation of the radiatively corrected effective potential from the quartic form is only allowed with the compensation by the metric and vector torsion dependencies of the effective potential. Because this constraint comes from the assumption that the bare action is conformally invariant except the potential term, if we consider non-conformal coupling in kinetic and interacting terms, such a constraint would not appear. Let us consider a reduction of the system. If the effective potential does not have the vector torsion dependency, i.e. $V_{eff}(\phi;g_{\beta\gamma})$, Eq.(\ref{box1}) allows the following conformally invariant reduction; \begin{equation} D_{\alpha}\phi = 0. \end{equation} This implies that the vector torsion is a gradient form; \begin{equation} S_{\alpha}=-\partial_{\alpha}ln(\phi/\phi_{o}) = - \partial_{\alpha}\sigma, ~~~ \phi \equiv \phi_{o}e^{\sigma}, \end{equation} where $\phi_{o}$ is a dimensional constant, and the field strength of the vector torsion vanishes $H_{\alpha\beta}=0$, which is consistent with Eq.(\ref{box1}). In this reduction, the bare action of Eq.(\ref{gen}) becomes \begin{equation} I =\int d^{4}x \sqrt{-g}\phi_{o}^{2}e^{2\sigma} \{ \frac{\xi}{2}R(\{\}) + 3\xi\partial_{\alpha}\sigma \partial^{\alpha}\sigma - \frac{\lambda}{4!}\phi_{o}^{2}e^{2\sigma} \}. \label{gen1} \end{equation} This is the form of conformal factor theory of dilaton gravity \cite{odintsov1,odintsov2}. If the antisymmetric torsion term is $\frac{1}{12}C_{\alpha\beta\gamma}C^{\alpha\beta\gamma}$ included, this action is the form of string gravity with some redefinition of fields except the quartic potential term \cite{Saa}. For the reduction of the effective action, the potential term is replaced by $e^{-2\sigma}V_{eff}(\phi_{o}e^{\sigma},g_{\alpha\beta})$. In this reduction, a dimensional constant $\phi_{o}$ is introduced, but the action is invariant under the global scaling $dx^{\mu} \rightarrow a dx^{\mu}$, $\phi_{o} \rightarrow \phi_{o}/a$ if no conformal anomaly is introduced in the effective action. However, the appearance of the conformal anomaly in the conformally induced gravity is not allowed due to the constraint Eq.(\ref{combix}) which is the requirement of conformal invariance \cite{yoon97} from the consistency of equations of motion for the bare and the effective action in the conformally induced gravity \cite{yoon95b}. In this reduction, the constraint which effective potential should satisfy is \begin{equation} 4V_{eff}(\phi;g_{\beta\gamma}) -\phi\frac{\partial V_{eff}(\phi;g_{\beta\gamma})}{\partial\phi} =2\frac{\partial V_{eff}(\phi;g_{\beta\gamma})} {\partial g^{\mu\nu}}g^{\mu\nu}~. \label{combix0} \end{equation} Therefore, only metric independent quartic potential with effective coupling $\lambda_{eff}$ seems to be allowed. But, in general, the quantum effect by quantum fluctuation of scalar field on the classical metric and torsion background \cite{odintsov1,odintsov2} gives the following corrections to the scalar mass $m^{2}$, the coupling $\lambda$ and the cosmological constant $\Lambda$ respectively \cite{Coleman,Dewitt}. \begin{equation} \delta m^{2} \propto \lambda m^{2},~~~ \delta\lambda \propto \lambda^{2},~~~ \delta\Lambda \propto a m^{2} + b \lambda. \end{equation} Therefore, $\lambda=0$ would be the only solution of Eq.(\ref{combix0}) as a trivial fixed point of the renormalization group. However, the definite claim about $\lambda=0$ is possible only after the consideration of full quantum effects of the theory, which is of course far beyond of our scope yet. Redefining the metric, \begin{equation} g_{\alpha\beta}e^{2\sigma} \rightarrow g_{\alpha\beta}, \end{equation} the above dilaton gravity action can be written in the following standard Einstein action; \begin{equation} I =\int d^{4}x \sqrt{-g} \{ \frac{\xi}{2}\phi_{o}^{2}R(\{\}) - \frac{\lambda}{4!}\phi_{o}^{4} \}, \label{gen2} \end{equation} where the gravitational constant $G_{N}= \frac{1}{8\pi\xi\phi_{o}^{2}}$, and the cosmological constant $\Lambda = \frac{\lambda}{4!}(\frac{1}{8\pi\xi G_{N}})^{2}$. If the effective potential deviates from the quartic form, then there would remain explicit $\sigma$ dependency in the action after this redefinition of metric even though it is not plausible due to the constraint Eq.(\ref{combix0}). In this Einstein frame the conformal symmetry is hidden and the cosmological constant term has the origin of the quartic potential term in the original induced gravity action. Reminding the constraint Eq.(\ref{combix0}) which the effective potential should satisfy, the non-zero coupling $\lambda$ can be hardly expected after the radiative correction by the scalar field in the original induced gravity action \cite{Coleman,Dewitt}. Therefore, we can say that if the Einstein gravity have the root in the conformally induced gravity, the non-zero cosmological constant is not plausible due to the constraint from the conformal coupling. In this discussion, we have not considered torsions generated by matter fields because vector torsion couplings are not expected in the standard minimal action for Dirac and gauge fields \cite{buch2}. ~\\~\noindent {\it Acknowledgments:} This work was supported in part by Korean Ministry of Education through the program BSRI-2441. The author wishes to acknowledge the financial support of Hanyang University, Korea, made in the program year of 1998.
\section{INTRODUCTION} In most physical problems, the degrees of freedom that are of interest are only a small fraction of the total number of degrees of freedom present. Some kind of reduction of the system has been performed. The way in which this reduction proceeds from quantum field theory to a moduli space of topological solitons is the subject of this paper. Soliton solutions can give rise to particle states \cite{coleman} and it has been predicted that their low-energy interations can be reduced to a simplified quantum mechanical problem in a set of collective coordinates \cite{gibbons,gauntlett}. One way in which this can be used is to indicate a duality between two quantum theories, where the soliton particle states in one theory become the fundamental particles in an equivalent quantum theory. A particular example can be seen in the monopole bound states which are important for the duality between the large and small coupling constant limits of the heterotic string theory \cite{sen}. In a range of different models there are classical multiple-soliton solutions that saturate Bogomol'nyi bounds and are parameterised by moduli. When the solitons move slowly and interact they deform adiabatically and the moduli trace out a path in a moduli space. The equations governing the path are equivalent to the equations for a point particle moving on a curved space \cite{manton,ward}. We would expect to be able to describe the interactions of quantized solitons by quantising the dynamical system induced on the moduli space. We will call this proceedure {\it truncation}. However, the quantisation rules of the truncated theory are often ambiguous without additional assumptions. If we impose general covariance, reflecting the freedom to choose coordinates on the moduli space, then the Hamiltonian operator takes the form \begin{equation} -\case1/2\Delta+\xi R+\dots\label{ambig} \end{equation} where $\Delta$ is the Laplacian on the moduli space and ambiguities are reduced to the coefficients $\xi$ of the Ricci scalar $R$ and other curvature invariants \cite{dewitt}. These ambiguities can be further reduced by introducing supersymmetry, in which case it has been argued that $\xi=0$ \cite{witten,alfaro}. In this paper we describe the alternative to truncating the full theory, which is to {\it reduce} it by a Born-Oppenheimer approximation. In a purely bosonic theory we will see that general covariance breaks down in the sense that a potential arises on the moduli space after integrating out the fluctuations that are orthogonal to the moduli space. In the basic situation we begin with a classical dynamical system in flat space which has a potential $V$ with a set of degenerate minima ${\cal M}$. We find that the original quantum theory reduces to a quantum theory on ${\cal M}$ with Hamiltonian \begin{equation} H_R=-\case1/2\Delta_{\cal M}+U +\case1/4R_{\cal M}-\case1/8|{\bf k}|^2+ +\dots \end{equation} The two purely geometrical terms depend on the intrinsic curvature scalar $R_{\cal M}$ and the (traced) extrinsic curvature ${\bf k}^I$. These geometrical terms where first written down by Maraner \cite{maraner}. The dynamical terms denoted by $U$ depend upon derivatives of the potential. Those depending on up to two extrinsic derivatives of the potential can be written \begin{equation} U=\case1/2{\rm tr}({\bf \omega})+\case1/{16}{\rm tr} \left({\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\cdot {\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\right)\label{vacuum} \end{equation} where $\omega$ is a matrix, ${\bf\omega}^2$ is the hessian matrix of $V$ and the covariant derivative is taken tangentially to ${\cal M}$. The application of these results to sigma models is considered in section 3. In these models the original configuration space, consisting of maps from physical space to a curved target space, is both infinite and curved. The infinities can be dealt with by regularisation. The curvature would lead in general to ambiguities even before the reduction. However, as shown in appendix A, these ambiguities are not present if the original space is a Riemannian symmetric space. Most models of interest fall into this class. The results are generalised to supersymmetric sigma models in section 4. The point at issue is whether the reduction leads to a superpotential on the moduli space and alters the hamiltonian operator. We find that, to leading order, the theory reduces to the Hodge Laplacian acting on $p-forms$, \begin{equation} H_R=-\case1/2\Delta_{\cal M}+\dots \end{equation} confirming previous expectations \cite{witten,alfaro,gauntlett}. Possible developments of this work are addressed in the conclusion, where we also discuss the advantages of the Hamiltonian approach over a Feynmann diagram approach. \section{BORN-OPPENHEIMER REDUCTION} Consider a classical dynamical system in flat space which has a degenerate set of stable points ${\cal M}$. If the potential is sufficiently steep, then energy spectrum of the quantum system typically falls into seperate bands. An adiabatic approximation scheme can then be used to reduce the quantum system to a quantum system on ${\cal M}$. We shall retrict attention to the case where the potential has an analytic expansion about its minima. We take a set of generalised coordinates $(x^a,x^I)$ such that $x^I=0$ on ${\cal M}$. In the lowest energy bands, the value of $x^I$ fluctuates over a narrow range of values determined by the eigenvalues of a matrix $\bf \omega$, where \begin{equation} {\bf \omega}^2=\partial_I\partial_JV. \end{equation} The expansion parameter of the approximation scheme is given by the width of an energy band divided by the smallest eigenvalue. The first step is to rewrite the Hamiltonian operator, \begin{equation} H=-\case1/2\Delta+V \end{equation} in the new coordinate system. We choose normal coordinates $x^I$ along the principal directions of $\partial_I\partial_JV$. It is always possible to find two mutually orthogonal sets of derivatives $\hat\partial_a$ and $\partial_I$, where \begin{equation} \hat\partial_a=\partial_a-N^I{}_a\partial_I. \end{equation} This implies that the metric can be written \begin{equation} {\bf g}=\gamma_{ab}dx^adx^b+\gamma_{IJ} (dx^I+N^I{}_adx^a)(dx^J+N^J{}_bdx^b). \end{equation} The metric components have normal coordinate expansions \begin{eqnarray} \gamma_{IJ}&=&\delta_{IJ}\label{nca}\\ N^I{}_a&=&-a_J{}^I{}_ax^J\\ \gamma_{ab}&=&\sigma_{ab}+2k_{I\,ab}x^I+k_{I\,ac}k_{J\,b}{}^cx^Ix^J\label{nce} \end{eqnarray} where the extrinsic curvature ${\bf k}^I$ and the torsion ${\bf a}_I{}^J$ are defined in appendix B. These expansions are exact if the space is flat, but we will keep in mind the possibility of generalising these results to curved spaces. In the new coordinate system, the Hamiltonian becomes \begin{equation} H=\case1/2|g|^{-1/2}\hat\partial_a|g|^{1/2}\gamma^{ab}\hat\partial_b+ \case1/2|g|^{-1/2}\partial_I|g|^{1/2}\delta^{IJ}\partial_J+V \end{equation} We regard the $x^I$ as small and expand the Hamiltonian as a series of terms $H=H_0+H_1+\dots$. The potential has a series expansion \begin{equation} V=\case1/2(\omega^2)_{IJ}x^Ix^J+\case1/6V_{IJK}x^Ix^Jx^K+\dots\label{pot} \end{equation} where $\omega_{IJ}$ is a diagonal matrix with eigenvalues $\omega_I$. The first few terms in the full Hamiltonian are then \begin{eqnarray} H_0&=&-\case1/2\delta^{IJ}\partial_I\partial_J+ \case1/2(\omega^2)_{IJ}x^Ix^J\label{unpert}\\ H_1&=&-\case1/2\,{\rm tr}({\bf k}^I)\partial_I+\case1/6V_{IJK}x^Ix^Jx^K \\ H_2&=&-\case1/2{\cal D}\cdot{\cal D} +\case1/2{\rm tr}({\bf k}_I{\bf k}^J)x^I\partial_J+ \case1/{24}V_{IJKL}x^Ix^Jx^Kx^L\label{htwo} \end{eqnarray} The dot denotes a scalar product using the intrinsic metric on ${\cal M}$. We have introduced a covariant derivative ${\cal D}_a$, where ${\cal D}_a\sigma_{ab}=0$ and \begin{equation} {\cal D}_af=(\partial_a+a_J{}^I{}_ax^J\partial_J)f\label{dcal} \end{equation} for scalar functions $f$. The next step is to apply degenerate perturbation theory. We take the unperturbed harmonic oscillator Hamiltonian $H_0$ together with a set of states $f_n$, where $H_0f_n=E_nf_n$. The perturbation theory is described in appendix C and gives a reduced Hamiltonian operator $H_R$ acting on the space of wave functions $\psi(x^a)$, \begin{equation} H_R=\langle H_0\rangle_{00}+\langle H_1\rangle_{00}+\langle H_2\rangle_{00}+ \sum_{m\ne0}\langle H_1\rangle_{0m}\left(E_0-E_m\right)^{-1}\langle H_1\rangle_{m0}+O(\omega^{-1})\label{degen} \end{equation} The matrix elements defined by \begin{equation} \langle H\rangle_{nm}=\int f_n^*\,H\,f_m \prod_Idx^I\label{matrix} \end{equation} are allowed to be operator-valued. The first term $\langle H_0\rangle_{00}$ consists of the vacuum energy of the system of unperturbed harmonic oscillators (\ref{unpert}), \begin{equation} E_0=\case1/2{\rm tr}({\bf \omega}). \end{equation} The second term $\langle H_1\rangle_{00}$ vanishes because the operator is odd under the inversion $x^I\to-x^I$. For the next term $\langle H_2\rangle_{00}$, with $H_2$ given by equation (\ref{htwo}), we set \begin{equation} \langle {\cal D}_a\rangle_{mn}= \delta_{mn}{}^\parallel\nabla_a+\langle {\cal A}_a\rangle_{mn} \end{equation} where ${}^\parallel\nabla_a$ is defined in appendix B and ${\cal A}$ arises form the action of ${\cal D}_a$ on the states $f_n$ in equation (\ref{matrix}). The matrix element of the derivative of an energy eigenstate is given by a standard identity, \begin{equation} \langle {\cal A}_a\rangle_{mn}={([{\cal D}_a,H_0])_{mn}\over E_n-E_m} \end{equation} After using equation (\ref{dcal}) for ${\cal D}$, and comparing with equation (\ref{grado}), the commutator produces a covariant derivative, \begin{equation} \langle {\cal A}_a\rangle_{mn}=\cases{ \case1/2{}^\parallel\nabla_a\omega_{IJ}\,\langle x^Ix^J\rangle_{mn}&$E_m<E_n$\cr 0&$m=n$\cr}. \end{equation} We can now evaluate \begin{equation} \langle {\cal D}\cdot{\cal D}\rangle_{00}=\Delta_{\cal M}- \case1/8{\rm tr}\left({\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\cdot {\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\right) \end{equation} where the covariant Laplacian $\Delta_{\cal M}= {}^\parallel\nabla\cdot{}^\parallel\nabla$. We also have, using (\ref{gauss}), \begin{equation} \langle{\rm tr}({\bf k}_I{\bf k}^J)x^I\partial_J\rangle_{00}= \case1/2{\rm tr}({\bf k}_I{\bf k}^I)= \case1/2|{\bf k}|^2-\case1/2R_{\cal M} \end{equation} where $|{\bf k}|^2={\rm tr}({\bf k}^I){\rm tr}({\bf k}_I)$. Therefore we get \begin{equation} \langle H_2\rangle_{00}=-\case1/2\Delta_{\cal M}-\case1/4|{\bf k}|^2 +\case1/4R_{\cal M}+\case1/{16} {\rm tr}\left({\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\cdot {\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\right) +\case1/{32}\sum_{IJ}\omega_I^{-1}\omega_J^{-1}V_{IJIJ} \end{equation} The last term given in equation (\ref{degen}) leads to \begin{equation} \case1/8\,|{\bf k}|^2 -\case1/{48}\sum_{IJK}(\omega_I+\omega_J+\omega_K)^{-1} \omega_I^{-1}\omega_J^{-1}\omega_K^{-1}(V_{IJK})^2 -\case1/{32}\sum_{IJK}\omega_I^{-2}\omega_J^{-1}\omega_K^{-1}V_{IJJ}V_{IKK} \end{equation} Putting the terms together gives a final expression for the reduced Hamiltonian, \begin{equation} H_R=-\case1/2\Delta_{\cal M}+U +\case1/4R_{\cal M}-\case1/8|{\bf k}|^2+O(\omega^{-1})\label{result} \end{equation} The terms are divided into purely geometrical terms and dynamical terms which are collected into $U$, where \begin{eqnarray} U&=&\case1/2{\rm tr}({\bf \omega})+\case1/{16}{\rm tr} \left({\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\cdot {\bf \omega}^{-1}{}^\parallel\nabla{\bf \omega}\right) +\case1/{32}\sum_{IJ}\omega_I^{-1}\omega_J^{-1}V_{IJIJ}\nonumber\\ &&-\case1/{48}\sum_{IJK}(\omega_I+\omega_J+\omega_K)^{-1} \omega_I^{-1}\omega_J^{-1}\omega_K^{-1}V_{IJK}{}^2 -\case1/{32}\sum_{IJK}\omega_I^{-2}\omega_J^{-1}\omega_K^{-1}V_{IJJ}V_{IKK} \end{eqnarray} The vacuum energy, which is of order $\omega$, is the dominant term in this expression. The next term depends on the gradient of the normal mode frequencies along the moduli space and includes the effects of the twisting, or torsion, of the normal mode directions. The remaining terms contain the effects of higher derivatives of the potential. The geometrical terms where discovered in the restricted situation of a surface embedded in $R^3$ by Jensen and Koppe \cite{jensen} and more generally by Maraner \cite{maraner}. The covariant forms for the dynamical terms beyond the obvious vacuum energy term are new, as far as we know. \section{BOSONIC FIELD THEORY} The non-linear sigma-model in two space dimension $x^\mu$ and one time dimension $t$ has fields $\phi^i$ taking values in a curved target space ${\cal T}$. We will take the Lagrangian to be $L=T-V$, where \begin{eqnarray} T&=&{1\over 2}\int g_{ij}\partial_t\phi^i\partial_t\phi^j\,d^2x\\ V&=&{1\over 2}\int g_{ij}\partial_\mu\phi^i\partial_\mu\phi^j\,d^2x \end{eqnarray} Depending upon the topology of ${\cal T}$, the classical field equations can have static solutions with finite energy, or topological solitons. These solutions minimise the potential energy for a given topological class. When ${\cal T}$ is a compact Kahler manifold the soliton solutions can be represented by holomorphic maps from the complex plane into ${\cal T}$ \cite{ward,din,ruback}. They depend on a continuous set of parameters, or moduli, which can be interpreted as the positions and charges of individual lumps. The initial situation is similar to the models considered in the previous section, except that the configuration space is no longer finite and flat. The introduction of curvature leads to the appearance of curvature invariants in the initial Schr\"odinger equation. However, we will take the target space to be a Riemannian symmetric space. The space of fields inherits this symmetry and is also a Riemannian symmetric space. In this case the curvature invariants are constants which can be absorbed into an overall phase factor. The soliton solutions $\phi_0$ will be parameterised by a set of coordinates $x^a$ belonging to the moduli space. We shall expand the Lagrangian about these solutions using standard background field methods \cite{luis} and express the result in terms of time-dependent coordinates $x^a$ and normal coordinates $x^I$. This will enable us to use the results of the previous section. The tangent vector $\xi$ to the geodesic from $\phi$ to $\phi_0$ provides a convenient measure of the displaced field. We also introduce the vectors ${\bf e}_\mu$ with target space components \begin{equation} {\bf e}_\mu{}^i=\partial_\mu\phi^i \end{equation} These can be parallel transported back from the tangent space at $\phi$ to the tangent space at $\phi_0$ to produce a field ${\bf e}'_\mu$. The tangent vector $\xi$ commutes with ${\bf e}_\mu$, and consequently $\nabla_\xi{\bf e}_\mu=D_\mu\xi$, where $D_\mu$ is the gradient along ${\bf e}_\mu$. This relation simplifies the Taylor series expansion for ${\bf e}'$, \begin{equation} {\bf e}'_\mu={\bf e}_\mu+D_\mu\xi+\case1/2R(\xi,{\bf e}_\mu)\xi +\case1/6R(\xi,D_\mu\xi)\xi+\dots\label{serie} \end{equation} We will define \begin{equation} ({\bf u},{\bf v})=\int g_{ij}(\phi_0)u^iv^j\,d^2x\label{product} \end{equation} By parallel transport of the Lagrangian density we can write the potential in the form \begin{equation} V=\case1/2({\bf e}'_\mu,{\bf e}'{}_\mu) \end{equation} The potential therefore has a series expansion in $\xi$ of the form $V=V_0+V_1+V_2\dots$ where \begin{eqnarray} V_0&=&\case1/2({\bf e}_\mu,{\bf e}_\mu)\label{vzero}\\ V_1&=&({\bf e}_\mu,D_\mu\xi)\label{vone}\\ V_2&=&\case1/2(D_\mu\xi,D_\mu\xi)+ \case1/2({\bf e}_\mu,R(\xi,{\bf e}_\mu)\xi)\label{vtwo}\\ V_3&=&\case2/3(D_\mu\xi,R(\xi,{\bf e}_\mu)\xi)\label{vthree}\\ V_4&=&\case1/6(D_\mu\xi,R(\xi,D_\mu\xi)\xi) +\case1/6(R(\xi,{\bf e}_\mu)\xi,R(\xi,{\bf e}_\mu)\xi) \end{eqnarray} Integration by parts in (\ref{vone}) produces the field equation \begin{equation} D_\mu{\bf e}_\mu= \partial_\mu e_\mu{}^i-\Gamma^i{}_{jk}e_\mu{}^je_\mu{}^k=0\label{fe} \end{equation} Integration by parts on (\ref{vtwo}) produces the operator \begin{equation} \Delta_f\xi=-D_\mu D_\mu\xi-R(\xi,{\bf e}_\mu){\bf e}_\mu,\label{fluc} \end{equation} which describes fluctuations about the moduli space. In the usual analysis of the sigma model, the inverse of this propagator would be the Green function in the background field approximation. The normalised positive modes of the fluctuation operator will be denoted by ${\bf u}_I$ and their eigenvalues by $(\omega_I)^2$. The fluctuation operator also has a set of zero modes ${\bf u}_a=\partial_a\phi_0$. All of the modes are parameterised by the collective coordinates $x^a$. We define the remaining coordinates in terms of the displacement vector, \begin{equation} \xi=x^I{\bf u}_I\label{xex} \end{equation} The potential then has a series expansion in the normal coordinates identical to equation (\ref{pot}) used in the previous section. From equation (\ref{vthree}), \begin{equation} V_{IJK}=4C_{(IJ)K}\label{pott} \end{equation} where \begin{equation} C_{IJK}=(D_\mu{\bf u}_I,R({\bf u}_J,{\bf e}_\mu){\bf u}_K).\label{cdef} \end{equation} The coefficients are symmetrical, since $C_{(IJ)K}=C_{(IJK)}$. They are also trace-free if the zero modes $u_a$ are included, $C^I{}_{IJ}+C^a{}_{aJ}=0$. The kinetic energy has a similar series expansion. Since \begin{equation} T=\case1/2({\bf e}'_t,{\bf e}'{}_t), \end{equation} we need only replace ${\bf e}_\mu$ by ${\bf e}_t$ and $D_\mu\xi$ by $D_t\xi$ in equations (\ref{vzero}-\ref{vthree}). However, both the $x^I$ and the $x^a$ depend on time, so that the time derivative of equation (\ref{xex}) implies \begin{equation} D_t\xi=\dot x^I\,{\bf u}_I+\dot x^a\,x^ID_a{\bf u}_I\label{dtxi} \end{equation} where the covariant derivative \begin{equation} D_a u^i=\partial_a u^i+u_a{}^k\Gamma^i{}_{jk}u^j. \end{equation} Also, since ${\bf u}_a=\partial_a\phi_0$, \begin{equation} {\bf e}_t=\partial_t\phi_0=\dot x^a{\bf u}_a.\label{et} \end{equation} Equations (\ref{dtxi}) and (\ref{et}) allow terms in the series expansion of the kinetic energy to be grouped as a quadratic polynomial in the generalised velocities, \begin{equation} T=\case1/2\gamma_{ab}\dot x^a\dot x^b+\case1/2\gamma_{IJ} (\dot x^I+N^I{}_a\dot x^a)(\dot x^J+N^J{}_b\dot x^b) \end{equation} To second order in $x^I$, \begin{eqnarray} \gamma_{IJ}&=&\delta_{IJ}-\case1/3R_{IKJL}x^Kx^L\\ N^I{}_a&=&-a_J{}^I{}_ax^J\\ \gamma_{ab}&=&\sigma_{ab}+2k_{I\,ab}x^I+k_{I\,ac}k_{J\,b}{}^cx^Ix^J -R_{aIbJ}x^Ix^J \end{eqnarray} The coefficients can be obtained by examining equations (\ref{vzero}-\ref{vthree}), \begin{eqnarray} \sigma_{ab}&=&({\bf u}_a,{\bf u}_b)\\ k^I{}_{ab}&=&({\bf u}_a, D_b{\bf u}_I)\\ a_I{}^J{}_a&=&({\bf u}_I, D_a{\bf u}_J)\\ R_{IJKL}&=&({\bf u}_I,R({\bf u}_K,{\bf u}_L){\bf u}_J) \end{eqnarray} These results recover the normal coordinate expansion of the previous section (\ref{nca})-(\ref{nce}), with extra terms due to the curvature of the original configuration space. We now have explicit expressions for the curvature and the extrinsic geometry of the space of soliton parameters in terms of eigenstates of the fluctuation operator. The results of the previous section still apply, except for the addition of terms depending on the target space curvature. A similar analysis to the previous section shows that two extra terms \begin{equation} \case1/4\sigma^{ac}\sigma^{bd}R_{abcd}+ \case1/6\sum_{IJ}\omega_I\omega_J^{-1}R_{IJIJ} \end{equation} should be included in the potential $U$. Expressions involving derivatives of the eigenmodes can be rewritten using the identity \begin{equation} ({\bf u}_I,D_a{\bf u}_J)={({\bf u}_I,[D_a,\Delta_f]{\bf u}_J)\over \omega_J^2-\omega_I^2} \end{equation} With the fluctuation operator $\Delta_f$ given explicitly in equation (\ref{fluc}), the commutator becomes \begin{equation} ({\bf u}_I,[D_a,\Delta_f]{\bf u}_J)=4C_{(IJ)a} \end{equation} where $C_{IJa}$ is defined as in equation (\ref{cdef}). This gives \begin{eqnarray} k_{I\,ab}&=&-{4\over\omega_I^2}C_{(Ia)b}\label{kexp}\\ {}^\parallel\nabla_a\omega_{IJ}&=&{4\over\omega_J+\omega_I} C_{(IJ)a}\label{goexp} \end{eqnarray} where equation (\ref{grado}) has been used. The dominant term in the reduced Hamiltonian should be the zero point energy $\case1/2{\rm tr}(\omega)$. In general, in order to evaluate this term it will be necessary to solve the eigenvalue problem for the fluctuation operator numerically. Once this has been done, equations (\ref{pott}), (\ref{kexp}) and (\ref{goexp}) could then be used to evaluate the geometrical terms in the reduced Hamiltonian (\ref{result}). \section{SUPERSYMMETRIC REDUCTION} We now turn to the reduction of an $N=2$ supersymmetric quantum system defined on a Riemannian symmetric space. As before, we supose that the system has a potential $V$ with a degenerate set of minima. The existence of two supersymmetries requires that the space is a Kahler manifold with a covariantly preserved complex structure $J$ \cite{zumino,alvarez}. We shall suppose that the moduli space of stable points is also a Kahler manifold. The variables consist of coordinates $x^\alpha$ and single complex component fermions $\psi^\alpha$. We also assume the existence of an operator $\pi_\alpha$, related to the momentum, and the following commutation relations \begin{eqnarray} \{\psi^\alpha,\psi^{*\beta}\}&=&g^{\alpha\beta}\\ \lbrack\pi_\alpha,x^\beta\rbrack&=&-i\delta_\alpha{}^\beta\\ \lbrack\pi_\alpha,\psi^\beta\rbrack &=&-i\Gamma^\beta{}_{\gamma\alpha}\psi^\gamma\\ \lbrack\pi_\alpha,\pi_\beta\rbrack&=& -iR_{\alpha\beta\gamma\delta}\psi^{*\gamma}\psi^\delta\label{coms} \end{eqnarray} involving the inverse metric $g^{\alpha\beta}$, connection components $\Gamma^\beta{}_{\gamma\alpha}$ and curvature components $R_{\alpha\beta\gamma\delta}$. It is possible to represent the commutator relations by covariant differential operators on a Hilbert space with the basis \begin{equation} \Psi(x)_{\alpha_1\dots\alpha_p} \psi^{*\alpha_1}\dots\psi^{*\alpha_p}|0\rangle, \end{equation} where $\psi^\alpha|0\rangle=0$. This gives a representation in which \begin{equation} \pi_\alpha=-i(\partial_\alpha- \Gamma_{\gamma\beta\alpha}\psi^{*\beta}\psi^\gamma)\label{pi} \end{equation} The action of $\pi$ is equivalent to $-i\nabla_\alpha$ acting on the $p$-forms $\Psi$, where $\nabla_\alpha$ is the metric connection. We shall assume that the classical system has two supersymmetries generated by the supercharges \begin{eqnarray} Q^{(1)}&=&\psi^\alpha\pi_\alpha-\psi^{*\alpha}W_\alpha(x)\\ Q^{(2)}&=&(J\psi)^\alpha\pi_\alpha-(J\psi^*)^\alpha W_\alpha(x) \end{eqnarray} The reason for choosing this particular form for these supercharges will become apparent when we come to discuss sigma models in the next section. We will only require the combination \begin{equation} Q=\case1/2(Q^{(1)}+iQ^{(2)})= \lambda^\alpha\pi_\alpha-\nu^\alpha W_\alpha(x)\label{sq} \end{equation} where the fermion fields \begin{eqnarray} \lambda&=&\case1/2(1+iJ)\psi\\ \nu&=&\case1/2(1+iJ)\psi^* \end{eqnarray} are purely holomorphic. The Hamiltonian is given (up to a constant) by $H=\{Q,Q^*\}$, which can be evaluated using the commutation relations (\ref{coms}), \begin{eqnarray} H&=&-\case1/2\nabla^2+g^{\alpha\beta}W_\alpha W^*_\beta +iW^*_{\alpha;\beta}\lambda^\beta\nu^{*\alpha} +iW_{\alpha;\beta}\lambda^{*\beta}\nu^\alpha\nonumber\\ &&-\case1/4R_{\alpha\beta\gamma\delta} [\lambda^\alpha,\lambda^{*\beta}][\lambda^{*\gamma},\lambda^\delta] -\case1/4R_{\alpha\beta\gamma\delta} [\lambda^\alpha,\lambda^{*\beta}][\nu^\gamma,\nu^{*\delta}]\label{ha} \end{eqnarray} The moduli space lies at the minimum of the bosonic potential and is parameterised by a set of complex coordinates $z^a$. We asume that the superpotential $W_\alpha$ has a series expansion in normal coordinates $z^I$, \begin{equation} W_{\alpha}=W_{\alpha\bar J}z^{\bar J} +W_{\alpha\bar J\bar K}z^{\bar J}z^{\bar K} +W_{\alpha\bar JK\bar L}z^{\bar J}z^Kz^{\bar L}+\dots\label{sup} \end{equation} The normal coordinates can be chosen to ensure that the boson mass matrix is diagonal, \begin{equation} g^{\alpha\beta}W_{\alpha\bar J}W^*_{\beta K}=(\omega^2)_{\bar J K}\label{bmass} \end{equation} For the fermion mass matrix $W_{\alpha;\beta}$, we need two seperate sets of basis vectors, in general, in order to put it into diagonal form. We can always choose one basis $({\bf e}_a,{\bf e}_I)$ to coincide with the coordinate basis at $z^I=0$ for the fields $\lambda$. Another basis $({\bf e}_{a'},{\bf e}_{I'})$ then has to be used for the fields $\nu$. (This proceedure is similar to the independent unitary rotations of the left and right chirality fermions in the standard model). With the choice \begin{equation} {\bf e}_{I'}=\omega_I^{-1}W^\alpha{}_I{\bf e}_\alpha, \end{equation} the fermion mass matrix becomes diagonal, \begin{equation} W_{I'\bar J}=\omega_{I\bar J}. \end{equation} The fermion fields $\lambda^a$ and $\nu^{a'}$ are massless and remain in the reduced quantum system. However, there is an important difference between the finite and infinite dimensional situations. In the infinite dimensional situation the basis ${\bf e}_{I'}$ may be complete (in which case there are no massless $\nu^{a'}$ fermions). We shall proceed assuming this to be the case. Previously, we used a series expansion for the Hamiltonian to arrive at the reduced theory. However, as shown in appendix C, for a supersymmetric theory we need only consider the supercharge. The supercharge (\ref{sq}) depends on the normal coordinates through the superpotential and through the connection coefficients in $\pi_\alpha$. In the basis ${\bf e}_\alpha=({\bf e}_a,{\bf e}_I)$, \begin{equation} \Gamma_{\alpha\beta a}\psi^{*\alpha}\psi^\beta\sim \Gamma_{b\bar ca}\lambda^{\bar c}\lambda^b +k_{\bar Iab}\lambda^{\bar I}\lambda^a +a_{I\bar Ja}\lambda^{\bar J}\lambda^I +a'{}_{\bar I'J'a}\nu^{J'}\nu^{\bar I'}\label{cterms} \end{equation} in the limit $z^I=0$. The first few terms in the series expansion of the supercharge (\ref{sq}) are now \begin{eqnarray} Q_0&=&-i\lambda^I\partial_I-\nu^{I'}W_{I'\bar J}z^{\bar J}\\ Q_1&=&-i\lambda^a{\cal D}_a-\nu^{I'}W_{I'\bar J\bar K}z^{\bar J}z^{\bar K}\\ Q_2&=&-i\lambda^\alpha R_{\beta\gamma\bar I\alpha}z^{\bar I} (\lambda^{*\gamma}\lambda^\beta-\nu^\gamma\nu^{*\beta})+ \nu^{I'}W_{I'\bar J\bar K\bar L}z^{\bar J}z^{\bar K}z^{\bar L} \end{eqnarray} where ${\cal D}_a$ now includes the connection terms (\ref{cterms}). The full quantum theory can now be reduced to the moduli space by following the methods described in appendix C. We define the fermion vacuum state $|0_T\rangle$ which is anihilated by two sets of anihilation operators \begin{equation} \lambda_-^I=\case1/{\sqrt{2}}(\lambda^I+i\nu^{I'}),\quad \lambda_-^{\bar I}=\case1/{\sqrt{2}} (\lambda^{\bar I}+i\nu^{\bar I'}) \end{equation} As in section 2, we build up from $|0_T\rangle$ a set of oscillator states $f_n$ which satisfy $H_0f_n=E_nf_n$, where $H_0$ is the unperturbed Hamiltonian. In the supersymmetric case, $E_0=0$. The reduced Hamiltonian is given in terms of the reduced supercharge by \begin{equation} H_R=\{Q_R,Q_R^*\} \end{equation} where \begin{equation} Q_R=\langle Q_1\rangle_{00}+ \sum_{n\ne0}\langle Q_1\rangle_{0n}(E_0-E_n)^{-1} \langle H_1\rangle_{n0}+O(\omega^{-1}) \end{equation} The covariant derivatives are treated as in section 2, that is we write \begin{equation} \langle{\cal D}_a\rangle_{0n}=\delta_{0n}{}^\parallel\nabla_a +\langle{\cal A}_a\rangle_{0n} \end{equation} In the supersymmetric case we include the fermion terms from equation (\ref{cterms}), \begin{equation} \langle{\cal A}_a\rangle_{0n}= {}^\parallel\nabla_a\omega_{\bar IJ}\langle z^{\bar I}z^J\rangle_{0n} -k_{\bar Iba}\langle \lambda^{\bar I}\lambda^b\rangle_{0n} -a_{I\bar Ja}\langle \lambda^{\bar J}\lambda^I\rangle_{0n} +a'{}_{I'\bar J'a}\langle \nu^{I'}\nu^{\bar J'}\rangle_{0n} \end{equation} At leading order we recover the same result that we would obtain by a trivial truncation of the theory, \begin{equation} Q_R=-i\lambda^a{}^\parallel\nabla_a+O(\omega^{-1}) \end{equation} As discussed in appendix A, the reduced states can be identified with antiholomorphic forms on the moduli space and the the reduced Hamiltonian can be identified with the Laplacian. At this order the reduced theory is identical with the truncated theory obtained under the simplest assumptions \cite{gauntlett}. \section{SUPERSYMMETRIC SIGMA MODELS} The supersymmetric sigma model in two space and one time dimension has Lagrangian \begin{equation} L={1\over 2}\int\left(-g_{ij}\partial^\mu\phi^i\partial_\mu\phi^j +ig_{ij}\bar\chi^i\gamma^\mu D_\mu\chi^j\right)d^2x \end{equation} where $\mu=0,1,2$, and $\chi$ is a two-component majorana spinor. If the target space is a K\"ahler manifold, with complex structure $J$, then the model has two supersymmetries. We shall consider the reduction of this theory. It is convenient to use a complex representation for the fermion fields, with \begin{equation} \chi=\pmatrix{\psi\cr\psi^*}, \end{equation} and a complex spatial coordinate $z=x+iy$. There are two complex supercharges in this representation, \begin{eqnarray} Q^{(1)}&=&\int\left(g_{ij}\psi^i\partial_t\phi^j -2g_{ij}\psi^{*i}\partial_z\phi^j\right)d^2z\\ Q^{(2)}&=&\int\left(g_{ij}(J\psi)^i\partial_t\phi^j -2g_{ij}(J\psi)^{*i}\partial_z\phi^j\right)d^2z \end{eqnarray} These can be used to generate the classical supersymmetry transformations by transforming to phase space and using Dirac brackets. The supercharges produce the Hamiltonian by Dirac brackets \begin{equation} H=\case{i}/2\{Q^{(1)},Q^{(1)*}\}_{\rm DB} =\case{i}/2\{Q^{(2)},Q^{(2)*}\}_{\rm DB} \end{equation} The combination $Q=(Q^{(1)}+iQ^{(2)})/2$ can also be used, but in this case \begin{equation} i\{Q,Q^*\}_{\rm DB}=H+E_T\label{qq} \end{equation} where \begin{equation} E_T=2\int g_{ij}(J\partial_z\phi)^i(\partial_{\bar z}\phi)^jd^2z. \end{equation} The integral is a constant for fields $\phi$ in the same topological class. We can still use (\ref{qq}) to generate the quantum Hamiltonian. The background field expansion used in section 3 can be used again here to expand the supercharge $Q$. The bosonic fluctuations about the background $\phi_0$ are described by a vector $\xi$ and a fluctuation operator $\Delta_f$. The fermion fields are simply parallel propagated to $\phi_0$ and decomposed into $\psi=\lambda+\nu^*$, where $\lambda$ and $\nu$ are holomorphic, i.e. $J\lambda=i\lambda$. Since the complex structure $J$ commutes with the fluctuation operator, we can choose holomorphic eigenmodes $u_\alpha$ and use these as a basis ${\bf e}_\alpha$. The supercharge for this theory is given by \begin{equation} Q=\int(\lambda^i\pi_i-2g_{ij}\nu^i\partial_z\phi^j)d^2z \end{equation} When the eingenmode expansions are used, we recover the expression (\ref{sq}) used in section 4, \begin{equation} Q=\lambda^\alpha\pi_\alpha-\nu^\alpha W_\alpha(\xi) \end{equation} However, now we have an explicit formula for the superpotential \begin{equation} W_\alpha=2(u_\alpha,\exp(\nabla_\xi){\bf e}_z) \end{equation} in terms of the product (\ref{product}). The series expansion for $W_{\alpha}$ in the coordinate $z^I$ is obtained using equation (\ref{serie}). The linear term is, \begin{equation} W_{\alpha \bar J}=2(u_\alpha,D_zu_J^*). \end{equation} where \begin{equation} D_zu_I=\partial_zu_I{}^i+(\partial_z\phi_0^j)\Gamma^i{}_{jk}u_I{}^k \end{equation} According to equation (\ref{bmass}), this implies that the fluctuation operator is \begin{equation} \Delta_fu_I=-4D_zD_{\bar z}u_I \end{equation} and also that the fermion mass matrix is diagonalised by choosing a basis \begin{equation} {\bf e}_{I'}=2\omega_I{}^{-1}D_{\bar z}u_I. \end{equation} These vectors actually form a complete basis, that is they form a normalisable set of eigenvectors of the (positive) operator $D_{\bar z}D_z$. The only massless fermions are the $\lambda^a$ and these remain as fermions on the moduli space. This is precisely the situation considered in the previous section. The reduced theory is therefore identical to the truncated theory at leading order. \section{CONCLUSION} The reduction of a classical dynamical system generally involves eliminating the internal forces and introducing generalised coordinates. In quantum theory, the uncertainty principal implies that the internal coordinates can never be frozen, but the reduction can be performed approximately when the internal degrees of freedom remain close to their ground state. We have considered the adiabatic reduction of various quantum systems onto a moduli space of collective coordinates. The results for a quantum mechanical system are quite general and depend on geometrical properties of the moduli space as well as the frequencies of the internal modes. The reduced Hamiltonian to leading order can be found at the end of section 1. The moduli space of sigma-model solitons has been quantised by taking the continuum limit of the quantum mechanical system. In the Bosonic case, the terms in the reduced Hamiltonian operator depend on the eigenfunctions of a fluctuation operator $\Delta_f$. The spectrum of the fluctuation operator is continuous and the eigenvalue sums of the quantum mechanical system have to be replaced by integrals and regularised. We could have attempted to quantise the sigma-model moduli space with the instanton techniques used in quantum field theory. In this approach, based on a path integral, the field is replaced by collective coordinates $x^a$ and field fluctuations $\xi$. The fluctuations are integrated out, leaving a path integral over the coordinates $x^a$ and a Jacobian factor. The classical action $S$ gets replaced by \begin{equation} W[x^a]=S[x^a]+\case1/2\log\det(\Delta_f-\partial_t^2)'+\dots \end{equation} where the prime indicates omission of any zero eigenvalues and the dots denote contributions from higher loop Feynman diagrams. An adiabatic reduction can now be obtained by taking the time derivatives to be small and evaluating some of the higher loop terms. In practice, we have found this approach less practical and not as rigourous as the Hamiltonian one presented earlier. The bosonic sigma model considered in section 3 can be generalised in various ways. An important possibility would be to include potentials of the form \begin{equation} \Phi=\int W(\phi)d^2x. \end{equation} The changes to the results brought about by introducing potentials are confined to the terms in equation (\ref{result}) involving $V_{IJK}$ and $V_{IJKL}$, which depend on functional derivatives of $\Phi$. The supersymetric sigma model is tightly constrained and reduces to the Hodge Laplacian acting on antiholomorphic $p$-forms, confirming a result that has been used previously with strong support but not derived rigourously \cite{witten,alfaro,gauntlett}. It is here, especially, that higher order terms in the adiabatic expansion might be of interest.
\section*{Acknowledgments} The author gratefully acknowledges the hospitality of University of Bielefeld where part of this work was done. He would also like to acknowledge useful discussions with Jean Cleymans and Francois Gelis. He is especially grateful to Haitham Zaraket for suggesting that the exact expression for the bremsstrahlung contribution be used. \bigskip
\chapter{Integrability Tests for Nonlinear Evolution Equations} \chaptitle{Integrability Tests} \section{Introduction} During the last three decades, the study of integrability of nonlinear ordinary and partial differential equations (ODEs and PDEs) has been the topic of major research projects (see, e.g., \cite{MAandPC91,AMetal91}). This chapter presents a few symbolic algorithms to illustrate how computer algebra systems (CASs) can be effectively used in integrability investigations. We work with {\it Mathematica} \cite{Wolfram96}, but our algorithms can be implemented in other languages. Among the many alternatives \cite{BGandAR97} for investigating the integrability of systems of PDEs with symbolic software, the search for conserved densities, generalized symmetries, and recursion operators is particularly appealing \cite{AMetal91}. Indeed, it turns out that these quantities can be computed without the use of sophisticated mathematical tools. As a matter of fact, not much beyond differentiation and solving of linear systems is needed. As a result, our algorithms are easy to implement. In fairness, our algorithms are restricted to the computation of polynomial quantities of polynomial equations. Yet, this covers the majority of the cases treated in the literature. The algorithms in this chapter are based on a common principle: scaling (or dilation) invariance. Indeed, we observed that many known integrable systems are invariant under dilation symmetry, which is a special Lie point symmetry. The dilation symmetry can be computed by solving a linear system. Using dilation invariance, the plan is to first produce candidates for the polynomial densities, symmetries, and recursion operators in an efficient way. Once the candidate expressions are computed, their unknown constant coefficients follow from solving a linear system. We focus our attention on explaining the strategy, at the cost of mathematical rigor and details, which can be found in \cite{UG98,UGandWH97a,UGandWH98b,WHetal98}. Rather than discussing the algorithms in general, we apply them to a few prototypical nonlinear evolution equations from nonlinear wave theory. Whenever appropriate, we address issues related to the implementation of the algorithms. For instance, we give explicit code for the Fr\'echet derivative, which is one of the key tools in our methods. Our package {\it InvariantsSymmetries.m} \cite{UGandWH98c} works for nonlinear evolution equations. Applied to a system with parameters, our package can determine the conditions on the parameters so that the system admits a sequence of conserved densities or generalized symmetries. Although we do not address it here, our package can also compute densities and symmetries of differential-difference equations (semi-discrete lattices). See \cite{UGandWH98a,UGandWH98b,UGetal97,WHetal98} for more information about that subject. Due to memory constraints, our software can only compute a limited number of conserved densities and symmetries (half a dozen for systems; at best a dozen for scalar equations). To prove integrability, one must show that infinitely many independent densities or symmetries exist. Alternatively, one could construct the operator that connects the symmetries, and prove that it is a true recursion operator. Such proofs involve mathematical methods \cite{AF87,AMetal91,Olver93,JPW98} that are beyond the scope of this article. Although it is not yet implemented, we also present an algorithm for the computation of recursion operators, based on the knowledge of a few conserved densities and symmetries. The computation of Lie point symmetries and generalized symmetries via prolongation techniques is purposely omitted. That topic and related software were covered extensively in \cite{WH96,WH97}. Space limitations also prevent the inclusion of the well-known Painlev\'e test, which is a widely applied and successful integrability detector for nonlinear ODEs and PDEs. We refer to \cite{WHetal98} for survey papers, books, and software related to Painlev\'e analysis. This chapter is organized as follows. In Section 2, we discuss scaling symmetries of PDEs and show how to compute them. Section 3 deals with conservation laws. We give the definition and the steps of our algorithm, and show how to implement and apply the algorithm. We do the same for generalized symmetries in Section 4. An algorithm to determine the recursion operator is given in Section 5. The leading examples in Sections 2 through 5 are the Korteweg-de Vries (KdV) and Sawada-Kotera (SK) equations, and a system of nonlinear Schr{\"o}dinger-% type equations. For the latter, we derive the recursion operator in Section 6. In Sections 7 and 8, we discuss our {\it Mathematica} package {\it InvariantsSymmetries.m} and review similar software. We draw some conclusions in Section 9. \section{Key Concept: Dilation Invariance} Our algorithms are based on the following observation: if a system of nonlinear evolution equations is invariant under a dilation (scaling) symmetry, then its conservation laws, generalized symmetries, and the recursion operator have the same scaling properties as the system. This is at least true for the polynomial case. As leading example, we use the ubiquitous Korteweg-de Vries (KdV) equation \cite{MAandPC91}, \begin{equation} \label{kdv} u_t = 6 u u_x + u_{3x}, \end{equation} which describes water and plasma waves and lattice dynamics. Throughout this chapter, we will use the notations \begin{equation} u_t = \frac{\partial u}{\partial t}, \quad u_{tx} = \frac{\partial^2 u}{\partial t \; \partial x}, \quad u_{nx} = \frac{\partial^n u}{\partial x^n}. \end{equation} Equation (\ref{kdv}) is invariant under the dilation (scaling) symmetry \begin{equation} \label{kdvdilation} (t, x, u) \rightarrow (t / {\lambda}^{3}, x / \lambda, {\lambda}^{2} u), \end{equation} where $\lambda$ is an arbitrary parameter. Indeed, replacement of $(t,x,u)$ according to (\ref{kdvdilation}) allows one to cancel a factor $\lambda^5$ in (\ref{kdv}). Note that, e.g., $\frac{\partial}{\partial t}$ is replaced by $\lambda^3 \frac{\partial}{\partial t}.$ Obviously, $u$ corresponds to two derivatives in $x$, i.e., $u \sim {\partial^2}/{\partial {x^2}}. $ Similarly, ${\partial}/{\partial t} \sim {\partial}^3/{\partial x}^3.$ We express all scalings in terms of $\frac{\partial}{\partial x}.$ Introducing {\it weights}, denoted by $w,$ we could say that $w(u) = 2$ and $w({\rm D}_t) = 3,$ if we set $w({\rm D}_x) = 1.$ We used ${\rm D}_t$ and ${\rm D}_x$ instead of $w(\frac{\partial}{\partial t})$ and $ w(\frac{\partial}{\partial x})$ to cover cases where densities and symmetries depend explicitly on $t$ and $x$ (see \cite{UGandWH98b,WHetal98} for examples). The {\it rank} $R$ of a monomial is equal to the sum of all of its weights. Observe that (\ref{kdv}) is {\it uniform in rank} since all the terms have rank $R=5,$ confirming that $\lambda^5$ was a common factor. \vskip 12pt \noindent {\bf Computation of scaling symmetries.}$\;\;$To compute the scaling symmetry of an equation, we compute the weights of all its terms, and {\it require} that the equation be uniform in rank. For (\ref{kdv}), with $w({\rm D}_x) = 1,$ this yields \begin{equation} w(u) + w({\rm D}_t) = 2 w(u) + 1 = w(u) + 3. \end{equation} The solution of this linear system is $w(u) = 2,$ and $w({\rm D}_t) = 3.$ \vskip 8pt As a second example, we consider a fifth-order PDE from soliton theory, \begin{equation} \label{sk} u_t = 5 u^2 u_x + 5 u_x u_{2x} + 5 u u_{3x} + u_{5x}, \end{equation} due to Sawada and Kotera \cite{KSandTK74}. Scaling invariance requires that \begin{equation} \label{balancesk} w(u) + w({\rm D}_t) = 3 w(u) + 1 = 2 w(u) + 3 = w(u) + 5. \end{equation} Hence $w(u) = 2$ and $w({\rm D}_t) = 5.$ \vskip 10pt \noindent {\bf Systems.}$\;\;$Single PDEs like (\ref{kdv}) are a special case of \begin{equation} \label{pdesys} {\bf u}_t = {\bf F}({\bf u}, {\bf u}_{x}, {\bf u}_{2x}, \ldots, {\bf u}_{mx}), \end{equation} where ${\bf u}$ and ${\bf F}$ are vector dynamical variables with $n$ components. The number of components, the order $m$ of the system, and its degree of nonlinearity are arbitrary. To determine the scaling symmetry, we require that each equation in (\ref{pdesys}) be uniform in rank, and solve the resulting linear system for the weights of all the variables. As an example, consider a vector nonlinear Schr{\"o}dinger equation, \begin{equation} \label{DMVvector} {\bf B}_t + (|{\bf B}|^2 {\bf B})_x + ({\bf B}_0 \cdot {\bf B}_x) {\bf B}_0 + {\bf e} \times {\bf B}_{xx} = 0, \end{equation} which occurs in plasma physics \cite{BDetal93a,BDetal93b}. With ${\bf B}_0 = (a,b)$ and ${\bf B} = (u,v)$ in the $(y,z)$-plane, and ${\bf e}$ along the $x$-axis, (\ref{DMVvector}) can be written as \begin{eqnarray} \label{DMV} && u_t + \left [ u ( u^2 + v^2 ) + \beta u + \gamma v - v_x \right ]_x = 0, \nonumber \\ && v_t + \left [ v ( u^2 + v^2 ) + \theta u + \delta v + u_x \right ]_x = 0 , \end{eqnarray} where $\beta = a^2, \gamma = \theta = a b,$ and $\delta = b^2$ are nonzero parameters. With reference to \cite{BDetal93a}, we call (\ref{DMVvector}) or (\ref{DMV}) the DMV equation. To start generally, we will consider the system (\ref{DMV}) for arbitrary nonzero parameters $\beta, \gamma, \theta $ and $\delta.$ System (\ref{DMV}) is not uniform in rank, unless we allow that the parameters $\beta$ through $\delta$ have weights. Doing so, with $w({\rm D}_x) = 1,$ we obtain \begin{eqnarray} \label{balanceDMV} w(u) + w({\rm D}_t) \!\!&\!=\!&\!\! 3 w(u) + 1 = w(u) + 2 w(v) + 1 = w(u) + w(\beta) + 1 \nonumber \\ \!\!&\!=\!&\!\! w(v) + w(\gamma) + 1 = w(v) + 2, \\ w(v) + w({\rm D}_t) \!&\!=\!&\! 2 w(u) + w(v) + 1 = 3 w(v) + 1 = w(u) + w(\theta) + 1 \nonumber \\ \!\!&\!=\!&\!\! w(v) + w(\delta) + 1 = w(u) + 2. \end{eqnarray} Hence, \begin{equation} \label{weightsDMV} w(u) = w(v) = \frac{1}{2}, \;\; w(\beta) = w(\gamma) = w(\theta) = w(\delta) = 1, \;\; w({\rm D}_t) = 2. \end{equation} \vskip 10pt \noindent {\bf Remark.}$\;\;$For scaling-invariant equations like (\ref{kdv}) and (\ref{sk}), it suffices to consider the dilation symmetry on the space of independent and dependent variables. For systems like (\ref{DMV}) that are inhomogeneous for scaling, we give weights to the parameters to circumvent the problem. For systems that lack scaling invariance and have no parameters, introducing one (or more) auxiliary parameter(s) with appropriate scaling provides a solution. The trick is to extend the action of the dilation symmetry to the space of independent and dependent variables, {\it including} the parameters. Doing so, our algorithms apply to a larger class of polynomial PDEs. The extra parameters are only used in the first step of the algorithms: that is, in producing the candidate densities and generalized symmetries. Beyond that first step, parameters are no longer treated as dependent variables! Details and examples are given in \cite{UGandWH97a,UGandWH98b,WHetal98}. \section{Conservation Laws} {\bf Definition.}$\;\;$A conservation law for (\ref{pdesys}), \begin{equation} \label{conslaw} {\rm D}_{t} \rho + {\rm D}_{x} J = 0, \end{equation} connects the {\em conserved density} $\rho$ and the {\em associated flux} $J.$ As usual, ${\rm D}_{t}$ and ${\rm D}_{x}$ are total derivatives, and (\ref{conslaw}) holds for all solutions of (\ref{pdesys}). Hence, density-flux pairs only depend on ${\bf u}, {\bf u}_x, $ etc., not on ${\bf u}_t.$ With a few exceptions, densities and fluxes do not explicitly depend on $t$ and $x.$ For the scalar case, $u_t = F,$ the computations are carried out as follows: \begin{equation} \label{Dtrho} {\rm D}_t \rho = \frac{\partial \rho}{\partial t} + \sum_{k=0}^{n} \frac{\partial \rho}{\partial u_{kx}} {\rm D}_x^k u_t, \end{equation} where $n$ is the order of $\rho.$ Upon replacement of $u_t, u_{xt},$ etc. from $u_t = F,$ one gets \begin{equation} \label{Dtrhonew} {\rm D}_t \rho = \frac{\partial \rho}{\partial t} + \rho'(u) [F], \end{equation} where $\rho'(u) [F]$ is the Fr\'echet derivative of $\rho$ in the direction of $F$ (see Section 1.4). Furthermore, \begin{equation} \label{DtJ} {\rm D}_x J = \frac{\partial J}{\partial x} + \sum_{k=0}^{m} \frac{\partial J}{\partial u_{kx}} u_{(k+1)x}, \end{equation} where $m$ is the order of $J.$ Integrating both terms in (\ref{conslaw}) with respect to $x$ yields \begin{equation} \label{} {\rm D}_t \int_{-\infty}^{+\infty} \rho \; dx = - J |_{-\infty}^{+\infty} = 0, \end{equation} provided that $J$ vanishes at infinity. In that case, \begin{equation} \label{conservedP} P = \int_{-\infty}^{+\infty} \rho \; dx = {\rm constant \; in \; time}. \end{equation} So, $P$ is the true conserved quantity. For ODEs, the quantities $P$ are called constants of motion. \vskip 8pt \noindent {\bf Examples.}$\;\;$The first three (of infinitely many) independent conservation laws \cite{MAandPC91,RMetal68} for (\ref{kdv}) are \begin{eqnarray} \label{kdvconslaw1} \!\!\!\!\!&&\!\!\!\! {\rm D}_t (u) - {\rm D}_x (3 u^2 + u_{2x} )=0, \\ \label{kdvconslaw2} \!\!\!\!\!&&\!\!\!\! {\rm D}_t (u^2) - {\rm D}_x ( 4 u^3 - u_x^2 + 2 u u_{2x} )=0,\\ \label{kdvconslaw3} \!\!\!\!\!&&\!\!\!\! {\rm D}_t \left(u^3 \!- \frac{1}{2} u_x^2\right) \!- {\rm D}_x \left({9\over 2} u^4 \!- 6 u u_x^2 \!+ 3 u^2 u_{2x} \!+ \frac{1}{2} u_{2x}^2 \!- u_x u_{3x}\right)\!=0. \end{eqnarray} The first two conservation laws correspond to conservation of momentum and energy. Note that the above conservation laws are indeed invariant under (\ref{kdvdilation}). The terms in the conservation laws have ranks $5, 7,$ and $9.$ The densities \begin{equation} \label{conskdv} \rho^{(1)} = u, \;\; \rho^{(2)} = u^2, \;\;\; {\rm and} \;\;\; \rho^{(3)} = u^3 - \frac{1}{2} u_x^2 \end{equation} have ranks $2, 4$ and $6$, respectively. The associated fluxes have ranks $4, 6$ and $8.$ Equation (\ref{kdv}) also has a density-flux pair that depends explicitly on $t$ and $x:$ \begin{eqnarray} \label{xtrhofluxkdv} {\tilde \rho} \!&\!=\!&\! t u^2 + \frac{1}{3} x u, \\ {\tilde J} \!&\!=\!&\! t (4 u^3 + 2 u u_{2x} - u_x^2) + x \left(u^2 + \frac{1}{3} u_{2x}\right) - \frac{1}{3} u_x. \end{eqnarray} ${\tilde \rho}$ has rank 1, ${\tilde J}$ has rank 3, since $w(t)=-3$ and $w(x) = -1.$ To accommodate this case, we used the total derivative notation ${\rm D}_t$ and ${\rm D}_x,$ in (\ref{conslaw}). For (\ref{sk}), the first two (of infinitely many) conserved densities \cite{UGandWH97a} are \begin{equation} \label{conssk} \rho^{(1)} = u \quad {\rm and} \quad \rho^{(2)} = {1 \over 3} u^3 - u_x^2. \end{equation} We will use them in the construction of the recursion operator for (\ref{sk}) in Section 5. We now describe how to compute {\it polynomial} densities that are (explicitly) independent of $t$ and $x.$ We refer to \cite{UG98,UGandWH97a} for the general algorithm covering systems as well as $(x,t)$-dependent densities. \vskip 8pt \vskip 5pt \noindent {\bf Algorithm for Polynomial Conserved Densities} \vskip 8pt \noindent {\bf Step 1: Determine the form of the density} \vskip 5pt \noindent Select the rank $R$ of $\rho ,$ say, $R=6.$ Make a list ${\cal L}$ of all monomials in the components of ${\bf u}$ and their $x$-derivatives that have rank $R.$ Remove from ${\cal L}$ all monomials where the power of the highest derivative is $1.$ This is done to remove terms in $\rho$ that are in ${\rm Image}\; ({\rm D}_x),$ and therefore belong to the flux $J.$ After all, densities are equivalent if they only differ by terms that are total derivatives with respect to $x.$ Make a linear combination with constant coefficients $c_i$ of the monomials that eventually remain in the list ${\cal L}.$ For (\ref{kdv}), ${\cal L} = \{ u^3, u_x^2, u u_{2x}, u_{4x} \}.$ Next, $u_{4x}$ and $u u_{2x}$ are removed. Obviously, $u_{4x} = {\rm D}_x u_{3x},$ and $u u_{2x} = \frac{1}{2} {\rm D}_x^2 u^2 - u_x^2.$ So, $u u_{2x}$ and $u_x^2$ only differ by a total $x$-derivative. From ${\cal L} = \{ u^3, u_x^2 \},$ one constructs $\rho = c_1 u^3 + c_2 u_x^2,$ which has rank $R=6.$ \vskip 8pt \noindent {\bf Step 2: Determine the unknown coefficients} \vskip 4pt \noindent Substitute $\rho$ into the conservation law (\ref{conslaw}), and compute ${\rm D}_t \rho$ via (\ref{Dtrho}). Use the PDE system to eliminate all $t$-derivatives of ${\bf u}$, and require the resulting expression $E$ to be a total $x$-derivative. To avoid integration by parts, apply the Euler operator (also called the variational derivative) \cite{Olver93} \begin{eqnarray} \label{euleroperator} L_u \!&\!=\!&\! \sum_{k=0}^{m} {(-{\rm D}_x)}^k \frac{\partial}{\partial u_{kx}} \nonumber \\ \!&\!=\!&\!\frac{\partial}{\partial{u}} - {\rm D}_x \left(\frac{\partial}{\partial{{u_x}}}\right)+ {\rm D}_{x}^2 \left( \frac{\partial }{\partial{{u_{2x}}}}\right)+\cdots+ (-1)^m {\rm D}_{x}^m \left(\frac{\partial }{\partial{u_{mx}}}\right) \end{eqnarray} to $E$ of order $m.$ If $L_u (E) = 0$ immediately, then $E$ is a total $x$-derivative. If $L_u (E) \ne 0,$ then the remaining expression must vanish identically. This yields a linear system for the constants $c_i.$ Solve the system. Carrying out these operations for (\ref{kdv}), one gets $c_1 = 1, c_2 = -\frac{1}{2}.$ \vskip 10pt \noindent {\bf Remark.}$\;\;$With (\ref{Dtrhonew}), the system for the $c_i$ follows from $L_u (\rho'(u) [F]) = 0$ by equating to zero the coefficients of monomials in $u$ and their $x$-derivatives. \vskip 10pt \noindent {\bf Implementation in Mathematica} \vskip 4pt \noindent In {\it Mathematica}, $\verb|D|$ is the total derivative operator and the variational derivative (Euler operator) can be found in the Standard Add-on Package {\it Calculus`VariationalMethods`}. For instance, returning to $\rho^{(3)}$ in (\ref{conskdv}), with (\ref{kdv}) and (\ref{Dtrho}), one computes \begin{eqnarray} \label{Dtrho3} E\!&\!=\!&\!{\rm D}_t \rho^{(3)} = {\rho^{(3)}}'(u) [u_t] = (3 u^2 - u_x {\rm D}) u_{t} \nonumber \\ \!&\!= \!&\!18 u^3 u_x - 6 u_x^3 - 6 u u_x u_{2x} + 3 u^2 u_{3x} - u_x u_{4x}. \end{eqnarray} Application of the variational derivative, \verb|VariationalD[E,u[x,t],{x,t}]| gives zero. That means that $E$ is a total $x$-derivative of a polynomial $J^{(3)}.$ Integration of $E = -{\rm D}_x J^{(3)}$ gives \begin{equation} J^{(3)} = -({9 \over 2} u^4 -6 u u_x^2 +3 u^2 u_{2x} +\frac{1}{2} u_{2x}^2 - u_x u_{3x}). \end{equation} \vskip 5pt \noindent {\bf Example.}$\;\;$With our package {\it InvariantsSymmetries.m}, we searched for conserved densities of (\ref{DMV}). Obviously, (\ref{DMV}) is a conservation law; thus, $\rho^{(1)} = u$ and $\rho^{(2)} = v,$ without conditions on the parameters. Additional conserved densities only exist if $\gamma = \theta.$ The first few are: \begin{eqnarray} \label{DMVdens3} \rho^{(3)}\!\!\!&\!=\!&\!\!u^2 + v^2 , \\ \label{DMVdens4} \rho^{(4)}\!\!\!&\!=\!&\!\!{1 \over 2} ( u^2 + v^2 )^2 + ( \beta - \delta ) u^2 + 2 \theta u v + 2 v u_x , \\ \label{DMVdens5} \rho^{(5)}\!\!\!&\!=\!&\!\!{1 \over 4} ( u^2 + v^2 )^3 + {1 \over 2} ( u_x^2 + v_x^2 ) + \theta u v ( u^2 + v^2 ) \nonumber \\ && + {1 \over 4} ( \beta - \delta ) ( u^4 - v^4 ) + 3 u^2 v u_x + v^3 u_x , \end{eqnarray} and \begin{eqnarray} \label{DMVdens6} \rho^{(6)}\!\!\!&\!=\!&\!\!{5 \over 32} ( u^2 + v^2 )^4 + {3 \over 4} ( u^2 + v^2 ) ( u_x^2 + v_x^2 ) + {1 \over 2} ( u u_x + v v_x )^2 \nonumber \\ && + {1 \over 8} ( \beta - \delta )^2 u^4 + {1 \over 4} ( \beta - \delta ) u^6 + {1 \over 2} ( \beta - \delta ) \theta u^3 v + {3 \over 4} ( \beta - \delta ) \theta^2 v^2 \nonumber \\ && + {3 \over 8} ( \beta - \delta ) u^4 v^2 - {1 \over 4} \theta^2 ( u^4 + v^4 ) + {3 \over 4} \theta ( u^5 v + u v^5) \nonumber \\ && - {1 \over 8} ( \beta - \delta ) v^6 - {3 \over 2} \theta^3 u v + {3 \over 2} \theta u^3 v^3 - {3 \over 2} \theta^2 v u_x \nonumber \\ && + {3 \over 2} ( \beta - \delta ) u^2 v u_x + {15 \over 4} u^4 v u_x + {3 \over 2} \theta u v^2 u_x + {5 \over 2} u^2 v^3 u_x \nonumber \\ && + {3 \over 4} v^5 u_x + {1 \over 4} ( \beta - \delta ) u_x^2 + {1 \over 2} \theta u_x v_x + {1 \over 2} v_x u_{2x}. \end{eqnarray} Via integration by parts, $\rho^{(4)}$ is equivalent to \begin{equation} \label{rho4alt} {\tilde \rho}^{(4)} = {1 \over 2} ( u^2 + v^2 )^2 + ( \beta - \delta ) u^2 + 2 \theta u v + v u_x - u v_x. \end{equation} Indeed, $\rho^{(4)} = {\tilde \rho}^{(4)} + D_x (u v).$ Judged from (\ref{DMVdens3})--(\ref{DMVdens6}), the complexity of the expressions dramatically increases as the rank increases. The fact that we were able to compute 6 independent densities for (\ref{DMV}) is an indicator that the system presumably is completely integrable, as was later proved in \cite{RWetal95}. Only the conserved densities $\rho^{(1)}$ through $\rho^{(3)}$ will be used in the construction of the recursion operator for (\ref{DMV}) in Section 6. \section{Generalized Symmetries} {\bf Definition.}$\;\;$A vector function ${\bf G} (x, t, {\bf u}, {\bf u}_{x}, {\bf u}_{2x}, \ldots)$ is called a {\it symmetry} of (\ref{pdesys}) if and only if it leaves (\ref{pdesys}) invariant under the replacement ${\bf u} \rightarrow {\bf u} + \epsilon {\bf G}$ within order $\epsilon.$ Hence, \begin{equation} \label{invariance} {\rm D}_t ({\bf u} + \epsilon {\bf G}) = {\bf F} ({\bf u} + \epsilon {\bf G}) \end{equation} must hold up to order $\epsilon$ on any solution of (\ref{pdesys}). Consequently, ${\bf G}$ must satisfy the linearized equation \cite{AF87,AMetal91} \begin{equation} \label{pdesymmetry} {\rm D}_t {\bf G} = {\bf F}'({\bf u})[{\bf G}], \end{equation} where ${\bf F}'$ is the Fr\'echet derivative of {\bf F}, i.e., \begin{equation} \label{pdefrechet} {\bf F}'({\bf u})[{\bf G}] = {\partial \over \partial{\epsilon}} {\bf F}({\bf u}+\epsilon {\bf G})_{|{\epsilon = 0}} . \end{equation} \vskip 3pt \noindent In (\ref{invariance}) and (\ref{pdefrechet}), we infer that ${\bf u}$ is replaced by ${\bf u} + \epsilon {\bf G},$ and ${\bf u}_{nx}$ by ${\bf u}_{nx} + \epsilon {\rm D}^n_x {\bf G}.$ As usual, ${\rm D}_{t}$ and ${\rm D}_{x}$ are total derivatives, and ${\bf G} =(G_1, G_2, \ldots, G_n)$ if the system (\ref{pdesys}) has $n$ components. Once higher-order symmetries have been found, these vector fields can be used to obtain fundamental information about the integrability of the equation. In many cases, conserved quantities, Hamiltonian structures, and recursion operators follow readily from the knowledge of generalized symmetries \cite{AF87}. \vskip 8pt \noindent {\bf Examples.}$\;\;$The first three (of infinitely many) symmetries \cite{Olver93} of (\ref{kdv}) are \begin{eqnarray} \label{symkdv} G^{(1)} \!&\!=\!&\! u_x, \;\;\;\;\; G^{(2)} = 6 u u_x + u_{3x}, \nonumber \\ G^{(3)} \!&\!=\!&\! 30 u^2 u_x + 20 u_x u_{2x} + 10 u u_{3x} + u_{5x}. \end{eqnarray} All the terms in these symmetries have rank $3, 5$ and $7,$ respectively. With higher-order symmetries one can generate new integrable PDEs. For example, $u_t = G^{(3)}$ is the Lax equation in the completely integrable KdV hierarchy \cite{MAandPC91}. Note that (\ref{kdv}) also admits symmetries that explicitly depend on $t$ and $x.$ Indeed, the symmetries \begin{equation} \label{xtsymkdv} {\tilde G}^{(1)} = 1 + 6 t u_x \;\;\;{\rm and}\;\;\; {\tilde G}^{(2)} = 4 u + 2 x u_x + 36 t u u_x + 6 t u_{3x} \end{equation} are of rank $0$ and $2.$ They linearly (and explicitly) depend on $t$ and $x.$ The algorithm presented in this paper can easily be extended to cover this type of symmetries (see \cite{UG98,UGandWH98b} for details). The situation for (\ref{sk}), which also has infinitely many polynomial symmetries, is more complicated. The symmetries of (\ref{sk}) originate from two distinct ``seeds": \begin{equation} \label{symsk} G^{(1)} = u_x \quad {\rm and} \quad G^{(2)} = 5 u^2 u_x + 5 u_x u_{2x} + 5 u u_{3x} + u_{5x}. \end{equation} We have also computed symmetries of higher rank, but we do not show them here. A detailed computer-aided study showed that the symmetries $G^{(2i-1)}$ with rank $6i-3$ come from the seed $G^{(1)},$ whereas $G^{(2i)}$ with rank $6i+1$ originate from $G^{(2)},$ where $i=1,2,\ldots$ (see \cite{JSandJW98b} for details). For systems of type (\ref{pdesys}), the symmetry ${\bf G}$ is a vector with $n$ components. Our computer search with {\it InvariantsSymmetries.m} revealed that (\ref{DMV}) is invariant under the transformation $(u,v) \rightarrow (v,-u)$, which is a Lie point symmetry, provided the conditions $\beta = \delta $ and $ \gamma = - \theta$ hold. However, these conditions do not lead to a hierarchy of integrable equations. We therefore continued our search with arbitrary nonzero parameters $\beta$ through $\delta.$ The first two symmetries of (\ref{DMV}) are ${\bf G}^{(1)} = (G_1^{(1)}, G_2^{(1)})$ and ${\bf G}^{(2)} = (G_1^{(2)}, G_2^{(2)})$, where \begin{eqnarray} \label{DMVsymm1} G_1^{(1)} \!&\!=\!&\! u_x , \nonumber \\ G_2^{(1)} \!&\!=\!&\! v_x , \\ \label{DMVsymm2} G_1^{(2)} \!&\!=\!&\! ( \beta - \delta ) u_x + 3 u^2 u_x + v^2 u_x + \gamma v_x + 2 u v v_x - v_{2x} , \nonumber \\ G_2^{(2)} \!&\!=\!&\! \theta u_x + 2 u v u_x + u^2 v_x + 3 v^2 v_x + u_{2x}. \end{eqnarray} Note that the sum of symmetries is still a symmetry. Remembering $(u_t, v_t) = (F_1, F_2)$ in (\ref{DMV}), we then have $G_1^{(2)} + \delta G_1^{(1)} = -F_1$ and $ G_2^{(2)} = -F_2. $ The next symmetry, ${\bf G}^{(3)} = (G_1^{(3)}, G_2^{(3)}),$ only exists if $\gamma = \theta.$ It is \begin{eqnarray} \label{DMVsymm3} G_1^{(3)} \!\!&\!\!\!=\!\!\!&\!\! 3 ( \beta - \delta ) u^2 u_x + {15 \over 2} u^4 u_x + 6 \theta u v u_x + 9 u^2 v^2 u_x + {3 \over 2} v^4 u_x + 3 \theta ( u^2 + v^2 ) v_x \nonumber \\ &&\!+\!6 ( u^3 v + u v^3 ) v_x - 3 ( u^2 \!+ v^2 ) v_{2x} - 6 u u_x v_x - 6 v v_x^2 \!- u_{3x} , \\ G_2^{(3)} \!\!&\!\!\!=\!\!\!&\!\! 3 \theta ( u^2 + v^2 ) u_x + 6 ( u^2 + v^2 ) u v u_x + 6 u u_x^2 + {3 \over 2} u^4 v_x + 6 \theta u v v_x \nonumber \\ &&\!+\!9 u^2 v^2 v_x - 3 ( \beta - \delta ) v^2 v_x + {15 \over 2} v^4 v_x + 6 v u_x v_x + 3 ( u^2 \!+ v^2 ) u_{2x} - v_{3x}. \end{eqnarray} The algorithm for symmetries is similar to the one for conserved densities. The only difference is that monomials that differ by a total $x$-derivative are no longer removed from the list ${\cal L}$. \vskip 15pt \noindent {\bf Algorithm for Polynomial Generalized Symmetries} \vskip 10pt \noindent {\bf Step 1: Determine the form of the symmetry} \vskip 6pt \noindent Select the rank $R$ of the symmetry. Make a list ${\cal L}$ of all monomials involving ${\bf u}$ and its $x$-derivatives of rank $R.$ To obtain the form of the symmetry, make a linear combination of these monomials with constant coefficients $c_i.$ For example, for (\ref{kdv}), $G = c_1 \, u^2 u_x + c_2 \, u_x u_{2x} + c_3 \, u u_{3x} + c_4 \, u_{5x}$ is the form of the generalized symmetry of rank $R=7.$ \vskip 10pt \noindent {\bf Step 2: Determine the unknown coefficients} \vskip 6pt \noindent Compute ${\rm D}_t {\bf G}.$ Use the PDE system to remove all $t$-derivatives. Equate the result to the Fr\'echet derivative ${\bf F}'({\bf u})[{\bf G}].$ Treat the different monomial terms in ${\bf u}$ and its $x$-derivatives as independent to get the linear system for the $c_i.$ Solve that system. For (\ref{kdv}), one obtains the symmetry of rank $7:$ \begin{equation} \label{laxsymmagain} G = 30 u^2 u_x + 20 u_x u_{2x} + 10 u u_{3x} + u_{5x}. \end{equation} Symmetries of lower or higher rank are computed similarly. See \cite{UG98,UGandWH98b,WHetal98} for details about the algorithm and its implementation. \vskip 12pt \noindent {\bf Remark.}$\;\;$ Starting with a conserved density $\rho$, the symmetries for a Hamiltonian system can be obtained from ${\rm D}_x (L_u (\rho )),$ where $L_u$ is defined in (\ref{euleroperator}). See, for example, \cite{VRandGK94} for a study of the connection between densities and symmetries for Lagrangian and non-Lagrangian systems. \vskip 15pt \noindent {\bf Implementation in Mathematica} \vskip 5pt \noindent The key tool to compute symmetries is the Fr\'echet derivative, which is implemented as follows: \vskip 4pt {\hfuzz=1pt \noindent \begin{verbatim} frechet[funcF_List,funcU_List,indVars_List,funcG_List]:= Module[{eps,resultlist={},i}, Do[resultlist=Append[resultlist, Expand[(D[Part[funcF,i] /. {(f_/; MemberQ[funcU,f])[Sequence @@ indVars]:> f[Sequence @@ indVars]+eps*funcG[[Flatten[Position[funcU,f]][[1]]]], Derivative[k_,0][f_][Sequence @@ indVars]:> D[f[Sequence @@ indVars]+eps*funcG[[Flatten[Position[funcU,f]][[1]]]], {indVars[[1]],k}]},eps]/. eps :> 0)]], {i,Length[funcF]}]; Return[resultlist]]; \end{verbatim} } \vskip 3pt \noindent To compute the Fr\'echet derivative of $u^3 - \frac{1}{2} u_x^2$ in the direction of $k(x,t),$ type \begin{verbatim} frechet[{u[x,t]^3-(1/2)*D[u[x,t],x]^2},{u},{x,t},{k[x,t]}]; \end{verbatim} This gives the answer $3 u^2 k - u_x k_x.$ \section{Recursion Operators for Scalar Equations} {\bf Definition.}$\;\;$A {\it recursion operator} is a linear operator ${\bf \Phi}$ on the space of differential functions with the property that whenever $\bf G$ is a symmetry of (\ref{pdesys}), so is $\bf \hat{G}$ with $ {\bf {\hat{G}}} = {\bf \Phi} {\bf G}.$ The equation for the recursion operator \cite{AF87,Olver93,JPW98} is \begin{equation} \label{recursion} {\rm D}_t {\bf \Phi} + [{\bf \Phi}, {\bf F}'(u) ] = \frac{\partial {\bf \Phi}}{\partial t} + {\bf \Phi}' [{\bf F}] + {\bf \Phi} \circ {\bf F}'(u) - {\bf F}'(u) \circ {\bf \Phi} = 0, \end{equation} where $[\; , \; ]$ means commutator, $\circ$ indicates for composition, and the variational derivative of the operator ${\bf \Phi}$ is defined in, e.g., \cite{AF87}. For $n$-component systems like (\ref{pdesys}), the symmetries ${\bf G}$ are vectors with $n$ components, and the recursion operator ${\bf \Phi}$ is an $n \times n$ matrix. \vskip 5pt \noindent {\bf Examples.}$\;\;$The recursion operator for (\ref{kdv}) is \begin{equation} \label{opkdv} \Phi_{\rm KdV} = {\rm D}^2 + 4 u + 2 u_x {\rm D}^{-1} = {\rm D}^2 + 2 u + 2 {\rm D} u {\rm D}^{-1} \;, \end{equation} where, for simplicity of notation, ${\rm D}={\rm D}_x$ and ${\rm D}^{-1}={\rm D}_x^{-1}.$ Here, $\Phi_{\rm KdV}\, G^{(1)} = G^{(2)}$ and $\Phi_{\rm KdV}\, G^{(2)} = G^{(3)}$ for the symmetries listed in (\ref{symkdv}). The recursion operator $\Phi_{\rm KdV}$ also connects the $(x,t)$-dependent symmetries (\ref{xtsymkdv}), i.e., $\Phi_{\rm KdV} {\tilde G}^{(1)} = {\tilde G}^{(2)},$ but $\Phi_{\rm KdV} {\tilde G}^{(2)}$ is no longer local. Since $w({\rm D}^{-1})=-w({\rm D})= -1,$ the three terms in (\ref{opkdv}) have rank $R=2.$ The recursion operator (\ref{opkdv}) is uniform in rank. Clearly, the rank of $\Phi_{\rm KdV}$ is the difference in rank between consecutive symmetries in (\ref{symkdv}). In view of the symmetries (\ref{symsk}), the recursion operator for (\ref{sk}) must have rank $6.$ Indeed, the recursion operator \cite{BFandWO1982,BFetal87} has rank $6:$ \begin{eqnarray} \label{opsk} \Phi_{\rm SK}\!\!\!&\!=\!&\!\!\!\! {\rm D}^6 + 2 u {\rm D}^4 + 2 {\rm D} u {\rm D}^3 + {\rm D}^2 u {\rm D}^2 + 3 u {\rm D} u {\rm D} + 3 u {\rm D}^2 u - 2 {\rm D} u {\rm D} u - 2 u^3 \nonumber \\ \!\!\!\!&&\!\!\!\!+ {\rm D}^5 u {\rm D}^{-1} + 5 {\rm D} u {\rm D}^2 u {\rm D}^{-1} + 5 u^2 {\rm D} u {\rm D}^{-1} + {\rm D} u {\rm D}^{-1} (u^2 - 2 u_x {\rm D}), \end{eqnarray} which can also be written as \begin{eqnarray} \label{finalopsk} \Phi_{\rm SK}\!\!\!&\!=\!&\!\!\!\! {\rm D}^6 + 3 u {\rm D}^4 - 3 {\rm D} u {\rm D}^3 + 11 {\rm D}^2 u {\rm D}^2 - 10 {\rm D}^3 u {\rm D} + 5 {\rm D}^4 u \nonumber \\ \!\!\!\!&&\!\!\!\!+ 12 u^2 {\rm D}^2 - 19 u {\rm D} u {\rm D} + 8 u {\rm D}^2 u + 8 {\rm D} u {\rm D} u + 4 u^3 \nonumber \\ \!\!\!\!&&\!\!\!\!+ u_x {\rm D}^{-1} (u^2 - 2 u_x {\rm D}) + G^{(2)} {\rm D}^{-1}, \end{eqnarray} with $G^{(2)}$ in (\ref{symsk}). Our algorithm for the computation of polynomial recursion operators is based on the following observations. \vskip 10pt \noindent {\bf Key observations.}$\;\;$All terms in (\ref{opkdv}) and (\ref{finalopsk}) are monomials in ${\rm D},{\rm D}^{-1}, u,$ and $u_x.$ Depending on the form of the recursion operator, $u_{2x}, u_{3x}, $ etc. can also appear, as is the case in (\ref{opsk}). Recursion operators split naturally in ${\Phi} = {\Phi}_0 + {\Phi}_1, $ where ${\Phi}_0$ is a differential operator (without ${\rm D}^{-1}$ terms), and ${\Phi}_1$ is an integral operator (with ${\rm D}^{-1}$ terms). Furthermore, application of $\Phi$ to any symmetry should not leave any integrals unresolved, since all symmetries are polynomial (see \cite{JSandJW97c}). This is where the connection between conserved densities and symmetries comes into play. For instance, for (\ref{kdv}) it is clear that ${\rm D}^{-1} (6 u u_x + u_{3x}) = 3 u^2 + u_{2x}$ is polynomial. Similarly, for (\ref{sk}), using (\ref{finalopsk}), we have \begin{equation} \label{skobs1} {\rm D}^{-1} (5 u^2 u_x + 5 u_x u_{2x} + 5 u u_{3x} + u_{5x} ) = {5 \over 3} u^3 + 5 u u_{2x} + u_{4x} \end{equation} and \begin{eqnarray} \label{skobs2} && {\rm D}^{-1} ( u^2 - 2 u_x {\rm D}) (5 u^2 u_x + 5 u_x u_{2x} + 5 u u_{3x} + u_{5x} ) \nonumber \\ && \; = u^5 - 10 u^2 u_x^2 + \cdots - 2 u_x u_{5x}. \end{eqnarray} The first two conserved densities of (\ref{sk}) are $\rho^{(1)} = u$ and $\rho^{(2)} = {1 \over 3} u^3 - u_x^2$. Thus, with (\ref{conslaw}), we get $ {\rm D}_t u = u_t = - {\rm D}_x J^{(1)} $ and \begin{equation} \label{piecessk} {\rm D}_t \left({1 \over 3} u^3 - u_x^2\right) = \rho'(u) [u_t] = (u^2 - 2 u_x {\rm D}) u_t = - {\rm D}_x J^{(2)}. \end{equation} So, the factor $(u^2 - 2 u_x {\rm D})$ in (\ref{finalopsk}) comes from $\rho^{(2)},$ and ${\rm D}^{-1} [(u^2 - 2 u_x {\rm D}) u_t]$ will be polynomial, namely $-J^{(2)}.$ A similar situation happens for (\ref{kdv}), where $\rho^{(1)} = u, \rho^{(2)} = u^2,$ and $\rho^{(3)} = u^3 - {1 \over 2} u_x^2.$ Then, with (\ref{conslaw}) and (\ref{Dtrhonew}), \begin{eqnarray} \label{pieceskdv} \!\!\!\!\!\!\!\!\! &&\!\!\! {\rm D}_t \rho^{(1)} = {\rm D}_t u = u_t = - {\rm D}_x J^{(1)}, \quad {\rm D}_t \rho^{(2)} = {\rm D}_t u^2 = 2 u u_t = - {\rm D}_x J^{(2)}, \;\;\; {\rm and} \nonumber \\ \!\!\!\!\!\!\!\!\! &&\!\!\! {\rm D}_t \rho^{(3)} = {\rm D}_t ( u^3 - {1 \over 2} u_x^2 ) = {\rho^{(3)}}' (u) [u_t] = (3 u^2 - u_x {\rm D}_x ) u_t = - {\rm D}_x J^{(3)} , \end{eqnarray} for polynomial $J^{(i)}, \; i=1,2,3.$ Thus, application of ${\rm D}^{-1},$ or ${\rm D}^{-1} u,$ or ${\rm D}^{-1} (3 u^2 - u_x {\rm D})$ to $ 6 u u_x + u_{3x}$ leads to a polynomial result. However, as will be shown below, the terms involving ${\rm D}^{-1} u$ and ${\rm D}^{-1} (3 u^2 - u_x {\rm D})$ are not needed in the construction of the recursion operator (\ref{opkdv}). Since our algorithm for recursion operators is the most elaborate, we give the steps that lead to (\ref{opkdv}) and (\ref{finalopsk}). We consider the scalar case first. Systems are dealt with in Section 6. \vskip 15pt \noindent {\bf Algorithm for Polynomial Recursion Operators} \vskip 8pt \noindent {\bf Step 1: Construct the form of the recursion operator} \vskip 7pt \noindent {\bf (i) Determine the rank of the operator} \vskip 3pt \noindent Compute the rank $R$ of the operator based on the known ranks of consecutive symmetries. For example, from (\ref{symkdv}), we compute \begin{equation} \label{ranksymkdv} R = {\rm rank} \; \Phi = {\rm rank}\; G^{(3)} - {\rm rank}\; G^{(2)} = {\rm rank}\; G^{(2)} - {\rm rank}\; G^{(1)} = 2. \end{equation} Obviously, ${\rm rank} \; {\Phi}_0 = {\rm rank} \; {\Phi}_1 = {\rm rank} \; {\Phi} = R.$ \vskip 8pt \noindent {\bf (ii) Determine the pieces of the operator ${\Phi}_0$} \vskip 3pt \noindent Make a list ${\cal L}$ of all permutations of ${\rm D}^j u^k,$ with $j$ and $k$ nonnegative integers, that have the rank $R.$ For (\ref{kdv}), \begin{equation} \label{permkdv} {\cal L} = \{ {\rm D}^2, u \}. \end{equation} \vskip 12pt \noindent {\bf (iii) Determine the pieces of the operator ${\Phi}_1$} \vskip 2pt \noindent It can be shown \cite{AB93,JPW98} that \begin{equation} \label{phi1part} {\Phi}_1 = \sum_j \sum_k G^{(j)} {\rm D}^{-1} {\rho^{(k)}}'(u), \end{equation} where the symmetries $G^{(j)}$ are combined with ${\rm D}^{-1}$ and ${\rho^{(k)}}'(u)$ in such a way that every term is exactly of ${\rm rank}\; {\Phi}_1 = R.$ That is, the indices $j$ and $k$ are taken so that ${\rm rank}\; (G^{(j)}) + {\rm rank}\;({\rho^{(k)}}'(u)) - 1 = R$ for every term in (\ref{phi1part}). Using the densities and symmetries, make a list ${\cal M}$ of the pieces involving ${\rm D}^{-1}.$ \vskip 0.001pt \noindent \begin{table \vspace{10pt} \caption{{\rm Building blocks of ${\Phi}_1$ for the KdV equation.}} \label{HGtable1} \vskip .001pt \noindent \begin{center} \begin{tabular}{| l | l | l | l |} \hline {\rule[-3mm]{0mm}{8mm} Rank} & Symmetry $G^{(j)}$& Density $\rho^{(k)}$ & ${\rho^{(k)}}'(u)$ \\ \hline & & & \\ 0 & --- & --- & $1$ \\ 1 & --- & --- & --- \\ 2 & --- & $u$ & $u$ \\ 3 & $u_x$ & --- & --- \\ 4 & --- & $u^2$ & $3 u^2 - u_x {\rm D}$\\ 5 & $6 u u_x + u_{3x}$ & --- & --- \\ 6 & --- & $u^3 - \frac{1}{2} u_x^2$ & \\ & & & \\ \hline \end{tabular} \end{center} \end{table} \vskip .001pt For (\ref{kdv}), from Table~\ref{HGtable1} it should be clear that ${\rm D}^{-1}$ can only be sandwiched between $u_x$ and $1.$ Any other combination would exceed rank $2$. Hence, \begin{equation} \label{mforkdv} {\cal M} = \{ u_x {\rm D}^{-1} \}. \end{equation} \vskip 5pt \noindent {\bf (iv) Build the operator ${\Phi}$} \vskip 2pt \noindent Next, produce ${\cal R} = {\cal L} \bigcup {\cal M},$ which has the building blocks of the recursion operator. To get $\Phi,$ linearly combine the pieces in ${\cal R}$ with constant coefficients $c_i.$ For (\ref{kdv}), we obtain \begin{equation} \label{rforkdv} {\cal R} = \{ {\rm D}^2, u, u_x {\rm D}^{-1} \}. \end{equation} Thus, \begin{equation} \label{formopkdv} \Phi_{\rm KdV} = c_1 \, {\rm D}^2 + c_2 \, u + c_3 \, u_x {\rm D}^{-1}. \end{equation} \vskip 6pt \noindent We now repeat steps (i)--(iv) for the SK equation (\ref{sk}). \vskip 2pt \noindent (i) Using the symmetries (\ref{symsk}), we get \begin{equation} \label{ranksymsk} {\rm rank}\; \Phi = {\rm rank}\; G^{(3)} - {\rm rank}\; G^{(1)} = {\rm rank}\; G^{(4)} - {\rm rank}\; G^{(2)} = 6. \end{equation} (ii) The operator ${\Phi}_0$ will be built from \begin{eqnarray} \label{permsk} {\cal L} \!&\!=\!&\! \{ {\rm D}^6 , u {\rm D}^4 , {\rm D} u {\rm D}^3 , {\rm D}^2 u {\rm D}^2 , {\rm D}^3 u {\rm D} , {\rm D}^4 u , u^2 {\rm D}^2 , u {\rm D} u {\rm D}, \nonumber \\ && u {\rm D}^2 u, {\rm D} u^2 {\rm D}, {\rm D} u {\rm D} u, {\rm D}^2 u^2, u^3 \}. \end{eqnarray} \begin{table}[h] \caption{{\rm Building blocks of ${\Phi}_1$ for the SK equation.}} \label{HGtable2} \vskip 0.001pt \noindent \begin{center} \begin{tabular}{| l | l | l | l |} \hline {\rule[-3mm]{0mm}{8mm} Rank} & Symmetry $G^{(j)}$ & Density $\rho^{(k)}$ & ${\rho^{(k)}}'(u)$ \\ \hline & & & \\ 0 & --- & --- & $1$ \\ 1 & --- & --- & --- \\ 2 & --- & $u$ & --- \\ 3 & $u_x$ & --- & --- \\ 4 & --- & --- & $u^2 - 2 u_x {\rm D}$\\ 5 & --- & --- & --- \\ 6 & --- & $\frac{1}{3} u^3 - u_x^2$ & --- \\ 7 & $5 u^2 u_x + 5 u_x u_{2x} + 5 u u_{3x} + u_{5x} $ & --- & --- \\ & & & \\ \hline \end{tabular} \end{center} \end{table} \vskip 0.001pt \noindent (iii) From Table~\ref{HGtable2}, which list the building blocks for ${\Phi}_1$ for (\ref{conssk}), we obtain \begin{equation} \label{integsk} {\cal M} = \{ u_x {\rm D}^{-1} (u^2 - 2 u_x {\rm D}), (5 u^2 u_x + 5 u_x u_{2x} + 5 u u_{3x} + u_{5x}) {\rm D}^{-1} \}. \end{equation} All other combinations of the form $ G^{(j)} {\rm D}^{-1} {\rho^{(k)}}'(u)$ exceed rank $6.$ \vskip 5pt \noindent (iv) Combining the monomials from ${\cal R} = {\cal L} \bigcup {\cal M},$ we get \begin{eqnarray} \label{formopsk} \Phi_{\rm SK}\!&\!=\!&\! c_1 {\rm D}^6 + c_2 u {\rm D}^4 + c_3 {\rm D} u {\rm D}^3 + c_4 {\rm D}^2 u {\rm D}^2 + c_5 {\rm D}^3 u {\rm D} \nonumber \\ \!&\!\!&\!+ c_6 {\rm D}^4 u + c_7 u^2 {\rm D}^2 + c_8 u {\rm D} u {\rm D} + c_9 u {\rm D}^2 u + c_{10} {\rm D} u^2 {\rm D} \nonumber \\ \!&\!\!&\!+ c_{11} {\rm D} u {\rm D} u + c_{12} {\rm D}^2 u^2 + c_{13} u^3 + c_{14} u_x {\rm D}^{-1} (u^2 - 2 u_x {\rm D}) \nonumber \\ \!&\!\!&\!+ c_{15} (5 u^2 u_x + 5 u_x u_{2x} + 5 u u_{3x} + u_{5x}) {\rm D}^{-1} . \end{eqnarray} \vskip 8pt \noindent {\bf Step 2: Determine the unknown coefficients} \vskip 4pt \noindent To determine the coefficients $c_i$, require that \begin{equation} \label{connection} \Phi G^{(k)} = G^{(k+s)}, \quad k = 1, 2, 3, \ldots , \end{equation} where $s$ is the number of seeds. In practice, it suffices to use $k=1$ and $2$ in (\ref{connection}) to fix all coefficients $c_i.$ Solve the resulting linear system(s) for the unknown $c_i.$ For (\ref{kdv}), $\Phi_{\rm KdV} G^{(2)} = G^{(3)}$ with $\Phi_{\rm KdV}$ in (\ref{formopkdv}) and its symmetries (\ref{symkdv}), we obtain \[ {\cal S}=\{ c_1 - 1 = 0, 18 c_1 + c_3 - 20 = 0, 6 c_1 + c_2 - 10 = 0, 2 c_2 + c_3 - 10 = 0 \}. \] The solution is $ c_1 = 1, c_2 = 4,$ and $c_3 = 2 $. Substituting it into (\ref{formopkdv}), we get \begin{equation} \label{finalopkdv} \Phi_{\rm KdV} = {\rm D}^2 + 4 \, u + 2 \, u_x {\rm D}^{-1}. \end{equation} The explicit computation on page 260 in \cite{AF87} shows that (\ref{finalopkdv}) satisfies (\ref{recursion}). For (\ref{sk}), we express that \begin{equation} \label{connsk} \Phi_{\rm SK} G^{(1)} = G^{(3)} \quad {\rm and} \quad \Phi_{\rm SK} G^{(2)} = G^{(4)}, \end{equation} with $\Phi_{\rm SK}$ in (\ref{formopsk}) and the symmetries (\ref{symsk}). Then we solve for the constants $c_i.$ This eventually yields \begin{eqnarray} \label{ouropsk} \Phi_{\rm SK}\!\!\!&\!=\!&\!\!\!\! {\rm D}^6 + 3 u {\rm D}^4 - 3 {\rm D} u {\rm D}^3 + 11 {\rm D}^2 u {\rm D}^2 - 10 {\rm D}^3 u {\rm D} + 5 {\rm D}^4 u \nonumber \\ \!\!\!\!&&\!\!\!\!+ 12 u^2 {\rm D}^2 - 19 u {\rm D} u {\rm D} + 8 u {\rm D}^2 u + 8 {\rm D} u {\rm D} u + 4 u^3 \nonumber \\ \!\!\!\!&&\!\!\!\!+ u_x {\rm D}^{-1} (u^2 - 2 u_x {\rm D}) + G^{(2)} {\rm D}^{-1}, \end{eqnarray} with $G^{(2)}$ in (\ref{symsk}). A lengthy computation shows that this recursion operator satisfies (\ref{recursion}). After integration by parts, (\ref{opsk}) or, equivalently, (\ref{ouropsk}) can also be written \cite{JSandJW98b} as \begin{eqnarray} \label{equivopsk} \!\!\!\!\Phi_{\rm SK} \!\!\!&\!=\!&\!\!\! {\rm D}^6 + 6 u {\rm D}^4 + 9 u_x {\rm D}^3 + 9 u^2 {\rm D}^2 + 11 u_{2x} {\rm D}^2 + 10 u_{3x} {\rm D} + 21 u u_x {\rm D} \nonumber \\ \!\!\!&\!\!&\!\!+ 4 u^3 \!+ 16 u u_{2x} \!+ 6 u_x^2 \!+ 5 u_{4x} \!+ u_x {\rm D}^{-1} ( u^2 \!+ 2 u_{2x}) \!+ G^{(2)} {\rm D}^{-1}. \end{eqnarray} \section{Recursion Operators for Systems} We show how to construct the recursion operators for systems (\ref{pdesys}) with $n$ components. The symmetry ${\bf G}$ has $n$ components and the recursion operator ${\bf \Phi}$ is a $n \times n$ matrix. We used {\it Mathematica} interactively to compute the recursion operator for (\ref{DMV}) with $\gamma = \theta.$ The recursion operator has the form \begin{equation} \label{opDMV} {\bf \Phi} = \left ( \begin{array}{cc} {\Phi}_{11} & {\Phi}_{12} \\ {\Phi}_{21} & {\Phi}_{22} \end{array} \right), \end{equation} >From ${\bf G}^{(2)} = {\bf \Phi} {\bf G}^{(1)}$, the rank of the entries ${\Phi}_{ij}$ is determined by \begin{eqnarray} \label{ranksystem} {\rm rank} \; G_1^{(2)} \!&\!=\!&\! {\rm rank} \; \Phi_{11} + {\rm rank}\; G_1^{(1)} = {\rm rank} \; \Phi_{12} + {\rm rank}\; G_2^{(1)}, \nonumber \\ {\rm rank} \; G_2^{(2)} \!&\!=\!&\! {\rm rank} \; \Phi_{21} + {\rm rank}\; G_1^{(1)} = {\rm rank} \; \Phi_{22} + {\rm rank}\; G_2^{(1)}. \end{eqnarray} For (\ref{DMV}), the difference in rank between consecutive symmetries is $1,$ so ${\rm rank} \; \Phi_{ij} = 1, \; i,j = 1,2. $ \vskip 5pt \noindent (i) As for the scalar case, we first construct the differential operator ${\bf \Phi}_0.$ In view of the weights (\ref{weightsDMV}), \begin{equation} \label{permDMV} {\cal L}_{ij} = \{ u^2, v^2, u v, {\rm D}, \beta, \delta, \theta \}. \end{equation} Hence, \begin{equation} \label{diffopDMV} {\bf \Phi}_0 \!=\! \left(\!\!\! \begin{array}{cc} c_1 u^2 \!+\! c_2 v^2 \!+\! c_3 u v \!+\! c_4 {\rm D} \!+\! c_5 & c_6 u^2 \!+\! c_7 v^2 \!+\! c_8 u v \!+\! c_9 {\rm D} \!+\! c_{10} \\ c_{11} u^2 \!+\! c_{12} v^2 \!+\! c_{13} u v \!+\!c_{14} {\rm D}\!+\!c_{15}& c_{16} u^2 \!+\! c_{17} v^2 \!+\! c_{18} u v \!+\! c_{19} {\rm D} \!+\!c_{20} \end{array} \!\!\!\right), \end{equation} where $c_5, c_{10}, c_{15}$, and $c_{20}$ will be linear in $\beta, \delta,$ and $\theta.$ \vskip 5pt \noindent (ii) Using the conserved densities $\rho^{(1)}=u, \rho^{(2)}=v,$ and $\rho^{(3)} = u^2 + v^2,$ we have \begin{eqnarray} \label{inner1DMV} {\rm D}_t \rho^{(1)}\!&\!=\!&\!{\rm D}_t u = \frac{\partial u}{\partial u} F_1 + \frac{\partial u}{\partial v} F_2 = (1, 0) \cdot (u_t, v_t) = - {\rm D}_x J^{(1)}, \\ {\rm D}_t \rho^{(2)}\!&\!=\!&\!{\rm D}_t v = \frac{\partial v}{\partial u} F_1 + \frac{\partial v}{\partial v} F_2 = (0, 1) \cdot (u_t, v_t) = - {\rm D}_x J^{(2)}, \\ {\rm D}_t \rho^{(3)}\!&\!=\!&\!{\rm D}_t (u^2 + v^2) = \frac{\partial (u^2 + v^2)}{\partial u} F_1 + \frac{\partial (u^2 + v^2) }{\partial v} F_2 \nonumber \\ \!&\!=\!&\!2 (u, v) \cdot (u_t, v_t) = - {\rm D}_x J^{(3)}, \end{eqnarray} where the dot $(\cdot)$ refers to the standard inner product of vectors. Therefore, introducing the symmetry ${(u_x, v_x)}^{\rm T}$ on the left of ${\rm D}^{-1}$ gives \begin{equation} \label{kforDMV} {\cal M} = \left \{ {(u_x, v_x)}^{\rm T} \odot {\rm D}^{-1} (u, v) \right \}, \end{equation} where $\odot$ stands for the tensor product of matrices and ${\rm T}$ for transpose. So, \begin{eqnarray} \label{phi1forDMV} {\bf \Phi}_1\!&\!=\!&\! c_{21}\; {(u_x, v_x)}^{\rm T} \odot {\rm D}^{-1} (u, v) \nonumber \\ \!&\!=\!&\! \left ( \begin{array}{cc} c_{21} u_x {\rm D}^{-1} u & c_{21} u_x {\rm D}^{-1} v \\ c_{21} v_x {\rm D}^{-1} u & c_{21} v_x {\rm D}^{-1} v \end{array} \right). \end{eqnarray} Note that ${(u_x, v_x)}^{\rm T} \odot {\rm D}^{-1} (1,0)$ and ${(u_x, v_x)}^{\rm T} \odot {\rm D}^{-1} (0,1)$ are of rank $\frac{1}{2}.$ They cannot be used in ${\bf \Phi}_1, $ where all pieces must have rank $1.$ To determine the unknown constants in ${\bf \Phi} = {\bf \Phi}_0 + {\bf \Phi}_1$, we use \begin{equation} \label{DMVrecopcond} {\bf \Phi} {\bf G}^{(k)} = {\bf G}^{(k+1)} + \sum_{l = 1}^{k} \alpha_{kl} {\bf G}^{(l)}, \quad k = 1, 2, \ldots, \end{equation} where $\alpha_{kl}$ are unknown coefficients (which can be zero). In contrast to the examples in the previous sections, the $\alpha_{kl}$ play a role when dealing with systems with weighted parameters like (\ref{DMV}). It suffices to take $k =1$ and $2$ in (\ref{DMVrecopcond}) to fix all coefficients $c_i,$ and the extra unknowns $\alpha_{11}, \alpha_{21}$ and $\alpha_{22}.$ By solving a linear system, we obtain $\alpha_{11}=0, \alpha_{21} = \theta^2$ and $\alpha_{22} = \beta - \delta,$ and the values for the coefficients $c_1$ through $c_{21}.$ The recursion operator then follows readily: \begin{eqnarray} \label{finopDMV} \!\!\!\!\!\!\!\!\! {\bf \Phi} \!&=\!&\! \left ( \begin{array}{cc} \beta-\delta + 2 u^2 + 2 u_x {\rm D}^{-1} u & \theta + 2 u v - {\rm D} + 2 u_x {\rm D}^{-1} v \\ \theta + 2 u v + {\rm D} + 2 v_x {\rm D}^{-1} u & 2 v^2 + 2 v_x {\rm D}^{-1} v \end{array} \right ). \end{eqnarray} The recursion operator for the case $\gamma=\theta=\delta=0$ was computed analytically in \cite{RWetal95}. \section[About the Integrability Package {\em InvariantsSymmetries.m}]{% About the Integrability Package InvariantsSymmetries.m} We briefly describe our package {\it InvariantsSymmetries.m}, which automatically performs the computation of conservation laws (invariants) and symmetries based on the algorithms in Sections 3 and 4. Users must have access to {\it Mathematica} 3.0 \cite{Wolfram96}. The files for our package are available from {\it MathSource} \cite{UGandWH98c}. The package includes instructions for installation, on-line help, documentation, and built-in examples. After proper installation, it is advisable to first run our notebook (called {\it Examples}), which is accessible through the browser as part of the Add-on Package {\it Integrability}. The interactive examples in the notebook will help familiarize the user with the syntax of our functions (see also \cite{UGandWH98c}). To use the package as part of a new notebook, start {\it Mathematica} and type \verb|In[1]:= <<Integrability`| to read in the package. You will get the following message: \begin{verbatim} Loading init.m for Integrability from AddOns. \end{verbatim} The key functions for conservation laws and symmetries of PDEs are {\verb|PDEInvariants|} and {\verb|PDESymmetries|}. These functions take the following arguments: the equations in the system, the dependent and independent variables, and the range for the rank. For example (\ref{DMV}), the first two lines below define the system (\ref{DMV}). The third line produces the densities (\ref{DMVdens3})--(\ref{DMVdens5}). The fourth line gives the symmetries (\ref{DMVsymm1})--(\ref{DMVsymm3}). \begin{verbatim} In[2]:= pde1 := D[u[x,t],t]+D[u[x,t]*(u[x,t]^2+v[x,t]^2)+ beta*u[x,t]+gamma*v[x,t]-D[v[x,t],x],x] == 0; In[3]:= pde2 := D[v[x,t],t]+D[v[x,t]*(u[x,t]^2+v[x,t]^2)+ theta*u[x,t]+delta*v[x,t]+D[u[x,t],x],x] == 0; In[4]:= PDEInvariants[{pde1,pde2}, {u,v}, {x,t}, {1,3}]; In[5]:= PDESymmetries[{pde1,pde2}, {u,v}, {x,t}, {3/2,5/2}]; \end{verbatim} Help about the functions and their options is available on-line. For instance, type \verb|In[6]:= ??PDEInvariants| to obtain the function description. Part of it reads: \vskip 0.1pt \noindent \begin{verbatim} PDEInvariants[eqn, u, {x,t}, R, opts] finds the invariant with rank R of a partial differential equation for the function u. PDEInvariants[{eqn1,eqn2,...}, {u1,u2,...}, {x,t}, {Rmin,Rmax}, opts] finds the invariants with rank Rmin through Rmax. x is understood as the space variable and t as the time variable. \end{verbatim} Typing \verb|In[7]:= ??PDESymmetries| produces descriptions like: \vskip 0.1pt \noindent \begin{verbatim} PDESymmetries[eqn, u, {x,t}, R, opts] finds the symmetry with rank R of a partial differential equation for the function u. \end{verbatim} Information about the options of PDESymmetries is obtained by typing \vskip 1pt \noindent {\verb|In[8]:= ??WeightedParameters|}. It returns: \begin{verbatim} WeightedParameters is an option that determines the parameters with weight. If WeightedParameters -> {p1,p2,...}, then p1, p2, ... are considered as constant parameters with weight. The default is WeightedParameters -> {}. \end{verbatim} The option {\verb|WeightedParameters|} is useful when working with systems that lack uniformity in rank. In such cases, our software tries to resolve the problem by itself and prints appropriate messages. When unsuccessful, the program will suggest the use of the {\verb|WeightedParameters|} option. Therefore, the option {\verb|WeightedParameters|} should not be used until it is explicitly recommended by the software. Rules for the weights of variables can be entered via the option {\verb|WeightRules|}: \begin{verbatim} WeightRules is an option that determines the rules for weights of the variables. If WeightRules -> {Weight[u] -> val,...}, then scaling properties are determined under these rules. There is a built in checking mechanism to see if the given rules cause inconsistency. \end{verbatim} For PDEs, the \verb|MaxExplicitDependency| option allows one to compute conserved densities or symmetries that explicitly depend on the independent variables: \begin{verbatim} MaxExplicitDependency is an option for finding the invariant and generalized symmetries of PDEs and DDEs. If MaxExplicitDependency -> Max_Integer, then the program allows for explicit dependency of independent variables of maximum degree Max. The default is MaxExplicitDependency -> 0. \end{verbatim} \section{Software Review} In this section, we briefly review software for the computation of conservation laws, higher-order symmetries and recursion operators. In Table~\ref{HGtable3}, we give a summary and contact information. Higher-order symmetries can be computed with prolongation methods, and numerous software packages are available that can aid in the tedious computations inherent to such methods. A 50 page survey of software for Lie symmetry computations, including generalized symmetries, can be found in \cite{WH96}, and a short update in \cite{WH97}. We will not repeat these software reviews here. A survey of packages for conservation laws was first given in \cite{UGandWH97a}. However, to keep this chapter self-contained, we present a summary of that survey. Based on dilation invariance, Ito's programs in REDUCE (see \cite{MI86,MI94,MIandFK85}) compute polynomial higher-order symmetries and conserved densities for systems of evolution equations that are uniform in rank (no weighted parameters can be introduced). Ito's latest program, called {\it SYMCD}, cannot be used to compute symmetries and densities that depend explicitly on the independent variables $t$ and $x$, nor can it handle systems with parameters. More details are given in \cite{UGandWH97a}. In \cite{BFetal97,BFetal87}, Fuchssteiner et al.\ present algorithms to compute higher-order {\it symmetries} of evolution equations. Their algorithm in \cite{BFetal97} is based on Lie-algebraic techniques and uses commutator algebra on the Lie algebra of vector fields. Their approach is different from the usual prolongation method in that no determining equations are solved. Instead, all necessary generators of the finitely generated Virasoro algebra are computed from one given element by direct Lie-algebraic methods. Their code is available in MuPAD. In \cite{BFetal87}, Fuchssteiner et al.\ give code to verify that recursion operators are hereditary. In \cite{BFetal97}, it is shown how to compute {\it mastersymmetries} from which the recursion operators can be retrieved. The REDUCE program {\it FS} for ``formal symmetries'' was written by Gerdt and Zharkov \cite{VGandAZ90} (see also \cite{VG93,VG96}). {\it FS} computes higher-order symmetries and conservation laws of polynomial type. The algorithm requires that the evolution equations be of order two or higher in the spatial variable. However, this approach does not require that the evolution equations be uniform in rank. With {\it FS}, one cannot compute symmetries that depend explicitly on the independent variables $t$ and $x.$ Applied to equations with parameters, {\it FS} computes the conditions on the parameters using the symmetry approach. The PC package {\it DELiA}, written in Turbo Pascal by Bocharov \cite{AB91} and co-workers, is a commercial computer algebra system for investigating differential equations using Lie's approach. The program deals with higher-order symmetries, conservation laws, integrability and equivalence problems. It has a special routine for systems of evolution equations. The program requires the presence of second- or higher-order spatial derivative terms in all equations. For systems with parameters, {\it DELiA} does not automatically compute the densities and symmetries corresponding to the (necessary) conditions on the parameters. One has to use {\it DELiA}'s integrability test first, to determine the conditions. Once the parameters are fixed, one can compute the densities and symmetries. Sanders and Wang have Maple and FORM code for the computation of symmetries in the scalar case, allowing zero and negative weights \cite{JSandJW98a,JSandJW98b,JSandJW98c} and nonpolynomial equations and symmetries. This code relies on the Maple package {\it diffalg} \cite{FBetal95} to do the reductions of solutions of ODEs (PDEs). See \cite{JSandJW97a} for theoretical foundations of the computation of conservation laws, and \cite{JSandJW97b} for the use of their algorithms in the integrability classification of KdV-type higher order PDEs. Wolf et al. \cite{TWetal98} have three packages, called {\it CONLAW 1/2/3}, in REDUCE for the computation of conservation laws. There is no limitation on the number of independent variables. The approach uses Wolf's program {\it CRACK} for solving overdetermined systems of PDEs (see \cite{WH96,WH97}). Wolf's algorithm is particularly efficient for showing the non-existence of conservation laws of high order. In contrast to our program, it also allows one to compute nonpolynomial conservation laws. Hickman \cite{MH98} at the University of Canterbury, Christchurch, New Zealand has implemented a slight variation of our algorithm for conserved densities in Maple. Instead of computing the differential monomials in the density by repeated differentiation, Hickman uses a tree structure combining the appropriately weighted building blocks. \section{Conclusions} The {\it Mathematica} package {\it InvariantsSymmetries.m} presented in this chapter can be used for computer-aided integrability detection of systems of nonlinear PDEs as they occur in various branches of science and engineering. More precisely, our package is a tool to search for the first half a dozen conservation laws and symmetries. If our programs succeed in finding a large set of independent conservation laws or symmetries, there is a good chance that the system has infinitely many of these quantities. For instance, if the number of conservation laws is 4 or less, most likely the system is not integrable---at least not in its current coordinate representation. Applied to a system with parameters, our package can determine the conditions on the parameters so that the system admits a sequence of conserved densities or generalized symmetries. An actual proof of integrability, by showing the existence of an infinity of conservations laws or symmetries, must be done analytically (see \cite{JSandJW98b} for results in this direction). On the other hand, constructing the recursion operator, and showing that it indeed satisfies the defining equation, provides conclusive proof of integrability. \section*{Acknowledgements} We acknowledge helpful discussions with Professors Jan Sanders and Frank Verheest, and thank them for careful reading of the manuscript. This research project was supported in part by the National Science Foundation of the United States of America under Grant CCR-9625421. \newpage \begingroup \font\prq=cmtt10 \prq \obeyspaces \hsize=7.5in \parindent=0pt \begin{tabular}{||l|l|l|l||} \hline \hline \multicolumn{4}{||c||} {\rule[-3mm]{0mm}{8mm} {\bf Table 3$\;\;\;$List of Software and Contact Information} } \\ \hline & & & \\ Name \& System & Scope & Developer(s) \& Address & Email Address \\ & & & \\ \hline & & & \\ CONLAW 1/2/3 & Conservation & T.$\!$ Wolf {\it et al.$\!$} & <EMAIL> \\ (REDUCE) & Laws & School of Math.$\!$ Sci.$\!$ & \\ & & Queen Mary \& & \\ & & Westfield College & \\ & & University of London & \\ & & London E1 4NS, U.K. & \\ & & & \\ \hline & & & \\ DELiA & Conservation & A.$\!$ Bocharov {\it et al.$\!$} & <EMAIL> \\ (Pascal) & Laws and & Saltire Software & \\ & Generalized & P.O. Box 1565 & \\ & Symmetries & Beaverton, OR 97075 & \\ & & U.S.A. & \\ & & & \\ \hline & & & \\ FS & Conservation & V.$\!$ Gerdt \& A.$\!$ Zharkov & <EMAIL> \\ (REDUCE) & Laws and & Laboratory of Computing & \\ & Generalized & Techniques \& Automation & \\ & Symmetries & Joint Institute for & \\ & & Nuclear Research & \\ & & 141980 Dubna, Russia & \\ & & & \\ \hline & & & \\ Invariants & Conservation & \"{U}.$\!$ G\"{o}kta\c{s} \& W.$\!$ Hereman & <EMAIL> \\ Symmetries.m & Laws and & Dept.$\!$ of Math.$\!$ Comp.$\!$ Sci.$\!$ & <EMAIL> \\ (Mathematica) & Generalized & Colorado School of Mines & \\ & Symmetries & Golden, CO 80401, U.S.A. & \\ & & & \\ \hline \hline \end{tabular} \vfil \newpage \begin{tabular}{||l|l|l|l||} \hline \hline \multicolumn{4}{||c||} {\rule[-3mm]{0mm}{8mm} {\bf Table 3 cont.$\;\;\;$List of Software and Contact Information}}\\ \hline & & & \\ Name \& System & Scope & Developer(s) \& Address & Email Address \\ & & & \\ \hline & & & \\ SYMCD & Conservation & M.$\!$ Ito & <EMAIL>. \\ (REDUCE) & Laws and & Dept.$\!$ of Appl.$\!$ Maths.$\!$ & hiroshima-u.ac.jp \\ & Generalized & Hiroshima University & \\ & Symmetries & Higashi-Hiroshima & \\ & & 724 Japan & \\ & & & \\ \hline & & & \\ symmetry \& & Generalized & B.$\!$ Fuchssteiner {\it et al.$\!$} & <EMAIL> \\ mastersymmetry & Symmetries & Dept.$\!$ of Mathematics & \\ (MuPAD) & & Univ.$\!$ of Paderborn & \\ & & D-33098 Paderborn & \\ & & Germany & \\ & & & \\ \hline & & & \\ Tests for & Conservation & J.$\!$ Sanders \& J.$\!$P.$\!$ Wang & <EMAIL> \\ Integrability & Laws, Genera-& Dept.$\!$ of Math.$\!$ & \\ (Maple \& FORM) & lized Symmetries, & \& Comp.$\!$ Sci.$\!$ & \\ & and Recursion & Vrije Universiteit & \\ & Operators & 1081 HV Amsterdam & \\ & & The Netherlands & \\ & & & \\ \hline & & & \\ Tools for & Conservation & M.$\!$ Hickman & M.$\!$Hickman \\ Conservation Laws & Laws & Dept.$\!$ of Maths.$\!$ \& Stats.$\!$ & @math.canterbury.ac.nz\\ (Maple) & & University of Canterbury & \\ & & Private Bag 4800 & \\ & & Christchurch & \\ & & New Zealand & \\ & & & \\ \hline \hline \end{tabular} \endgroup
\section{Introduction} Faint radio sources provide important information about global star formation history. Sensitive radio observations of the Hubble Deep Field (HDF) (Richards~\etal 1998, \cite{richards98a}) and other fields well-studied at optical wavelengths (Windhorst~\etal 1995, \cite{windhorst95a}, Fomalont~\etal 1991, \cite{fomalont91a}) have shown that sub-mJy radio sources are predominantly associated with star formation activity rather than active galactic nuclei (AGN). The radio luminosity of a galaxy is a reliable predictor of the star formation rate (SFR) for local galaxies (Condon~1992, \cite{condon92a} Cram~\etal 1998, \cite{cram98a}). Estimates of star formation based on radio observations also have the advantage of being independent of extinction by dust, which has caused much difficulty in the determination of star formation history from optical data. In section 2, we make use of the tight correlation between radio and FIR luminosity for star forming galaxies to compare the FIR and radio backgrounds and to study the sources producing both. In section 3, we determine the evolving radio luminosity function from the observed redshift distribution of faint radio sources, and then estimate the history of star formation to a redshift of about 3. Throughout, we assume $\Omega_m=1$, $\Omega_\Lambda=0$, and $H_0$=50\kmsmpc. We define the radio spectral index as $S_\nu\propto\nu^{-\alpha}$. \section{FIR vs. Radio Backgrounds} This section follows our recent paper (Haarsma~\& Partridge 1998, \cite{haarsma98a}). The far infrared (FIR) background was recently detected with DIRBE (Hauser \etal1998, \cite{hauser98a}; Dwek \etal1998, \cite{dwek98a}), and is most likely the collective emission of star forming galaxies. We use the radio-FIR correlation for individual galaxies (Helou, Soifer, \&~Rowan-Robinson 1985, \cite{helou85a}) to calculate the radio background associated with the FIR background, assuming that the bulk of emission is from $z\sim 1$. We find the radio background associated with the FIR background has a brightness temperature of $T_{40\unit{cm}}=0.31$~K, or $T_{170\unit{cm}}\sim15$~K (scaled using a spectral index of $\alpha=0.7$). At 170\unit{cm}\ (178~MHz), the observed radio background is $T_{170\unit{cm}}=30\pm7$~K (Bridle 1967, \cite{bridle67a}). This allows us to draw several conclusions about the faint sources making up the FIR background: \begin{enumerate} \item The radio emission from these sources makes up about half of the observed extragalactic radio background. (The other half is the summed radio emission of AGN.) \item Since (i) is in agreement with other radio observations (Condon 1989, \cite{condon89a}), the FIR-radio correlation appears to hold even for the very faint sources making up the FIR background. This confirms the assumption that the FIR background between about 140 and 240\unit{\mu m}\ is dominated by star-formation, not AGN activity. \item By quantitatively comparing the radio and FIR backgrounds, we find a relationship for the sources contributing to the background, \begin{equation} A \left(\frac{ 1+z}{8.5}\right)^\alpha = 0.20\pm0.05, \end{equation} where $\alpha$ is their radio spectral index, $A$ is the fraction of the radio background they produce (from (i), $A\sim0.5$), and $z$ is their mean redshift. This function is plotted in Figure~\ref{fig.Azalpha}. Note that the redshift $z$ is the mean redshift of the sources dominating the FIR and radio backgrounds, which is not necessarily the redshift of peak star formation activity (see \S3). \item By extrapolating the 3.6\unit{cm}\ $\log N-\log S$ curve to fainter flux densities, we estimate that most of the FIR background is produced by sources whose 3.6\unit{cm}\ flux density is greater than about 1\unit{\mu Jy}. This lower limit is consistent with other work (Windhorst~\etal 1993, \cite{windhorst93a}), but has more interesting observational consequences. An RMS sensitivity of 1.5\unit{\mu Jy}\ has already been reached in VLA observations (Partridge~\etal 1997, \cite{partridge97a}). The $\log N-\log S$ curve indicates that the number density of $S\geq1$\unit{\mu Jy}\ sources is about $25/\unit{arcmin^2}$, similar to some model predictions (Guiderdoni \etal 1998, \cite{guiderdoni98a}). At this density, these sources will cause SIRTF to encounter confusion problems at 160\unit{\mu m}. \end{enumerate} \section{Radio Star Formation History} In this section, we use the redshift distribution of faint radio sources to determine the evolution of the radio luminosity function, and the evolution of the star formation rate density. \subsection{Data} Three fields have been observed to microJy sensitivity at centimeter wavelengths and also have extensive photometric and spectroscopic data: the Hubble Deep Field (HDF), the Medium Deep Survey (MDS), and the V15 field. Table~\ref{tab.data} gives the details of the three fields and references. For the first time we have a sample of microJy radio sources with nearly complete optical identifications and about 50\% complete redshift measurements. We assume that all sources detected at these flux levels are star-forming galaxies, since optical identifications indicate that $\sim$80\% of these radio sources have spiral or irregular counterparts (Richards \etal 1998, \cite{richards98a}). The known quasars (two in the MDS sample, none in the HDF or V15 samples) were removed. In the flanking fields of the HDF, we have used the relationship between redshift and K-band magnitude (Lilly, Longair, \&~Allington-Smith 1985, \cite{lilly85a}) to estimate redshifts for 10 sources without spectroscopic values. For the remaining sources without redshifts, we arbitrarily selected redshifts to fill in gaps in the redshift distribution, in order to illustrate the total number of sources that will ultimately appear on the plot. Photometric redshifts for these sources are currently being calculated (Waddington~\& Windhorst, in preparation), and will be included in future work. To compare these data to the model, we calculate $n(z)$, the average number of sources per arcmin$^2$ in each redshift bin. This requires a correction for the varying sensitivity across the primary beam of the radio observations (Katgert, Oort, \& Windhorst 1988, \cite{katgert88a}; Martin, Partridge, \& Rood 1980, \cite{martin80a}). For example, a faint source which could only be detected at the center of the field contributes more to $n(z)$ than a strong source which could be detected over the entire primary beam area. The resulting redshift distributions are plotted in Figures~\ref{fig.modelC.nz} and \ref{fig.modelS.nz}. It is interesting how different the MDS and HDF distributions are, even though the surveys were both performed at 8~GHz with similar flux limits. The average source density (including all sources) is 1.26 sources/arcmin$^2$ in the HDF, but 2.63 sources/arcmin$^2$ in the MDS field (the V15 field at 5~GHz has 0.736 sources/arcmin$^2$). The density of sources in the MDS field is over twice that of the HDF field, possibly due to galaxy clustering. In the analysis below, we fit the model to the data in all three fields simultaneously. \begin{table}[t] \begin{center} \begin{tabular}[]{lllllll} \hline \hline Field & Location & Band & Flux limit & N & N$_z$ & Reference \\ \hline Hubble Deep Field & 12h+62d & 8 GHz & 9\unit{\mu Jy} & 29 & 13 (+10)& \cite{richards98a} \\ Medium Deep Survey & 13h+42d & 8 GHz & 8.8\unit{\mu Jy} & 19 & 10 & \cite{windhorst95a} \\ V15 field & 14h+52d & 5 GHz & 16\unit{\mu Jy} & 35 & 18 & \cite{fomalont91a,hammer95a} \\ \hline \end{tabular} \vspace{3mm} \caption{Summary of three deep radio surveys. The flux limit of the complete catalog for each field is approximately 5 times the RMS noise of the observation, but varies across the field. N is the total number of sources above the flux limit, and $N_z$ is the number of those sources with spectroscopic redshifts. An additional 10 redshifts were estimated in the HDF from their K-band magnitudes. } \label{tab.data} \end{center} \end{table} \subsection{Calculations} In order to determine the star formation history, we must first determine the evolving radio luminosity function. We used two versions of the local 1.4~GHz luminosity function for star-forming/spiral galaxies. We define the luminosity function $\phi(L_{e,1.4})$ as the number per comoving Mpc$^3$ per $d\log_{10}L$ of star-forming radio sources with emitted luminosity $L_{e,1.4}$(W/Hz)at 1.4~GHz. Condon (1989, \cite{condon89a}) uses the following form for the luminosity function (but different notation), \begin{eqnarray} \log_{10}[\phi(L_{e,1.4})] & d\log_{10}L = 28.43 + Y - 1.5\log_{10}L_{e,1.4} \nonumber \\ & - \left[ B^2 + \frac{1}{W^2} (\log_{10}L_{e,1.4} - X)^2 \right]^{1/2} d\log_{10}L, \label{eq.lumfunc.C89} \end{eqnarray} with the fitted parameters for star-forming galaxies of $Y=2.88$, $X=22.40$, $W=2/3$, and $B=1.5$. Serjeant \etal (1998, \cite{serjeant98a}) use the standard Schechter form, \begin{equation} \phi(L_{e,1.4})d\log_{10}L = \phi_{\ast} \ln10 \left(\frac{L_{e,1.4}}{L_{\ast}}\right)^{(1+\alpha_l)} \exp\left(-\frac{L_{e,1.4}}{L_{\ast}}\right) d\log_{10}L \label{eq.lumfunc.S98} \end{equation} where a factor of $L\ln10$ has been included to convert the function from d$L$ to $d\log_{10}L$. Serjeant \etal find fitted parameters of $\phi_{\ast}=4.9\times10^{-4}\unit{Mpc^{-3}}$, $L_{\ast}=2.8\times10^{22}\unit{W/Hz}$, and $\alpha_l = -1.29$. To describe the evolution of the luminosity function, we use the functional form suggested by Condon (1984a, \cite{condon84a}, eq.~24), a power-law in $(1+z)$ with an exponential cut-off at high redshift. The luminosity evolves as \begin{equation} f(z) = (1+z)^Q \exp\left[ -\left(\frac{z}{z_q}\right)^q \right], \label{eq.lumevol} \end{equation} and the number density evolves as \begin{equation} g(z) = (1+z)^P \exp\left[ -\left(\frac{z}{z_p}\right)^p \right]. \label{eq.numevol} \end{equation} This gives six parameters $\{Q,q,z_q,P,p,z_p\}$ to use in describing the evolution. When fitting for the parameters, we constrained the functions $g(z)$ and $f(z)$ to the physically reasonable ranges of $1<g(z)<100$, $1<f(z)<100$ for $0<z<3$. The general expression for the evolving luminosity function is then (Condon 1984b, \cite{condon84b}) \begin{equation} \phi(L_{e,1.4},z) = g(z) \phi\left( \frac{L_{e,1.4}}{f(z)}, 0 \right). \label{eq.evolvedlumfunc} \end{equation} To use this expression at an arbitrary observing frequency $\nu$ and redshift $z$, we must convert the observed luminosity $L_{o,\nu}$ to 1.4~GHz and do the K-correction, i.e. \begin{equation} L_{e,1.4} = L_{o,\nu} \left( \frac{\nu }{ 1.4\unit{GHz}} \right)^\alpha (1+z)^\alpha \end{equation} where $\alpha$ is the radio spectral index, as defined in \S1. We have assumed $\alpha=0.4$ for all calculations in \S3 (Windhorst \etal 1993, \cite{windhorst93a}). The evolving luminosity function can be used to predict the observed redshift distribution. The number of sources per redshift bin $\Delta z$ that could be detected in a survey of angular area $\Delta\Omega$ and flux limit $S_{lim}$ at frequency $\nu$ is \begin{equation} n(z) = V_c(z,\Delta z,\Delta\Omega) \int_{L'(z)}^{\inf} \phi(L_{e,1.4},z) d\log_{10}L \end{equation} where the lower limit of the integral is \begin{equation} L'(z) = 9.5\times10^{12}\unit{\frac{W}{Hz}} \left( \frac{S_{lim} }{\unit{\mu Jy} } \right) (1+z)^\alpha \left( \frac{\nu }{1.4\unit{GHz} } \right)^\alpha 4 \pi \left( \frac{D_L(z) }{\unit{Mpc} } \right)^2 \end{equation} and $D_L$ is the luminosity distance. The comoving volume in a shell from $z$ to $z + \Delta z$ and angular size $\Delta\Omega$ is \begin{eqnarray} V_c(z, \Delta z, \Delta\Omega) & = & \int d\Omega \int r^2(z) dr \nonumber \\ & = & \frac{\Delta\Omega }{ \unit{ster} } \left(\frac{\unit{ster} }{ 1.18\times10^7\unit{arcmin^2}}\right) \frac{[r^3(z+\Delta z) - r^3(z)] }{ 3 } \label{eq.covol} \end{eqnarray} where the comoving distance is \begin{equation} r(z) = \frac{2c}{H_0} \left( 1 - \frac{1}{\sqrt{1+z}} \right) \end{equation} for our assumed cosmology (see \S1). We have used $\Delta \Omega = 1$~arcmin$^2$ for comparison to the data in Figures~\ref{fig.modelC.nz} and \ref{fig.modelS.nz}. The evolving luminosity function also allows us to calculate the star formation history. For an individual galaxy, the star formation rate is directly proportional to its radio luminosity (Condon 1992, \cite{condon92a}): \begin{equation} \unit{SFR} = Q \left( \frac{L_\nu /\unit{\frac{W}{Hz}} }{5.3\times10^{21} \left(\frac{\nu}{\unit{GHz}}\right)^{-0.8} + 5.5\times10^{20} \left(\frac{\nu}{\unit{GHz}}\right)^{-0.1} } \right)\unit{\frac{M_{\odot}}{yr}} \label{eq.sfrtoLpergal} \end{equation} The radio luminosity is primarily due to synchrotron emission from supernova remnants (the first term in the denominator) plus a small thermal component (the second term). Both components are proportional to the formation rate of high-mass stars which produce supernova ($M>5M_{\odot}$), so the factor $Q$ is included to account for the mass of all stars ($0.1-100M_{\odot}$), \begin{equation} Q = \frac{ \int_{0.1M_{\odot}}^{100M_{\odot}} M \psi(M) dM }{ \int_{ 5M_{\odot}}^{100M_{\odot}} M \psi(M) dM }, \end{equation} where $\psi(M)\propto M^{-x}$ is the initial mass function (IMF). We have assumed throughout a Salpeter IMF ($x=2.35$), for which $Q=5.5$. If an upper limit of 125$M_{\odot}$ is used, then $Q=5.9$. In order to use eq.~\ref{eq.sfrtoLpergal} at high redshift, both $L_\nu$ and $\nu$ in the equation must be K-corrected to the emitted luminosity at the emission frequency. Are there other ways in which this relation evolves? The thermal term is much smaller than the synchrotron term, so evolution in the thermal term will have little effect. In the synchrotron term, the dependence on the supernova environment is weak. One component that might cause significant evolution in eq.~\ref{eq.sfrtoLpergal} is an evolving IMF, entering through the factor $Q$. In active starbursts, the IMF may be weighted to high-mass stars (Elmegreen 1998, \cite{elmegreen98a}), which would result in a smaller value of $Q$. However, the smallest $Q$ is unity (when virtually all mass occurs in high-mass stars), so the strongest decrease due to IMF evolution would be roughly a factor of five. To determine the star formation rate per comoving volume, we simply substitute the radio luminosity density for $L_\nu$ in eq.~\ref{eq.sfrtoLpergal}. The star formation rate depends on the emitted (rather than observed) luminosity density. The luminosity density emitted at 1.4~GHz can be easily found from the evolving luminosity function, \begin{equation} \rho_{e,1.4}(z) = \int_{-\inf}^{\inf} L_{e,1.4} \phi(L_{e,1.4},z) d\log_{10}L. \end{equation} Thus the predicted star formation history is \begin{equation} \psi(z) = Q \left(\frac{\rho_{e,1.4}(z)}{ 4.6\times10^{21} \unit{\frac{W}{Hz Mpc^3}}}\right) \unit{\frac{\msun}{yr Mpc^3}} \label{eq.sfrthist} \end{equation} where 1.4\unit{GHz} is used in the denominator of eq.~\ref{eq.sfrtoLpergal} (no K-correction is needed because the luminosity density is the emitted value). \subsection{Results} We use the formulation of \S3.2 to determine the star formation history from the evolving luminosity function. To determine the evolution parameters, we compare the model to the observed $n(z)$ for the three surveys. We immediately found that pure luminosity evolution [$f(z)=(1+z)^3$ and $g(z)=1$], as often suggested in the literature, is a poor fit for the faint star-forming galaxy population (the predicted $n(z)$ is too small and has a very long high-redshift tail). The model fit of Condon (1984a, \cite{condon84a}), $\{Q=3.5, P=1.75, p=1.8, z_p=1 \}$ with no exponential cut off in luminosity evolution, is much better (more reasonable redshift dependence, but $n(z)$ is still too low). To improve on these models, we adjust the evolution parameters $\{Q, q, z_q, P, p, z_p\}$ to improve the model fit to the $n(z)$ data, using a downhill simplex algorithm (Press \etal 1992, \cite{recipes2}) to find the global $\chi^2$ minimum. We performed this fit using both the Condon (1989, \cite{condon89a}) luminosity function (Model~C, see eq.~\ref{eq.lumfunc.C89}) and the Serjeant \etal (1998, \cite{serjeant98a}) luminosity function (Model~S, see eq.~\ref{eq.lumfunc.S98}). In Model~C we use the luminosity function of Condon (1989, \cite{condon89a}). The fitted evolution parameters were $\{ Q=7.6, q=1.3, z_q=0.48, P=1.6, p=1.2, z_p=1.8 \}$. The resulting evolution factors $f(z)$ and $g(z)$ are plotted in Figure~\ref{fig.modelC.evolfactor} and the resulting luminosity function is shown in Figure~\ref{fig.modelC.lumfuncevolve}. Although the term $(1+z)^{7.6}$ seems extreme, when combined with the exponential cut-off the luminosity evolution $f(z)$ is reasonable. The fit to the the redshift distribution is shown in Figure~\ref{fig.modelC.nz}. The fit significantly underestimates the total number of sources in the MDS field, but only slightly underestimates the other two fields. The V15 survey has the largest total number of sources and thus has the most weight during fitting, so the result is a better fit for V15 than the other fields. Finally, Figure~\ref{fig.modelC.sfr} shows our predicted star formation history (heavy line) along with model predictions from several others (thin lines). The vertical lines indicate the $1/\sqrt{N}$ uncertainty, where $N$ is the sum of galaxies at that redshift from the three surveys. The Model~C prediction is in good agreement with other models at low redshift (the curve follows closely the prediction of Pei \&~Fall 1995, \cite{pei95a}, as plotted in Dwek \etal 1998, \cite{dwek98a}, figure 3), which is impressive given that no free parameters were adjusted to fit the $z=0$ value. The predicted star formation history peaks around a redshift of 1, and falls off more quickly than other models at high redshift. In Model~S we use the luminosity function of Serjeant \etal (1998, \cite{serjeant98a}) (see eq.~\ref{eq.lumfunc.S98} above). The fitted evolution parameters were $\{ Q=4.3, q=2.1, z_q=1.3, P=1.3, p=1.7, z_p=2.3 \}$. The resulting evolution factors $f(z)$ and $g(z)$ are plotted in Figure~\ref{fig.modelS.evolfactor} and the resulting luminosity function is shown in Figure~\ref{fig.modelS.lumfuncevolve}. Despite very different individual parameters ($Q=4.3$ vs. $Q=7.6$), the two fits have similar functions $f(z)$ and $g(z)$. The predicted redshift distribution (Figure~\ref{fig.modelS.nz}) is peaked at a slightly lower redshift and has a slightly longer tail then Model~C. The predicted star formation history (Figure~\ref{fig.modelS.sfr}) has a larger local value than Model~C, but still less than that predicted by Baugh \etal (1998, \cite{baugh98a}) (thin solid line). The peak is around a redshift of 1.4, and falls off less rapidly than Model~C at high redshift. \subsection{Discussion} The star formation histories predicted by Model~C and Model~S both fall off more quickly at high redshift than model predictions by others. However, we are considering several refinements to our model that might modify this result. We are currently determining additional photometric redshifts (Waddington \& Windhorst, in preparation), which will make the modeling more reliable particularly at high redshift. The predicted shape of the star formation history is limited by the functional form we chose for evolution (eq.~\ref{eq.lumevol} and \ref{eq.numevol}), and we plan to experiment with other functions. If the IMF is evolving, or is dependent on environment, this would also affect our results. The relationship between star formation rate and radio luminosity (eq.~\ref{eq.sfrtoLpergal}) might be evolving in addition to its dependence on an evolving IMF. Finally, we have not explored the dependence of our results on cosmological parameters. This method has the potential to be an important indicator of star formation history. Radio luminosity is a reliable indicator of star formation rate in local galaxies, and is not affected by dust extinction. While others are performing similar calculations (Cram \etal 1998, \cite{cram98a}; Cram 1998, \cite{cram98b}; Mobasher \etal 1999, \cite{mobasher99a}, Serjeant \etal 1998, \cite{serjeant98a}), the survey data used here are complete to a substantially lower flux limit, with nearly complete knowledge of optical counterparts and $\sim$50\% completeness in redshifts. This allows us to place stronger constraints on the evolving radio luminosity function and to probe star formation activity to much higher redshifts. \section*{Acknowledgments} We are grateful to Eric Richards for helpful discussions. D.H. thanks the National Science Foundation for travel support for this Symposium. D.H. and B.P. acknowledge the support of NSF AST 96-16971. \section*{References}
\section{INTRODUCTION} An important class of binary systems has been identified in which a low-mass secondary (companion) star orbits a probable black hole; see White, Nagase, \& Parmar (1995), van Paradijs \& McClintock (1995), and Tanaka \& Lewin (1995) for extensive reviews. In all cases they were first observed in outburst as ``X-ray novae" (or ``soft X-ray transients"). Their X-ray spectra are generally characterized by a prominent, ``soft" thermal component ($kT \approx 1$ keV) as well as a ``hard" power-law tail extending to very high energies (0.1--1 MeV). During outburst, the radiation from these and other low-mass X-ray binaries (LMXBs; the ``low mass" refers to the secondary star) is emitted predominantly by the accretion disk surrounding the primary star. A lower limit to the mass of the primary ($M_1$) in a given X-ray transient can be measured when it returns to quiescence. At that time, light from the secondary star contributes significantly to (or even dominates) the visible spectrum, and the secondary's radial-velocity curve can be determined with a series of time-resolved spectra (see Cowley 1992 for a review). The orbital period ($P$) and semiamplitude ($K_2$) of the secondary yield the mass function of the primary, $f(M_1) = PK_2^3 /2\pi G = M_1^3\, {\rm sin}^3 i/(M_1 + M_2)^2$, where $i$ is the inclination of the orbital plane to our line of sight. Clearly, $f(M_1)$ provides an absolute lower limit to the mass of the primary; only if $M_2 = 0$ and $i = 90^\circ$ is $M_1 = f(M_1)$. It is generally acknowledged that if the primary is dark and has $f(M_1) \gtrsim 3.2~M_\odot$, it is probably a black hole, since the theoretical upper limit to the mass of a normal neutron star is $\sim$ 3.0--3.2~$M_\odot$ (Friedman, Ipser, \& Parker 1986; but see Friedman \& Ipser 1987, as well as Bahcall, Lynn, \& Selipsky 1990, for conditions that might allow neutron stars to exceed this nominal limit). The six best examples of LMXBs whose mass function exceeds (or is close to) $\sim 3~M_{\odot}$, along with the derived values of $f(M_1)$ (in units of $M_{\odot}$) and references, are as follows in chronological order of the mass-function measurement: A0620--00 = V616 Mon ($3.18 \pm 0.16$, McClintock \& Remillard 1986; $2.72 \pm 0.06$, Marsh, Robinson, \& Wood 1994; $2.91 \pm 0.08$, Orosz et al. 1994), GS~1124--68 = Nova Mus 1991 ($3.1 \pm 0.4$, Remillard, McClintock, \& Bailyn 1992; $3.01 \pm 0.15$, Orosz et al. 1996), GS~2023+338 = V404 Cyg ($6.26 \pm 0.31$, Casares, Charles, \& Naylor 1992; $6.08 \pm 0.06$, Casares \& Charles 1994), GRO J1655--40 = Nova Sco 1994 ($3.16 \pm 0.15$, Bailyn et al. 1995; $3.24 \pm 0.09$, Orosz \& Bailyn 1997), GS~2000+25 = Nova Vul 1988 ($5.02 \pm 0.47$, Casares, Charles, \& Marsh 1995; $4.97 \pm 0.10$, Filippenko, Matheson, \& Barth 1995a, slightly revised to $5.01 \pm 0.15$ [not $\pm 0.12$] by Harlaftis, Horne, \& Filippenko 1996), and Nova Oph 1977 ($4.0 \pm 0.8$, Remillard et al. 1996; $4.86 \pm 0.13$, Filippenko et al. 1997, slightly revised to $4.65 \pm 0.21$ by Harlaftis et al. 1997). Here we add a seventh object to this list: Nova Vel 1993 = GRS 1009--45, with a derived mass function of $3.17 \pm 0.12~M_\odot$. Nova Vel 1993 was discovered on 12 September 1993 with the WATCH all-sky monitor aboard {\it Granat} (Lapshov, Sazanov, \& Sunyaev 1993; Lapshov et al. 1994) and with BATSE on the {\it Compton Gamma-Ray Observatory} (Harmon et al. 1993). Its spectrum exhibited an ``ultrasoft" hump at low energies ($\lesssim 1$ keV) and a power-law tail out to at least 100 keV (Kaniovsky, Borozdin, \& Sunyaev 1993), typical of X-ray binaries in which the compact object is a black hole. Two months later (17 November), Della Valle \& Benetti (1993) discovered a blue optical counterpart at $V \approx 14.6$ mag, but reasonable estimates suggest that the magnitude at the time of outburst was $V = 13.8 \pm 0.3$ (Della Valle et al. 1997). Optical photometry conducted by Bailyn \& Orosz (1995) about half a year after the primary outburst showed the presence of a secondary outburst and several mini-outbursts, again reminiscent of black-hole X-ray novae. Della Valle et al. (1997) suggested an orbital period of about 4 hours, but a more reliable period of $6.86 \pm 0.12$ hours was obtained by Shahbaz et al. (1996). The spectral type of the secondary star in the binary system was estimated to be late-G/early-K by Shahbaz et al. (1996), and later than G5--K0 by Della Valle et al. (1997). On 1998 January 25 (UT dates are used throughout this paper), we obtained several $R$-band images of the field of Nova Vel 1993 with the Low Resolution Imaging Spectrometer (LRIS; Oke et al. 1995) at the Cassegrain focus of the Keck-II telescope. As can be seen in Figure 1, which shows a subset of one image (seeing $0.65''$), the nova was in quiescence by this time, with $R = 21.2 \pm 0.2$ mag.\footnote[2]{We adopt the magnitudes of comparison stars quoted in Table 1 of Della Valle et al. (1997). There appears to be a numbering mismatch, or errors in the photometry, of some stars in Table 1 of Shahbaz et al. (1996): for example, their Stars 2, 5, and 7 should have comparable magnitudes, yet they are listed as being very different.} To determine whether Nova Vel 1993 should be considered a good dynamical black-hole candidate, we decided to obtain a radial-velocity curve with LRIS. Our group had already obtained excellent results in this way for the black-hole candidates GRO J0422+32 (Filippenko, Matheson, \& Ho 1995b; Harlaftis et al. 1999), GS~2000+25 (Filippenko et al. 1995a; Harlaftis et al. 1996), and Nova Oph 1977 (Filippenko et al. 1997; Harlaftis et al. 1997), all of which are comparably faint. \section{OBSERVATIONS AND REDUCTIONS} Nova Vel 1993 was observed with LRIS in 1998 during the nights of January 25, February 1, March 5--6, and May 2, as well as in 1999 during the night of January 21. A journal of useful observations is given in Table~1. (The spectra obtained on 1998 May 2, and a few spectra on other nights, were of marginal quality and are not considered here.) Given the object's far southerly declination ($-45^\circ$), and the restrictive southwest azimuth limit on Keck-II ($\lesssim 185^\circ$), observing was restricted to $\sim 1.5$ hours per night; hence, a number of different observing runs separated by a range of intervals was needed to avoid serious aliasing of the orbital period. Conditions were always clear, and the seeing was about $1.2^{\prime\prime}$, reasonably good considering the high airmass ($\sim 2.4$). Typical exposure times were 1100~s. The long slit of width $1^{\prime\prime}$ was oriented at a position angle (PA) of $160^\circ$ for all observations except that of 1998 January 25. This was close to the parallactic angle at the time of observation, thereby reducing differences in the relative amount of light lost at different wavelengths (Filippenko 1982). At such a PA, the slit went directly through Star A, which is much brighter than the quiescent nova and only $\sim 1.6^{\prime\prime}$ SE of it. (Della Valle et al. 1997 incorrectly state that Star A is SW of the nova.) However, since the direction of atmospheric dispersion (i.e., the parallactic angle) at the time of observation coincided with the nova/Star-A orientation, the degree of contamination from Star A was nearly independent of wavelength. The slit also partially intersected Stars B and C (Fig. 1), but their light did not contaminate that of the nova; see Figure 2, which plots the intensity of light along the slit at the wavelength of H$\alpha$ (thereby accentuating the nova's contribution). We used a Tektronix $2048 \times 2048$ pixel CCD with a scale of $0.215^{\prime\prime}$ pixel$^{-1}$ ($0.43^{\prime\prime}$ per binned pixel in the spatial direction). The 1200 grooves mm$^{-1}$ grating, blazed at 7500~\AA, resulted in a wavelength range of $\sim$ 5650--6950~\AA, and a full-width at half-maximum (FWHM) spectral resolution of $\sim 2.5$~\AA\ ($\sim 120$ km s$^{-1}$) with the $1^{\prime\prime}$ slit, essentially identical to what we used in our previous studies of black-hole X-ray novae. Thirteen velocity standards with spectral types in the range G5~V--M2~V were observed with the same setup, given the estimated classification of the secondary star in Nova Vel 1993 (Shahbaz et al. 1996; Della Valle et al. 1997). Also, spectra of some sdF stars (Oke \& Gunn 1983) were obtained for flux calibration and removal of telluric absorption lines. Cosmic rays were eliminated from the two-dimensional spectra through comparison of pairs of consecutive exposures. The two-dimensional spectra were bias-subtracted and flattened in the usual manner. The wavelength scale was determined from polynomial fits to the positions of emission lines in spectra of Hg-Ne lamps obtained with the telescope at (or near) the position of each object. To ensure accurate wavelength calibration, final corrections ($\sim 0.0$--0.3~\AA) to the wavelength solution were obtained from night-sky emission lines in the spectra of the nova. We used the APALL task in IRAF\footnote[3]{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation.} to optimally extract (Horne 1986) one-dimensional, sky-subtracted spectra of Nova Vel 1993. However, to minimize contamination from Star A, we extracted only a few rows of the CCD centered on the position of the nova; as shown in Figure 2, these were rows 3--5 from the center of Star A, in the left wing of its spatial profile, corresponding to a displacement of 1.3--2.1$^{\prime\prime}$. To further remove contamination, we also extracted the same rows (3--5) in the right wing of Star A's spatial profile, and subtracted this spectrum from that of the nova.\footnote[4]{This extraction determined the flux scale of the subtracted spectrum, but its signal-to-noise (S/N) ratio was improved by using an extraction of the entire right half (rows 0--5) of Star A's spatial profile and scaling to the spectrum of rows 3--5.} We found that Star A typically contributed 40\% (65\% on January 21) of the flux in the original extraction of Nova Vel 1993. In general (but, surprisingly, not on January 21), the subsequent analysis worked best on these fully decontaminated spectra of the nova, whose typical S/N ratio per final 0.75~\AA\ bin is $3.3 \pm 0.4$ in the continuum. \section{RESULTS} \subsection{Unphased Average Spectra} Two unphased average spectra of Nova Vel 1993 are shown in Figure 3, with that of 6 March 1998 offset by 2.5 units. In the region of overlap, these resemble the spectra shown by Shahbaz et al. (1996). The strongest emission line is H$\alpha$, and there may be weak He~I $\lambda$5876 on 6 March, though the latter line partially coincides with the Na~I~D absorption. Despite being in quiescence, Nova Vel 1993 still exhibits variability; the equivalent width (EW) of the H$\alpha$ emission line, which is unaffected by errors in flux calibration, was somewhat larger on March 6 (76~\AA) than on January 21 (60~\AA). Note that this emission is considerably weaker than in GRO J0422+32 (EW $\approx 250$~\AA; Filippenko et al. 1995b), but comparable to that in GS~2000+25 (EW $\approx 40$~\AA; Filippenko et al. 1995a) and Nova Oph 1977 (EW = 25--85~\AA; Filippenko et al. 1997). As in the previous objects we have studied, the H$\alpha$ line has two peaks ($\Delta v \approx 1200$ km s$^{-1}$), more obvious in March than in January. \subsection{Cross-Correlations} Following the procedure discussed in Filippenko et al. (1995b), we employed the FXCOR package (``Release 9/13/93") in IRAF to cross-correlate the spectra of Nova Vel 1993 with the 13 velocity standards (shown in Fig. 4, along with two others from our previous studies). The correlation was done over the ranges 5980--6270~\AA\ and 6320--6500~\AA\ to avoid the H$\alpha$ and He~I $\lambda$5876 emission lines, Na~I~D absorption, the 6270~\AA\ interstellar line, and poorly subtracted [O~I] $\lambda$6300 night-sky emission. In almost all cases a definitive correlation peak was obvious. Typical values of the Tonry \& Davis (1979) significance threshold were quite high, $R \gtrsim 3$, with a few as high as $\sim 6$. As in our previous studies, we adopted the FXCOR $1\sigma$ uncertainties reduced by a factor of 2.77; the Fourier transform properties of Gaussians (the functions used in fitting the cross-correlation peaks) were used to determine that FXCOR overestimates the Tonry \& Davis uncertainties by this amount. The strongest formal correlation was obtained with BD+00~3090, which is officially listed as an M0~V star (Upgren et al. 1972), but over our spectral range it also looks very similar to K7~V and K8~V stars; see Figure 4. Indeed, the correlation was insignificantly lower with K7~V and K8~V stars, but far inferior with K5~V and M1~V stars. (We did not observe K6~V and K9~V stars.) Thus, we conclude that the secondary star lies somewhere in the range K7~V through M0~V, and possibly as early as K6~V. This is a slightly later spectral type than preferred by Shahbaz et al. (1996; late-G/early-K) and Della Valle et al. (1997; later than G5--K0), though the former authors note that their derived mean density for the secondary (2.4 g cm$^{-3}$) suggests a K5~V star. Our phased-average spectrum of the secondary star (see below) supports the late-K classification. Using the radial velocities evaluated from the correlations with BD+00~3090 (corrected for the radial velocity of BD+00~3090 itself; 46.6 km s$^{-1}$, Evans 1967), we conducted a non-linear least-squares fit (i.e., a $\chi^2$ fit; Press et al. 1986, p. 521) to obtain the best cosine curve to match the data (Fig. 5). The four-parameter fit (zero point, semiamplitude, period, and phase) yielded a systemic velocity of $\gamma_2 = 40.7 \pm 4.0$ km s$^{-1}$, a semiamplitude of $K_2 = 475.4 \pm 5.9$ km s$^{-1}$, a period of $P = 0.285206 \pm 0.0000014$~d (6.84~hr), and a starting time (heliocentric Julian day) for the phase of $T_0 =$ HJD 2,450,835.0661 $\pm$ 0.0007, where $T_0$ is defined as the point of maximum {\it redshifted} velocity. All the uncertainties are the formal $1\sigma$ values derived from the $\chi^2$ fit, but that of $T_0$ may be an underestimate because the choice of the range in which to search for $T_0$ is unconstrained by the data. A better measurement of the systemic velocity is $\gamma_2 = 30.1 \pm 5.0$ km s$^{-1}$, the weighted average of values obtained with the nine best standard stars. Note that when Star A (Fig. 1) was cross-correlated with the velocity standards, its derived velocities were constant to within $\pm 7$ km s$^{-1}$ (full range); thus, our results for Nova Vel 1993 are not an artifact of telescope position or instrument orientation, and our removal of the contamination by Star A (spectral type $\sim$K1) was effective. It is interesting to examine the distribution of possible periods resulting from our series of observations. In Figure 6, which shows ${\chi}^2$ versus trial period, there are several major groups of possible periods; the structure within each group results from ambiguities in the counting of cycles between widely separated epochs of observation (for example, from January through March 1998). Until we obtained the observations of 1999 January 21, the periods near 0.22~d and 0.285~d were equally probable, with those near 0.18~d and 0.4~d distinctly inferior. With the additional data, however, we are now reasonably confident that 0.285~d (6.84~hr) is correct. Based on observations of photometric modulation observed during the decline from outburst, Bailyn \& Orosz (1995) had speculated that the orbital period might be 1.6~hr or $\sim 3$~d, but these are clearly excluded by our radial velocity measurements. The possible presence of ``superhumps" in the optical light curve obtained 4 months after the primary outburst led Della Valle et al. (1997) to deduce an orbital period of 4~hr, closer yet still incorrect. Our period is essentially identical that found by Shahbaz et al. (1996; $6.86 \pm 0.12$~hr) from Gunn $R$-band ellipsoidal modulations in quiescence. Note that the secondary in Nova Vel 1993 must be a dwarf; with an orbital period of only 6.84~hr, the compact primary would be inside a giant or subgiant. The formal reduced ${\chi}^2$ (${\chi}_{\nu}^2$) for the best fit is 1.55 (17 points, 13 degrees of freedom). Given that the velocity uncertainties given by the Tonry \& Davis (1979) method don't reflect external errors such as miscentering of the object in the slit, they could easily be too small. In our case, increasing them by only 25\% yields a reduced ${\chi}_{\nu}^2$ of 1.0. Moreover, had we assigned uncertainties to the adopted times (i.e., phases) of the observations, and performed the fit to simultaneously minimize residuals in both velocity and phase, we would have obtained a smaller value of ${\chi}_{\nu}^2$. (The effective times of the observations can differ from the calculated midpoints due to variations in observing conditions --- seeing, transparency, etc.) \subsection{The Phased Average Spectrum} Having determined the orbital parameters of the secondary star, we obtained its master ``rest-frame" spectrum by averaging the 17 spectra after Doppler shifting each one to zero velocity. This represents a total integration time of $\sim$5~hr, but of course the S/N ratio is lower than for a single 5-hr exposure (ignoring cosmic rays) due to the increase in readout noise. Note that the H$\alpha$ and He~I emission lines are smeared out by this process, since they are produced in the accretion disk around the compact primary star. Similarly, interstellar absorption lines become less distinct. Figure 7 shows the phased average spectrum of Nova Vel 1993, in comparison with its typical spectrum (obtained at 11:14 UT on 1998 Feb. 1), the unphased average around H$\alpha$, and the spectrum of the M0~V velocity standard star. Scrutiny of the phased average spectrum reveals stellar absorption lines that are also present in the M0~V star (e.g., the Ca + Fe blend at 6498~\AA), but those of Nova Vel 1993 are weak. While this could indicate that the spectral type of the secondary star is substantially earlier than M0, tests show that several other factors are likely to be more important: (1) rotational broadening, perhaps 50--100 km s$^{-1}$ (e.g., Wade \& Horne 1988; Harlaftis et al. 1996); (2) orbital broadening, typically $4K_2 T/P = 80$--90 km s$^{-1}$ (Filippenko et al. 1995b); and, most importantly, (3) contamination by the featureless continuum of an accretion disk. We attempted to quantify the contribution of the accretion disk in Nova Vel 1993 by comparing its phased average spectrum with that of the M0~V star, broadening and diluting the latter by various amounts. A good match to the depths of the narrow absorption lines was found with accretion disk contamination fractions of 60\%--70\% of the total flux density at $\sim 6300$~\AA. Because of the uncertainties in the adopted parameters (broadening, spectral type) and the relatively low S/N ratio of the Nova Vel 1993 spectrum, we cannot confidently exclude contributions somewhat outside this range, but it is clear that the accretion disk dominates the spectrum even in quiescence. No absorption line of Li~I $\lambda$6708 is visible in our phased-average spectrum, to a $3\sigma$ upper limit of $\sim 0.1$~\AA. An unexpectedly strong Li~I line (EW = 0.25--0.48~\AA) is seen in the spectra of several X-ray novae (Mart\'\i n et al. 1994, and references therein; Filippenko et al. 1995a), but not in others (such as Nova Oph 1977; Filippenko et al. 1997; Harlaftis et al. 1997). \subsection{H$\alpha$ Measurements} To investigate the motion of the compact primary in Nova Vel 1993, we fit a Gaussian to the high-velocity wings ($v \approx$ 800--2000 km s$^{-1}$) of the H$\alpha$ line in each individual spectrum using the IRAF task SPECFIT. The regions of the fit (including the continuum) were 6400--6545~\AA\ and 6580--6710~\AA; the double-horned core of the line (Fig. 3) was excluded. The velocity of the Gaussian peak and its formal $1\sigma$ uncertainty were adopted for each spectrum (Table 1). As shown in Figure 8, the derived velocity tends to be low in the first half of the orbit and high in the second half, perhaps suggesting periodic behavior. We used a least-squares fit to determine the H$\alpha$ radial-velocity cosine curve, forcing the data to have the period found for the secondary star (0.285206~d) but allowing all other parameters to vary. The {\it formal} results (Fig. 8; ${\chi}_{\nu}^2 = 1.34$) are as follows: $\gamma_1 = 4.6 \pm 6.2$ km s$^{-1}$, $K_1 = 65.3 \pm 7.0$ km s$^{-1}$, and a zero point in the phase of HJD $2,450,834.9686 \pm 0.0093$~d. The value of $\gamma_1$ is inconsistent with that of $\gamma_2$, but this is often the case in X-ray novae (e.g., Nova Oph 1977, Filippenko et al. 1997). Once again, the zero point in the phase is the maximum {\it redshifted} velocity; it implies that the compact object is 237$^\circ$ out of phase with the companion star, rather than the expected 180$^\circ$. This is comparable to the offsets in GRO J0422+32 ($253^\circ$; Filippenko et al. 1995b) and GS~2000+25 ($260^\circ$; Filippenko et al. 1995a), as well as in A0620--00 and Nova Mus 1991 (Orosz et al. 1994). To date, there is no satisfactory quantitative explanation for these distortions, but they suggest that the accretion disk often has a nonaxisymmetric distribution of surface brightness, noncircular velocities, or a warp. They also cast some doubt on the use of H$\alpha$ radial-velocity curves to determine the motion and mass of the primary star. On the other hand, the mass ratios determined in this manner are frequently quite consistent with those obtained with independent techniques (e.g., A0620--00, Orosz et al. 1994; GS 2000+25, Harlaftis et al. 1996; Nova Oph 1977, Harlaftis et al. 1997; GRO J0422+32, Harlaftis et al. 1999). Thus, here we will cautiously adopt the ratio of semiamplitudes as an estimate of the mass ratio: $q = M_2/M_1 = K_1/K_2 = 0.137 \pm 0.015$. \section{THE MASS OF THE COMPACT PRIMARY} From the semiamplitude ($K_2 = 475.4 \pm 5.9$ km s$^{-1}$) and period ($P = 0.285206 \pm 0.00000138$~d) of the radial velocity curve of the secondary star, we find a mass function $f(M_1) = PK_2^3 /2\pi G = 3.17 \pm 0.12~M_\odot$. This corresponds to the absolute minimum mass of the compact primary, and it is close to the maximum gravitational mass of a slowly rotating neutron star (3.0--3.2~$M_\odot$; see Chitre \& Hartle 1976, and the discussion in Filippenko et al. 1995b). No evidence for eclipses is seen in the data of Bailyn \& Orosz (1995), Shahbaz et al. (1996), or Della Valle et al. (1997); similarly, we do not see significant variations in the apparent brightness of the secondary star over the orbital period. Hence, it is likely that the orbital inclination $i \lesssim 80^\circ$, and the relation $f(M_1) = M_1^3\, {\rm sin}^3 i/(M_1 + M_2)^2$ then implies that $M_1 \gtrsim$ 4.2--4.4~$M_\odot$ for nominal secondary-star masses of 0.5--0.65~$M_\odot$ (M0--K6~V; Allen 1976). Even if the secondary is quite undermassive (e.g., $M_2 = 0.3~M_\odot$), as in some X-ray binaries (e.g., van den Heuvel 1983), we derive $M_1 \gtrsim 3.9~M_\odot$. The primary star is therefore almost certainly a black hole rather than a neutron star. Adopting the mass ratio derived from the measured semiamplitude of the radial velocity curve of the primary ($q = M_2/M_1 = K_1/K_2 = 0.137 \pm 0.015$), we find $M_1 =$ 3.64--4.74~$M_\odot$ if $M_2 =$ 0.5--0.65~$M_\odot$. The constraints from $q$ and the mass function yield $M_1 = 4.4~M_\odot$ and $i \approx 78^\circ$ if we use a normal K7--K8 secondary ($M_2 \approx 0.6~M_\odot$). Indeed, consistency cannot be achieved for $M_2 \lesssim 0.59~M_\odot$ if the maximum inclination estimate ($80^\circ$) is correct. We conclude that the secondary star cannot be substantially undermassive, and that the orbital inclination is probably rather large, almost making Nova Vel 1993 an eclipsing binary. Our suggested inclination is significantly higher than the nominal value derived by Shahbaz et al. (1996) from their observed $R$-band ellipsoidal modulations ($i = 44^\circ \pm 7^\circ$). However, these authors admit that when contamination by light from the accretion disk is included, their allowed range for the inclination is $37^\circ$--$80^\circ$. Of course, in view of the unexplained phase offset between the expected and observed H$\alpha$ radial velocity curves (\S~3.4), it is also possible that our derived value of $q$ is erroneous, thereby affecting our estimate of $i$. Further studies are needed to accurately determine the inclination. Recently, Bailyn et al. (1998) found that the distribution of masses of putative black holes in LMXBs is very strongly peaked at $\sim 7~M_\odot$, only V404 Cyg being a clear high-mass deviant. The number of objects in the sample is still quite small, but if our mass estimate for Nova Vel 1993 is correct, then it appears to be a low-mass counterexample ($M_1 \approx 4.4~M_\odot$). Another possible exception is GRO J0422+32 ($M_1 \approx 5~M_\odot$; Harlaftis et al. 1999). An independent measure of the mass ratio of Nova Vel 1993 from the rotational broadening of the secondary star's absorption lines (e.g., Nova Oph 1977, Harlaftis et al. 1997; GRO J0422+32, Harlaftis et al. 1999), together with better constraints on the inclination derived from near-infrared ellipsoidal modulations (e.g., GS 2000+25, Callanan et al. 1996), would provide a very useful check on our estimated mass for the primary star. When we calculate the effective Roche lobe radius ($R_L$) of the companion star from the relation of Paczy\' nski (1971; see also Eggleton 1983) and Kepler's third law, we find that $R_L = 0.7~R_\odot$ if $M_1 \approx 4.4~M_\odot$ and $M_2 \approx 0.6~M_\odot$. This is only slightly larger than the expected radius of a typical K8 dwarf ($R = 0.67~R_\odot$; Allen 1976). Thus, the secondary star may be just starting its evolution off the main sequence. \section{OBSERVATIONS OF MXB 1659--29 IN QUIESCENCE} As part of our effort to determine the mass functions of X-ray binaries, we observed the field of MXB 1659--29. This burst source was discovered in 1976 October with SAS~3 (Lewin et al. 1976). It was initially considered unusual because of its very stable burst intervals (Lewin 1977) and apparent absence of constant emission, but such emission was found a year later (Lewin et al. 1978; Share et al. 1978). An optical counterpart (now known as V2134 Oph) was discovered by Doxsey et al. (1979) at $V = 18.3$ mag; it was quite blue, and possibly exhibited emission lines of He~II $\lambda$4686 and C~III/N~III $\lambda\lambda$4640--4650. Figure 2 of Doxsey et al. (in which N is up and E to the left, although this isn't stated) shows $U$-band and $B$-band finder charts for the object, from images taken on 1978 May 30 and June 1 with the CTIO 4-m telescope. We obtained three dithered $R$-band images (exposure times of 60, 60 and 30~s) of the MXB 1659--29 field on 1999 February 9 with LRIS/Keck-II, at airmass 1.8--1.9. These were bias-subtracted, flattened, registered, and combined in the usual manner. A small subset of the resulting crowded image (Galactic latitude 7.3$^\circ$) is shown in Figure 9a; the FWHM of stars is measured to be $\sim 0.9''$. Star A is close to the apparent position of V2134 Oph indicated in the relatively shallow charts published by Doxsey et al. (1979). However, seven LRIS/Keck-II long-slit spectra (PA = 160$^\circ$) of this star, each with a typical exposure time of 900--1000~s, do not reveal any H$\alpha$ emission characteristic of accretion disks. Moreover, cross-correlation of the individual spectra with the 13 velocity standards (G5 providing the best match) reveals no clear variability beyond the $\pm 15$ km s$^{-1}$ level, and no systematic trend among consecutive exposures, casting further doubt on Star A as the secondary. H$\alpha$ emission is also weak or absent in our noisy spectrum of Star B (Fig. 9a), and Star E has a spectral type of M. Shortly before the completion of this paper, MXB 1659--29 went into outburst again, after a hiatus of 21 years. During the interval 1999 April 2.06--3.47, the Wide Field Camera on BeppoSAX detected a transient X-ray source coincident with the position of MXB 1659--29 (in 't Zand et al. 1999). This object was confirmed on April 5.83--6.05 with RXTE (Markwardt et al. 1999). Optical observations (Augusteijn, Freyhammer, \& in 't Zand 1999) on April 3.41 revealed a bright new source ($V = 18.3 \pm 0.1$) at that location, with a spectrum typical of LMXBs in an X-ray bright phase (emission lines of H~I, He~II, C~III, and N~III). An image (exposure time 30~s) obtained on April 18 with LRIS/Keck-II is shown in Figure 9b (stellar FWHM = $0.65''$); the optical counterpart ($R = 18.2 \pm 0.05$) is marked ``F." This star is also visible at the center of the circle in Figure 9a, barely at the detection limit. We measured the $R$ magnitude of V2134 Oph in quiescence (Fig. 9a) with the technique of point-spread-function (PSF) fitting, where the PSF was determined iteratively by subtracting faint stars near the ones chosen for the PSF. The zero point was obtained from twilight-sky observations of PG1525--071A,C (Landolt 1992). Star F, which we identify with V2134 Oph in quiescence, has $R = 23.6 \pm 0.4$ mag. If Star F is just a chance superposition, then the true optical counterpart is even fainter. Thus, any future attempts to obtain the mass function of MXB 1659--29 (V2134 Oph) will be extremely difficult to perform! As an aid for future photometry of this object, we note that the final magnitudes for Stars A, B, C, D, and E are 19.6, 22.5, 23.1, 23.3, and 21.0, respectively. The $1\sigma$ uncertainty is about 0.05 mag at the bright end (primarily due to the dearth of photometric standards) and perhaps 0.2 mag for stars at $R \approx 23$. \section{CONCLUSIONS} Our observations of Nova Vel 1993 provide a definitive mass function of $3.17 \pm 0.12~M_\odot$, and a likely mass of around $4.4~M_\odot$ for the primary star. Thus, Nova Vel 1993 joins the small but growing list of secure Galactic black holes first identified as X-ray novae. However, its mass seems to be lower than that of other objects in its class, bridging the apparent gap between $\sim 3~M_\odot$ (the theoretical maximum mass of a neutron star, though observed masses almost always yield $M = 1.0$--1.8~$M_\odot$; Thorsett et al. 1993) and $\sim 7~M_\odot$ (the mass of most Galactic black holes in binary systems with well-determined parameters). It will be important to measure the mass ratio and orbital inclination of the system with techniques independent of the indirect ones used here, to confirm our estimates ($q = 0.137 \pm 0.015$; $i \approx 78^\circ$) and our derived mass. A similar study of MXB 1659--29 (V2134 Oph) eliminates several candidate stars as the optical counterpart. The recent new outburst, which occurred just prior to the submission of this paper, allows us to identify the quiescent nova at $R = 23.6 \pm 0.4$ mag, unless this is an unrelated star superposed along the line of sight. \acknowledgments Data presented herein were obtained at the W. M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California, and NASA. The Observatory was made possible by the generous financial support of the W. M. Keck Foundation. We thank T. Bida, R. Campbell, R. Goodrich, D. Lynn, G. Puniwai, H. Rodriguez, C. Sorensen, W. Wack, T. Williams, and other members of the Keck staff for their able assistance. We are also grateful to N. Vogt, A. C. Phillips, and D. C. Koo for obtaining a Keck image of MXB 1659--29 on April 18, after its recent outburst. This work was supported by the NSF through Grant AST--9417213 to A.V.F. \clearpage \begin{deluxetable}{rccclrr} \tablewidth{500pt} \tablecaption{Journal of Observations and Radial Velocities \label{tbl-1}} \tablehead{\colhead{HJD\tablenotemark{a}} & \colhead{UT Date} & \colhead{Exposure (s)} & \colhead{Airmass\tablenotemark{b}} & \colhead{Phase\tablenotemark{c}} & \colhead{$v_2$ (km s$^{-1}$)\tablenotemark{d}} & \colhead{$v_1$ (km s$^{-1}$)\tablenotemark{e}} } \startdata 838.96576 & 1998 Jan 25 & 900 & 2.53 & 0.67402 & $-236.7 \pm 21.7$ & $123.0 \pm 62.3$ \nl 845.95655 & 1998 Feb 01 & 1200 & 2.46 & 0.18531 & $219.4 \pm 18.9$ & $-36.10 \pm 21.7$ \nl 845.97085 & 1998 Feb 01 & 1200 & 2.39 & 0.23545 & $52.09 \pm 17.1$ & $-21.29 \pm 25.1$ \nl 877.85456 & 1998 Mar 05 & 900 & 2.58 & 0.026917 & $448.5 \pm 13.4$ & $-61.26 \pm 28.3$ \nl 877.86708 & 1998 Mar 05 & 1200 & 2.48 & 0.070815 & $430.6 \pm 15.0$ & $-79.79 \pm 19.7$ \nl 877.88132 & 1998 Mar 05 & 1200 & 2.41 & 0.12074 & $328.9 \pm 8.62$ & $-34.98 \pm 18.8$ \nl 877.89554 & 1998 Mar 05 & 1200 & 2.36 & 0.17060 & $248.4 \pm 17.5$ & $-51.49 \pm 17.6$ \nl 877.90868 & 1998 Mar 05 & 1000 & 2.35 & 0.21667 & $123.3 \pm 15.3$ & $-71.38 \pm 21.1$ \nl 877.92060 & 1998 Mar 05 & 1000 & 2.35 & 0.25847 & $-61.19 \pm 15.5$ & $-86.64 \pm 21.0$ \nl 878.85228 & 1998 Mar 06 & 1100 & 2.58 & 0.52515 & $-476.1 \pm 12.0$ & $75.89 \pm 25.1$ \nl 878.89168 & 1998 Mar 06 & 1100 & 2.37 & 0.66329 & $-229.7 \pm 17.2$ & $63.71 \pm 22.3$ \nl 878.90475 & 1998 Mar 06 & 1100 & 2.35 & 0.70912 & $-120.3 \pm 15.8$ & $40.81 \pm 20.7$ \nl 878.91785 & 1998 Mar 06 & 1100 & 2.36 & 0.75505 & $ 11.61 \pm 21.8$ & $86.20 \pm 18.5$ \nl 1199.97900 & 1999 Jan 21 & 1200 & 2.52 & 0.46778 & $-476.9 \pm 12.0$ & $14.13 \pm 41.8$ \nl 1200.00574 & 1999 Jan 21 & 1100 & 2.38 & 0.56154 & $-451.1 \pm 8.61$ & $27.51 \pm 50.5$ \nl 1200.01881 & 1999 Jan 21 & 1100 & 2.35 & 0.60736 & $-370.1 \pm 12.8$ & $31.60 \pm 66.0$ \nl 1200.03255 & 1999 Jan 21 & 1200 & 2.35 & 0.65554 & $-272.4 \pm 27.6$ & $14.54 \pm 63.6$ \nl \enddata \tablenotetext{a}{HJD--2,450,000 at midpoint of exposure.} \tablenotetext{b}{Airmass at midpoint of exposure.} \tablenotetext{c}{Using $P = 0.285206$~d and $T_0 = 2,450,835.0661$.} \tablenotetext{d}{Secondary star radial velocity.} \tablenotetext{e}{H$\alpha$ centroid radial velocity, from fit to emission-line wings.} \end{deluxetable} \clearpage
\section*{Abstract (Not appropriate in this style!)}% \else \small \begin{center}{\bf Abstract\vspace{-.5em}\vspace{\z@}}\end{center}% \quotation \fi }% }{% }% \@ifundefined{endabstract}{\def\endabstract {\if@twocolumn\else\endquotation\fi}}{}% \@ifundefined{maketitle}{\def\maketitle#1{}}{}% \@ifundefined{affiliation}{\def\affiliation#1{}}{}% \@ifundefined{proof}{\def\proof{\noindent{\bfseries Proof. }}}{}% \@ifundefined{endproof}{\def\endproof{\mbox{\ \rule{.1in}{.1in}}}}{}% \@ifundefined{newfield}{\def\newfield#1#2{}}{}% \@ifundefined{chapter}{\def\chapter#1{\par(Chapter head:)#1\par }% \newcount\c@chapter}{}% \@ifundefined{part}{\def\part#1{\par(Part head:)#1\par }}{}% \@ifundefined{section}{\def\section#1{\par(Section head:)#1\par }}{}% \@ifundefined{subsection}{\def\subsection#1% {\par(Subsection head:)#1\par }}{}% \@ifundefined{subsubsection}{\def\subsubsection#1% {\par(Subsubsection head:)#1\par }}{}% \@ifundefined{paragraph}{\def\paragraph#1% {\par(Subsubsubsection head:)#1\par }}{}% \@ifundefined{subparagraph}{\def\subparagraph#1% {\par(Subsubsubsubsection head:)#1\par }}{}% \@ifundefined{therefore}{\def\therefore{}}{}% \@ifundefined{backepsilon}{\def\backepsilon{}}{}% \@ifundefined{yen}{\def\yen{\hbox{\rm\rlap=Y}}}{}% \@ifundefined{registered}{% \def\registered{\relax\ifmmode{}\r@gistered \else$\m@th\r@gistered$\fi}% \def\r@gistered{^{\ooalign {\hfil\raise.07ex\hbox{$\scriptstyle\rm\RIfM@\expandafter\text@\else\expandafter\mbox\fi{R}$}\hfil\crcr \mathhexbox20D}}}}{}% \@ifundefined{Eth}{\def\Eth{}}{}% \@ifundefined{eth}{\def\eth{}}{}% \@ifundefined{Thorn}{\def\Thorn{}}{}% \@ifundefined{thorn}{\def\thorn{}}{}% \def\TEXTsymbol#1{\mbox{$#1$}}% \@ifundefined{degree}{\def\degree{{}^{\circ}}}{}% \newdimen\theight \def\Column{% \vadjust{\setbox\z@=\hbox{\scriptsize\quad\quad tcol}% \theight=\ht\z@\advance\theight by \dp\z@\advance\theight by \lineskip \kern -\theight \vbox to \theight{% \rightline{\rlap{\box\z@}}% \vss }% }% }% \def\qed{% \ifhmode\unskip\nobreak\fi\ifmmode\ifinner\else\hskip5\p@\fi\fi \hbox{\hskip5\p@\vrule width4\p@ height6\p@ depth1.5\p@\hskip\p@}% }% \def\cents{\hbox{\rm\rlap/c}}% \def\miss{\hbox{\vrule height2\p@ width 2\p@ depth\z@}}% \def\vvert{\Vert \def\tcol#1{{\baselineskip=6\p@ \vcenter{#1}} \Column} % \def\dB{\hbox{{}} \def\mB#1{\hbox{$#1$} \def\nB#1{\hbox{#1} \@ifundefined{note}{\def\note{$^{\dag}}}{}% \defLaTeX2e{LaTeX2e} \ifx\fmtnameLaTeX2e \DeclareOldFontCommand{\rm}{\normalfont\rmfamily}{\mathrm} \DeclareOldFontCommand{\sf}{\normalfont\sffamily}{\mathsf} \DeclareOldFontCommand{\tt}{\normalfont\ttfamily}{\mathtt} \DeclareOldFontCommand{\bf}{\normalfont\bfseries}{\mathbf} \DeclareOldFontCommand{\it}{\normalfont\itshape}{\mathit} \DeclareOldFontCommand{\sl}{\normalfont\slshape}{\@nomath\sl} \DeclareOldFontCommand{\sc}{\normalfont\scshape}{\@nomath\sc} \fi \def\alpha{{\Greekmath 010B}}% \def\beta{{\Greekmath 010C}}% \def\gamma{{\Greekmath 010D}}% \def\delta{{\Greekmath 010E}}% \def\epsilon{{\Greekmath 010F}}% \def\zeta{{\Greekmath 0110}}% \def\eta{{\Greekmath 0111}}% \def\theta{{\Greekmath 0112}}% \def\iota{{\Greekmath 0113}}% \def\kappa{{\Greekmath 0114}}% \def\lambda{{\Greekmath 0115}}% \def\mu{{\Greekmath 0116}}% \def\nu{{\Greekmath 0117}}% \def\xi{{\Greekmath 0118}}% \def\pi{{\Greekmath 0119}}% \def\rho{{\Greekmath 011A}}% \def\sigma{{\Greekmath 011B}}% \def\tau{{\Greekmath 011C}}% \def\upsilon{{\Greekmath 011D}}% \def\phi{{\Greekmath 011E}}% \def\chi{{\Greekmath 011F}}% \def\psi{{\Greekmath 0120}}% \def\omega{{\Greekmath 0121}}% \def\varepsilon{{\Greekmath 0122}}% \def\vartheta{{\Greekmath 0123}}% \def\varpi{{\Greekmath 0124}}% \def\varrho{{\Greekmath 0125}}% \def\varsigma{{\Greekmath 0126}}% \def\varphi{{\Greekmath 0127}}% \def{\Greekmath 0272}{{\Greekmath 0272}} \def\FindBoldGroup{% {\setbox0=\hbox{$\mathbf{x\global\edef\theboldgroup{\the\mathgroup}}$}}% } \def\Greekmath#1#2#3#4{% \if@compatibility \ifnum\mathgroup=\symbold \mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}% \else \mathchar"#1#2#3# \fi \else \FindBoldGroup \ifnum\mathgroup=\theboldgroup \mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}% \else \mathchar"#1#2#3# \fi \fi} \newif\ifGreekBold \GreekBoldfalse \let\SAVEPBF=\pbf \def\pbf{\GreekBoldtrue\SAVEPBF}% \@ifundefined{theorem}{\newtheorem{theorem}{Theorem}}{} \@ifundefined{lemma}{\newtheorem{lemma}[theorem]{Lemma}}{} \@ifundefined{corollary}{\newtheorem{corollary}[theorem]{Corollary}}{} \@ifundefined{conjecture}{\newtheorem{conjecture}[theorem]{Conjecture}}{} \@ifundefined{proposition}{\newtheorem{proposition}[theorem]{Proposition}}{} \@ifundefined{axiom}{\newtheorem{axiom}{Axiom}}{} \@ifundefined{remark}{\newtheorem{remark}{Remark}}{} \@ifundefined{example}{\newtheorem{example}{Example}}{} \@ifundefined{exercise}{\newtheorem{exercise}{Exercise}}{} \@ifundefined{definition}{\newtheorem{definition}{Definition}}{} \@ifundefined{mathletters}{% \newcounter{equationnumber} \def\mathletters{% \addtocounter{equation}{1} \edef\@currentlabel{\arabic{equation}}% \setcounter{equationnumber}{\c@equation} \setcounter{equation}{0}% \edef\arabic{equation}{\@currentlabel\noexpand\alph{equation}}% } \def\endmathletters{% \setcounter{equation}{\value{equationnumber}}% } }{} \@ifundefined{BibTeX}{% \def\BibTeX{{\rm B\kern-.05em{\sc i\kern-.025em b}\kern-.08em T\kern-.1667em\lower.7ex\hbox{E}\kern-.125emX}}}{}% \@ifundefined{AmS}% {\def\AmS{{\protect\usefont{OMS}{cmsy}{m}{n}% A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}}{}% \@ifundefined{AmSTeX}{\def\AmSTeX{\protect\AmS-\protect\TeX\@}}{}% \def\@@eqncr{\let\@tempa\relax \ifcase\@eqcnt \def\@tempa{& & &}\or \def\@tempa{& &}% \else \def\@tempa{&}\fi \@tempa \if@eqnsw \iftag@ \@taggnum \else \@eqnnum\stepcounter{equation}% \fi \fi \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@eqnswtrue \global\@eqcnt\z@\cr} \def\@ifnextchar*{\@TCItagstar}{\@TCItag}{\@ifnextchar*{\@TCItagstar}{\@TCItag}} \def\@TCItag#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{(#1)}} \def\@TCItagstar*#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{#1}} \def\tfrac#1#2{{\textstyle {#1 \over #2}}}% \def\dfrac#1#2{{\displaystyle {#1 \over #2}}}% \def\binom#1#2{{#1 \choose #2}}% \def\tbinom#1#2{{\textstyle {#1 \choose #2}}}% \def\dbinom#1#2{{\displaystyle {#1 \choose #2}}}% \def\QATOP#1#2{{#1 \atop #2}}% \def\QTATOP#1#2{{\textstyle {#1 \atop #2}}}% \def\QDATOP#1#2{{\displaystyle {#1 \atop #2}}}% \def\QABOVE#1#2#3{{#2 \above#1 #3}}% \def\QTABOVE#1#2#3{{\textstyle {#2 \above#1 #3}}}% \def\QDABOVE#1#2#3{{\displaystyle {#2 \above#1 #3}}}% \def\QOVERD#1#2#3#4{{#3 \overwithdelims#1#2 #4}}% \def\QTOVERD#1#2#3#4{{\textstyle {#3 \overwithdelims#1#2 #4}}}% \def\QDOVERD#1#2#3#4{{\displaystyle {#3 \overwithdelims#1#2 #4}}}% \def\QATOPD#1#2#3#4{{#3 \atopwithdelims#1#2 #4}}% \def\QTATOPD#1#2#3#4{{\textstyle {#3 \atopwithdelims#1#2 #4}}}% \def\QDATOPD#1#2#3#4{{\displaystyle {#3 \atopwithdelims#1#2 #4}}}% \def\QABOVED#1#2#3#4#5{{#4 \abovewithdelims#1#2#3 #5}}% \def\QTABOVED#1#2#3#4#5{{\textstyle {#4 \abovewithdelims#1#2#3 #5}}}% \def\QDABOVED#1#2#3#4#5{{\displaystyle {#4 \abovewithdelims#1#2#3 #5}}}% \def\tint{\mathop{\textstyle \int}}% \def\tiint{\mathop{\textstyle \iint }}% \def\tiiint{\mathop{\textstyle \iiint }}% \def\tiiiint{\mathop{\textstyle \iiiint }}% \def\tidotsint{\mathop{\textstyle \idotsint }}% \def\toint{\mathop{\textstyle \oint}}% \def\tsum{\mathop{\textstyle \sum }}% \def\tprod{\mathop{\textstyle \prod }}% \def\tbigcap{\mathop{\textstyle \bigcap }}% \def\tbigwedge{\mathop{\textstyle \bigwedge }}% \def\tbigoplus{\mathop{\textstyle \bigoplus }}% \def\tbigodot{\mathop{\textstyle \bigodot }}% \def\tbigsqcup{\mathop{\textstyle \bigsqcup }}% \def\tcoprod{\mathop{\textstyle \coprod }}% \def\tbigcup{\mathop{\textstyle \bigcup }}% \def\tbigvee{\mathop{\textstyle \bigvee }}% \def\tbigotimes{\mathop{\textstyle \bigotimes }}% \def\tbiguplus{\mathop{\textstyle \biguplus }}% \def\dint{\mathop{\displaystyle \int}}% \def\diint{\mathop{\displaystyle \iint }}% \def\diiint{\mathop{\displaystyle \iiint }}% \def\diiiint{\mathop{\displaystyle \iiiint }}% \def\didotsint{\mathop{\displaystyle \idotsint }}% \def\doint{\mathop{\displaystyle \oint}}% \def\dsum{\mathop{\displaystyle \sum }}% \def\dprod{\mathop{\displaystyle \prod }}% \def\dbigcap{\mathop{\displaystyle \bigcap }}% \def\dbigwedge{\mathop{\displaystyle \bigwedge }}% \def\dbigoplus{\mathop{\displaystyle \bigoplus }}% \def\dbigodot{\mathop{\displaystyle \bigodot }}% \def\dbigsqcup{\mathop{\displaystyle \bigsqcup }}% \def\dcoprod{\mathop{\displaystyle \coprod }}% \def\dbigcup{\mathop{\displaystyle \bigcup }}% \def\dbigvee{\mathop{\displaystyle \bigvee }}% \def\dbigotimes{\mathop{\displaystyle \bigotimes }}% \def\dbiguplus{\mathop{\displaystyle \biguplus }}% \ifx\ds@amstex\relax \message{amstex already loaded}\makeatother\endinpu \else \@ifpackageloaded{amsmath}% {\message{amsmath already loaded}\makeatother\endinput} {} \@ifpackageloaded{amstex}% {\message{amstex already loaded}\makeatother\endinput} {} \@ifpackageloaded{amsgen}% {\message{amsgen already loaded}\makeatother\endinput} {} \fi \let\DOTSI\relax \def\RIfM@{\relax\ifmmode}% \def\FN@{\futurelet\next}% \newcount\intno@ \def\iint{\DOTSI\intno@\tw@\FN@\ints@}% \def\iiint{\DOTSI\intno@\thr@@\FN@\ints@}% \def\iiiint{\DOTSI\intno@4 \FN@\ints@}% \def\idotsint{\DOTSI\intno@\z@\FN@\ints@}% \def\ints@{\findlimits@\ints@@}% \newif\iflimtoken@ \newif\iflimits@ \def\findlimits@{\limtoken@true\ifx\next\limits\limits@true \else\ifx\next\nolimits\limits@false\else \limtoken@false\ifx\ilimits@\nolimits\limits@false\else \ifinner\limits@false\else\limits@true\fi\fi\fi\fi}% \def\multint@{\int\ifnum\intno@=\z@\intdots@ \else\intkern@\fi \ifnum\intno@>\tw@\int\intkern@\fi \ifnum\intno@>\thr@@\int\intkern@\fi \int \def\multintlimits@{\intop\ifnum\intno@=\z@\intdots@\else\intkern@\fi \ifnum\intno@>\tw@\intop\intkern@\fi \ifnum\intno@>\thr@@\intop\intkern@\fi\intop}% \def\intic@{% \mathchoice{\hskip.5em}{\hskip.4em}{\hskip.4em}{\hskip.4em}}% \def\negintic@{\mathchoice {\hskip-.5em}{\hskip-.4em}{\hskip-.4em}{\hskip-.4em}}% \def\ints@@{\iflimtoken@ \def\ints@@@{\iflimits@\negintic@ \mathop{\intic@\multintlimits@}\limits \else\multint@\nolimits\fi \eat@ \else \def\ints@@@{\iflimits@\negintic@ \mathop{\intic@\multintlimits@}\limits\else \multint@\nolimits\fi}\fi\ints@@@}% \def\intkern@{\mathchoice{\!\!\!}{\!\!}{\!\!}{\!\!}}% \def\plaincdots@{\mathinner{\cdotp\cdotp\cdotp}}% \def\intdots@{\mathchoice{\plaincdots@}% {{\cdotp}\mkern1.5mu{\cdotp}\mkern1.5mu{\cdotp}}% {{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}% {{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}}% \def\RIfM@{\relax\protect\ifmmode} \def\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\RIfM@\expandafter\RIfM@\expandafter\text@\else\expandafter\mbox\fi@\else\expandafter\mbox\fi} \let\nfss@text\RIfM@\expandafter\text@\else\expandafter\mbox\fi \def\RIfM@\expandafter\text@\else\expandafter\mbox\fi@#1{\mathchoice {\textdef@\displaystyle\f@size{#1}}% {\textdef@\textstyle\tf@size{\firstchoice@false #1}}% {\textdef@\textstyle\sf@size{\firstchoice@false #1}}% {\textdef@\textstyle \ssf@size{\firstchoice@false #1}}% \glb@settings} \def\textdef@#1#2#3{\hbox{{% \everymath{#1}% \let\f@size#2\selectfont #3}}} \newif\iffirstchoice@ \firstchoice@true \def\Let@{\relax\iffalse{\fi\let\\=\cr\iffalse}\fi}% \def\vspace@{\def\vspace##1{\crcr\noalign{\vskip##1\relax}}}% \def\multilimits@{\bgroup\vspace@\Let@ \baselineskip\fontdimen10 \scriptfont\tw@ \advance\baselineskip\fontdimen12 \scriptfont\tw@ \lineskip\thr@@\fontdimen8 \scriptfont\thr@@ \lineskiplimit\lineskip \vbox\bgroup\ialign\bgroup\hfil$\m@th\scriptstyle{##}$\hfil\crcr}% \def\Sb{_\multilimits@}% \def\endSb{\crcr\egroup\egroup\egroup}% \def\Sp{^\multilimits@}% \let\endSp\endSb \newdimen\ex@ \[email protected] \def\rightarrowfill@#1{$#1\m@th\mathord-\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$}% \def\leftarrowfill@#1{$#1\m@th\mathord\leftarrow\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill\mkern-6mu\mathord-$}% \def\leftrightarrowfill@#1{$#1\m@th\mathord\leftarrow \mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$}% \def\overrightarrow{\mathpalette\overrightarrow@}% \def\overrightarrow@#1#2{\vbox{\ialign{##\crcr\rightarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \let\overarrow\overrightarrow \def\overleftarrow{\mathpalette\overleftarrow@}% \def\overleftarrow@#1#2{\vbox{\ialign{##\crcr\leftarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \def\overleftrightarrow{\mathpalette\overleftrightarrow@}% \def\overleftrightarrow@#1#2{\vbox{\ialign{##\crcr \leftrightarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \def\underrightarrow{\mathpalette\underrightarrow@}% \def\underrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil $\crcr\noalign{\nointerlineskip}\rightarrowfill@#1\crcr}}}% \let\underarrow\underrightarrow \def\underleftarrow{\mathpalette\underleftarrow@}% \def\underleftarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil $\crcr\noalign{\nointerlineskip}\leftarrowfill@#1\crcr}}}% \def\underleftrightarrow{\mathpalette\underleftrightarrow@}% \def\underleftrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th \hfil#1#2\hfil$\crcr \noalign{\nointerlineskip}\leftrightarrowfill@#1\crcr}}}% \def\qopnamewl@#1{\mathop{\operator@font#1}\nlimits@} \let\nlimits@\displaylimits \def\setboxz@h{\setbox\z@\hbox} \def\varlim@#1#2{\mathop{\vtop{\ialign{##\crcr \hfil$#1\m@th\operator@font lim$\hfil\crcr \noalign{\nointerlineskip}#2#1\crcr \noalign{\nointerlineskip\kern-\ex@}\crcr}}}} \def\rightarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@ $#1\copy\z@\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\box\z@\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$} \def\leftarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@ $#1\mathord\leftarrow\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\copy\z@\mkern-2mu$}\hfill \mkern-6mu\box\z@$} \def\qopnamewl@{proj\,lim}{\qopnamewl@{proj\,lim}} \def\qopnamewl@{inj\,lim}{\qopnamewl@{inj\,lim}} \def\mathpalette\varlim@\rightarrowfill@{\mathpalette\varlim@\rightarrowfill@} \def\mathpalette\varlim@\leftarrowfill@{\mathpalette\varlim@\leftarrowfill@} \def\mathpalette\varliminf@{}{\mathpalette\mathpalette\varliminf@{}@{}} \def\mathpalette\varliminf@{}@#1{\mathop{\underline{\vrule\@depth.2\ex@\@width\z@ \hbox{$#1\m@th\operator@font lim$}}}} \def\mathpalette\varlimsup@{}{\mathpalette\mathpalette\varlimsup@{}@{}} \def\mathpalette\varlimsup@{}@#1{\mathop{\overline {\hbox{$#1\m@th\operator@font lim$}}}} \def\stackunder#1#2{\mathrel{\mathop{#2}\limits_{#1}}}% \begingroup \catcode `|=0 \catcode `[= 1 \catcode`]=2 \catcode `\{=12 \catcode `\}=12 \catcode`\\=12 |gdef|@alignverbatim#1\end{align}[#1|end[align]] |gdef|@salignverbatim#1\end{align*}[#1|end[align*]] |gdef|@alignatverbatim#1\end{alignat}[#1|end[alignat]] |gdef|@salignatverbatim#1\end{alignat*}[#1|end[alignat*]] |gdef|@xalignatverbatim#1\end{xalignat}[#1|end[xalignat]] |gdef|@sxalignatverbatim#1\end{xalignat*}[#1|end[xalignat*]] |gdef|@gatherverbatim#1\end{gather}[#1|end[gather]] |gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]] |gdef|@gatherverbatim#1\end{gather}[#1|end[gather]] |gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]] |gdef|@multilineverbatim#1\end{multiline}[#1|end[multiline]] |gdef|@smultilineverbatim#1\end{multiline*}[#1|end[multiline*]] |gdef|@arraxverbatim#1\end{arrax}[#1|end[arrax]] |gdef|@sarraxverbatim#1\end{arrax*}[#1|end[arrax*]] |gdef|@tabulaxverbatim#1\end{tabulax}[#1|end[tabulax]] |gdef|@stabulaxverbatim#1\end{tabulax*}[#1|end[tabulax*]] |endgroup \def\align{\@verbatim \frenchspacing\@vobeyspaces \@alignverbatim You are using the "align" environment in a style in which it is not defined.} \let\endalign=\endtrivlist \@namedef{align*}{\@verbatim\@salignverbatim You are using the "align*" environment in a style in which it is not defined.} \expandafter\let\csname endalign*\endcsname =\endtrivlist \def\alignat{\@verbatim \frenchspacing\@vobeyspaces \@alignatverbatim You are using the "alignat" environment in a style in which it is not defined.} \let\endalignat=\endtrivlist \@namedef{alignat*}{\@verbatim\@salignatverbatim You are using the "alignat*" environment in a style in which it is not defined.} \expandafter\let\csname endalignat*\endcsname =\endtrivlist \def\xalignat{\@verbatim \frenchspacing\@vobeyspaces \@xalignatverbatim You are using the "xalignat" environment in a style in which it is not defined.} \let\endxalignat=\endtrivlist \@namedef{xalignat*}{\@verbatim\@sxalignatverbatim You are using the "xalignat*" environment in a style in which it is not defined.} \expandafter\let\csname endxalignat*\endcsname =\endtrivlist \def\gather{\@verbatim \frenchspacing\@vobeyspaces \@gatherverbatim You are using the "gather" environment in a style in which it is not defined.} \let\endgather=\endtrivlist \@namedef{gather*}{\@verbatim\@sgatherverbatim You are using the "gather*" environment in a style in which it is not defined.} \expandafter\let\csname endgather*\endcsname =\endtrivlist \def\multiline{\@verbatim \frenchspacing\@vobeyspaces \@multilineverbatim You are using the "multiline" environment in a style in which it is not defined.} \let\endmultiline=\endtrivlist \@namedef{multiline*}{\@verbatim\@smultilineverbatim You are using the "multiline*" environment in a style in which it is not defined.} \expandafter\let\csname endmultiline*\endcsname =\endtrivlist \def\arrax{\@verbatim \frenchspacing\@vobeyspaces \@arraxverbatim You are using a type of "array" construct that is only allowed in AmS-LaTeX.} \let\endarrax=\endtrivlist \def\tabulax{\@verbatim \frenchspacing\@vobeyspaces \@tabulaxverbatim You are using a type of "tabular" construct that is only allowed in AmS-LaTeX.} \let\endtabulax=\endtrivlist \@namedef{arrax*}{\@verbatim\@sarraxverbatim You are using a type of "array*" construct that is only allowed in AmS-LaTeX.} \expandafter\let\csname endarrax*\endcsname =\endtrivlist \@namedef{tabulax*}{\@verbatim\@stabulaxverbatim You are using a type of "tabular*" construct that is only allowed in AmS-LaTeX.} \expandafter\let\csname endtabulax*\endcsname =\endtrivlist \def\endequation{% \ifmmode\ifinner \iftag@ \addtocounter{equation}{-1} $\hfil \displaywidth\linewidth\@taggnum\egroup \endtrivlist \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@ignoretrue \else $\hfil \displaywidth\linewidth\@eqnnum\egroup \endtrivlist \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@ignoretrue \fi \else \iftag@ \addtocounter{equation}{-1} \eqno \hbox{\@taggnum} \global\@ifnextchar*{\@tagstar}{\@tag}@false% $$\global\@ignoretrue \else \eqno \hbox{\@eqnnum $$\global\@ignoretrue \fi \fi\fi } \newif\iftag@ \@ifnextchar*{\@tagstar}{\@tag}@false \def\@ifnextchar*{\@TCItagstar}{\@TCItag}{\@ifnextchar*{\@TCItagstar}{\@TCItag}} \def\@TCItag#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{(#1)}} \def\@TCItagstar*#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{#1}} \@ifundefined{tag}{ \def\@ifnextchar*{\@tagstar}{\@tag}{\@ifnextchar*{\@tagstar}{\@tag}} \def\@tag#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{(#1)}} \def\@tagstar*#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{#1}} }{} \makeatother \endinput
\section{Introduction} The state of a quantum system composed of $N$ subsystems is {\it separable\/} if it can be written as a classical mixture or classical {\it ensemble\/} of tensor product states. Although it is straightforward to decide whether a pure state is separable, the same question is in general unsolved for mixed states. A simple separability criterion is known only for two subsystems, and even then only if one of the subsystems is a two-level system (qubit) and the other one is at most a three-level system (qutrit) \cite{Peres1996a,Horodecki1996a} Recently it has been shown that there exists a neighborhood of the maximally mixed state in which all states are separable \cite{Zyczkowski1998}. For $N$ qubits the size of this neighborhood decreases exponentially with the number of qubits \cite{Braunstein1999a}. These results are interesting, among other reasons, because NMR quantum computers \cite{Cory1997,Gershenfeld1997,Cory1998} operate with states near the maximally mixed state. In this paper we develop a general method for constructing pure-product-state representations for $N$-qubit density operators; when the expansion coefficients in such a representation are nonnegative, it provides an explicit product ensemble for the density operator, and the state is separable. In Sec.~\ref{sec:super}, we introduce a superoperator formalism that allows us to treat different discrete and continuous representations on the same footing. In Sec.~\ref{sec:pauli}, we apply the formalism to a discrete overcomplete operator basis, leading to a simple rederivation of the separability bound found in \cite{Braunstein1999a}. Section~\ref{sec:continuous} introduces a ``canonical'' continuous representation for an arbitrary density operator, and Sec.~\ref{sec:discrete} develops a method for deriving discrete representations from the continuous representation. In Sec.~\ref{sec:spherical}, we show that the canonical continuous representation is the unique representation containing only scalar and vector spherical-harmonic components; all other representations can be obtained by adding arbitrary higher-order spherical harmonics. Section~\ref{sec:epcat} applies the canonical continuous representation to mixtures of the $N$-qubit ``cat state'' with the maximally mixed state. In Sec.~\ref{sec:GHZ}, we derive the conditions for separability of a mixture of the Greenberger-Horne-Zeilinger state \cite{Greenberger1990} with the maximally mixed state. Section~\ref{sec:discuss} concludes with a brief discussion. \section{Product-state representations and separable ensembles} \label{sec:prodrep} \subsection{Superoperator formalism} \label{sec:super} The set of linear operators acting on a $D$-dimensional Hilbert space ${\cal H}$ is a $D^2$-dimensional complex vector space ${\cal L(H)}$. Let us introduce operator ``kets'' $\Ket{A}=A$ and ``bras'' $\Bra{A}=A^\dagger$, distinguished from vector kets and bras by the use of round brackets \cite{Caves1999a}. Then the natural inner product on ${\cal L(H)}$, the trace-norm inner product, can be written as $\Inner A B={\rm tr}(A^\dagger B)$. The notation ${\cal S}=\Outer A B$ defines a superoperator ${\cal S}$ acting like \begin{equation} \label{eq:super} {\cal S}\Ket X = \Ket A \Inner B X ={\rm tr}(B^\dagger X) A \;. \end{equation} Now let the set $\{\Ket{N_j}\}$ constitute a (complete or overcomplete) operator basis; i.e., let the operator kets $\Ket{N_j}$ span the vector space ${\cal L(H)}$. It follows that the superoperator ${\cal G}$ defined by \begin{equation} \label{eq:g} {\cal G} \equiv \sum_j \Proj{N_j} \end{equation} is invertible. The operators \begin{equation} Q_j\equiv{\cal G}^{-1}\Ket{N_j} \end{equation} form a dual basis, which gives rise to the following resolutions of the superoperator identity: \begin{equation} \label{eq:resid} {\bf 1} = \sum_j \Outer{Q_j}{N_j} = \sum_j \Outer{N_j}{Q_j} \;. \end{equation} An arbitrary operator $A$ can be expanded as \begin{equation} \label{eq:expansion1} A = \sum_j \Outer{N_j}{Q_j} A) = \sum_j N_j {\rm tr}(Q_j^\dagger A) \end{equation} and \begin{equation} \label{eq:expansion2} A = \sum_j \Outer{Q_j}{N_j} A) = \sum_j Q_j {\rm tr}(N_j^\dagger A) \;. \end{equation} These expansions are unique if and only if the operators $N_j$ are linearly independent. \subsection{A discrete representation for the 6 cardinal directions} \label{sec:pauli} The Hermitian operators $P_{j\mu}$, $j=1,2,3$, $\mu=0,1$, defined by \begin{equation} \label{eq:pauli} P_{j\mu} \equiv {1\over2}\Bigl(1+(-1)^\mu\sigma_j\Bigr)\;, \end{equation} are the pure-state projectors corresponding to the 6 cardinal directions on the Bloch sphere. They form an overcomplete basis in the space of operators acting on a qubit. The superoperator \begin{equation} {\cal G} = \sum_{j,\mu} \Proj{P_{j\mu}} = {1\over2} \Biggl( 3 \Proj1 + \sum_j \Proj{\sigma_j}\Biggr) \label{eq:sixG} \end{equation} has normalized eigenoperators $\Ket1/\sqrt2$ and $\Ket{\sigma_j}/\sqrt2$, with eigenvalues 3 and 1, respectively, from which it follows that \begin{equation} {\cal G}^{-1} = {1\over2}\Biggl( {1\over3}\Proj1 + \sum_j\Proj{\sigma_j}\Biggr)\;. \end{equation} The last equation allows us to find the dual-basis operators, \begin{equation} \label{eq:paulidual} Q_{j\mu} = {\cal G}^{-1} \Ket{P_{j\mu}} = {1\over6}\Bigl(1 + 3(-1)^\mu\sigma_j\Bigr) \;. \end{equation} Any density operator $\rho={1\over2}(1+\vec S\cdot\vec\sigma)$ for one qubit can be represented in the form \begin{equation} \label{eq:rhopauli} \rho = \sum_{j,\mu} \Ket{P_{j\mu}}\Inner{Q_{j\mu}}\rho = {1\over6} \, \sum_{j,\mu} P_{j\mu}\Bigl(1+3(-1)^\mu S_j\Bigr)\;. \end{equation} For an $N$-qubit density operator, the analogous representation is \begin{equation} \label{eq:rhonpauli} \rho = \sum_{j_1,\mu_1,\ldots,j_N,\mu_N} w^{\rm can}(j_1,\mu_1,\ldots,j_N,\mu_N) P_{j_1\mu_1}\otimes\cdots\otimes P_{j_N\mu_N} \;, \end{equation} where \begin{eqnarray} w^{\rm can}(j_1,\mu_1,\ldots,j_N,\mu_N) &\equiv& {\rm tr}(\rho\,Q_{j_1\mu_1}\otimes\cdots\otimes Q_{j_N\mu_N}) \nonumber \\ &=& {1\over6^N}{\rm tr}\biggl(\rho \Bigl(1+3(-1)^{\mu_1}\sigma_{j_1}\Bigr) \otimes\cdots\otimes \Bigl(1+3(-1)^{\mu_N}\sigma_{j_N}\Bigr)\biggr)\;. \nonumber \\ \label{eq:wnpauli} \end{eqnarray} This representation exists for arbitrary density operators, entangled or separable. If the coefficients $w^{\rm can}(j_1,\mu_1,\ldots,j_N,\mu_N)$ are nonnegative, the density operator $\rho$ is separable, and the representation (\ref{eq:rhonpauli}) provides an explicit product-state ensemble for the density operator. The expansions~(\ref{eq:rhopauli}) and (\ref{eq:rhonpauli}) are not unique because the 6 operators $P_{j\mu}$ are not linearly independent. Throughout the paper we use the superscript ``can,'' standing for {\it canonical}, to denote the natural, but generally not unique expansion coefficients that come from using the procedure of Sec.~\ref{sec:super}. For any $N$-qubit density operator $\rho$, the coefficients~(\ref{eq:wnpauli}) obey the bound \begin{equation} \label{eq:bound1} w^{\rm can}(j_1,\mu_1,\ldots,j_N,\mu_N) \ge \pmatrix{\hbox{smallest eigenvalue of}\cr Q_{j_1\mu_1}\otimes\cdots\otimes Q_{j_N\mu_N}} = -{2^{2N-1}\over6^N} \;. \end{equation} This follows from the fact that $Q_{j\mu}$ has eigenvalues $2/3$ and $-1/3$. The most negative eigenvalue of the product operator $Q_{j_1\mu_1}\otimes\cdots\otimes Q_{j_N\mu_N}$ is therefore $(-1/3)(2/3)^{N-1}=-2^{2N-1}/6^N$, from which Eq.~(\ref{eq:bound1}) follows. Consider now an $N$-qubit density operator that is a mixture of the maximally mixed density operator, $1/2^N$, and an arbitrary density operator $\rho_1$: \begin{equation} \rho_\epsilon = {1-\epsilon\over2^N}1 + \epsilon\rho_1 \;. \label{eq:rhoeps} \end{equation} Any density operator can be written in this form \cite{Braunstein1999a}. The coefficients~(\ref{eq:wnpauli}) for $\rho_\epsilon$ are given by \begin{eqnarray} w^{\rm can}_\epsilon(j_1,\mu_1,\ldots,j_N,\mu_N) &=& {1-\epsilon\over6^N} + \epsilon w^{\rm can}_1(j_1,\mu_1,\ldots,j_N,\mu_N) \nonumber\\ &\ge& {1-\epsilon(1 + 2^{2N-1})\over6^N} \;, \end{eqnarray} which is nonnegative for \begin{equation} \label{eq:bound2} \epsilon \le {1\over1+2^{2N-1}} \;. \end{equation} This bound, derived in \cite{Braunstein1999a} using a different route, implies that, for $\epsilon \le {1/(1+2^{2N-1})}$, all density operators of the form (\ref{eq:rhoeps}) are separable. It sets a lower bound on the size of the separable neighborhood surrounding the maximally mixed state. \subsection{Canonical continuous representation} \label{sec:continuous} The set of all pure-state projectors on the Bloch sphere, \begin{equation} P_{\vec n} \equiv |\vec n\rangle\langle\vec n| = {1\over2}(1+\vec\sigma\cdot\vec n) \;, \end{equation} forms an overcomplete basis. The corresponding superoperator \begin{equation} \label{eq:contG} {\cal G} = \int d\Omega\,\Proj{P_{\vec n}} = \pi \Biggl( \Proj1 + {1\over3} \sum_j \Proj{\sigma_j}\Biggr) \end{equation} is proportional to the superoperator of Eq.~(\ref{eq:sixG}). The normalized eigenoperators of ${\cal G}$ are $\Ket1/\sqrt2$ and $\Ket{\sigma_j}/\sqrt2$, with eigenvalues $2\pi$ and $2\pi/3$, respectively. Thus ${\cal G}^{-1}$ can be expressed as \begin{equation} {\cal G}^{-1} = {1\over4\pi}\!\left( \Proj1 + 3\sum_j\Proj{\sigma_j} \right) \;. \end{equation} The dual-basis operators are then given by \begin{equation} \label{eq:contdual} Q_{\vec n} = {\cal G}^{-1} \Ket{P_{\vec n}} = {1\over4\pi} (1+3\vec\sigma\cdot\vec n) \;. \end{equation} Any density operator $\rho={1\over2}(1+\vec S\cdot\vec\sigma)$ for one qubit can be represented in the form \begin{equation} \label{eq:rhocont} \rho = \int d\Omega\, \Ket{P_{\vec n}}\Inner{Q_{\vec n}}\rho = {1\over4\pi}\int d\Omega\,P_{\vec n}(1+3\vec S\cdot\vec n) \;. \end{equation} For $N$ qubits, we define the pure-product-state projector \begin{equation} \label{eq:prodproj} P(\tilde n)\equiv P_{\vec n_1}\otimes\cdots\otimes P_{\vec n_N} = {1\over2^N}(1+\vec\sigma\cdot\vec n_1)\otimes\cdots \otimes(1+\vec\sigma\cdot\vec n_N) \;, \end{equation} where $\tilde n$ stands for the collection of unit vectors $\vec n_1,\ldots,\vec n_N$. Any $N$-qubit density operator can be expanded as \begin{equation} \rho = \int d\Omega_{\tilde n}\, w^{\rm can}(\tilde n) P(\tilde n) \equiv\int d\Omega_{n_1}\cdots d\Omega_{n_N}w^{\rm can}(\tilde n) P(\tilde n) \;, \label{eq:rho} \end{equation} where \begin{equation} w^{\rm can}(\tilde n) \equiv {\rm tr}\Bigl(\rho Q(\tilde n)\Bigr) \label{eq:wn} \end{equation} is the canonical expansion function, with \begin{equation} Q(\tilde n) \equiv Q_{\vec n_1}\otimes\cdots\otimes Q_{\vec n_N} = {1\over(4\pi)^N} (1+3\vec\sigma\cdot\vec n_1)\otimes\cdots \otimes(1+3\vec\sigma\cdot\vec n_N) \;. \end{equation} If the coefficients $w^{\rm can}(\tilde n)$ are nonnegative, the density operator $\rho$ is separable, and the representation (\ref{eq:rho}) provides an explicit product-state ensemble for the density operator. Using the same argument as in the preceding subsection, one can show that for any density operator $\rho$ \cite{Braunstein1999a}, \begin{equation} \label{eq:bound3} w^{\rm can}(\tilde n) \ge \pmatrix{\hbox{smallest eigenvalue}\cr \hbox{of $Q(\tilde n)$}} = -{2^{2N-1}\over(4\pi)^N} \;. \end{equation} Applying this result to density operators of the form~(\ref{eq:rhoeps}), one arrives again at the lower bound~(\ref{eq:bound2}) on the size of the separable neighborhood of the maximally mixed density operator \cite{Braunstein1999a}. \subsection{Other discrete representations} \label{sec:discrete} The expansion coefficients~(\ref{eq:wnpauli}) for the discrete representation of Sec.~\ref{sec:pauli} can be obtained by evaluating the expansion coefficients~(\ref{eq:wn}) for the canonical continuous representation along the cardinal directions for each qubit and then renormalizing the resulting function. This is no accident. It is easy to characterize a class of discrete representations whose expansion coefficients are similarly related to the continuous representation. Consider a discrete set of projection operators, \begin{equation} P_{\vec n_\alpha}=|\vec n_\alpha\rangle\langle\vec n_\alpha|= {1\over2}(1+\vec\sigma\cdot\vec n_\alpha)\;,\;\;\; \alpha=1,\ldots,K. \end{equation} The corresponding superoperator \begin{eqnarray} {\cal G} &=& \sum_{\alpha=1}^K \Proj{P_{\vec n_\alpha}} \nonumber \\ &=& {1\over4}\!\left( K\Proj 1 + \sum_\alpha\Bigl(\vec n_\alpha\cdot\Outer{\vec\sigma}{1} +\Outer{1}{\vec\sigma}\cdot\vec n_\alpha\Bigr) +\sum_{j,k}\Outer{\sigma_j}{\sigma_k} \sum_\alpha(n_\alpha)_j(n_\alpha)_k \right) \label{eq:discreteG} \end{eqnarray} generates dual-basis operators and expansion coefficients proportional to those for the continuous representation if and only if ${\cal G}$ is proportional to the superoperator~(\ref{eq:contG}) for the continuous representation, i.e., if and only if \begin{eqnarray} 0&=&\sum_\alpha\vec n_\alpha \;, \label{eq:cond1} \\ {1\over3} \, \delta_{jk}&=&{1\over K}\sum_\alpha(n_\alpha)_j(n_\alpha)_k \;. \label{eq:cond2} \end{eqnarray} The trace of Eq.~(\ref{eq:cond2}) is satisfied by any collection of unit vectors. When these conditions are satisfied, the superoperator~(\ref{eq:discreteG}) simplifies to \begin{equation} {\cal G} = {K\over4} \Biggl( \Proj1 + {1\over3} \sum_j \Proj{\sigma_j}\Biggr) \;, \end{equation} with an inverse \begin{equation} {\cal G}^{-1}= {1\over K} \Biggl( \Proj1 + 3 \sum_j \Proj{\sigma_j}\Biggr) \end{equation} that generates dual-basis operators \begin{equation} Q_{\vec n_\alpha} = {\cal G}^{-1}\Ket{P_{\vec n_\alpha}} = {1\over K}(1 + 3\vec\sigma\cdot\vec n_\alpha) \;. \end{equation} Any density operator $\rho={1\over2}(1+\vec S\cdot\vec\sigma)$ for one qubit can be represented in the form \begin{equation} \label{eq:rhodiscrete} \rho = \sum_\alpha \Ket{P_\alpha}\Inner{Q_\alpha}\rho = {1\over K} \, \sum_{\alpha} P_{\alpha}(1+3\vec S\cdot\vec n_\alpha)\;. \end{equation} The conditions~(\ref{eq:cond1}) and (\ref{eq:cond2}) are satisfied by unit vectors that point to the vertices of any regular polyhedron inscribed within the Bloch sphere---i.e., a tetrahedron, octahedron, cube, icosahedron, or dodecahedron. The 4 projectors for a tetrahedron are linearly independent, making the corresponding tetrahedral representation~(\ref{eq:rhodiscrete}) unique. An octahedron gives rise to the 6-cardinal-direction representation of Sec.~\ref{sec:pauli}. The regular polyhedra by no means exhaust the possibilities for representations of this sort. One can use simultaneously the vertices from any number of polyhedra. Moreover, one can produce an appropriate set of unit vectors by the following procedure: starting with any set of unit vectors in the first octant, reflect the set successively through the three Cartesian planes to produce a set of vectors that occupies all eight octants. For more than one qubit, the possibilities for discrete representations that follow from the continuous representation are even more diverse. The simplest possibility is to use the same set of pure-state projectors for each qubit, but this is not necessary. One can use a different set of projectors for each qubit, or more generally, one can use a ``tree structure,'' in which there is a different set of projectors for the $n$th qubit for each projector for the first $n-1$ qubits. There are possibilities more general even than those with a tree structure, analogous to the ``domino states'' introduced for two qutrits in \cite{Bennett1999a}. We do not dwell here on the myriad of discrete representations of this sort, merely noting that since the expansion coefficients in such representations follow from the expansion coefficients for the continuous representation, all such representations lead to the same general lower bound~(\ref{eq:bound2}) on the size of the separable neighborhood surrounding the maximally mixed state. Both the continuous and these discrete representations can, however, give better bounds on the separability of {\it particular\/} states of the form~(\ref{eq:rhoeps}), a point to which we return in Sec.~\ref{sec:epcat}. \section{Spherical-harmonic expansions} \label{sec:spherical} In this section we explore the relationship between product-state representations and sphe\-ri\-cal-harmonic expansions. \subsection{Spherical-harmonic expansions and the Pauli representation} \label{sec:sphpauli} We begin by noting that if $\vec n$ is a unit vector, then \begin{eqnarray} \vec\sigma\cdot\vec n & = & \sigma_3 \cos\theta + \sigma_1\sin\theta\cos\phi + \sigma_2\sin\theta\sin\phi \nonumber\\ & = & \sqrt{4\pi\over3}\Bigl(\sigma_3{Y_1^0}^\ast + \sigma_1{1\over\sqrt2}(-{Y_1^{+1}}^\ast+{Y_1^{-1}}^\ast) - i\sigma_2{1\over\sqrt2}({Y_1^{+1}}^\ast+{Y_1^{-1}}^\ast) \Bigr) \\ & = & \sigma_1^0{Y_1^0}^\ast+ \sigma_1^{+1}{Y_1^{+1}}^\ast+\sigma_1^{-1}{Y_1^{-1}}^\ast \;, \nonumber \end{eqnarray} where \begin{equation} Y_0^0={1\over\sqrt{4\pi}}\;,\;\;\; Y_1^0=\sqrt{3\over4\pi}\cos\theta\;,\;\;\; Y_1^{\pm1}=\mp\sqrt{3\over8\pi}\sin\theta\,e^{\pm i\phi} \end{equation} are spherical harmonics and, in the last line, we define \begin{equation} \sigma_0^0\equiv\sqrt{4\pi}\,1\;,\;\;\; \sigma_1^0\equiv\sqrt{4\pi\over3}\sigma_3\;,\;\;\; \sigma_1^{\pm1}\equiv\mp\sqrt{2\pi\over3}(\sigma_1\pm i\sigma_2)\;. \end{equation} We can then write any pure-state projector on the Bloch sphere in the form \begin{equation} P_{\vec n} = \proj{\vec n} = {1\over2}(1+\vec\sigma\cdot\vec n) = {1\over2}\, \sum_{l=0}^1\sum_{m=-l}^l \sigma_l^mY_l^{m*} \;. \end{equation} Any $N$-qubit density operator can be written in the form \begin{equation} \label{eq:rhow} \rho = \int d\Omega_{\tilde n}\, w(\tilde n) P(\tilde n) \;, \end{equation} where the expansion coefficients $w(\tilde n)\equiv w(\vec n_1,\ldots,\vec n_N)$ are not necessarily nonnegative [the expansion coefficients $w(\tilde n)$ need not be those of the canonical representation~(\ref{eq:wn})]. A density operator $\rho$ is separable if and only if it can be written in the form (\ref{eq:rhow}) for some nonnegative probability density $w(\tilde n)$. Now expand the product projector~(\ref{eq:prodproj}) in terms of spherical harmonics, \begin{equation} P(\tilde n) = {1\over2^N}\, \sum_{\{l_j\le1,m_j\}} Y_{l_1}^{m_1*}\cdots Y_{l_N}^{m_N*} \sigma_{l_1}^{m_1}\otimes\cdots\otimes\sigma_{l_N}^{m_N} \;, \label{eq:projexpand} \end{equation} where \begin{equation} \sum_{\{l_j\le1,m_j\}} \equiv \; \sum_{l_1=0}^1\sum_{m_1=-l_1}^{l_1} \cdots \sum_{l_N=0}^1\sum_{m_N=-l_N}^{l_N} \end{equation} denotes a sum that contains only scalar and vector terms ($l=0,1$), and also expand the expansion coefficients, \begin{equation} w(\tilde n) = \sum_{\{l_j,m_j\}} a_{l_1\ldots l_N}^{m_1\ldots m_N}Y_{l_1}^{m_1}\cdots Y_{l_N}^{m_N} \;, \label{eq:wexpand} \end{equation} where \begin{equation} \sum_{\{l_j,m_j\}} \equiv \; \sum_{l_1=0}^\infty\sum_{m_1=-l_1}^{l_1} \cdots \sum_{l_N=0}^\infty\sum_{m_N=-l_N}^{l_N} \end{equation} denotes an unrestricted sum over spherical harmonics. Inserting Eqs.~(\ref{eq:projexpand}) and (\ref{eq:wexpand}) into Eq.~(\ref{eq:rhow}) and using the orthonormality of the spherical harmonics yields the (unique) Pauli-operator representation of $\rho$: \begin{equation} \rho= {1\over2^N} \sum_{\{l_j\le1,m_j\}} a_{l_1\ldots l_N}^{m_1\ldots m_N} \sigma_{l_1}^{m_1}\otimes\cdots\otimes\sigma_{l_N}^{m_N} \;. \label{eq:rho2} \end{equation} We see that $\rho$ depends on and determines only the scalar and vector parts of $a_{l_1\ldots l_N}^{m_1\ldots m_N}$; the higher-order-spherical-harmonic (HOSH) content of $w(\tilde n)$ corresponds to the freedom in constructing pure-product-state representations of $\rho$. It is straightforward to generate all pure-product-state representations: given the Pauli-operator representation~(\ref{eq:rho2}) of $\rho$, any $w(\tilde n)$ is given by Eq.~(\ref{eq:wexpand}), where the coefficients $a_{l_1\ldots l_N}^{m_1\ldots m_N}$ are arbitrary except for $l_1,\ldots,l_N\le1$. The canonical continuous representation $w^{\rm can}(\tilde n)$ of Eq.~(\ref{eq:wn}) is the unique pure-product-state representation given by the $l=0$ and $l=1$ components. We can see this directly in the following way. Write the Pauli representation of $\rho$ in the form \begin{equation} \rho= {1\over2^N } \sum_{\alpha_1,\ldots,\alpha_N}c_{\alpha_1\ldots\alpha_N}\sigma_{\alpha_1}\otimes\cdots\otimes\sigma_{\alpha_N} \;, \label{eq:pauli2} \end{equation} where \begin{equation} \label{eq:paulic} c_{\alpha_1\ldots\alpha_N}={\rm tr}(\rho\sigma_{\alpha_1}\otimes\cdots\otimes\sigma_{\alpha_N})=\langle\sigma_{\alpha_1}\otimes\cdots\otimes\sigma_{\alpha_N}\rangle \;. \end{equation} Here $\sigma_0=1$, and the sums over Greek indices run from 0 to 3. Now we have that \begin{eqnarray} w^{\rm can}(\tilde n)&=& {1\over(4\pi)^N} \Bigl\langle (1+3\vec\sigma\cdot\vec n_1)\otimes\cdots\otimes(1+3\vec\sigma\cdot\vec n_N) \Bigr\rangle \nonumber \\ \label{eq:probability} &=& \left({3\over4\pi}\right)^N \sum_{\alpha_1,\ldots,\alpha_N} \Bigl\langle (n_1)_{\alpha_1}\sigma_{\alpha_1} \otimes\cdots\otimes (n_N)_{\alpha_N}\sigma_{\alpha_N} \Bigr\rangle \\ &=& \left({3\over4\pi}\right)^N \sum_{\alpha_1,\ldots,\alpha_N} c_{\alpha_1\ldots\alpha_N}(n_1)_{\alpha_1}\cdots (n_N)_{\alpha_N} \;, \nonumber \end{eqnarray} where $(n_j)_{\alpha_j}\equiv1/3$ if $\alpha_j=0$ and $(n_j)_{\alpha_j}$ is a Cartesian component of $\vec n_j$ if $\alpha_j=1,2$, or 3. Thus any pure-product-state representation of a density operator $\rho$ can be obtained by adding HOSH to the canonical continuous representation $w^{\rm can}(\tilde n)$. The discrete representations discussed in Sec.~\ref{sec:discrete} are examples---though by no means the only examples---of representations with HOSH content. The continuous canonical representation can be generated directly from the Pauli representation of $\rho$ by making the substitutions \begin{equation} \label{eq:rule1} 2^N\rho\longrightarrow w^{\rm can}(\tilde n) \;, \end{equation} \begin{equation} \label{eq:rule2} 1\longrightarrow{1\over4\pi}\;,\;\;\; \sigma_j\longrightarrow{3\over4\pi}n_j \;. \end{equation} These rules are equivalent to $\sigma_l^m\longrightarrow Y_l^m$ and to $\sigma_\alpha\longrightarrow(3/4\pi)n_\alpha$. \subsection{The $\epsilon$-cat state} \label{sec:epcat} As an example, consider the following $N$-qubit generalization of the Werner \cite{Werner1989a} state, \begin{equation} \rho_\epsilon = {1-\epsilon\over2^N}1 + \epsilon\proj{\psi_{\rm cat}} \;, \label{eq:epcat} \end{equation} where \begin{equation} \label{eq:cat} |\psi_{\rm cat}\rangle \equiv {1\over\sqrt2}\Bigl(\ket{0\ldots0}+\ket{1\ldots1}\Bigr) \end{equation} is the $N$-qubit ``cat state.'' We call the mixed state~(\ref{eq:epcat}) the {\it $\epsilon$-cat state}. Now let $\ket0$and $\ket1$ be the eigenstates of $\sigma_3$, with $\ket0$ at the top of the Bloch sphere and $\ket1$ at the bottom, so that $\sigma_1=\outer 0 1 +\outer 1 0$, $\sigma_2=-i\outer 0 1 +i\outer 1 0$, and $\sigma_3=\proj 0 -\proj 1$. It follows that the Pauli representation of the $\epsilon$-cat state is \begin{eqnarray} \rho_\epsilon &=& {1\over2^N}\biggl( (1-\epsilon)1 \nonumber \\ &\mbox{}&\hphantom{{1\over2^N}} +{\epsilon\over2}(1+\sigma_3)\otimes\cdots\otimes(1+\sigma_3) +{\epsilon\over2}(1-\sigma_3)\otimes\cdots\otimes(1-\sigma_3) \label{eq:pauliepcat} \\ &\mbox{}&\hphantom{{1\over2^N}} +{\epsilon\over2} (\sigma_1+i\sigma_2)\otimes\cdots\otimes(\sigma_1+i\sigma_2) +{\epsilon\over2} (\sigma_1-i\sigma_2)\otimes\cdots\otimes(\sigma_1-i\sigma_2) \biggr) \;. \nonumber \end{eqnarray} Applying the rules~(\ref{eq:rule1}) and (\ref{eq:rule2}) yields the continuous canonical representation of $\rho_\epsilon$: \begin{eqnarray} w^{\rm can}_\epsilon(\tilde n) &=& {1\over(4\pi)^N} \biggl( (1-\epsilon) \nonumber \\ &\mbox{}&\hphantom{{1\over(4\pi)^N}} +{\epsilon\over2}(1+3\cos\theta_1)\cdots(1+3\cos\theta_N) +{\epsilon\over2}(1-3\cos\theta_1)\cdots(1-3\cos\theta_N) \nonumber \\ &\mbox{}&\hphantom{{1\over(4\pi)^N}} +\epsilon\,3^N \sin\theta_1\cdots\sin\theta_N\cos(\phi_1+\ldots+\phi_N) \biggr) \;. \label{eq:wcat} \end{eqnarray} The minimum value of $w^{\rm can}_\epsilon(\tilde n)$ occurs either (i)~when one qubit is evaluated at the north (south) pole, and the other $N-1$ qubits are evaluated at the south (north) pole, or (ii)~when all the qubits are evaluated on the equator with $\cos(\phi_1+\ldots+\phi_N)=-1$. For $N=2$ and $N=4$, both the poles and the equator give the same minimum value, which shows that the $\epsilon$-cat state is separable for $\epsilon\le1/9$ for $N=2$ and for $\epsilon\le1/81$ for $N=4$. For $N=3$ and $N=5$, the equator gives the minimum value, which shows that the $\epsilon$-cat state is separable for $\epsilon\le1/27$ for $N=3$ \cite{Braunstein1999a} and for $\epsilon\le1/243$ for $N=5$. For $N\ge6$, the poles give the minimum value, which shows that $\rho_\epsilon$ is separable for \begin{equation} \epsilon\le{1\over1\pm 2^N+2^{2N-2}} \;, \label{eq:sepNcat} \end{equation} the upper (lower) sign applying when $N$ is even (odd). Notice that except for $N=2$, all these bounds are better than the general-purpose bound of Eq.~(\ref{eq:bound2}). These bounds on the separability of the $\epsilon$-cat state can be improved by using tetrahedral ensembles of the sort discussed in Sec.~\ref{sec:discrete}. The tetrahedral ensembles are chosen to have a tree structure such that the vertices avoid the places where $w^{\rm can}_\epsilon(\tilde n)$ has its smallest values. We do not discuss these tetrahedral bounds here, preferring instead to consider in detail the separability of the $\epsilon$-GHZ state. \section{The $\epsilon$-GHZ state} \label{sec:GHZ} In this section we consider the 3-qubit version of the $\epsilon$-cat state introduced in Sec.~\ref{sec:epcat}. In the case of 3 qubits, the cat state of Eq.~(\ref{eq:cat}) is called the Greenberger-Horne-Zeilinger (GHZ) state \cite{Greenberger1990}, \begin{equation} |\psi_{\rm GHZ}\rangle = {1\over\sqrt2}\Bigl(\ket{000}+\ket{111}\Bigr) \;; \end{equation} we call the mixture $\rho_\epsilon$ of Eq.~(\ref{eq:epcat}) the {\it $\epsilon$-GHZ state}. We show in this section that {\it the $\epsilon$-GHZ state is separable if and only if $\epsilon\le1/5$}. The class of $\epsilon$-GHZ states has been considered previously in \cite{Murao1998}. \subsection{Separability of the Werner state} \label{sec:werner} Before proceeding to the proof of the separability of the $\epsilon$-GHZ state, we review a similar, but simpler proof for the Werner state \cite{Werner1989a}, the 2-qubit version of the $\epsilon$-cat state: \begin{eqnarray} \rho_W &=& {1-\epsilon\over4}1\otimes1+ {\epsilon\over2} \Bigl(|00\rangle+|11\rangle\Bigr)\Bigl(\langle00|+\langle11|\Bigr) \nonumber \\ &=& {1\over4}\Bigl(1\otimes1+\epsilon\, (\sigma_3\otimes\sigma_3+\sigma_1\otimes\sigma_1-\sigma_2\otimes\sigma_2) \Bigr)\;. \label{eq:rhowerner} \end{eqnarray} The Werner state is separable if and only if $\epsilon\le1/3$ \cite{Bennett1996a,RHorodecki1996a}, a result we now show in the following way. A density operator $\rho$ for 2 qubits, $A$ and $B$, is separable if and only if it can be written as \begin{equation} \rho=\intd\Omega_{\n_A}d\Omega_{\n_B}\,w(\n_A,\n_B)P_{\n_A}\otimes P_{\n_B} \;, \end{equation} with $w(\n_A,\n_B)$ being a nonnegative probability distribution. If a state is separable, the coefficients~(\ref{eq:paulic}) in the Pauli representation~(\ref{eq:pauli2}) of $\rho$ become classical expectations over the distribution $w(\n_A,\n_B)$, expectations that we denote by $E[\cdot]$: \begin{equation} c_{\alpha\beta}= {\rm tr}(\rho\,\sigma_\alpha\otimes\sigma_\beta) =\intd\Omega_{\n_A}d\Omega_{\n_B}\,(n_A)_\alpha(n_B)_\beta\,w(\n_A,\n_B)\equiv E[(n_A)_\alpha(n_B)_\beta] \;. \end{equation} For the Werner state, the coefficients $c_{\alpha\beta}$ can be read off the Pauli representation in Eq.~(\ref{eq:rhowerner}): \begin{eqnarray} &\mbox{}&E[(n_A)_j]=E[(n_B)_j]=0 \;, \\ &\mbox{}&E[(n_A)_j(n_B)_k]=0 \;,\;\;\mbox{for $j\ne k$,} \\ \label{eq:diag} &\mbox{}&E[(n_A)_1(n_B)_1]=-E[(n_A)_2(n_B)_2]=E[(n_A)_3(n_B)_3]=\epsilon \;. \end{eqnarray} If $w(\n_A,\n_B)$ is a probability distribution, we have that \begin{equation} \epsilon=\Bigl|E[(n_A)_j(n_B)_j]\Bigr|\le {1\over2}E[(n_A)_j^2+(n_B)_j^2] \;. \end{equation} Adding over the three values of $j$ gives \begin{equation} 3\epsilon \le {1\over2} E[(n_A)_1^2+(n_A)_2^2+(n_A)_3^2+(n_B)_1^2+(n_B)_2^2+(n_B)_3^2]=1 \;, \end{equation} which shows that if $\epsilon>1/3$, the Werner state is nonseparable. We still have the task of designing a product ensemble that gives the Werner state for $\epsilon\le1/3$. Defining three density operators \begin{eqnarray} \rho_1 &\equiv& {1\over2}\bigl( P_{\vec e_3}\otimes P_{\vec e_3}+P_{-\vec e_3}\otimes P_{-\vec e_3} \bigr) ={1\over4}\bigl( 1\otimes1+\sigma_3\otimes\sigma_3 \bigr) \;, \\ \rho_2 &\equiv& {1\over2}\bigl( P_{\vec e_1}\otimes P_{\vec e_1}+P_{-\vec e_1}\otimes P_{-\vec e_1} \bigr) ={1\over4} \bigl(1\otimes1+\sigma_1\otimes\sigma_1 \bigr) \;, \\ \rho_3 &\equiv& {1\over2}\bigl( P_{\vec e_2}\otimes P_{-\vec e_2}+P_{-\vec e_2}\otimes P_{\vec e_2} \bigr) ={1\over4}\bigl( 1\otimes1-\sigma_2\otimes\sigma_2 \bigr) \;, \end{eqnarray} we see that each takes care of one of the correlations in Eq.~(\ref{eq:diag}). An equal mixture gives the $\epsilon=1/3$ Werner state: \begin{equation} {1\over3}(\rho_1+\rho_2+\rho_3)={1\over4}\left( 1\otimes1+{1\over3}\bigl( \sigma_3\otimes\sigma_3+\sigma_1\otimes\sigma_1-\sigma_2\otimes\sigma_2 \bigl)\right) \;. \end{equation} Smaller values of $\epsilon$ can be handled by adding on a product ensemble for the maximally mixed state. \subsection{Separability of the $\epsilon$-GHZ state} \label{sec:epGHZ} Consider now the $\epsilon$-GHZ state \begin{eqnarray} \rho_{\epsilon{\rm GHZ}} &=& {1-\epsilon\over8}1\otimes1\otimes1+ {\epsilon\over2} \Bigl(|000\rangle+|111\rangle\Bigr)\Bigl(\langle000|+\langle111|\Bigr) \nonumber \\ \label{eq:rhoepGHZ} &=& {1\over8}\biggl(1\otimes1\otimes1+ \epsilon\Bigl( 1\otimes\sigma_3\otimes\sigma_3 +\sigma_3\otimes1\otimes\sigma_3 +\sigma_3\otimes\sigma_3\otimes1 \\ &\mbox{}&\hphantom{{1\over8}\Bigl(1} +\sigma_1\otimes\sigma_1\otimes\sigma_1 -\sigma_1\otimes\sigma_2\otimes\sigma_2 -\sigma_2\otimes\sigma_1\otimes\sigma_2 -\sigma_2\otimes\sigma_2\otimes\sigma_1 \Bigr)\biggr) \;. \nonumber \end{eqnarray} A density operator $\rho$ for 3 qubits, $A$, $B$, and $C$, is separable if and only if it can be written as \begin{equation} \rho=\intd\Omega_{\n_A}d\Omega_{\n_B}d\Omega_{\n_C}\,w(\n_A,\n_B,\n_C)P_{\n_A}\otimes P_{\n_B}\otimes P_{\n_C} \;, \end{equation} with $w(\n_A,\n_B,\n_C)$ being a nonnegative probability distribution. If a state is separable, the coefficients~(\ref{eq:paulic}) in the Pauli representation~(\ref{eq:pauli2}) of $\rho$ become classical expectations over the distribution $w(\n_A,\n_B,\n_C)$: \begin{equation} c_{\alpha\beta\gamma} = {\rm tr}(\rho\,\sigma_\alpha\otimes\sigma_\beta\otimes\sigma_\gamma) = E[(n_A)_\alpha(n_B)_\beta(n_C)_\gamma] \;. \end{equation} For the $\epsilon$-GHZ state, the coefficients $c_{\alpha\beta\gamma}$ can be read off the Pauli representation in Eq.~(\ref{eq:rhoepGHZ}): \begin{eqnarray} &\mbox{}&E[(n_A)_j]=E[(n_B)_j]=E[(n_C)_j]=0 \;, \\ \label{eq:corr3} &\mbox{}&E[(n_A)_j(n_B)_k]=E[(n_A)_j(n_C)_k]=E[(n_B)_j(n_C)_k] = \epsilon\,\delta_{j3}\delta_{k3} \;, \\ \label{eq:corr12} &\mbox{}&E[(n_A)_j(n_B)_k(n_C)_l]=\epsilon\,( \delta_{j1}\delta_{k1}\delta_{l1}-\delta_{j1}\delta_{k2}\delta_{l2} -\delta_{j2}\delta_{k1}\delta_{l2}-\delta_{j2}\delta_{k2}\delta_{l1}) \;. \end{eqnarray} Now define a ``vector'' \begin{equation} \vec N\equiv \Bigl((n_B)_1(n_C)_1-(n_B)_2(n_C)_2\Bigr)\vec e_1 -\Bigl((n_B)_1(n_C)_2+(n_B)_2(n_C)_1\Bigr)\vec e_2 +(n_B)_3\vec e_3 \;, \end{equation} whose magnitude satisfies \begin{eqnarray} \vec N\cdot\vec N &=& \Bigl((n_B)_1(n_C)_1-(n_B)_2(n_C)_2\Bigr)^2+ \Bigl((n_B)_1(n_C)_2+(n_B)_2(n_C)_1\Bigr)^2+(n_B)_3^2 \nonumber \\ &=& \Bigl((n_B)_1^2+(n_B)_2^2\Bigr)\Bigl((n_C)_1^2+(n_C)_2^2\Bigr) +(n_B)_3^2 \\ &\le& \vec n_B\cdot\vec n_B=1 \;. \nonumber \end{eqnarray} If $w(\n_A,\n_B,\n_C)$ is a probability distribution, we have that \begin{equation} 5\epsilon= \Bigl|E\Bigl[\vec n_A\cdot\vec N\,\Bigr]\Bigr| \le E\Bigl[\Bigl|\vec n_A\cdot\vec N\Bigr|\Bigr] \le {1\over2}E\Bigl[\vec n_A\cdot\vec n_A+\vec N\cdot\vec N\,\Bigr] \le 1 \;, \end{equation} which shows that if $\epsilon>1/5$, the $\epsilon$-GHZ state is nonseparable. To finish the proof, we need to construct a product ensemble for the $\epsilon$-GHZ state when $\epsilon\le1/5$. We begin by defining five density operators \begin{eqnarray} \rho_1 &\equiv& {1\over2}\Bigl( \Pe{}{3}{}{3}{}{3}+\Pe{-}{3}{-}{3}{-}{3} \Bigr) \nonumber \\ &=& {1\over8}\Bigl( 1\otimes1\otimes1+1\otimes\sigma_3\otimes\sigma_3 +\sigma_3\otimes1\otimes\sigma_3+\sigma_3\otimes\sigma_3\otimes1 \Bigr) \;, \\ \rho_2 &\equiv& {1\over4}\Bigl( \Pe{}{1}{}{1}{}{1}+ \Pe{}{1}{-}{1}{-}{1}+ \Pe{-}{1}{}{1}{-}{1}+ \Pe{-}{1}{-}{1}{}{1} \Bigr) \nonumber \\ &=&{1\over8}\bigl( 1\otimes1\otimes1+\sigma_1\otimes\sigma_1\otimes\sigma_1 \Bigr) \;, \\ \rho_3 &\equiv& {1\over4}\Bigl( \Pe{}{1}{}{2}{-}{2}+ \Pe{}{1}{-}{2}{}{2}+ \Pe{-}{1}{}{2}{}{2}+ \Pe{-}{1}{-}{2}{-}{2} \Bigr) \nonumber \\ &=& {1\over8}\bigl( 1\otimes1\otimes1-\sigma_1\otimes\sigma_2\otimes\sigma_2 \Bigr) \;, \\ \rho_4 &\equiv& {1\over4}\Bigl( \Pe{}{2}{}{1}{-}{2}+ \Pe{-}{2}{}{1}{}{2}+ \Pe{}{2}{-}{1}{}{2}+ \Pe{-}{2}{-}{1}{-}{2} \Bigr) \nonumber \\ &=& {1\over8}\Bigl( 1\otimes1\otimes1-\sigma_2\otimes\sigma_1\otimes\sigma_2 \Bigr) \;, \\ \rho_5 &\equiv& {1\over4}\Bigl( \Pe{}{2}{-}{2}{}{1}+ \Pe{-}{2}{}{2}{}{1}+ \Pe{}{2}{}{2}{-}{1}+ \Pe{-}{2}{-}{2}{-}{1} \Bigl) \nonumber \\ &=& {1\over8}\Bigl( 1\otimes1\otimes1- \sigma_2\otimes\sigma_2\otimes\sigma_1 \Bigr) \;, \end{eqnarray} each of which takes care of one of the nonzero correlations in Eqs.~(\ref{eq:corr3}) and (\ref{eq:corr12}). An equal mixture gives the ($\epsilon=1/5$)-GHZ state: \begin{eqnarray} &\mbox{}&{1\over5}(\rho_1+\rho_2+\rho_3+\rho_4+\rho_5) \nonumber \\ &\mbox{}&\hphantom{{1\over5}(\rho} ={1\over8}\biggl(1\otimes1\otimes1+ {1\over5}\Bigl(1\otimes\sigma_3\otimes\sigma_3 +\sigma_3\otimes1\otimes\sigma_3+\sigma_3\otimes\sigma_3\otimes1 \\ &\mbox{}&\hphantom{{1\over5}(\rho={1\over8}\biggl(1} +\sigma_1\otimes\sigma_1\otimes\sigma_1 -\sigma_1\otimes\sigma_2\otimes\sigma_2 -\sigma_2\otimes\sigma_1\otimes\sigma_2 -\sigma_2\otimes\sigma_2\otimes\sigma_1 \Bigr)\biggr) \;, \nonumber \end{eqnarray} thus completing the proof. For both the Werner state and the $\epsilon$-GHZ states, we are able to construct an optimal ensemble based on the 6 cardinal directions, but notice that this ensemble is not the canonical representation of Sec.~\ref{sec:pauli}. \section{Conclusion} \label{sec:discuss} In this paper we introduce a canonical procedure for constructing representations of density operators in terms of complete or overcomplete sets of pure product projectors. The procedure is useful for investigating the separability of density operators: if the expansion coefficients in such a representation are nonnegative, they can be interpreted as a probability distribution, with the result that the representation then provides an explicit product ensemble for the density operator. Generally, however, the canonical representations have negative values even when the density operator is separable. The canonical continuous representation is the unique representation that contains only scalar and vector spherical-harmonic terms, and we outline how to obtain all possible pure-product-state representations from the canonical continuous representation by adding arbitrary higher-order-spherical-harmonic content. Searching the space of spherical-harmonic expansions for nonnegative representations would be quite difficult, but we provide an optimal ensemble in the particular case of the $\epsilon$-GHZ state, thereby identifying the separability boundary for this class of states. \section*{Acknowledgments} Thanks to G.~J. Milburn and R.~Pranaw for discussions of the method used to prove the separability of the $\epsilon$-GHZ state. This work was supported in part by the US Office of Naval Research (Grant No.~N00014-93-0-0116) and by the UK Engineering and Physical Sciences Research Council. \section*{Note added} The separability bound for the $\epsilon$-GHZ state derived in Sec.~\ref{sec:GHZ} has also been found by D\"ur, Cirac, and Tarrach \cite{Duer1999} as part of a more general result. They prove that the $N$-qubit $\epsilon$-cat state (\ref{eq:epcat}) is separable if and only if $\epsilon<1/(1+2^{N-1})$, which supersedes our Eq.~(\ref{eq:sepNcat}).
\section{Introduction} Double neutron star (DNS) systems observed as binary radio pulsars have provided striking empirical confirmation of general relativity with the measurement of orbital decay accompanied with gravitational-wave emission in the prototypical system PSR B1913+16~\cite{TW82}. This orbital period decrease is expected to lead eventually to the merger of the two neutron stars. DNS coalescence events represent one of the most promising sources for the {\em direct} detection of gravitational waves by the new generation of laser-interferometer detectors currently under construction~\cite{A92} (e.g., LIGO, VIRGO). The final coalescence of DNS has also been discussed as a possible central engine of gamma-ray bursts~\cite{P86}$^{,}$\cite{R97} and has been suggested as the possibly dominant source of r-process elements in the universe~\cite{R99}. Estimates of the rate of DNS coalescence in our Galaxy are crucial for assessing the prospects for gravitational-wave detection and the possible connection of DNS mergers to gamma-ray bursts and r-process elements. Formation rates of {\em coalescing} DNS (systems with tight enough orbits that merge within a Hubble time) have been calculated so far either theoretically, based on evolutionary models of DNS formation, or empirically, based on the observed DNS sample. In this paper I present a critical review of the current coalescence-rate estimates addressing the uncertainties involved and I discuss new ways of constraining these estimates using all available observational information and current theoretical understanding. It is useful to provide a scale for the various estimates in the context of gravitational-wave detection: based on the expected performance of the second-generation LIGO observatories~\cite{T96}, for a detection rate of one merger event per year, a Galactic DNS coalescence rate of $\sim 10^{-5}$\,yr$^{-1}$ is required. For comparison, the estimated~\cite{W98} rate of gamma-ray bursts per galaxy is $\sim 10^{-7}$\,yr$^{-1}$. \section{Theoretical Estimates} Theoretical calculations of the formation rate of coalescing DNS are possible, given a sequence of evolutionary stages that leads from primordial binaries to DNS formation. Over the years a relatively standard picture has been formed describing the birth of DNS~\cite{H76}, although more recently variations of it have also been discussed~\cite{B95}. In all versions of the DNS formation path the main picture remains the same. The initial binary progenitor consists of two binary members massive enough to eventually collapse into neutron stars. Its evolution involves multiple phases of stable or unstable mass transfer, common-envelope evolution, and accretion onto neutron stars, as well as two supernova explosions. Such theoretical modeling of DNS formation has been undertaken by various authors by means of population syntheses. This provides us with {\em ab initio} predictions of the coalescence rate. The evolution of an ensemble of primordial binaries with assumed initial properties is followed through specific evolutionary stages until a coalescing DNS is formed. The changes in the properties of the binaries at the end of each stage are calculated based on our current understanding of the various processes involved: wind mass loss from massive hydrogen- and helium-rich stars, mass and angular-momentum losses during mass transfer phases, dynamically unstable mass transfer and common-envelope evolution, effects of highly super-Eddington accretion onto neutron stars, and supernova explosions with kicks imparted to newborn neutron stars. Given that several of these phases are not very well understood, the results of population synthesis are expected to depend on the assumptions made in the treatment of the various processes. Therefore, exhaustive parameter studies are required by the nature of the problem. Recent studies of DNS formation and calculations of coalescence rates~\cite{L97}$^{,}$\cite{F98}$^{,}$\cite{PZ98}$^{,}$\cite{BB98}$^{,}$ \cite{BB99} have explored the input parameter space and the robustness of the results at different levels of (in)completeness. Almost all have studied the sensitivity of the coalescence rate to the average magnitude of the kicks imparted to newborn neutron stars. The range of predicted Galactic rates from {\em all} these studies obtained by varying the kick magnitude is $5\times 10^{-7}~-~5\times 10^{-4}$\,yr$^{-1}$. This large range indicates the importance of supernovae (two in this case) in binaries. Variations in the assumed mass-ratio distribution for the primordial binaries can {\em further} change the predicted rate by about a factor of $10$, while assumptions of the common-envelope phase add another factor of about $10-100$. Variation in other parameters typically affects the results by factors of two or less. It is evident that recent theoretical predictions for the DNS coalescence rate cover a disappointingly wide range of values (typically 3-4 orders of magnitude), which actually includes the ``nominal'' value of $\sim 10^{-5}$\,yr$^{-1}$. Note that DNS properties other than the coalescence rate, such as orbital sizes, eccentricities, center-of-mass velocities, are much less sensitive to the various input parameters and assumptions; the latter affect more severely the absolute normalization (birth rate) of the population. Given these results it seems fair to say that population synthesis calculations have very limited predictive power and provide fairly loose constraints on the DNS coalescence rate. Overall, we cannot use them to make a robust statement about the prospects for detection of merger events by the upcoming gravitational-wave observatories. \section{Empirical Estimates} Another way to estimate the coalescence rate is to use the properties of the observed coalescing DNS (only two systems: PSR B1913+16 and PSR B1534+12) combined with models of selection effects in radio pulsar surveys. For each observed object, a scale factor is calculated based on the fraction of the Galactic volume within which pulsars with properties identical to those of the observed pulsar could be detected, in principle, by any of the radio pulsar surveys, given their detection thresholds. This scale factor is a measure of how many more pulsars like those detected in the coalescing DNS systems exist in our galaxy. The coalescence rate can then be calculated based on the scale factors and estimates of detection lifetimes summed up for all the observed systems. This basic method was first used by Phinney~\cite{P91} and Narayan et al.~\cite{N91} who estimated the Galactic rate to be $\sim 1-3\times 10^{-5}$\,yr$^{-1}$. Since then, estimates of the coalescence rate have decreased significantly primarily because of (i) the increase of the Galactic volume covered by radio pulsar surveys with no additional coalescing DNS being discovered~\cite{CL95}, (ii) the increase of the distance estimate for PSR B1534+12 based on measurements of post-Newtonian parameters~\cite{S98} (see also contribution by I.\ Stairs, this volume), (iii) changes in the lifetime estimates for the observed systems~\cite{HL96}$^{,}$\cite{A99}. The most recently published study~\cite{A99} gives a lower limit of $2\times 10^{-7}$\,yr$^{-1}$ and a ``best'' estimate of $\sim 6-10\times 10^{-7}$\,yr$^{-1}$. Some of the assumptions made in obtaining the above estimates, are not clearly justifiable or testable. In particular, one assumes that the sample of the two observed coalescing DNS is representative of the total Galactic population~\cite{K99}, that the detection volume for each object and its lifetime are independent and separable (see contribution by T.A.\ Prince, this volume, for arguments against this in the presence of pulsar luminosity evolution and subsequent further reduction of the estimated rate), and that the DNS pulsar luminosity function is similar to that of young, non-recycled pulsars. Additional uncertainties arise from estimates of pulsar ages and distances, the pulsar beaming fraction, the spatial distribution of DNS in the Galaxy, and the number of undetectable pulsars with luminosities below the detection limits of surveys. Despite all these uncertainties the empirical estimates of the DNS coalescence rate appear to span a range of $\sim 1-2$ orders of magnitude, which is relatively narrow compared to the range covered by the theoretical estimates. \subsection{Small-number Sample} One important limitation of empirical estimates of the coalescence rates is that they are derived based on {\em only two} observed DNS systems, under the assumption that the observed sample is representative of the true population. Therefore, assessing the effect of small-number statistics on the results of the above studies is necessary. Assuming that DNS pulsars follow the radio luminosity function of young pulsars and that therefore their population is dominated in number by low-luminosity pulsars, it can be shown that the current empirical estimates underestimate the true coalescence rate. If a small-number sample is drawn from a parent population dominated by low-luminosity (hence hard to detect) objects, it is statistically more probable that the sample will actually be dominated by objects from the high-luminosity end of the population. Consequently, the empirical estimates based on such a sample will tend to overestimate the detection volume for each observed system, and therefore underestimate the scale factors and the resulting coalescence rate (this effect is partly related to the Malmquist bias). This effect can be clearly demonstrated with a Monte Carlo experiment (R. Narayan, private communication) with the use of simple models for the pulsar luminosity function and the survey selection effects. As a first step, the average observed number of pulsars is calculated given a known true total number of pulsars in the Galaxy (thick-solid line in Figure 1). As a second step, a large number of sets of a given number of `observed' (simulated) pulsars with assigned luminosities according to the assumed luminosity function are realized using Monte Carlo methods. Based on each of these sets, one can estimate the total number of pulsars in the Galaxy using empirical scale factors, as is done for the real observed sample. The many (simulated) `observed' samples can then be used to obtain the distribution of the estimated total Galactic numbers of pulsars. The median and 25\% and 75\% percentiles of this distribution are plotted as a function of the assumed number of systems in the (fake) `observed' samples in Figure 1 (thin-solid and dashed lines, respectively). \begin{figure} \centerline{\epsfig{figure=ramesh.ps,height=3.5in}} \caption{Bias of the empirical estimates of the DNS coalescence rate because of the small-number observed sample. See text for details.} \end{figure} It is evident from Figure 1 that, in the case of small-number observed samples (less than $\sim 10$ objects), it is highly probably that the estimated total number, and hence the estimated coalescence rate, is underestimated by a significant factor. For a two-object sample, for example, the true rate maybe higher by more than a factor of ten. \section{Limits on Coalescence Rates} One way to circumvent the uncertainties involved in the estimates of the DNS coalescence rate is to focus on obtaining upper or lower limits to this rate. Depending on how their value compares to the value of $\sim 10^{-5}$\,yr$^{-1}$ needed for one LIGO II event per year, such limits can provide us with valuable information about the prospects of gravitational-wave detection. Bailes~\cite{B96} used the absence of any young pulsars detected in DNS systems and obtained a rough upper limit to the rate of $\sim 10^{-5}$\,yr$^{-1}$, while recently Arzoumanian {\it et al}~\cite{A99} reexamined this in more detail and claimed a more robust upper limit of $\sim 10^{-4}$\,yr$^{-1}$. An upper bound to the rate can also be obtained by combining our theoretical understanding of orbital dynamics (for supernovae with neutron-star kicks occurring in binaries) with empirical estimates of the birth rates of {\em other} types of pulsars related to DNS formation~\cite{KLN99}. Binary progenitors of DNS systems experience two supernova explosions when the neutron stars are formed. The second supernova explosion (forming the neutron star that is {\em not} observed as a pulsar) provides a unique tool for the study of DNS formation, since the post-supernova evolution of the system is simple, driven only by gravitational-wave radiation. There are three possible outcomes after the second supernova: (i) a coalescing DNS is formed (CB), (ii) a wide DNS (with a coalescence time longer than the Hubble time) is formed (WB), or (iii) the binary is disrupted (D) and a single pulsar similar to the ones seen in DNS (DNS-like) is ejected. Based on supernova orbital dynamics we can calculate the probability branching ratios for these three outcomes, $P_{\rm CB}$, $P_{\rm WB}$, and $P_{\rm D}$. For a given kick magnitude, we can calculate the maximum ratio $(P_{\rm CB}/P_{\rm D})^{\rm max}$ for the complete range of pre-supernova parameters defined by the necessary constraint $P_{\rm CB}\neq 0$ (Figure 2). Given that the two types of systems have a common parent progenitor population, the ratio of probabilities is equal to the ratio of the birth rates $(BR_{\rm CB}/BR_{\rm D})$. \begin{figure} \centerline{\epsfig{figure=Vk.ps,height=3.45in}} \caption{Maximum probability ratio for the formation of coalescing DNS and the disruption of binaries as a function of the kick magnitude at the second supernova.} \end{figure} We can then use (i) the absolute maximum of the probability ratio ($\approx 0.26$ from Figure 2) and (ii) an empirical estimate of the birth rate of single DNS-like pulsars based on the current observed sample to obtain an upper limit to the DNS coalescence rate. The selection of this small-number sample involves some subtleties~\cite{KLN99}, but a preliminary analysis shows $BR_{\rm CB} > 0.5-3\times 10^{-5}$\,yr$^{-1}$~\cite{KLN99}. Note that this number could be increased because of the small-number sample bias affecting this time the empirical estimate of $BR_{\rm D}$ (see \S\,3.1). This is an example of how we can use observed systems other than DNS to improve our understanding of their coalescence rate. A similar calculation can be done using the wide DNS systems instead of the single DNS-like pulsars~\cite{KLN99}. \section{Conclusions} A comparison of the various results on the DNS coalescence rate indicates that theoretical estimates based on modeling of DNS formation have a quite limited predictive power (range of at least 3-4 orders of magnitude), whereas empirical estimates based on the observed DNS sample appear to be more robust (within a factor of $\sim 100$). Nevertheless, current estimates, which fall right around the range of interest for evaluating the prospects of gravitational-wave detection, still suffer from uncertainties and systematic effects (e.g., small-number statistics, distances, luminosity function, beaming). Therefore, the need for additional constraints and alternative methods for estimates is clear. As an example, a fairly robust upper bound to the rate can be obtained making use all of the available information for other systems evolutionarily linked to DNS. Apart from DNS systems, binaries with two black-holes or a black-hole and a neutron star are also of interest in the context of gravitational-wave detection and possibly gamma-ray bursts. Since at present there are {\em no} observed systems of this type, we have to rely on theoretically predicted formation rates for them, keeping in mind the normalization uncertainties associated with them. We note, however, that the absence of such systems from the observed samples combined with some basic understanding of their formation relative to DNS could provide valuable information for the frequency of their formation in our Galaxy. \section*{Acknowledgments} I would like to thank T.A.\ Prince for discussions, R.\ Narayan for discussions and for providing me with the results of his Monte Carlo simulations, and D.R.\ Lorimer for providing me with an empirical estimate of the birth rate of single DNS-like pulsars. I would also like to thank D.\ Psaltis and F.\ Rasio for comments on an initial manuscript. I am grateful to the organizers for inviting me to the meeting and acknowledge full support by the Smithsonian Institute in the form of a CfA Post-doctoral Fellowship. \section*{References}
\section{Introduction} In a last decade we have witnessed a tremendous flow of applications of $1(+1$)-dimensional exactly soluble models in 2-d CFT, 2-dimensional gravity --- both continuum and matrix model approach ---, (super) string theories, etc. Specifically, KdV, KP hierarchy and their modified cousins played important and remarkable roles in various occasions. We owe this success mainly to the existence of fascinating mathematical structures underlying such exactly soluble models, {\it i.e.}\ an infinite dimensional symmetry, known as ${\cal W}$-algebras, including Virasoro algebra. In the case of higher dimensional exactly soluble models (see \eg\ \cite{abl,kono93}), however, even though their importance was pointed out some time ago \cite{ky}, due to their mathematical complexity, not much application has been explored until recently. Veselov-Novikov (VN) hierarchy \cite{VN} and its modified cousin (mVN) \cite{bog} are demonstrated as another type of $2(+1)$-dimensional extension of KdV and mKdV, as compared to the well-known KP-hierarchy. One interesting feature of this higher dimensional generalization is that in the mVN case one deals with a deformation problem of Dirac operators in 2 dimensions \cite{bog}, rather than that of quadratic differential operators as in ordinary cases. In addition, thanks to the contributions of the authors of the recent literature \cite{kono96,kt,taim} we now know the important relevance of mVN to conformally Euclidean immersion of 2-surfaces into 3 (or higher) dimensional Euclidean (or Minkowski) manifold \cite{ken,hoff,eisen}. There the potential term in the mVN equation is interpreted as mean curvature of the immersed surface (times $\sqrt[4]{g}$), and the first integral of the mVN equation ({\it i.e.}\ 1-st member of the hierarchy) is shown \cite{carkon,kt} to be in agreement with Polyakov's extrinsic string action \cite{poly} (Willmore functional \cite{will}) in Euclidean signature at classical level. Purpose of this note is to show explicitly that the first integral is also invariant under the deformations associated to the rest of the members of the hierarchy. This confirms the statement suggested in the literature \cite{carkon} obtained from general argument of the soluble system. The derived transformation laws for the potential will also be useful, following similar methods to ordinary (intrinsic) string cases, to pin down the algebraic structure of the infinite symmetries in the extrinsic strings.\footnote[2]{ As a decade-old subject, there are a huge number of works/contributions to the subject. The references cited in this Letter may not reflect all of them.} (For the current status of the extrinsic strings in connection with QCD, see \eg\ \cite{ph} and references therein.) \section{The mVN and generalized Weierstrass inducing} We first review generalized Weierstrass inducing and see how the mVN hierarchy is involved there. The Weierstrass representation is the construction of the conformally Euclidean minimal surface ({\it i.e.}\ that with vanishing mean curvature) in Euclidean 3-space. (See also \cite{par}.) The generalized Weierstrass inducing considered here is the extension of this construction to non-minimal surfaces. To explain this, here we follow the notations of Kenmotsu \cite{ken}. Let $x_j:\Sigma \rightarrow \mathbb{R}^3\ (j=1,\dots,3)$ be a conformally Euclidean immersion of an oriented 2-surface $\Sigma$ (coordinatized locally by $z,\overline{z}\in \mathbb{C}$) into $\mathbb{R}^3$. Then following \cite{ken} we have: \begin{eqnarray} \partial\overline{\partial}x_j &=& \frac{\lambda^2}{4}(h_{11}+h_{22})\,e_{3j},\\ {\overline{\partial}}^2 x_j &=& \overline{\partial}\lambda\cdot(e_1+i\, e_2)_j +\frac{\lambda^2}{4}(h_{11}-h_{22}+2i\,h_{12})\,e_{3j},\\ {\partial}^2 x_j &=& \partial\lambda\cdot(e_1-i\, e_2)_j +\frac{\lambda^2}{4}(h_{11}-h_{22}-2i\,h_{12})\,e_{3j}. \end{eqnarray} Here $\lambda(>0)$ is related to the conformal factor of the induced metric \begin{equation} ds^2\equiv \sum_{j=1}^3(dx_j)^2=\lambda^2dz\,d\overline{z}\ ,\quad\ {\rm or}\quad \ \lambda^2=2\sum_j\left|\frac{\partial x_j}{\partial z}\right|^2 =2\sqrt{|g|}\ , \end{equation} and $h_{ij}$ are related to the mean $H$ and Gaussian $K(\sim$ Ricci scalar) curvatures \begin{eqnarray} H &=& \frac{1}{2} (h_{11}+h_{22})\\ K &=& H^2 - |\phi|^2\ ,\quad \phi\equiv\frac{1}{2} (h_{11}-h_{22})-i\,h_{12}\ , \end{eqnarray} respectively. The quantities $e_\alpha (\alpha=1,\dots,3)$ are normalized tangent and normal vectors of the immersed surface $\Sigma$, whose components are defined as follows: \begin{equation} e_{1j}=\frac{1}{\lambda}\left(\frac{\, \partial x_j}{\partial z}+ \frac{\ \partial x_j}{\partial\overline{z}}\right)\, ,\quad e_{2j}=\frac{i}{\lambda}\left(\frac{\, \partial x_j}{\partial z}- \frac{\ \partial x_j}{\partial\overline{z}}\right)\, ,\quad e_{3i}=\epsilon_{ijk}\,e_{1j}\,e_{2k}\, . \end{equation} Next we introduce $\psi_{1,2}$ \cite{taim} \begin{eqnarray} \psi_1 &\equiv& \left[\,\,\,\overline{\partial}(x_2+i\,x_1)\,\right]^{1/2}\ , \label{def1}\\ \psi_2 &\equiv& \left[-\partial(x_2+i\,x_1)\,\right]^{1/2}\ . \label{def2} \end{eqnarray} Then by direct calculation, using above formulas, we find that $\psi_i$'s have to satisfy \begin{eqnarray} \partial\psi_1 &=& \frac{\lambda H}{2}\psi_2 \label{vn1}\\ \overline{\partial}\psi_2 &=& -\frac{\lambda H}{2}\psi_1 \label{vn2} \end{eqnarray} and \begin{eqnarray} \overline{\partial}(\psi_1/\lambda) &=& \frac{\overline{\phi}}{2}\psi_2\\ {\partial}(\psi_2/\lambda) &=& -\frac{\phi}{2}\psi_1 \ . \end{eqnarray} It is also easy to show \begin{equation} \lambda=|\psi_1|^2+|\psi_2|^2\, . \end{equation} It is remarkable that eqs.(\ref{vn1}),(\ref{vn2}) guarantee the integrability of forms $\Omega_{\pm}, \Omega_3$ defined by \begin{equation} \Omega_+=\overline{\psi}_1^{\ 2}dz -\overline{\psi}_2^{\ 2}d\overline{z}\, ,\quad \Omega_-=\overline{(\Omega_+)}\, ,\quad \Omega_3=-(\psi_2\overline{\psi}_1\,dz + \psi_1\overline{\psi}_2\,d\overline{z})\, , \end{equation} namely, we have $d\Omega_{\bullet}=0$. Now we can consider a converse problem, given a solution to eqs.(\ref{vn1}),(\ref{vn2}) as a system of differential equations with respect to $\psi_i$'s. Since forms $\Omega_{\bullet}$ are all integrable, comparing with (\ref{def1}),(\ref{def2}), we observe that eqs.(\ref{vn1}),(\ref{vn2}) induces an immersion $X_j(z,\overline{z})\ (j=1,\dots,3)$ of conformally Euclidean 2d surface in $\mathbb{R}^3$ via relations \begin{equation} X_2\mp i\, X_1=\int_\Gamma\Omega_\pm\, ,\quad X_3=\int_\Gamma\Omega_3\, . \end{equation} Here $\Gamma$ is an appropriate integration contour ending at $(z,\overline{z})$. The last relation comes from imposed conformally Euclidean property $g_{zz}= (\partial X_3)^2+(\partial X_1)^2+(\partial X_2)^2=0$, and $g_{\overline{z}\, \overline{z}}=0$. The original Weierstrass inducing corresponds to the special case $H=0$, {\it i.e.}\ induction for minimal surfaces. An important observation made in the recent literature \cite{kono96,kt,taim,kono98} is the relevance of eqs.(\ref{vn1}),(\ref{vn2}) to the $2(+1)$ dimensional exactly soluble mVN system. The mVN hierarchy is defined as a deformation problem associated to 2-dimensional Dirac operator (times some $\gamma$-matrix\footnote[2]{We would rather use following form (\ref{Dirac}) of the operator for computational simplicity. We could have used ordinary form for the Dirac operator. Only $B_n$ in equation (\ref{defo}) get changed under such redefinition. In any event essential point is unaffected. }) ${\cal L}$ with a potential $p=p(z,\overline{z})$: \begin{equation}\label{Dirac} {\cal L}=\left(\begin{array}{cr} \partial & -p \\ p & {\overline{\partial}}\end{array}\right)\, ,\quad {\cal L}\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right) = 0\, . \end{equation} We note that eqs.(\ref{vn1}),(\ref{vn2}) correspond to taking special potential $p=\lambda H/2$. The $n$-th deformation in the hierarchy is defined via \begin{equation} \frac{\delta}{\delta t_n}\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right)= A_n\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right)\, , \end{equation} where the deformation operator $A_n$ takes the form \begin{equation}\label{Xs} A_n=\partial^{2n+1}+\sum_{i=0}^{2n-1}X^{(i)}\partial^{i} +\overline{\partial}^{\, 2n+1}+\sum_{i=0}^{2n-1}\tilde{X}^{(i)}\overline{\partial}^{\, i} \end{equation} with $X^{(i)}$, $\tilde{X}^{(j)}$ ($i,j=0,\dots,2n-1$) being $2\times2$ matrices. These matrices are completely determined, together with the other matrix-valued differential operator $B_n$ in eq.(\ref{defo}), from the compatibility condition: \begin{equation}\label{defo} \left[\ \frac{\delta}{\delta t_n} -A_n\ , \ {\cal L}\ \right]=B_n{\cal L}\, . \end{equation} The operator $B_n$ has a similar expression as in the case $A_n$: \begin{equation} B_n=\sum_{i=0}^{2n-1}S^{(i)}\partial^{i} +\sum_{i=0}^{2n-1}\tilde{S}^{(i)}\overline{\partial}^{\, i} \end{equation} with $2\times2$ matrices $S^{(i)}$, $\tilde{S}^{(j)}$ ($i,j=0,\dots,2n-1$). The compatibility condition (\ref{defo}) also gives a deformation equation for the potential $p$ in the form $$ \frac{\delta p}{\delta t_n}= \partial^{2n+1}p +\overline{\partial}^{\, 2n+1}p+\cdots\, . $$ The first case $n=1$ is known to yield modified Veselov-Novikov equation. Here we just write down the result. We will provide more technical details for higher mVN case in later sections. \begin{equation} \frac{\delta p}{\delta t_1}= \partial^{3}p + 3\omega\, \partial p+ \frac{3}{2}p\,\partial\omega+ \overline{\partial}^{\, 3}p+ 3\overline{\omega}\, \overline{\partial}p+ \frac{3}{2}p\, \overline{\partial}\overline{\omega}\ ,\quad \overline{\partial}\omega\equiv\partial p^2\, . \end{equation} It is remarkable that we have a simple first integral of this deformation, which is obtained from the relation \begin{equation} \frac{\delta p^2}{\delta t_1}=\partial\Bigl(\partial^{2}p^2 - 3(\partial p)^2+ 3p^2\omega\Bigr)+ \overline{\partial}\Bigl(\overline{\partial}^{\,2}p^2 - 3(\overline{\partial}p)^2+ 3p^2\overline{\omega}\Bigr) . \end{equation} Namely, the integral $S$ \begin{equation}\label{will} S=2\int p^2\, dz\,d\overline{z} \end{equation} does not change its value under the first deformation ${\delta }/\!{\delta t_1}$ (if $p$ is localized). This conserved quantity has special meaning in the generalized Weierstrass inducing discussed before. Substituting $p=\lambda H/2$, we find that $S$ is nothing but Polyakov's extrinsic string action $$ S=\int\sqrt{|g|}\ H^2\, d^2\!x. $$ (That is known as a Willmore functional in the mathematics literature.) To sum, Polyakov's extrinsic string action is invariant under the deformation associated to the (1st) mVN equation. \section{Second mVN deformation} We write deformation operators for the 2nd mVN as \begin{equation} \frac{\delta}{\delta t_2}\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right)= A_2\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right)\, ,\quad A_2= A_2^{(+)} + A_2^{(-)}\, , \end{equation} where $2\times2$ matrix-valued operators $A_2^{(\pm)}$ are defined as \begin{equation} A_2^{(+)}= \partial^5+V\partial^3 +W\partial^2 + X\partial +Z\, ,\quad A_2^{(-)}= \overline{\partial}^{\,5}+\tilde{V} \overline{\partial}^{\,3} +\tilde{W} \overline{\partial}^{\,2} + \tilde{X}\overline{\partial} +\tilde{Z}\, . \end{equation} Then all we have to do is to work out the compatibility condition (\ref{defo}) for $B_2=B_2^{(+)} + B_2^{(-)}$, here represented as $$ B_2^{(+)}= Q\partial^3 +R\partial^2 + S\partial +T\, ,\quad B_2^{(-)}= \tilde{Q} \overline{\partial}^{\,3} +\tilde{R} \overline{\partial}^{\,2} + \tilde{S}\overline{\partial} +\tilde{T}\, . $$ We will perform this independently on $(+)$ and $(-)$ parts of the compatibility conditions \begin{equation}\label{comp} \left[\ \frac{\delta}{\delta t_2^\pm} -A_2^{(\pm)}\ , \ {\cal L}\ \right]=B_2^{(\pm)}{\cal L}\, ,\quad\quad \frac{\delta}{\delta t_2}= \frac{\delta}{\delta t_2^+} + \frac{\delta}{\delta t_2^-}\, . \end{equation} Some cares must be taken with regards to matrix components $V_{11},W_{11}, X_{11}$, and $\tilde{V}_{22}, \tilde{W}_{22},\tilde{X}_{22}$. For instance, $X_{11}$ and $\tilde{X}_{22}$ give rise to a term in the deformation \begin{equation} \frac{\delta}{\delta t_2}\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right)\sim \left(\begin{array}{c} X_{11}(\partial\psi_1-p\psi_2) \\ \tilde{X}_{22}(\overline{\partial}\psi_2+p\psi_1) \end{array} \right)\, , \end{equation} which vanishes on-shell, so we set $X_{11}=\tilde{X}_{22}=0$ by hand. Similarly we set $V_{11}=W_{11}=\tilde{V}_{22}=\tilde{W}_{22}=0$. In the course of calculation we also get $Z_{11}=\,\,\stackrel{\circ}{Z}_{11}\,\,=\hbox{\rm const}$, and $Z_{22}=\,\,\stackrel{\circ}{Z}_{11}+\cdots$. This type of deformation $$ \frac{\delta}{\delta t_2}\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right)\sim \ \stackrel{\circ}{Z}_{11}\!\!\left(\begin{array}{c} \psi_1 \\ \psi_2 \end{array} \right)\, $$ is merely an overall constant scalar transformation, so we set $\stackrel{\circ}{Z}_{11}=0$ as well. Same is true for $\tilde{Z}_{22}$. After all these done, the deformation matrices are uniquely determined. For the operator $A_2^{(+)}$ we obtain $$ V=\left(\begin{array}{cc} 0 & -5\partial p\\ 0 & 5\omega \end{array} \right)\ ,\quad\quad W=\left(\begin{array}{cc} 0 & -5\partial^2\! p+5p\,\omega\\ 0 & 15\,\partial\omega\!/2 \end{array} \right)\ , $$ \begin{equation}\label{result} X=\left(\begin{array}{cc} 0 & \frac{5}{2}(p\partial\omega-2\omega\partial p-2\partial^3\!p)\\ 0 & \frac{5}{2}(2\omega^2+3\,\partial^2\!\omega+2\zeta) \end{array} \right)\ ,\quad \overline{\partial}\zeta\equiv\partial(p^2\omega-(\partial p)^2)\, , \end{equation} $$ Z=\left(\begin{array}{cc} 0 & 5(p\,(\omega^2+\zeta+\partial^2\omega)+\omega\partial^2\! p+\frac{1}{2}\partial p\,\partial\omega)\\ 0 & \frac{5}{2}\partial(\omega^2+\zeta+\partial^2\omega) \end{array} \right)\ . $$ For the operator $B_2^{(+)}$ the result is $$ Q=\left(\begin{array}{cc} 0 & -5\partial p\\ 5\,\partial p & 0 \end{array} \right)\ ,\quad\quad R=\left(\begin{array}{cc} 0 & ((W_{12}))\\ 10\partial^2\! p+5p\,\omega & 0 \end{array} \right)\ , $$ \begin{equation} S=\left(\begin{array}{cc} 0 & ((X_{12}))\\ \frac{5}{2}(3p\partial\omega+6\omega\partial p+4\partial^3\!p) & 0 \end{array} \right)\ ,\quad \end{equation} $$ T=\left(\begin{array}{cc} 0 & ((Z_{12}))\\ \frac{5}{2}p\,(2\omega^2+2\zeta+3\partial^2\omega)+15\omega\partial^2\! p+15\partial p\,\partial\omega + 5\partial^4\!p & 0 \end{array} \right)\ . $$ The `12'-components should be taken from those in eq.(\ref{result}). The results for operators $A_2^{(-)}$ and $B_2^{(-)}$ are given by the general rule \begin{equation} A_2^{(-)}=\left(\begin{array}{cc} 0 & -1\\ 1 & 0 \end{array} \right)\overline{ A_2^{(+)}}\left(\begin{array}{cc} 0 & 1\\ -1 & 0 \end{array} \right)\, ,\quad B_2^{(-)}=\left(\begin{array}{cc} 0 & -1\\ 1 & 0 \end{array} \right)\overline{ B_2^{(+)}}\left(\begin{array}{cc} 0 & 1\\ -1 & 0 \end{array} \right)\, , \end{equation} where $\overline{ A_2^{(+)}}$ and $\overline{ B_2^{(+)}}$ imply taking complex conjugate of respective components of the matrix operators. Finally the deformation equation for the potential $p$ is obtained as \begin{equation}\label{2nd} \frac{\delta p}{\delta t^+_2}= \partial^5\!p+ 5\omega\,\partial^3\!p + \frac{15}{2}\partial\omega\,\partial^2\!p + \frac{5}{2}\partial p\,( 2\omega^2+3\,\partial^2\!\omega+2\zeta) + \frac{5}{2}p\,\partial( \omega^2+\zeta+\partial^2\omega)\, , \end{equation} and $\delta p/\delta t^-_2=\overline{\delta p/\delta t^+_2}$, respectively. At this stage we can check explicit invariance of our first integral (\ref{will}). From our deformation equation (\ref{2nd}) we immediately have \begin{eqnarray} \lefteqn{\frac{\delta p^2}{\delta t^+_2}= \partial\Biggl[\partial^4\!p^2- 5\partial^2(\partial p)^2+ } \nonumber\\ & & +5(\partial^2p)^2 +5\omega\partial^2p^2- 15\omega(\partial p)^2 + \frac{5}{2}\partial\omega\,\partial p^2 + 5 p^2\,(\omega^2+\zeta+\partial^2\omega) \Biggr] ,\label{res2} \end{eqnarray} and similarly for $\delta p^2/\delta t^-_2=\overline{\delta p^2/\delta t^+_2}$. Thus we find that for localized $p$ the integral (\ref{will}) remains invariant under deformations associated to our 2nd member of mVN hierarchy. Actually, if the purpose is only to show $\delta p^2/\delta t^\pm_2=$ (total divergence), the following is more straightforward. We just inspect `21'- and `12'-components of every terms from the compatibility conditions (\ref{comp}), which provide us \begin{eqnarray} 5\partial p + V_{12}&=&0 \nonumber\\ 10\partial^2p +\partial V_{12}-p\,V_{22} + W_{12}&=&0 \nonumber \\ 10\partial^3p +\partial W_{12}-p\,W_{22} + X_{12}&=&0 \\ 5\partial^4p +\partial X_{12}-p\,X_{22} + Z_{12}&=&0 \nonumber \\ 2\partial^5p+\partial Z_{12} +V_{22}\partial^3p +W_{22}\partial^2p+ X_{22}\partial p &=& 2\delta p/\delta t^+_2\, .\nonumber \end{eqnarray} Eventually we end up with a simpler expression \begin{equation} \frac{\delta p^2}{\delta t_2^+}=\partial\left(\partial^4p^2 +V_{12}\partial^3p +W_{12}\partial^2p + X_{12}\partial p +Z_{12}p \right). \end{equation} This agrees with the previous result (\ref{res2}) after substitutions of the matrix-components from eqs.(\ref{result}). The argument for the other deformation $\delta p^2/ \delta t_2^-$ goes completely in parallel. From these calculations we can naturally infer that the similar structure persists in higher mVN deformations, and we can state quite safely that we have generally \begin{equation} \frac{\delta p^2}{\delta t_n}=\partial\left(\partial^{2n}p^2+\sum_{i=0}^{2n-1} X^{(i)}_{12}\partial^ip\right) + \overline{\partial}\left(\overline{\partial}^{\,2n}p^2+\sum_{i=0}^{2n-1} \overline{X}^{(i)}_{12}\overline{\partial}^ip\right) \end{equation} in our original notation (\ref{Xs}). On the physics side this indicates the invariance of the extrinsic action \`a la Polyakov under all deformations associated to mVN hierarchy. \section{Discussions} We have derived 2nd member of the mVN hierarchy, and checked its consistency to the first integral (one of the so-called Kruskal integrals) derived from the 1st member of the hierarchy. Even though the result is not unexpected from the general argument of exactly soluble models of this sort, writing down a correct form of explicit deformation equation is very important in various respects. First of all, we would like to know the complete symmetry structure of the Polyakov's extrinsic string. In order to pin down such an infinite symmetry we have to be aware of the algebraic structure of the Poisson algebra associated to this system. In the case of KdV, KP we have successfully identified their Poisson structures \cite{ds} to be $W_{\infty}$-algebra family \cite{ky}. The latter has played an important role in other string analyses. Higher Kruskal integrals are also important there. The derived deformation equations here are crucial to deduce such Poisson structures. Secondly, though the immediate relevance found is to Polyakov's extrinsic string, we have pointed out several years ago \cite{ky} the importance of 2- (or higher) dimensional exactly soluble system (that reduces to KdV in one dimension) in its application to 4-dimensional self-dual gravity. The (m)VN is one of the simplest extension of (m)KdV, other than KP. Clarifying its symmetry structure through associated Poisson algebra is an important step towards that goal as well. We would like to know the possible relevance of mVN to that problem, before proceeding to more complex Davey-Stewartson hierarchy. Though calculations become more involved in higher dimensional exactly soluble system, we wish to report on our analyses of these issues in future publications.
\section{Introduction} Direct evidence for the first-order nature of the wetting transition, as originally predicted by Cahn\cite{cahn}, together with the characteristic prewetting jumps away from the coexistence line, has been obtained in several different systems: quantum liquid films physisorbed on heavy alkali metals\cite{taborek}, complex organic liquids\cite{bonn}, near-critical liquid mercury\cite{fisher} and binary liquid crystal mixtures\cite {lucht,esteve}. Quite generally, wetting transitions at temperatures above the triple point are expected for weakly attractive substrates. In particular, it was argued that alkali metals provide the weakest adsorption potentials of any surfaces for He atoms and therefore that wetting transitions could be observed on such substrates \cite{cheng1}. This hypothesis has subsequently been confirmed and the corresponding transition studied in many laboratories \cite {nacher,rutledge,ketola}. More recently, similar phenomena have also been predicted and/or seen for H$_{2}$, Ne, and Hg on various surfaces \cite {cheng2,ross1,mistura1,hess,hensel,kozhevnikov}. Theory dictates \cite{nakanishi,dietrich} that when a first order transition from partial to complete wetting occurs at a temperature $T_{w}$ above the triple-point (and below the bulk critical temperature), then a locus of first-order surface phase transitions must extend away from the vapor-fluid coexistence curve, on the vapour side. At $T<T_{w}$ the thickness of the adsorbed liquid film increases continuously with pressure,\ but the film remains microscopically thin up to the coexistence pressure $P_{sat}(T)$, and becomes infinitely (macroscopically) thick just above it. The adsorption isotherms reach $% P_{sat}(T)$ with a finite slope. At temperatures $T_{w}<T<T_{c}^{pw}$ ($% T_{c}^{pw}$ being the prewetting critical temperature) the thin film grows as the pressure is increased until a transition pressure $% P_{tr}(T)<P_{sat}(T)$ is reached. At this pressure a thin film is in equilibrium with a thicker one, and a jump in coverage occurs as $P$ increases through $P_{tr}(T)$. At still higher pressures this first order transition is followed by a continuous growth of the thick film which becomes infinitely thick at $P_{sat}(T)$. The discontinuity (i.e. the difference in thickness between the two films in equilibrium) becomes smaller and smaller as $T$ increases above $T_{w}$, until at a temperature $% T_{c}^{pw}$ the two films are no longer distinguishable from one another. Finally, when $T>T_{c}^{pw},$ the adsorbed liquid film increases continuously with pressure to become infinitely thick at $P_{sat}(T)$. Noble-gas fluids (other than He) adsorbed on weakly attractive substrates such as the surface of alkali metals, are good candidates for showing prewetting transitions. There is however a limited experimental knowledge for these systems. The measurements of Hess et al.\cite{hess} indicate that Ne undergoes a wetting transition at a temperature within $2\%$ of $T_{c}$ on a Rb surface and that it undergoes a drying transition on Cs. Non-wetting of Ar on a Cs surface has been proposed on the basis of desorption experiments \cite{friess}. No experimental information is available at present on the wetting properties of Ar and Ne on other alkali surfaces. In the absence of experimental observations, numerical simulations may play an important role in understanding the wetting properties of fluids and may be used as a useful guide in the choice of systems to be studied experimentally. We present in this paper one such calculation, based on a Density Functional approach, where the adsorption properties of simple classical fluids (namely Ar and Ne) on Li and Na surfaces are studied. Ne on Rb has also been studied as a test case, where the wetting transition has experimentally been observed \cite{hess}. Recently, results of extensive Monte Carlo (MC) simulations on the Ne/Li system have been presented\cite{bojan}, which seem to show that this system is nonwetting for all T between the triple point and the critical point. As shown in the following, our calculations predict quite a different scenario, and are in fact consistent with prewetting transitions below the critical point. We will discuss the discrepancies between our findings and the MC results of Ref.\cite{bojan} in Section 3. \section{Computational Method} In recent years Density Functional (DF) methods have become increasingly popular because of their ability to describe inhomogeneous fluids and phase equilibria. Comparison with simulation results shows that, once the long-range (van der Waals) attractive forces exerted by a surface on the gas atoms are included in the free energy density functional, the method provides a qualitatively (and most often quantitatively) good description of the thermodinamics of gas adsorption on a solid surface and in particular it is able to predict correctly a large variety of phase transitions (wetting, prewetting,layering,etc.). In addition, also the microscopic structure of the fluid in the vicinity of a solid surface, e.g. the oscillatory behavior of the density profile due to ''packing'' effects, may described accurately by the more sophisticated Density Functionals, which treat in a non-local way the short-ranged repulsive part of the intermolecular potential \cite{tarazona_all,meister,sokolowsky,vanderlick,kroll}. In our calculations we use the DF proposed in Ref.\cite{rosinberg} which has proven to describe successfully the structure of a classical liquid in very demanding situations, e.g. in the presence of hard-walls or strongly attractive surfaces, where rapid variations of the liquid density occur in the close vicinity of the adsorbing surface. Although this functional does not seem to describe correctly inhomogeneities of infinite spatial extent such as those occurring in a freezing transition, it is fairly accurate in describing density inhomogeneities of finite extent. In the Density Functional approach the free energy of the fluid is written in terms of the density $\rho (\vec{r})$ of the fluid as: \begin{eqnarray} F[\rho ] &=&F_{HS}[\rho ]+{\frac{1}{2}}\int \int \rho (\vec{r})\rho (\vec{r}% ^{\,\prime })u_{a}(|\vec{r}-\vec{r}^{\,\prime }|)d\vec{r}d\vec{r}^{\,\prime } \nonumber \\ &&+\int \rho (\vec{r})V_{s}(\vec{r})d\vec{r}. \label{energy} \end{eqnarray} Here $F_{HS}$ is the free-energy functional for an inhomogeneous hard-sphere reference system, the second term is the usual mean-field approximation for the attractive part of the fluid-fluid intermolecular potential $u_{a}$ \cite{allen}, while $V_{s}(\vec{r})$ is the external adsorption potential due to the surface. For $F_{HS}$ we use the non-local functional of Ref. \onlinecite {rosinberg} written in terms of a suitable coarse-grained density obtained by averaging the true fluid density over an appropriate local volume. This scheme predicts good triplet correlation functions $c^{(3)}({\bf r},{\bf r}% ^{\prime })$ for a bulk one-component fluid. In addition, applied to liquid-solid interfaces it gives fairly good results, when compared to ''exact'' Monte-Carlo computer simulations, if the nonuniformities are not too large \cite{rosinberg1}. For instance, both DF and MC simulations agree in predicting and locating the prewetting transitions in the case of Ar/CO$% _{2}$ system\cite{monson}. As an even more stringent test, the two-dimensional limit of the theory of Ref.\cite{rosinberg} has been studied \cite{kr1}: the behavior of a 3D DF theory in this limit is a good test of its performances for describing adsorption phenomena at low coverages. The overall agreement is surprisingly accurate \cite{kr1}. One major weakness of the free-energy functional described above is its mean-field treatment (i.e. its neglect of correlation effects) of the long-range attractive part of the adatom-adatom interaction appearing in the second term of Eq.(\ref{energy}). This term gives the correct non-local dependence of the free energy with respect to variations of the density at two points ${\bf r}$,${\bf r}% ^{\prime }$ away from each other, but it neglects the correlation effects which should enhance the attractive interactions when $|{\bf r}-{\bf r}% ^{\prime }|$ is close to the minimum of the adatom-adatom potential. If one includes this effect within perturbation theory, an expression for the free energy similar to Eq. (\ref{energy}) should be used, but with the pair correlation function of the adsorbed phase multiplying the interaction in the integral of the second term. Rather than following this approach, which is cumbersome and impractical from a computational point of view, we follow Ref. \onlinecite{tarazona} and continue to use the much simpler form (\ref{energy}) but regard $u_{a}$ as an {\it effective} attractive interaction which incorporates the main effect of the pair correlation. This way of proceeding is found to improve the prediction of the DF theory in many cases \cite{monson,tarazona}. We follow previous prescriptions \cite{monson,tarazona,bruno} and use in the second term in Eq.(\ref{energy}) an effective interaction: \begin{eqnarray} u_{a}(r) &=&0\,\,\,\,\,,\,\,\,r\leq \lambda ^{\frac{1}{6}}\tilde{\sigma} \nonumber \\ &=&4\epsilon \{\lambda (\tilde{\sigma}/r)^{12}-(\tilde{\sigma}% /r)^{6}\}\,\,\,\,\,\,,\,\,\,r>\lambda ^{\frac{1}{6}} \tilde{\sigma} \label{pot} \end{eqnarray} which for $\lambda =1$, corresponds to the bare attractive interaction. For $% \lambda <1$ this effective potential has the same long range properties as the bare interaction, which is an important factor when studying wetting phenomena, but its value at the minimum is increased by a factor $\lambda ^{-1}$, to simulate qualitatively the pair correlation effects described above. At any temperature, we treat the HS diameter $\tilde{\sigma}$ \ and the enhancing factor $\lambda $ as free parameters determined by requiring that the experimental values of the liquid and vapor densities at coexistence, $\rho _{l}$ and $\rho _{v}$, are reproduced for the bulk fluid, as explained in detail in the following Section. As for the interaction with the substrate, we make the usual approximation of treating it as an inert, planar surface acting on the the fluid as an external potential $V_{s}(z)$ ($z$ is the coordinate normal to the surface plane). Accurate {\it ab initio} potentials are now available\cite{chizmeshya}, which describe the interaction between noble gases and the alkali surfaces. Approximating the true surface by an ideal plane should not be a very bad approximation for these systems, given the uniformity and almost complete lack of corrugation of clean alkali surfaces. The equilibrium density profile $\rho (z)$ of the fluid adsorbed on the surface is determined by direct minimization of the functional in Eq. (\ref{energy}) with respect to density variations. In practice, at a given temperature, we fix the coverage $\Gamma =\int (\rho (z)-\rho _{v})dz$ and solve iteratively the Euler equation $\mu =\delta F/\delta \rho (z)$ by using a fictitious dynamics, with the value of the chemical potential $\mu $ fixed by $\Gamma $. The results of such minimizations are reported and discussed in the following Section. \section{Results and Discussion} Our results are based on the analysis of adsorption isotherms, calculated by using the DF described in the previous Section, for the case of Ar and Ne adsorption on planar surfaces representing Li and Na surfaces, respectively. The bare Ne-Ne and Ar-Ar interactions are usually given in the form of a Lennard-Jones (LJ) 12-6 potential, with parameters \cite{hirsch,bojan} $\epsilon _{Ne-Ne}/k=33.9\,\,K$, $% \sigma _{Ne-Ne}=2.78\,\AA $ and $\epsilon _{Ar-Ar}/k=119.8\,\,K$, $\sigma _{Ar-Ar}= 3.405\,\AA $. We use in Eq.(\ref{pot}) the above values for $\epsilon $, while we determine the two adjustable parameters entering the functional (\ref{energy}), i.e. the HS diameter $% \tilde{\sigma}$ and the corrective factor to the potential well depth $% \lambda $, by imposing that the experimental bulk phase diagram of the adsorbed fluid (either Ar or Ne) is correctly reproduced. In particular, it is crucial to verify that the choice of these two parameters leads to the correct thermodynamic equilibrium conditions. We have calculated several isotherms for a bulk system. If the temperature is lower than the critical temperature then a blip in the $P-V$ plane, including region of positive slope, $(\partial P/\partial V)_{T}>0$ , develops. As usual, the densities of the coexisting liquid and gas and the equilibrium pressure are found by applying a Maxwell (equal-area) construction in the P-V plane. For a particular choice of $\tilde{\sigma}$ and $\lambda $, this construction gives the vapor pressure $P$ and the liquid and vapor densities $\rho _{v}$ and $\rho _{l}$ at a given temperature. At any $T$, we determine the parameters $\tilde{\sigma}$ and $\lambda $ giving the best fit to the experimental \ values of $P(T)$, $\rho _{v}(T)$ and $\rho _{l}(T)$). A summary of our best fit parameters are given in Tables I and II, while our results for the coexistence line in the $(\rho ^\ast ,P^\ast)$ plane are shown in Fig. (\ref{fig1}) (we use reduced units for the quantities considered. These are defined as: $P^\ast =P\sigma ^3/\epsilon $; $\rho ^\ast =\rho \sigma ^3 $, where $\epsilon $ and $ \sigma $ are the LJ parameters of the bare atom-atom interaction). We notice that the value of $\lambda$ is such to increase the depth of the effective potential $u_a$ over the depth of the corresponding bare potential by a factor $\simeq 1.78$ for Ar and $\simeq 1.8$ for Ne. These values are of the same order of magnitude as that obtained in Ref. \cite{bruno} by fitting $T_c$ for Ar and are reasonable in order to describe the effect of the peak in $g(r)$ at the intermediate densities of interest. A second comment refers to the fitted values of the $\tilde \sigma$'s, which turn out to be about $5 \%$ smaller than the LJ values quoted above, usually employed in Monte-Carlo calculations \cite{bojan,bruno}. As a consequence, the coefficients of the long range attraction, $C_6=4 \epsilon {\tilde\sigma}^6$ are reduced by about $30 \%$ with respect to the corresponding above mentioned LJ values, and are in a much better agreement with the most accurate theoretical values \cite{tang}. As remarked in Ref.\cite{allen}, however, the LJ parameters have nothing of fundamentals and are only determined to provide reasonable agreement between "exact" results (say, from Montecarlo simulations) and the experimental results in the bulk liquids. In other words they are not the values which would apply to an isolated pair of Ar or Ne atoms. Similarly, our $\tilde\sigma$'s are fitted so that the values predicted by our DF calculation for the coexistence pressure at a given T and for the corresponding densities agree with their experimental counterparts. In view of this, our effective interaction entering Eq.(\ref{pot}) should not be regarded as a pair potential and we do not expect that, when inserted in a Montecarlo simulation, it would necessarily lead to the same results as those found from our DF treatment. Once the bulk properties are optimized in the way described above, we switch on the adatom-surface potential in Eq. (\ref{energy}) and compute the chemical potential $\mu $ corresponding to the equilibrium density profile for a given coverage, as described in Section 2. The coverage is expressed in nominal layers $l=\rho _{l}^{-2/3}\int_{0}^{\infty }[\rho (z)-\rho _{v}]dz$. The collection of pairs ($\mu ,l$) at equilibrium, at a given $T$, represents an adsorption isotherm. As an example the isotherm of Ar/Li at $T=128 K$ is shown in Fig.(\ref{fig2}), where the chemical potential, measured with respect to its value at coexistence, is plotted as a function of the film thickness . For small coverages the chemical potential increases approaching the saturation value from below. For certain temperatures like the one considered in Fig.(\ref{fig2}), however, a sudden change in slope is observed as the coverage is further increased. After reaching a minimum value the chemical potential rises again, generating in the $\mu -l$ plane a van der Waals-like loop. Since films for which $\mu $ has a negative slope are unstable, this structure reveals the existence of first-order transitions between films of different thickness. The amplitude of the transition (i.e. the amplitude of the discontinuous jump in coverage) is determined by making an equal-area Maxwell construction. The results of such a construction are shown in Fig.(\ref{fig2}) with a dashed line: two equilibrium thicknesses are thus identified, $l_{1}$ and $l_{2}$, and the jump between them represents a prewetting transition. A few equilibrium density profiles are shown in Fig.(\ref{fig3}), for coverages smaller and larger, respectively, than the thin-thick film values $% l_{1}$ and $l_{2}$ determined above. We summarize our results for the Ar/Li system in Fig.(\ref{fig4}), where the calculated adsorption isotherms for this system are shown for a number of temperatures. It appears from Fig.(\ref{fig4}) that a wetting transition occurs at $T\simeq 124\,K$, accompanied by first-order prewetting transitions at higher $T$. The amplitude of the prewetting transition decreases with temperature, until it vanishes at a critical value $% T_{c}^{pw}\simeq 130\,K$. At temperatures higher than $T_{pw}^{c}$ a continuous wetting of the surface takes place. In the $(\mu - T)$ plane the occurrence of these first order wetting transitions is represented by a line leaving smoothly the liquid-gas coexistence curve at $T_w$ and ending at the prewetting critical temperature $% T^c_{pw}$. On general grounds one expects $\Delta \mu \equiv \mu - \mu _{coex} \sim -a(T-T_w)^{3/2} $, where the exponent $3/2$ is related to the exponent of the van der Waals tail $-C/z^3$ of the surface potential. We show in Fig.(% \ref{fig5}) our calculated values for $\Delta \mu $, together with a fit with the expected analytical form. Fig.(\ref{fig6}),(\ref{fig7}),(\ref{fig8}) show the calculated adsorption isotherms for Ar/Na, Ne/Li and Ne/Na, respectively. In all cases a sequence of prewetting transitions is found, below the bulk critical temperature ($T_c\,(Ar)=150.9 K$ and $T_c\,(Ne)=44.4 K$). In Table III we summarize the wetting and critical pre-wetting temperatures obtained from our calculated adsorption isotherms. In parenthesis we report the wetting temperature as predicted by Eq.(\ref{criterion}) (see the following). The last line in Table III refers to the system Ne/Rb experimentally investigated by Hess et al. \cite{hess}. We calculated two isotherms for this system, at $T=43\,K$ and $T=44\,K$. We find non-wetting behavior at $T=43\,K$, while at $T=44\,K$ complete wetting occurs. We thus estimate $T_w\sim 44\,K$ for this system, but we are not able to resolve any prewetting line of transitions between these two temperatures. Both the wetting temperature and the temperature interval $|T_w-T_{c}^{pw}|\sim 1\,K$ separating the nonwetting from the complete continuous wetting regimes are in excellent agreement with experiments \cite{hess}. Although the almost perfect agreement between theoretical and experimental values of $T_w$ should be regarded as fortuitous in view of the many approximations contained in our calculations ( planar surface, DF treatment, effective potential fit etc.), nonetheless it is rewarding to find that our treatment correctly predicts the observed wetting transition. A simple heuristic model has been proposed by Cheng et al. \cite{cheng} where the energy cost of forming a thick film is compared with the benefit due to the gas-surface attractive interaction, resulting in an implicit equation for the wetting temperature \begin{eqnarray} (\rho _l-\rho _v)\int_{z_0}^\infty V_s(z)dz=-2\gamma \label{criterion} \end{eqnarray} Here $\rho _{l}$ and $\rho _{v}$ are the densities of the adsorbate liquid and vapor at coexistence, $\gamma $ is the surface tension of the liquid and $z_{0}$ is the equilibrium distance of the gas-surface interaction potential $V_{s}$. If one uses in Eq.(\ref{criterion}) the potential $V_{s}$ used in our calculations, and the experimental values for $\rho _{l}(T)$, $\rho _{v}(T)$ and $\gamma (T)$, values of the wetting temperature, $T_{w}$ systematically lower than those found in our calculations are predicted, as shown in the column $T_{w}^{C}$ of Table III. One issue raised by the present results is thus an apparent failure of this simple model to describe wetting behavior in the case of ultraweak potentials, such as those investigated here. Prewetting transitions in classical fluids are notoriously elusive. On the basis of model calculations, it has been suggested \cite{sen} that one possible reason for this elusiveness is that they lie so close to the adsorbate bulk coexistence line as to render them difficult to detect. This happens because the adatom-substrate potential is comparable in strength with the adatom-adatom potential. The model calculations of Ref.\cite{sen} show that, for a wide range of interaction parameters, the prewetting line lies at a chemical potential which differs from that of coexistence by an amount on the order of $10^{-3}KT_{c}$. Our results show no exception to this behavior, although the proximity to coexistence, which makes prewetting transitions so difficult to observe experimentally, is slightly lower than expected. In the case of Ne/Li the values of the chemical potential for which the discontinuities in coverage are observed are only $1K $ (at most) below the saturation value (see Fig. 7), while for Ar/Li they are $2K$ (at most) below (see Fig. 4). Thus $\Delta \mu \sim 10^{-2}KT_{c}$. The vapor pressure at which the prewetting jumps should be observed for the system we have investigated, may be approximately estimated by assuming ideal gas behavior for the vapor phase (see Ref. \cite{bojan} for a better estimate). One finds $P=P_{sat}exp(\Delta \mu /T)\sim 0.985P_{sat}$ for $% \Delta \mu \sim 2\,K$, thus in a range which may be accessible to experiments. We mention at this point recent results\cite{bojan} based on extensive Monte Carlo simulations on the behavior of \ Ne on a Li surface (the gas-surface potential used in these simulations is the same as used here ) where nonwetting behavior up to the critical point has been found, at variance with the results presented here. One may invoke various explanations for these different findings. For example one could think that metastability in the region close to the critical point may affect seriously MC simulations and that reliable results may require a number of MC moves prohibitively large to reach the regime where the wetting transition is found. Another possible explanation on why Boyan et al. do not see wetting of Ne on Li is that the prewetting regime is so close to saturation that their simulation cannot discern it. In fact at 40 K we find a prewetting jump of slightly more than two layers at a chemical potential only $0.5 K$ below saturation. At higher temperatures, where the jump is higher, it occurs even closer to saturation. The situation is even worse for Ne/Na. It seems to be more promising for Ar/Li, where one should see a $\sim 2-3$ layers jump $\sim 1.5 K$ below saturation at $T=128 K$, so this is the system were we suggest to make experiments. We remark at this point the extreme sensitivity of the results presented above to the details of the gas-surface interaction potential. It has been shown recently\cite {ancilotto}, for the case of Ar/CO$_{2}$ system, that minor changes in the overall shape of the adsorption potential may alter dramatically the properties of the adsorbed film. In order to evaluate the sensitivity of the present results to the shape of the substrate potential, we re-calculated the wetting diagram of Ar/Li by using a 9-3 Lennard-Jones potential to describe the Ar-surface interaction, instead than the ab-initio potentials used above \cite{chizmeshya}, and adjusting the two parameters entering the 9-3 potential in order to have the same well depth $D$ and minimum position as the ab-initio potential. We found that the wetting temperature calculated with this interaction shifts upward from 123 to 136 K, while the critical wetting temperature changes from 130 to 138. This is not surprising, given the dependence of the wetting temperature, as clear from the implicit definition Eq. (\ref{criterion}), from the global shape of the potential, and not just from the well depth and position of the minimum. Of course the wetting properties will also be very sensitive to the adatom-adatom potential, which determines the surface tension explicitly entering Eq. (\ref{criterion}). Thus, on these very weak substrates, wetting or non wetting above the triple point, but below $T_c$, is the result of a delicate balance between adatom-substrate and adatom-adatom interactions. Another possible explanation of the discrepancy between our findings and the MC results of Bojan et al. \cite{bojan} may thus be ascribed to the different long range behavior of our effective interaction $u_a$ (see Eq.(2) and the ensuing discussion) and of the bare LJ potential used in the Montecarlo simulations of Ref.\cite{bojan}. Experiments are currently in progress \cite{mist1}, which will hopefully be able to verify the predictions contained in this paper. \acknowledgments We thank G.Mistura, L.Bruschi, M.W.Cole and S.Curtarolo for useful discussions and critical comments. \bigskip \bigskip \bigskip
\section{Introduction} Kondo insulators form a group of compounds that at high temperatures behave as dirty metals but at very low temperatures have their thermodynamic and transport properties determined by the existence of a small gap (10-100K) that arises from the hybridization of local electrons and the conduction band~\cite{fisk1,fisk2}. This family of compounds can be characterized as strongly correlated semiconductors due to the $f$ or $d$ character of the relevant electrons and includes $FeSi$~\cite{c0}, $Ce_{3}Bi_{4}Pt_{3} $~\cite {c1a,c1b,c1c}, $SmB_{6}$~\cite{c2}, $YbB_{12}$~\cite{c3} and $CeFe_{4}P_{12}$% ~ \cite{c4}. Many theoretical models have been used to describe Kondo insulators. Some of them considered a two-band system, a large uncorrelated band of $s$ electrons and another narrow correlated band which describes the $f$ or $d$ electrons~\cite{mucio1,mucio2,mucio3}. Others considered an Anderson lattice, a conduction band and localized $f$ states, with the correlation in these states generally treated within the slave-boson method~\cite {slave1,slave2}. Although Kondo insulators do not present long range magnetic order, the occurrence of short range antiferromagnetic correlations in these materials has been observed by inelastic neutron scattering \cite{anti}, suggesting that a good model to describe Kondo insulators must include antiferromagnetic correlations between neighbors $f$-moments. We also point out that the magnetic susceptibility of Kondo insulators in the Curie-Weiss regime has a negative Curie temperature, as shown in Table 1 for some compounds. This is a clear indication of the presence of antiferromagnetic correlations in these materials. \begin{table}[bbp] \centering \par \begin{tabular}{|l|l|} \hline\hline Compound & $~~~~~\Theta$ \\ \hline $Ce_{3}Bi_{4}Pt_{3}$ & - 125 K ~\cite{c1c} \\ $YbB_{12}$ & -79 K ~\cite{japon} \\ $FeSi$ & -1030 K ~\cite{c0} \\ \hline\hline \end{tabular} \caption{Table of Curie Temperature for some compounds} \end{table} In this letter we study some physical properties of the Kondo lattice taking into account short-range antiferromagnetic correlations between the localized $f$-electrons (only first neighbors) as first proposed by Coqblin \textit{et al.}\cite{coqblin1}. These authors have used this approach to describe metallic heavy-fermions while here we shall use it to investigate Kondo insulators. We calculate the density of states and study its behavior with increasing temperature and increasing magnetic correlations. We evaluate some physical properties such as optical conductivity, magnetic susceptibility and electrical resistivity. The format of the present letter is as follows. In section \ref{s2} we present the Hamiltonian and describe the approach used. In section \ref{s3} we calculate the density of states and some physical quantities in order to analyze the influence of magnetic correlations in the system. \section{Model Hamiltonian\label{s2}} We consider the following Hamiltonian to describe the system \cite{coqblin1}: \begin{equation} H=\sum_{k,\sigma }{\varepsilon _{k}n_{k\sigma }^{c}} + E_{0}\sum_{i,\sigma } {n_{i\sigma }^{f}} - {J_{K}} \sum_{i}{\vec{s}_{i}\cdot{\vec{S}_{i}}} - {J_{H}% }{\ \sum_{i,\delta }} {\vec{S}_{i}}\cdot{\vec{S}_{i+\delta}}, \label{hini} \end{equation} where $n_{k\sigma }^{c}=c_{k\sigma }^{\dagger }c_{k\sigma }$, $n_{i\sigma }^{f}=f_{i\sigma }^{\dagger }f_{i\sigma }$. $\vec{S_{i}}$ are the spin operators associated to localized $f$-moments and $\vec{s_{i}}$ to conduction electrons at site $i$. The first term represents the conduction band. The operators $c_{k\sigma }^{\dagger }$ and $c_{k\sigma }$, respectively, create and annihilate electrons in the conduction band state labeled by the wave vector $\vec{k}$ and spin $\sigma$. The second term, proportional to $E_{0}$, represents the binding energy of the $f$-electrons. The third term is the $s-f$ exchange interaction $J_{K}$ which gives rise to the Kondo coupling. In this model, following Coqblin \textit{et al.} \cite {coqblin1}, we add a Heisenberg-like interaction $J_{H}$ between nearest neighboring $f$-moments. We examine the model above, specifically in the situation where the ground state of the system is an insulating state. First we construct an effective Hamiltonian, rewriting the original one, Eq.(\ref{hini}), in terms of new operators, as performed by Ruppenthal \textit{et. al}~\cite{rs}. Since we intend to describe the Kondo effect, we choose an operator $\hat{\lambda} _{i,\sigma }$ that couples $c$ and $f$ electrons. In order to describe short-range magnetic correlations we introduce another operator $\hat{\Gamma}% _{i,\sigma }$ that couples $f$ electrons in neighboring sites. These Hermitian operators are given by: \begin{equation} \hat{\lambda}_{i\sigma }\equiv \frac{1}{\sqrt{2}}(c_{i\sigma}^{\dagger } f_{i\sigma}+f_{i\sigma}^{\dagger }c_{i\sigma}) \end{equation} and \begin{equation} \hat{\Gamma}_{i+\delta,\sigma }\equiv \frac{1}{\sqrt{2}}(f_{i\sigma}^{% \dagger } f_{i+\delta,\sigma}+f_{i+\delta,\sigma}^{\dagger }f_{i\sigma}). \end{equation} The spin components can be written in terms of the new operators. Using an additional constraint, that excludes the double $f$-occupancy, we get: \begin{eqnarray} s_{i}^{x}S_{i}^{x}+s_{i}^{y}S_{i}^{y} = -\frac{1}{2}\sum_{\sigma}{\hat{% \lambda} _{i\sigma }\hat{\lambda}_{i,-\sigma }}, \nonumber \\ s_{i}^{z}S_{i}^{z} =\frac{1}{4}({n_{i}^{c}}+{n_{i}^{f}})- \frac{1}{4}{% n_{i}^{c}}{n_{i}^{f}}-\frac{1}{2}\sum_{\sigma}{\ \hat{\lambda}_{i,\sigma }^{2}}, \nonumber \\ S_{i}^{x}S_{i+\delta }^{x}+S_{i}^{y}S_{i+\delta }^{y} = -\frac{1}{2}\sum_{ \sigma}{\hat{\Gamma}_{i+\delta ,\sigma }{\hat{\Gamma}_{i+\delta ,-\sigma }}}, \\ S_{i}^{z}S_{i+\delta }^{z} =\frac{1}{4}({n_{i}^{f}}+{n_{i+\delta}^{f}}) -% \frac{1}{4}{n_{i}^{f}}{n_{i+\delta}^{f}}- \frac{1}{2}\sum_{\sigma}{\hat{% \Gamma}_{i+\delta ,\sigma }^{2}}, \nonumber \end{eqnarray} where $n^{\alpha}_{i}=n^{\alpha}_{i\uparrow}+n^{\alpha}_{i\downarrow}$. Since we are describing an insulating state at low temperatures, we use the constraints $<n_{i}^{f}>=<n_{i}^{c}>=1$. Furthermore applying the decoupling, \begin{equation} <n_{i\sigma}^{\alpha }n_{j\sigma}^{\beta }>\approx <n_{i\sigma}^{\alpha }>n_{j% \sigma }^{\beta }+<n_{j\sigma}^{\beta }>n_{i\sigma}^{\alpha }, \end{equation} we find that the terms containing the number operators cancel out. Note that we are interested in non-magnetic solutions such that $<n_{i\sigma}^{\alpha }>=<n_{i-\sigma}^{\alpha }>$. Finally we can write the Kondo and Heisenberg parts of the Hamiltonian as: \begin{equation} H_{K}=\frac{1}{2}J_{K}\sum_{i\sigma}{(\hat{\lambda}_{i\sigma }+\ \hat{\lambda% } _{i,-\sigma })\hat{\lambda}_{i\sigma }}, \label{lam} \end{equation} \begin{equation} H_{H}=\frac{1}{2}J_{H}\sum_{i+\delta \sigma}{(\hat{\Gamma}_{i+\delta,\sigma } +\hat{\Gamma}_{i+\delta ,-\sigma})\hat{\Gamma}_{i+\delta ,\sigma}}. \label{gam} \end{equation} Consistently with the decoupling used above, Eq.(5), we deal with the product of operators in equations (\ref{lam}) and (\ref{gam}) introducing the mean fields $\lambda _{i}=<\hat{\lambda}_{i,\sigma}>$ and $\Gamma _{i}=<% \hat{\Gamma} _{i+\delta ,\sigma}>$. Furthermore due to translational invariance, $\lambda _{i}=\lambda $ and $\Gamma _{i}=\Gamma $. Performing a Fourier transform, using $\varepsilon _{k}=-\frac{W}{z}\sum_{R}{\cos {(kR)}}$ where $z$ is the number of neighbors and $2W$ is the band width, we obtain the effective Hamiltonian: \begin{eqnarray} H^{\prime}=\sum_{k,\sigma}{\varepsilon_{k\sigma}n_{k\sigma}^{c}} +J_{K}\lambda \sum_{k\sigma}{(c^{\dagger}_{k\sigma}f_{k\sigma}+ f^{\dagger}_{k\sigma}c_{k\sigma})} +\sum_{k,\sigma}{(E_{0}-b\varepsilon_{k})n_{k\sigma}^{f}} \\ ~~~~~~-2J_{K}\lambda^{2}-2J_{H}\Gamma^{2}, \nonumber \label{hamiltonfinal} \end{eqnarray} where \begin{equation} b=\frac{J_{H}}{W}\Gamma. \end{equation} We have then arrived at a new picture where the magnetic correlations $J_{H}$ give rise to a narrow band of $f$-electrons of effective width $2bW$ (b is temperature dependent). The interaction $J_{K}$ (using $J_{K}<0$, since we are describing the Kondo effect) introduces an effective hybridization between the conduction and the renormalized $f$ bands. The new quasi-particles associated with the Hamiltonian (\ref{hamiltonfinal}% ) can now be obtained. This is done by calculating the one-electron Green's functions which are found from their equations of motion. These Green's functions are given by, \begin{equation} \mathit{G_{k,\sigma }^{ff}(\omega )}=\frac{A^{f}(\varepsilon _{k},\omega )}{% \omega -\omega _{1}(\varepsilon _{k})}-\frac{A^{f}(\varepsilon _{k},\omega )% }{\omega -\omega _{2}(\varepsilon _{k})}, \label{Gffb} \end{equation} \begin{equation} \mathit{G^{cf}_{k,\sigma}(\omega)}=\frac{J_{K}\lambda}{(\omega_{1} (\varepsilon_{k})-\omega_{2}(\varepsilon_{k}))}\biggl(\frac{1}{\omega -\omega_{1}(\varepsilon_{k})}-\frac{1}{\omega-\omega_{2}(\varepsilon_{k})}% \biggr), \label{Gsfb} \end{equation} \begin{equation} \mathit{G_{k,\sigma }^{cc}(\omega )}=\frac{A^{c}(\varepsilon _{k},\omega )}{% \omega -\omega _{1}(\varepsilon _{k})}-\frac{A^{c}(\varepsilon _{k},\omega )% }{\omega -\omega _{2}(\varepsilon _{k})}. \label{Gssb} \end{equation} where $G_{k,k^{\prime }}^{\alpha \beta }=<<\alpha _{k\sigma}^{\dagger }|\beta _{k^{\prime }\sigma }>>$ for $\alpha ,\beta =c,,f$ and \begin{equation} A^{c}(\varepsilon _{k},\omega )=\frac{\omega -(b\varepsilon _{k}+E_{0})}{% \omega _{1}(\varepsilon _{k})-\omega _{2}(\varepsilon _{k})}, \end{equation} \begin{equation} A^{f}(\varepsilon _{k},\omega )=\frac{\omega -\varepsilon _{k}}{\omega _{1}(\varepsilon _{k})-\omega _{2}(\varepsilon _{k})}. \end{equation} We have expressed the Green's functions in terms of simple poles corresponding to two hybridized quasi-particle bands with dispersion relations given by \begin{equation} \omega _{1;2}(\varepsilon _{k})=\frac{1}{2}[(1+b)\varepsilon _{k}+E_{0}]\pm \frac{1}{2}\sqrt{[(1-b)\varepsilon _{k}-E_{0}]^{2}+4J_{K}^{2}\lambda ^{2}}. \label{polos} \end{equation} \begin{figure}[!h] \epsfxsize=2.7 truein \centerline{\epsfbox{fig1.eps}} \caption{The energy dispersion relations $\omega_{1;2}(k)$ for the hybridized bands found in the mean-field approximation using $\varepsilon_{k}% =-Wcos(ka/\pi) $, where $G/2=\pi/2a$ is half a reciprocal lattice vector. Solid line is used for $J_{H}=0$ and dashed line for $J_{H}<0$.} \label{fig1} \end{figure} As shown in figure~\ref{fig1} the two bands are separated by an indirect gap between the lower band at $k=G/2$ and the upper band at $k=0$. For increasing values of the magnetic correlation $|J_{H}|$, the indirect gap gets larger while the direct gap remains constant. The direct gap $\Delta_{dir}$ is governed only by $J_{K}$. It is easy to see that $\Delta_{dir}=2J_{K}\lambda$ and it occurs at $E_{0}=(1-b)\varepsilon_{k}$. An insulating ground state, as observed in $Ce_{3}Bi_{4}Pt_{3}$\cite{c1a} or $YbB_{12}$\cite{c3}, is obtained when the lower band is completely filled. This corresponds to the half-filled band case where the occupation numbers assume the values $<n_{i}^{f}>=<n_{i}^{c}>=1$. Furthermore, we take $% E_{0}=\mu =0$ and symmetric bands with respect to the chemical potential $% \mu $. Since the $f$-moments are correlated antiferromagnetically, which corresponds to $J_{H}<0$, for $J_{K}\neq 0$ we always have an insulating ground state with the chemical potential fixed in the middle of the gap of the density of states. The parameters $\lambda$ and $\Gamma$ are obtained from the two coupled equations that have to be solved self-consistently for each temperature: \begin{eqnarray} 1={J_{K}}\sum_{k}{\frac{1}{(\omega_{1}-\omega_{2})} \left(\frac{f(\omega_{1})% }{\omega-\omega_{1}}-\frac{f(\omega_{2})} {\omega-\omega_{2}}\right)} \\ \Gamma=-\sum_{k}{\frac{1}{(\omega_{1}-\omega_{2})}\left (\frac{(\omega_{1}-% \varepsilon_{k})f(\omega_{1})}{\omega-\omega_{1}}-\frac{(\omega_{2}-\varepsilon_{k})f(\omega_{2})} {% \omega-\omega_{2}}\right)} \end{eqnarray} where $\mathit{f}(\omega)=\frac{1}{\exp{\beta(\omega-\mu)}+1}$ is the Fermi distribution and $\omega_{1;2}=\omega_{1;2}(\varepsilon_{k})$ is given by Eq.(\ref{polos}% ) The present method has many similarities with the slave-boson mean field approach and like it, presents a critical temperature $T_{c}$ where the conduction and the $f$ electrons become decoupled~\cite{slave1,slave2}. In our model, this arises for $\lambda(T_{c})=\Gamma(T_{c})=0$ which implies the vanishing of the effective hybridization term $\lambda$. So, these mean-field methods are useful to study the properties of Kondo insulators only in the low temperature regime, below the critical temperature $T_{c}$ of the spurious phase transition. We shall present results here for $T\ll T_{c}$ such that the present approach is valid. \section{Theoretical Results and Comparison with Experiments\label{s3}} As discussed previously, the magnetic correlations modify the band structure of the system, increasing the indirect gap. In this section we analyze the influence of these correlations in some physical properties, calculating $% \lambda$ and $\Gamma$ self-consistently for each value of temperature. For simplicity we consider a uniform density of states of width $2W$ for the conduction electrons. \subsection{Density of states} Using that $<n^{\alpha,\beta}_{k,\sigma}>=\frac{1}{\pi} Im[G^{\alpha,% \beta}_{k,\sigma}(\omega)]$ we obtain the $c$ and $f$ contributions to the density of states. In figures \ref{denst} and \ref{densjh}, we show the density of states which consists of two hybridized bands separated by a gap. The density of states is very sharp near the band edges independently of the form of the unperturbed bands. \begin{figure}[!h] \epsfysize=2.7truein \centerline{\epsfbox{fig2.eps}} \caption{$f$ contribution to the density of states for different temperatures. $J_{K}/W=-0.4$, $J_{H}/W=-0.1$.} \label{denst} \end{figure} \begin{figure}[!h] \epsfysize=2.7truein \centerline{\epsfbox{fig3.eps}} \caption{$f$ contribution to the density of states for two different values of $J_{H}$ and $k_{B}T/W=0$.} \label{densjh} \end{figure} Temperature renormalizes the density of states decreasing the gap (figure \ref{denst}). With increasing temperatures the density of states become more peaked near the band edges. The density of states also depends on the magnetic correlations represented here by the parameter $b$. The last term in Eq.(\ref{hamiltonfinal}) produces an effective $f$-band and its weight renormalizes with $b$. In a system where the Fermi level is inside the gap, the main effects of short range antiferromagnetic correlations are to increase the gap and enhance the density of states near the band edges, as we can see in figure \ref{densjh}. \subsection{Optical conductivity} The optical conductivity can be written as~\cite{cond}: \begin{eqnarray} \sigma(\nu)=\frac{\pi}{2W}\sum_{\sigma}{\int{d\varepsilon\int{d\omega % \rho^{c}_{\sigma}(\varepsilon,\omega)\rho^{c}_{\sigma}( \varepsilon,\omega+\nu)}}} {{\frac{[f(\omega)-f(\omega+\nu)]}{\nu}}} \end{eqnarray} where $\rho^{c}_{\sigma}(\varepsilon,\omega)$ is the one particle spectral density of the conduction electrons. \begin{figure}[h] \epsfysize=3.0truein \centerline{\epsfbox{fig4.eps}} \caption{Optical conductivity for $J_{K}/W=-1.0$, $J_{H}/W=-0.1$ varying the temperature. The frequency h$\nu $ is in units of the bandwidth $W$.} \label{ocond} \end{figure} In the optical conductivity, shown in figure \ref{ocond}, the behavior of the gap involves two characteristic temperature scales as observed experimentally for $Ce_{3}Bi_{4}Pi_{3}$~\cite{ocondce} and $FeSi$~\cite{c0}. Firstly the high temperature regime, compared to the width of the conductivity gap $\Delta_c$, where the gap $\Delta_{c}$ itself is strongly renormalized with $T$, as we can see in figure \ref{ocond}. The optical conductivity also presents a Drude-like feature at low frequencies, as observed recently by Gorshunov \textit{et al}~\cite{ocon3} in dynamical conductivity measurements in $SmB_{6}$. This peak indicates that the low-energy transport is determined by free charge carriers. Then the second temperature scale is set by the width of this Drude peak. Only at small temperatures the gap is completely opened. At these low temperatures the variation in the intensity of this peak, whose weight is transferred to the high frequency region of the spectrum, is the main effect of temperature. The gap itself remaining unchanged as the low-energy behavior changes from being dominated by free carriers to more localized carriers. We point out that the interaction $J_{H}$ has no effect in the gap size of the optical conductivity. \subsection{Magnetic susceptibility and electrical resistivity} In this subsection we calculate the electrical resistivity and the magnetic susceptibility for different values of $J_{H}$ to analyze the influence of antiferromagnetic correlations in these properties. The electric conductivity is obtained from the limit $\nu \rightarrow 0$ of the optical conductivity. \begin{equation} \sigma (0)=\frac{\pi \beta }{2W}\int {d\varepsilon \int {d\omega [\rho _{\sigma }^{c}\ (\varepsilon ,\omega )]^{2}f(\omega )[1-f(\omega )]}}, \label{condu} \end{equation} where $\beta \equiv 1/T$. The resistivity is given by $\rho =1/\sigma (0)$. Notice that this expression must be used with care. In the problem considered here, there is translational invariance and consequently\textbf{\ k }is a good quantum number. However in real systems impurity scattering is always present and this limits the electron mean free path. In the calculations presented below, this is taken into account by including a finite lifetime for the conduction electrons. Formally this is done by replacing, $\omega \rightarrow \omega +i\Gamma $, in the Green 's functions, where $\Gamma $ is temperature independent. We point out that here we are interested in the temperature dependence of the conductivity which is not affected by the magnitude of $\Gamma $. If an external magnetic field $h$ with its direction along the z-axis is applied, the quasi-particle energies become \begin{equation} \omega _{i,\sigma }(\varepsilon _{k})=\omega _{i}(\varepsilon _{k})-\sigma h. \label{eh} \end{equation} The free energy in the presence of the field, using the new quasi-particles energies (\ref{eh}), is written as \begin{equation} F(\lambda ,\Gamma ,T)=-\frac{1}{\beta }\sum_{i=1,2}\sum_{k,\sigma }{{\ln {% \biggl(1+\exp {({-\beta \omega _{i,\sigma }(\varepsilon _{k}))}\biggr) }}}}% +2J_{H}\lambda ^{2}+2J_{H}\Gamma ^{2} \label{free} \end{equation} and the magnetic susceptibility is given by \begin{equation} \chi (T)=\frac{\partial ^{2}F}{\partial h^{2}}_{h=0}=\frac{1}{4}\beta \sum_{i=1,2}\sum_{k,\sigma }{{sech^{2}\left( \frac{1}{2}\beta \omega _{i,\sigma }(\varepsilon _{k})\right) }}. \end{equation} \begin{figure}[h] \epsfysize=2.3truein \centerline{a)\epsfbox{fig5a.eps} b)\epsfysize =2.3truein \epsfbox{fig5b.eps}} \caption{Temperature behavior of the calculated susceptibility (a) and resistivity (b) showing activation behavior. The fitting using expressions given in the text are shown as full lines for different values of $J_{H}$ and $ J_{K} /W=-0.4$.} \label{ajustes} \end{figure} The magnetic susceptibility and electrical resistivity can be obtained from the expressions above for different values of $J_{H}$ and $J_{K}$ fixed. The calculated low temperature resistivity and susceptibility curves were fitted using the activated forms $\chi (T)=(C/T)\exp {(-\Delta _{\chi }/kT)}$ and $% \rho (T)=\rho _{0}\exp {(-\Delta _{\rho }/kT)}$, which describe the experimental data in Kondo insulators. As we can see in figure \ref{ajustes}% , these analytical expressions give also a good description of the theoretical results, indicating that our model describes very well the low temperature properties of these systems. The fittings of the theoretical results yield values for the transport (${\Delta _{\rho }}$) and magnetic (${% \Delta _{\chi }}$) gaps \cite{c0}. In figure \ref{gaps} we show these gaps for different values of $J_{H}$. $\Delta _{\chi }$ and $\Delta _{\rho }$ increase for increasing values of $|J_{H}/W|$ but their ratio remains the same ($\Delta _{\chi }/\Delta _{\rho }\approx 2.2$) suggesting that magnetic correlations do not influence the relation between them. \begin{figure}[!h] \epsfysize=3.0truein \centerline{\epsfbox{fig6.eps}} \caption{Magnetic (full squares) and transport (full circles) gaps versus $ J_{H}$ for $J_{K}/W=-0.4$.} \label{gaps} \end{figure} \subsection{Analysis} Experimental results yield different values for the gaps of Kondo insulators. The gap measured in optical conductivity $\Delta_{c}$ has larger values than the gaps measured in neutron scattering and susceptibility measurements,$\Delta_{\chi}$ or in transport measurements $\Delta_{\rho}$% \cite{ocondce}. On the other hand, comparing the spin and transport gaps, experimental results show that $\Delta_{\chi}\geq \Delta_{\rho}$~\cite {c0,c1c}. The present theory accounts for these observations and introduces a new discussion about the roll of short-range magnetic correlations in the properties of these gaps. We have discussed in section 2 the influence of antiferromagnetic correlations in the indirect and direct gaps in the density of states. Although in the calculations of section 3 we used a square band, without a defined dispersion relation, we point out that, since our results show that the frequency gap $\Delta _{c}$ in the optical conductivity does not renormalize as $J_{H}$ varies, we may conclude that this quantity is determined by the direct gap. This is expected, since due to the negligeable momentum of the photon, optical excitations involve essentially an energy transfer. On the other hand the results in figure (\ref{gaps}) show that the transport and susceptibility gaps are strongly renormalized by magnetic correlations indicating the indirect nature of the gap that determines these physical properties. \section{Conclusions} In this work we have discussed a model for Kondo insulators which takes into account the influence of short-range magnetic correlations in these systems. We used a Kondo lattice with an additional Heisenberg term and we treated the problem within a mean-field approach in order to obtain an effective Hamiltonian where the additional term gives rise to an effective band width for the $f$ electrons. Although the Kondo interaction leads to an indirect coupling between the local electrons, we have added explicitly to the Hamiltonian an interaction between them. In this way we can make a straightforward analysis of the influence of these magnetic correlations on the properties of the system. We have obtained the density of states of the new quasi-particles and observed that their energy bands are renormalized by temperature and antiferromagnetic correlations. We have calculated the magnetic susceptibility and electric resistivity to get the magnetic and transport gaps which turn out to be strongly renormalized by magnetic correlations. Calculating the optical conductivity we find a Drude-like peak at very low frequencies, as observed recently in dynamical conductivity measurements of $% SmB_{6}$\cite{ocon3}. However, the gap in the optical conductivity remains the same for different values of the strength of magnetic correlations. This gap is determined by the direct gap in the band structure, while the magnetic and transport gaps arise mostly from the indirect gap. The antiferromagnetic correlations change substantially the physical properties related to the indirect gap but do not change the quantities related to the direct one. These observations can produce an insight for future experimental analysis. This letter discusses explicitly the short-range magnetic correlations which have not been explored previously in the theoretical study of Kondo insulators. The model is able to reproduce their physical properties, including their temperature dependence, and is in good qualitative agreement with experimental results. \section*{Acknowledgments} The authors like to thank M. A. Gusm\~{ao} for useful discussions and sending a pre-print of his work. This work was supported by The Brazilian agencies Conselho Nacional de Desenvolvimento Cientifico (CNPq) and CAPES.
\section{Introduction} Let $B$ be a projective algebraic curve over an algebraically closed field $k$ of characteristic zero, and let $S \subseteq B$ be a finite set of points. Shafarevich's conjecture for families of curves, proved by Parshin and Arakelov (see \cite{A}), states that \begin{quote} (I) There are only finitely many isomorphism classes of smooth non-isotrivial families of curves over $B-S$. \end{quote} \begin{quote} (II) If $2q - 2 + \# S \leq 0$, then there are no such families. \end{quote} Similar questions can be asked for smooth families of higher dimensional manifolds. (II) was recently verified for families of minimal surfaces of general type and for canonically polarized manifolds, by Kov\'acs \cite{K 2}, \cite{K 1}, \cite{K 3}, Migliorini \cite{M 1} and Zhang \cite{Z 1}. As a byproduct we reprove their result, but the reader familiar with their articles will recognize strong similarities between their and our approach. Unfortunately our method does not work for families whose relative dualizing sheaf is big and numerically effective on the smooth fibers. In this case, for non-isotrivial families over $\P^1$ or over an elliptic curve, Kov\'acs (Thesis, Utah, 1995, see \cite{K 2}) showed the existence of at least one degenerate fibre. In the higher dimensional case (I) might be too much to hope for. For fixed $B$ and $S$ there exist non-trivial deformations of families of abelian varieties over $B-S$ (see \cite{F}). So (I) splits up in two questions: boundedness and rigidity. To be more precise, let us fix some polynomial $h \in {\mathbb Q} [t]$ with $h ({\mathbb Z}) \subset {\mathbb Z}$. If ${\rm deg} (h) =2$, we define $M_h$ to be the moduli scheme of minimal surfaces $S$ of general type with $h(\mu) = \chi (\omega^{\mu}_{S})$, or allowing singularities, the moduli scheme of canonically polarized normal surfaces with at most rational double points and with Hilbert polynomial $h$ (see \cite{V 1}, for example). If ${\rm deg} (h) > 2$, we denote by $M_h$ the moduli scheme of canonically polarized manifolds, with Hilbert polynomial $h$. In both cases (I) should be replaced by two sub-problems: \begin{quote} (B) The non-trivial morphisms $B - S \to M_h$, which are induced by smooth projective maps $g_0 : Y_0 \to B - S$, are parameterized by some scheme of finite type. \end{quote} \begin{quote} (R) Under which additional conditions are the morphisms $B - S \to M_h$ in (B) rigid. \end{quote} The corresponding questions, for abelian varieties and their moduli scheme, were solved by Faltings \cite{F}. In this note we prove (B) for surfaces of general type, and for canonically polarized manifolds, in case $S = \emptyset$. The only obstruction to extend (B) to arbitrary families of canonically polarized manifolds, is the lack of a proof for the existence of relative minimal models for semi-stable families of such varieties over curves. We have nothing to contribute to problem (R). For smooth families $g : Y \to B$ of surfaces, i.e. for $S = \emptyset$, (B) has been proved by the first author. This note benefited from discussions between the first author and A. Parshin and between the second author and Qi Zhang. We thank both of them, and K. Oguiso for remarks and comments. \section{Families of canonically polarized manifolds and of surfaces of general type} Let $B$ be a curve and $Y$ a variety of dimension $n+1$, both non-singular, projective and defined over an algebraically closed field $k$ of characteristic zero. Let $g : Y \to B$ be a flat morphism with connected fibers $Y_b = g^{-1} (b)$. Let $S \subset B$ be a finite set of points and $D = g^{-1} (S)$. Frequently we will denote the divisors $\sum_{s \in S} s$ and $\sum_{s \in S} g^{-1} (s)$ by $S$ and $D$, as well. We want to study both, families of surfaces of general type and of canonically polarized manifolds, and $D$ will be supposed to be the set of bad fibers. To cover both cases we formulate the following assumption: \begin{assumption} \label{fam-ass} \ \begin{enumerate} \item[a)] $g|_{Y-D} : Y - D \>>> B - S$ is smooth and $\omega_{Y/B}|_{Y-D}$ is relatively semi-ample, i.e. for some $\nu \gg 0$ the natural map $$ \Phi_{\nu} : g^* g_* \omega^{\nu}_{Y/B} \>>> \omega^{\nu}_{Y/B} $$ is surjective over $Y-D$. \item[b)] The fibers $Y_b$ should be manifolds of general type and the $\nu$-canonical map should contract at most curves. Hence if $$ \pi_{\nu} : Y - D \twoheadrightarrow V \subseteq \P (g_* \omega^{\nu}_{Y/B} |_{B-S} ) $$ is the morphism, induced by $\Phi_{\nu}$, then $\pi_{\nu}$ should be birational and the maximal fibre dimension of $\pi_{\nu}$ should be one, for some $\nu \gg 0$, satisfying the condition a). \end{enumerate} \end{assumption} Recall from \cite{EV 2}, \S \ 2, or \cite{V 1}, \S \ 5, that for an invertible sheaf ${\mathcal L} $ on a normal projective variety $F$ the integer $e ({\mathcal L}) $ is defined to be the smallest positive integer $e$ for which the multiplier sheaf $\omega_F \left\{ - \frac{D}{e} \right\}$ is isomorphic to $\omega_F$, for all divisors $D$ of global sections of ${\mathcal L}$. For $e ({\mathcal L})$ to exist, one has to require $F$ to have at most rational singularities. Then, by \cite{EV 2}, 2.3, or \cite{V 1}, 5.11, 5.12 and 5.21, $e ({\mathcal L})$ exists and in some cases there are explicit bounds: \begin{lemma} \label{multiplier} Assume $F$ is projective with at most rational singularities and let ${\mathcal L}$ be an ample invertible sheaf on $F$. \begin{enumerate} \item[a)] If $F$ is non-singular and ${\mathcal L}$ very ample, then $e({\mathcal L}) \leq c_1 ({\mathcal L})^{\dim F} +1$. \item[b)] In general, let ${\mathcal L}$ be very ample and assume that there exists a desingularization $\sigma : F' \to F$ and an effective divisor $E$ such that $\sigma^* {\mathcal L} \otimes {\mathcal O}_{F'} (-E)$ is very ample. Then $e ({\mathcal L}) \leq c_1 ({\mathcal L})^{\dim F} +1$. \item[c)] If $F$ has rational Gorenstein singularities and if $Z = F \times \ldots \times F$ and ${\mathcal M} = \otimes pr^{*}_{i} {\mathcal L}$, then $e ({\mathcal M}) = e ({\mathcal L}). $ \end{enumerate} \end{lemma} Recall that $g: Y \to B$ is called isotrivial, if for some variety $F$, defined over $k$, there is a birational map $Y{\times_B} \overline{k(B)} \to F{\times_k} \overline{k(B)}$. In \cite{V 1}, 2.7 and 2.9, we defined a locally free sheaf ${\mathcal G} $ on $B$ to be numerically effective (or nef), if for all $\mu >0$ and for a point $p \in B$ the sheaf $$ S^{\mu} ({\mathcal G}) \otimes {\mathcal O}_B (p) $$ is ample. We will use: \begin{proposition} \label{pos}\ \begin{enumerate} \item[a)] If $g: Y \to B$ is a projective morphism between non-singular varieties, then $g_* \omega^{\nu}_{Y/B}$ is nef, for all $\nu >0$. \item[b)] If $g$ satisfies the assumptions made in \ref{fam-ass}, a), if the general fibre of $g$ is of general type and if $g$ is non-isotrivial, then $\kappa (\omega_{Y/B}) = n+1$ and ${\rm det} (g_* \omega^{\eta}_{Y/B})$ is ample, provided $g_* \omega^{\eta}_{Y/B} \neq 0$ and $\eta > 1$. \end{enumerate} \end{proposition} a) is a special case of \cite{V 2}, Theorem III, and b) can be found in \cite{V 3}, Theorem II. In fact, there is the ampleness of ${\rm det} (g_* \omega^{\eta}_{Y/B})$ is shown for some $\eta$, but as in \cite{EV 2}, 3.1, or \cite{V 1} this implies that $g_* \omega^{\eta}_{Y/B}$ is ample for all $\eta >1$. \begin{thm} \label{bounds} Let $g : Y \to B$ be a non-isotrivial morphism, satisfying the assumptions made in \ref{fam-ass}. Let us fix some $\nu >1$, such that \ref{fam-ass} a) and b) hold true. We write \begin{itemize} \item $q = g (B)$ for the genus of $B$. \item $s = \# S$ for the number of degenerate fibers. \item $e (\nu) = e (\omega^{\nu}_{F})$ for some general fibre $F$ of $g$. \item $r (\nu) = {\rm rank} (g_* \omega^{\nu}_{Y/B})$. \end{itemize} Then one has: \begin{enumerate} \item[a)] $2q - 2 + s > 0$ \item[b)] If $g : Y \to B$ is semi-stable, then $$n \cdot (2q - 2 +s) \cdot \nu \cdot e (\nu) \cdot r (\nu) \geq {\rm deg} (g_* \omega^{\nu}_{Y/B}).$$ \item[c)] In general $$ (n \cdot (2q - 2 +s) +s) \cdot \nu \cdot e (\nu) \cdot r (\nu) \geq {\rm deg} (g_* \omega^{\nu}_{Y/B}).$$ \end{enumerate} \end{thm} a) will follow from \ref{pos} and b). This part of theorem \ref{bounds} is due to Kov\'acs \cite{K 2}, \cite{K 1}, \cite{K 3} and to Migliorini \cite{M 1}. Qi Zhang \cite{Z 1} gave an elegant proof in case $B$ is an elliptic curve. His proof easily extends to $B = \P^1$, as he and the second author found out discussing his result. In this note we take up their approach, together with positivity properties of direct image sheaves, as stated in \cite{EV 2}. \begin{remark} \label{bounds 2} The constants $\nu$ and $r (\nu)$ are determined, using Matsusaka's big theorem, by the Hilbert polynomial of the fibers $Y_b$ for $b \in B - S$. For canonically polarized manifolds, $e(\nu) \leq \nu^n \cdot c_1 (\omega_F)^n +1$, as we have seen in \ref{multiplier}, a). For surfaces of general type, the canonical model $\tilde{F}$ of $F$ has $A - D - E$ singularities, and the number of $(-2)$-curves on $F$ is bounded by $\dim H^1 (F, \Omega^{1}_{F})$, a number which is constant in families. One can use \ref{multiplier}, b) to bound $e (\omega^{\nu}_{F}) =e (\omega^{\nu}_{\tilde{F}} )$. Such a bound exists for different reasons. By \cite{V 1}, the moduli space of normal canonically polarized surfaces with at most rational double points and with given Hilbert polynomial is quasi-projective, in particular it has only finitely many irreducible components. Moreover, Koll\'ar \cite{Ko} (see also \cite{V 1}, 9.25) constructed a finite covering $Z$ of the moduli scheme, together with a ``universal family''. By \cite{V 1}, 5.17, $e(\omega^{\nu}_{\tilde{F}})$ is bounded on each irreducible component of $Z$. Hence there exists some constant $e$ depending on $h$ with $e(\omega^{\nu}_{\tilde{F}})\leq e$, for all surfaces $F$ with Hilbert polynomial $h$. \end{remark} \section{Ampleness and vanishing theorems} Let $X$ be a projective manifold, $U \subset X$ an open dense submanifold, and let ${\mathcal L}$ be an invertible sheaf on $X$. \begin{definition} \label{def-sa}\ \begin{enumerate} \item[a)] ${\mathcal L}$ is semi-ample with respect to $U$ if for some $\eta > 0$ $$ \iota_{\eta} : H^0 (X, {\mathcal L}^{\eta}) \otimes_k {\mathcal O}_X \>>> {\mathcal L}^{\eta} $$ is surjective over $U$. \item[b)] ${\mathcal L}$ is $\ell$-ample and big with respect to $U$, if one can choose $\eta$ in a) such that the morphism $$ \Phi_{\eta} : U \twoheadrightarrow V \subset \P (H^0 (X, {\mathcal L}^{\eta})) $$ induced by $\iota_{\eta}$ is proper, birational and $$ {\rm Max} \{ \dim \Phi^{-1}_{\eta} (v) ; v \in V \} \leq l. $$ \end{enumerate} \end{definition} A slight modification of the argument used to prove 6.4 in \cite{EV 1}, yields a generalization of the Akizuki-Kodaira-Nakano vanishing theorem, similar to the one used in the proof of \cite{K 1}, 1.1 and \cite{K 3}, 1.1. \begin{proposition} \label{vanishing} Let ${\mathcal L}$ be $\ell$-ample and big with respect to $U$. Assume that for $\eta$ as in \ref{def-sa}, b), the image of $V$ of $\Phi_{\eta}$ allows a projective flat morphism $\gamma : V \to W$ to a non-singular affine variety $W$. Then there exists a blowing up $\tau : X' \to X$ with centers in $\Delta = X = U$, such that $\Delta' = \tau^{-1} (X - U)$ is a normal crossing divisor, and an effective divisor $\Gamma'$ with $\Gamma'_{{\rm red}} \leq \Delta'$, such that for all numerically effective invertible sheaves ${\mathcal N}$ and for $p +q < \dim X - {\rm Max} \{0,\ell-1\}$, $$ H^p (X', \Omega^{q}_{X'} ({\rm log} \ \Delta' ) \otimes \tau^* ({\mathcal L}^{-1} \otimes {\mathcal N}^{-1}) \otimes {\mathcal O}_{X'} (\Gamma')) =0. $$ \end{proposition} \begin{proof} If $\Delta = \emptyset$, this is \cite{EV 1}, 6.6. Hence we will assume that $\Delta \neq 0$, and fix some $\eta$ for which the assumption \ref{def-sa} b) on $\Phi_{\eta}$ hold true. Let $\tau : X' \to X$ be a blowing up, such that $X'$ is non-singular, $\Delta' = \tau^{-1} (\Delta)$ a normal crossing divisor and such that, for ${\mathcal L}' = \tau^* {\mathcal L}$, the image of $$ \iota'_{\eta} : H^0 (X', {\mathcal L}^{'\eta} ) \otimes_k {\mathcal O}_{X'} \>>> {\mathcal L}^{'\eta} $$ is an invertible sheaf, isomorphic to ${\mathcal L}^{'\eta}$ over $U' = \tau^{-1} (U)$. So ${\rm Im} (\iota'_{\eta}) = {\mathcal L}^{'\eta} (- \Gamma_1)$, for some divisor $\Gamma_1$, supported in $\Delta'$. Let $$ \Phi' : X ' \twoheadrightarrow Z \subset \P (H^0 (X', {\mathcal L}^{'\eta} (- \Gamma_1 ))) $$ be the induced morphism. $\Phi' |_{U'}$ is a proper morphism with image $V$. Let $I$ be the ideal sheaf of $Z - V$ and $$J =\Phi^{'*} I/_{\rm torsion}.$$ Blowing up again, we may assume that $J = {\mathcal O} (- \Gamma_2)$ for some divisor $\Gamma_2$ with $(\Gamma_2)_{{\rm red}} = \Delta'_{{\rm red}}$. For all $\mu \gg 0$ the sheaf ${\mathcal L}^{'\mu \cdot \eta} ( - \mu \cdot \Gamma_1 - \Gamma_2)$ will be generated by global sections. For some $\mu \gg 0$, the number $\mu \cdot \eta$ will not divide the multiplicity of any of the components of $\mu \cdot \Gamma_1 + \Gamma_2$. Allowing $\Delta'$ to have multiplicities, we thereby got to the following situation: \begin{assumption} \label{ass-van} For some $\eta$ and some normal crossing divisor $\Delta'$ with $$\Delta'_{{\rm red}} = X' - U' = (\Delta' - \eta \cdot \Bigl[ \frac{\Delta'}{\eta} \Bigr])_{{\rm red}},$$ the sheaf ${\mathcal L}^{'\eta} (-\Delta') = {\mathcal L}^{'\eta} \otimes {\mathcal O}_{X'} (- \Delta') $ is generated by global sections, the induced morphism $\Phi' : X' \twoheadrightarrow Z$ is birational and the fiber-dimension of $U' = \Phi^{'-1} (V) \to V$ is at most $\ell$. Moreover there exists a projective flat morphism $\gamma : V \to W$ to a non-singular affine variety $W$. Let ${\mathcal N}'$ be any numerically effective invertible sheaf on $X'$ and choose $[\frac{\Delta'}{\eta}] = \Gamma '$. \end{assumption} We will prove, by induction on $\dim X' - \dim W$, that the assumptions \ref{ass-van} imply \begin{gather} \label{eq-van} H^p (X', \Omega^{q}_{X'} ({\rm log} \ \Delta') \otimes {\mathcal L}^{'-1} \otimes {\mathcal N}^{'-1} \otimes {\mathcal O}_{X'} (\Gamma')) =0 \\ \mbox{for \ \ } \ p + q < \dim X - {\rm Max} \{ 0,r-1\}. \notag \end{gather} If $\dim X' = \dim V > \dim W$ we may replace $\eta$ by $\mu \cdot \eta$ and $\Delta'$ by $\mu \cdot \Delta'$ for $\mu \gg 0$. Thereby we are allowed to assume that $({\mathcal L}' \otimes {\mathcal N}')^{\eta} \otimes {\mathcal O}_{X'} (- \Delta')$ is generated by global sections, and that for the zero divisor $H$ of a general section of this sheaf $\Phi' (H) \cup V \to W$ is again projective and flat. Moreover $\Delta' + H$ is a normal crossing divisor and $\Gamma' |_{H} = [\frac{\Delta'|_{H}}{\eta}]$. Therefore $H$, ${\mathcal L}'|_H$ and $\Delta'|_H$ satisfy again the assumptions \ref{ass-van}. If $\dim V = \dim W$, we choose $H = 0$. In both cases, the morphism $$V- \Phi'(H) \>>> W$$ is affine and $V - \Phi' (H)$ is an affine variety. $({\mathcal L}' \otimes {\mathcal N}')^{\eta}$ has a section with zero divisor $\Delta' +H$. By \cite{EV 1}, \S \ 3, $$ ({\mathcal L}' \otimes {\mathcal N}')^{-1} \otimes {\mathcal O}_{X'} (\Gamma') =({\mathcal L}' \otimes {\mathcal N}')^{-1} \otimes {\mathcal O}_{X'} (\Bigl[ \frac{\Delta'+H}{\eta} \Bigr]) $$ has an integrable connection which satisfies the $E_1$-degeneration. The assumption $\Delta'_{{\rm red}} = (\Delta' - \eta \cdot [ \frac{\Delta'}{\eta} ])_{{\rm red}}$ allows to apply \cite{EV 1}, 4.12, and to obtain $$ H^p(X', \Omega^{q}_{X'} ({\rm log} \ \Delta' + H') \otimes ({\mathcal L}' \otimes {\mathcal N}')^{-1} \otimes {\mathcal O}_{X'} (\Gamma')) =0, $$ for $p+q < n - {\rm Max} \{ 0,\ell-1\}$. If $\dim V = \dim W$, we are done. If $\dim V > \dim W$ one uses the exact sequence $$ 0 \to \Omega^{p}_{X'} ({\rm log} \ \Delta') \to \Omega^{p}_{X'} ({\rm log} \ \Delta' + H') \to \Omega^{p-1}_{H'} ({\rm log} \ \Delta' |_{H'} ) \to 0 $$ to obtain \ref{eq-van} by induction. \end{proof} \section{The proof of theorem \ref{bounds}} We will assume that $D$ is a normal crossing divisor. Enlarging $S$ and $D$ we may assume that $2q - 2 + s \geq 0$, hence that $\omega_B (S)$ is nef. If $2q - 2 + s =0$, then either $B$ is elliptic and $g$ smooth, or $B = \P^1$ and $S = \{ b_1, b_2\}$. In the second case, there exists a finite covering $\P^1 \to \P^1$, totally ramified in $S$, such that the pullback family has stable reduction. Altogether, part a) of \ref{bounds} follows from \ref{pos} and part b) of \ref{bounds}. If the fibers of $Y -D \to B - S$ are canonically polarized manifolds, the next proposition follows from \cite{EV 2}, 2.4. Since we will use it for families of minimal models of surfaces of general type, as well, we will recall the proof. \begin{proposition} \label{ample} Under the assumptions made in \ref{bounds} let ${\mathcal N}$ be an invertible sheaf on $B$ with ${\rm deg} \ {\mathcal N} < {\rm deg} \ g_* \omega^{\nu}_{Y/B}.$ Then the sheaf $$ \omega^{\nu \cdot e (\nu) \cdot r(\nu)}_{Y/B} \otimes g^* {\mathcal N}^{-1} $$ is 1-ample and big with respect to $Y-D$. \end{proposition} \begin{proof} By the assumptions a) and b) in \ref{fam-ass} it is sufficient to show, that for some $\mu >0$ $$ S^{\mu \cdot r (\nu) \cdot e(\nu)} (g_* \omega^{\nu}_{Y/B} ) \otimes {\mathcal N}^{- \mu} $$ is generated by global sections, or that $$ S^{r (\nu) \cdot e(\nu)} (g_* \omega^{\nu}_{Y/B}) \otimes {\mathcal N}^{-1} $$ is ample. By definition of nef, this hold true, if the sheaf $$ S^{r(\nu) \cdot e(\nu)} (g_* \omega^{\nu}_{Y/B} ) \otimes {\rm det} (g_* \omega^{\nu}_{Y/B} )^{-1} $$ is nef. To this aim (see \cite{V 1}, 2.8) we can replace $B$ by a covering, unramified in $S$, and $Y$ by the pullback family. Thereby we may assume that $$ {\rm det} (g_* \omega^{\nu}_{Y/B} ) = {\mathcal A}^{e(\nu)} $$ for an invertible sheaf ${\mathcal A}$ on $B$. For $r = r (\nu)$ let $$ f : X = Y \times_B \ldots \times_B Y \>>> B $$ be the $r$-fold fibre product, let $\sigma: X' \to X$ be a desingularization, and $$f' = f \circ \sigma : X' \to B.$$ We write ${\mathcal M} = \sigma^* (\otimes^{r}_{i=1} pr^{*}_{i} \omega_{Y/B})$. The morphism $f$ is Gorenstein and the general fibre is non-singular. Hence there are natural injective maps \begin{gather} \otimes^r g_* \omega^{\nu}_{Y/B} = f_{*} \otimes^{r}_{i=1} pr^{*}_{i} \omega^{\nu}_{Y/B} \>>> f'_{*} {\mathcal M}^{\nu} \label{eq1} \\ \mbox{and} \ \ \ f'_* {\mathcal M}^{\nu -1} \otimes \omega_{X'/B} \>>> f_* \omega^{\nu}_{X/B} = \otimes^r g_* \omega^{\nu}_{Y/B}, \label{eq2} \end{gather} both isomorphisms on some open dense subset of $B$. (\ref{eq1}) induces $$ {\mathcal A}^{e(\nu)} = {\rm det} (g_* \omega^{\nu}_{Y/B}) \>>> \otimes^r g_* \omega^{\nu}_{Y/B} \>>> f'_* {\mathcal M}^{\nu} , $$ hence a section of ${\mathcal M}^{\nu} \otimes f^{'*} {\mathcal A}^{-e(\nu)}$ with zero-divisor $\Gamma$. Blowing up $X'$, with centers in a finite number of fibers, we may assume that the image ${\mathcal M}^{\nu} \otimes J$ of $$ f^{'*} \otimes^r g_* \omega^{\nu}_{Y/B} \>>> f^{'*} f'_* {\mathcal M}^{\nu} \>>> {\mathcal M}^{\nu} $$ is invertible. The ideal sheaf $J$ is trivial in a neighborhood of the general fibre. By \ref{pos} $g_* \omega^{\nu}_{Y/B}$ is nef, hence ${\mathcal M}^{\nu} \otimes J$ is nef, as well. Let us write ${\mathcal L} = {\mathcal M}^{\nu-1} \otimes f^{'*} {\mathcal A}^{-1}$. Then $$ {\mathcal L}^{e(\nu) \cdot \nu}(-\nu\cdot\Gamma) = {\mathcal M}^{e(\nu) \cdot \nu \cdot (\nu-1)}\otimes f^{'*} {\mathcal A}^{-\nu\cdot e(\nu)} \otimes {\mathcal O}_{X'}(-\nu\cdot\Gamma) = {\mathcal M}^{e(\nu) \cdot \nu \cdot (\nu-1) - \nu^2} $$ contains a nef subsheaf, isomorphic to ${\mathcal L}^{e(\nu) \cdot \nu}(-\nu\cdot\Gamma)$ in a neighborhood of the general fibre. Moreover, by \ref{pos}, b), $$\kappa ({\mathcal L}^{e (\nu) \cdot \nu} (-\nu \cdot \Gamma)) = \dim X'.$$ \cite{EV 2}, 1.7 implies that the sheaf $f'_* {\mathcal L} \otimes \omega_{X'/B} \{ - \frac{\Gamma}{e(\nu)} \}$ is nef. By the definition of $e(\nu)$ and by \ref{multiplier}, c), the natural inclusion $$ f'_* {\mathcal L} \otimes \omega_{X'/B} \Bigl\{ - \frac{\Gamma}{e (\nu)} \Bigr\} \>>> f'_* {\mathcal L} \otimes \omega_{X'/B} = f'_* ({\mathcal M}^{\nu-1} \otimes \omega_{X'/B}) \otimes {\mathcal A}^{-1} $$ is an isomorphism on some open dense subset of $B$. Using (\ref{eq2}) one obtains a nef subsheaf of $$ (\otimes^r g_* \omega^{\nu}_{Y/B} ) \otimes {\mathcal A}^{-1} $$ of full rank, hence the latter is nef, as well. Then $$ S^{e(\nu)} (\otimes^r g_* \omega^{\nu}_{Y/B}) \otimes {\mathcal A}^{-e(\nu)} = S^{e(\nu)} (\otimes^r g_* \omega^{\nu}_{Y/B} ) \otimes {\rm det} (g_* \omega^{\nu}_{Y/B})^{-1} $$ as well as its quotient $S^{e (\nu) \cdot r} (g_* \omega^{\nu}_{Y/B}) \otimes {\rm det} (g_* \omega^{\nu}_{Y/B} )^{-1}$ are nef. \end{proof} Let us return to the proof of \ref{bounds} b) and c). If $g$ is semi-stable, i.e. if $D$ is reduced, we choose $\delta=0$, otherwise $\delta=1$. Let us assume that \ref{bounds} b) or c) are wrong. Hence \begin{equation} \label{equality} (n\cdot (2q -2+s) +\delta \cdot s) \cdot \nu \cdot e (\nu) \cdot r(\nu) < {\rm deg} (g_* \omega^{\nu}_{X/B}) \end{equation} and \ref{ample} implies that for $$ {\mathcal A} = \omega_{B} (S)^n \otimes {\mathcal O}_B (\delta \cdot S) $$ the sheaf $\omega^{\nu \cdot e (\nu) \cdot r(\nu)}_{Y/B} \otimes {\mathcal A}^{-\nu \cdot e(\nu) \cdot r (\nu)}$, hence ${\mathcal L} = \omega_{Y/B} \otimes g^* {\mathcal A}^{-1}$, is 1-ample and big with respect to $Y-D$. Since we assumed $\omega_{B} (S)$ to be nef, \ref{vanishing} implies that: \begin{claim} \label{van-am} There exists a blowing up $\tau : X \to Y$ with centers in $D$, such that $\Delta = \tau^* D$ is a normal crossing divisor, and there exists an effective divisor $\Gamma$, supported in $\Delta$, with $$ H^p (X, \Omega^{q}_{X} ({\rm log} \ \Delta) \otimes \tau^* \omega^{-1}_{Y/B} \otimes g^* (\omega_B (S)^{n-m} \otimes {\mathcal O}_{B} (\delta \cdot S)) \otimes {\mathcal O}_X (\Gamma)) = 0 $$ for $p + q < \dim X = n+1$ and for all $m \geq 0$. \end{claim} The morphism $f = g \circ \tau$ is smooth outside of $\Delta$, and one has an exact sequence of locally free sheaves \begin{equation} \label{seq1} 0 \to f^* \omega_B (S) \to \Omega^{1}_{X} ({\rm log} \Delta) \to \Omega^{1}_{X/B} \to 0. \end{equation} Comparing the determinants one finds $$\Omega^{n}_{X/B} = {\rm det} (\Omega^{1}_{X/B}) = \omega_X (\Delta_{{\rm red}}) \otimes f^* \omega_B (S)^{-1} = \omega_{X/B} (\Delta_{{\rm red}} - \Delta).$$ \begin{claim} \label{effective} $$ H^0 (X, \Omega^{n}_{X/B} \otimes \tau^* \omega^{-1}_{Y/B} \otimes g^* {\mathcal O}_B (\delta \cdot S)) \neq 0 $$ \end{claim} \begin{proof} If $\delta =1$, this holds true with $\tau^* \omega^{-1}_{Y/B}$ replaced by the smaller sheaf $\omega^{-1}_{X/B}$, since $$ \Omega^{n}_{X/B} \otimes \omega^{-1}_{X/B} \otimes g^* {\mathcal O}_B (S) = {\mathcal O} (\Delta_{{\rm red}}). $$ If $\delta =0$, i.e. if $g$ is semi-stable, then $\Omega^{n}_{Y/B} = \omega_{Y/B}$. By \cite{EV 1}, 3.21, one has an inclusion $\tau^* \Omega^{1}_{Y} ({\rm log} \ D) \to \Omega^{1}_{X} ({\rm log} \ \Delta)$, hence $\tau^* \omega_{Y/B} = \tau^* \Omega^{n}_{Y/B} \subset \Omega^{n}_{X/B}$. \end{proof} The $m$-th wedge product, applied to the sequence (\ref{seq1}) induces an exact sequence \begin{equation} 0 \to f^* \omega_B (S) \otimes \Omega^{m-1}_{X/B} \to \Omega^{m}_{X} ({\rm log} \ \Delta) \to \Omega^{m}_{X/B} \to 0 . \tag{$\Sigma^{\bullet}_{m}$} \end{equation} \ref{van-am} implies, that the $(n - m)$-th cohomology of the sheaf in the middle of the sequence $$ \Sigma^{\bullet}_{m} \otimes \tau^* \omega^{-1}_{Y/B} \otimes g^* \omega^{n-m}_{B} ((n-m + \delta) \cdot S) \otimes {\mathcal O}_X (\Gamma) $$ is zero, hence $$ H^{n-m} (X, \Omega^{m}_{X/B} \otimes \tau^* \omega^{-1}_{Y/B} \otimes g^* \omega^{n-m}_{B} ((n-m + \delta) \cdot S) \otimes {\mathcal O}_X (\Gamma)) $$ injects into $$ H^{n-m+1} (X, \Omega^{m-1}_{X/B} \otimes \tau^* \omega^{-1}_{Y/B} \otimes g^* \omega^{n-m+1}_{B} ((n-m+1+\delta) \cdot S) \otimes {\mathcal O}_X (\Gamma)). $$ Altogether $$ H^0 = H^0 (X, \Omega^{n}_{X/B} \otimes \tau^* \omega^{-1}_{Y/B} \otimes g^* {\mathcal O}_B (\delta \cdot S) \otimes {\mathcal O}_X (\Gamma)) $$ is a subspace of $$ H^n = H^n (X, \tau^* \omega^{-1}_{Y/B} \otimes g^* \omega^{n}_{B} ((n+\delta) \cdot S) \otimes {\mathcal O}_X (\Gamma)). $$ Using \ref{van-am} again, one finds $H^n$ and thereby $H^0$ to be zero. Since $\Gamma$ is effective, this contradicts \ref{effective}. Hence the inequality stated in (\ref{equality}) does not hold true, and we obtain \ref{bounds} b) and c). \section{Moduli schemes and boundedness} Let ${\mathcal M}_h$ denote the moduli functor of minimal surfaces of general type, if ${\rm deg} \ h = 2$, and of canonically polarized manifolds, if ${\rm deg} \ h > 2$, both times with fixed Hilbert polynomial $h\in {\mathbb Q}[t]$. The corresponding moduli scheme is denoted by $M_h$. If ${\rm deg} \ h = 2$, i.e. in the surface case, Koll\'ar and Shepherd-Barron \cite{KS} defined stable surfaces, Alexeev \cite{Al} proved that the index of the singularities is bounded in terms of the coefficients of the Hilbert polynomial, which implies by \cite{Ko} that $M_h$ has a compactification $\bar{M}_h$, parameterizing families of stable surfaces (see also \cite{V 1}, section 9.6). For ${\rm deg} \ h \geq 2$, those results were generalized by Karu \cite{KK}, assuming the minimal model conjecture for semi-stable families of $n$-folds over curves $(MMP (n+1))$. Let again $\bar{M}_h$ be the compactification and $\bar{{\mathcal M}}_h$ the corresponding moduli functor. The existence of $\bar{M}_h$ implies that for some $N_0$, depending on $h$, the reflexive hull $\omega^{[N_0]}_{X}$ of $\omega^{N_0}_{X}$ is invertible for all $x \in \bar{{\mathcal M}}_h (k)$. In both cases, for some $\eta\gg 0$ the sheaf $\omega^{[\eta]}_{X}$ is very ample $({\rm deg} \ h > 2)$ or semi-ample and the induced morphism the contradiction of $(-2)$ curves $({\rm deg} \ h = 2)$. Koll\'ar \cite{Ko} (see also \cite{V 1}) has shown, that for $\eta\gg 0$ and for some $p >0$ there exists a very ample invertible sheaf $\lambda$ on $\bar{M}_h$ with: \begin{quote} For $f: X \to Z \in \bar{{\mathcal M}}_h (Z)$ and for the induced morphism $\varphi : Z \to \bar{M}_h$, $$\varphi^* \lambda = {\rm det} (f_* \omega^{[\eta]}_{X/Z})^p.$$ \end{quote} \begin{corollary} \label{b_s} Assume ${\rm deg} \ h = n = 2$. Let $B$ be a projective non-singular curve, $S \subset B$ a finite subset, and let $g_0 : Y_0 \to B - S$ be a smooth projective morphism, whose fibers are surfaces of general type with Hilbert polynomial $h$. Let $\Phi : B \to \bar{M}_h$ be the induced morphism. Then ${\rm deg} \ \Phi^* \lambda$ is bounded above by a constant, depending on $h, g (B)$ and $\# S$. \end{corollary} \begin{addendum} \label{b_cp} Assuming $MMP ({\rm deg} \ h + 1 )$, \ref{b_s} remains true for ${\rm deg} \ h > 2$ and for $g_0 : Y_0 \to B - S$ a family of canonically polarized manifolds with Hilbert polynomial $h$. \end{addendum} \begin{proof} Choose a non-singular projective compactification $Y$ of $Y_0$ such that $g_0$ extends to $g: Y \to B$. The assumptions \ref{fam-ass} hold true for $\nu = \eta$. By \ref{bounds} the degree of ${\rm det} (g_* \omega^{\eta}_{Y/B})$ is smaller than a constant depending on $h, g (B)$ and $\# S$. There exists a finite covering $\gamma : C \to B$ and $f: X \to C \in \bar{{\mathcal M}}_h (C)$ which induces $$ C \> \gamma >> B \> \Phi >> \bar{M}_h . $$ We may assume, in addition, that $f$ has a semi-stable model $f' : X' \to C$, with $X'$ non-singular. By the definition of stable surfaces in \cite{KS} (or of stable canonically polarized varieties in \cite{KK}), $$ f_* \omega^{[\eta]}_{X/C} = f'_* \omega^{\eta}_{X'/C}. $$ \cite{V 2}, 3.2, gives an injective map $$ f'_* \omega^{\eta}_{X'/C} \to \gamma^* g_* \omega^{\eta}_{Y/B} $$ which is an isomorphism over $\gamma^{-1}(B-S)$. Hence $$ {\rm deg} \ \Phi^* \lambda = \frac{{\rm deg} (f_* \omega^{[\eta]}_{X/C} )^p}{{\rm deg} \ \gamma} \leq {\rm deg} (g_* \omega^{\eta}_{Y/B})^p $$ is bounded, as well. \end{proof} Let us denote by ${\rm \bf H} = {\rm \bf Hom} ((B, B-S), (\bar{M}_h , M_h))$ the scheme parameterizing morphism $\Phi : B \to \bar{M}_h$ with $ \Phi (B-S) \subset M_h. $ Since $\lambda$ is ample, \ref{b_s} and \ref{b_cp} imply: \begin{corollary} \label{b_maps} Under the assumptions made in \ref{b_s} (or \ref{b_cp}) there exists a subscheme $T \subset {\rm \bf H}$, of finite type over $k$, which contains all points $[\Phi] \in {\rm \bf H}$, induced by smooth morphisms $$ g_0 : Y_0 \to B -S \in {\mathcal M}_h (B - S). $$ \end{corollary} Without assuming the minimal model conjecture $MMP ({\rm deg} \ h + 1)$ \ref{b_cp} and \ref{b_maps} remain true, if $S = \emptyset$. In fact, the existence of the very ample sheaf $\lambda$ on $M_h$ has been shown in \cite{V 1}. For the corresponding embedding $M_h \to \P^m$ choose $\bar{M}_h$ to be the closure of $M_h$ in $\P^m$. Then for complete curves $B$ and for morphisms $B \to \bar{M}_h$ with image in $M_h$, the arguments used to prove \ref{b_cp} remain valid. In \cite{V 1}, section 8.5, one finds the definition of a moduli functor ${\mathcal D}^{[N_0]}_{h}$ of canonically polarized normal varieties with canonical singularities of index $N_0$. Kawamata \cite{Kaw} has shown that this moduli functor is locally closed. Unfortunately it is not known to be bounded, i.e. whether Matsusaka's big theorem holds true. As in \cite{V 1}, 1.20, one can enforce boundedness by considering the submoduli functor ${\mathcal D}^{[N_0] (M)}_{h}$ with $$ {\mathcal D}^{[N_0] (M)}_{h} (k) = \{ X \in {\mathcal D}^{[N_0]}_{h} ; \omega^{[N_0 \cdot M]}_{X} \ \ \mbox{very ample} \}. $$ Then there exists a moduli scheme $D^{[N_0](M)}_{h}$ for ${\mathcal D}^{[N_0](M)}_{h}$. For $\nu = N_0 \cdot M$ and some $p > 0$ there exists again a very ample invertible sheaf $\lambda $ on $D^{[N_0](M)}_{h}$ which induces $$ {\rm det} (f_* \omega^{[\nu]}_{X/Z})^p \mbox{ \ \ \ for \ \ \ } f: X \to Z \in {\mathcal D}^{[N_0](M)}_{h} (Z). $$ Choosing $\bar{M}_h$ again as the closure of $D^{[N_0](M)}_{h}$ for an embedding $\phi:D^{[N_0](M)}_{h}\to\P^m$ with $\phi^*{\mathcal O}_{\P^m}(1)=\lambda$, one obtains: \begin{corollary} \label{b_maps2} For ${\rm deg} \ h > 2$, there exists a subscheme $T \subset {\rm \bf H}$, of finite type over $k$, which contains all points $[\Phi] \in {\rm \bf H}$, induced by morphisms $$ g: Y \to B \in {\mathcal D}^{[N_0](M)}_{h} (B), $$ smooth over $B - S$. \end{corollary} \bibliographystyle{plain}
\section{Introduction} Electron-phonon interactions directly affect numerous physical properties in solids. It is generally the case, however, that electrons and phonons are considered independently, with each system a weak perturbation on the other. This distinction is usually clear in consideration of collective modes of the electron system. Plasma oscillations, in particular, have been studied extensively by considering only the effects of Coulomb interactions between electrons. Of particular interests to this work is the limit in which the assumption of independent electron and phonon systems is not valid. If the interactions between electrons and phonons are strong enough, phonon mediated electron-electron (e-e) scattering could support a new collective mode. A recent theoretical calculation\cite{bonsager} proposed this type of electron-phonon collective mode in a double layer two-dimensional electron gas (2DEG) system, where both Coulomb and phonon mediated interactions are important. It was shown that a mode similar to plasmons emerges when the excitation energy $\omega$ approaches $sq$, where $s$ is the sound velocity and $q$ is the wavevector. The work predicted that this mode could be detected in measurements of phonon-mediated e-e scattering between two 2DEG layers. This coupled mode was argued to dominate and enhance such scattering when the phonon mean free path, $l_{ph}$, is larger than a critical value, $l_{c}$. The unexpectedly large magnitude observed in experiments\cite{gramila,nara,gramila2} provides clear motivation for investigating its existence, which would indicate a breakdown of the assumption of independent systems. When $l_{ph} < l_{c}$, this new mode was predicted to be negligible in the scattering process, and electrons and phonons could instead be treated independently. It was also established that the layer spacing and density dependences would show different behavior in the regimes of strongly coupled or independent electron-phonon systems. For small $l_{ph}$, i.e. the independent system regime, the scattering decreases logarithmically with increasing layer spacing, $d$, until $d \sim l_{ph}/2k_{F}L$, beyond which it decreases exponentially ($k_{F}$ and $L$ are the Fermi wavevector and the width of the quantum well respectively). In the coupled mode regime, i.e. large $l_{ph}$, the scattering also varies logarithmically with $d$ but exhibits a local maximum at $d \sim \sqrt{l_{ph}/k_{F}}$, beyond which it also rapidly declines. The dependence on the density ratio of the two layers is expected to show almost identical behavior in either regime, characterized by a peak at matched density. It was argued that distinct behaviors were present in the dependence on total density with the individual densities matched. For temperatures, $T$, between 2 K and 4 K, the scattering is predicted to increase with density in the coupled mode regime, while it remains nearly constant or decreases in the short mean free path regime. In both regimes, the scattering decreases with density at 1 K. The matched density dependence provides the central experimental test for the existence of the coupled collective mode. In this paper, we report measurements of phonon-mediated drag examining the dependence on temperature, layer spacing, relative density, and matched densities. In addition to testing for the existence of the proposed coupled mode, these measurements explore the general properties of phonon drag in detail. New measurements on remotely spaced layers clearly confirm the fundamental elements of the phonon scattering process which is the basis for phonon drag, including the dominance of $2k_{F}$ scattering. These new measurements include density dependence measurements which are the first performed in samples in which all Coulomb scattering is absent. Additional measurements probe specific aspects of phonon drag, focusing on elements related to the existence of a coupled electron-phonon mode. The layer spacing dependence confirm the logarithmic behavior predicted theoretically for closely spaced layers. The value of $l_{ph}$ determined by fitting to the theoretical layer spacing dependence is found to be small and within the independent system regime. The dependence on total matched density for a sample in this regime closely mimics theoretical predictions for $l_{ph} < l_{c}$, providing additional evidence for independent systems. Contradictory evidence is provided by the magnitude of the drag signal, which is substantially larger than predicted for small $l_{ph}$, and by the value determined for $l_{ph}$ from the density dependence, which is much smaller than that found in thermal conductivity measurements on other samples\cite{eisenstein}. These apparent contradictions raise questions regarding phonon drag which have not been addressed in theoretical studies to date. Measurements of phonon drag in samples with the largest layer spacing, 5200 \AA, raise additional questions regarding the underlying mechanism of phonon drag. The overall magnitude of the measured signal, in agreement with earlier measurements\cite{nara,gramila2}, lies well below the logarithmic dependence which applies for smaller layer spacings. Measurements of the dependence on matched densities at various temperatures, performed to test whether the deviation may be related to the independent or strongly coupled regime, provides no clear guidance, as the dependence observed is inconsistent with the predictions for either regime. In the following section, the background of phonon drag and experimental details are presented. General properties of phonon-mediated drag are explored through measurements discussed in the following section. The subsequent section directly examines the possible existence of the coupled electron-phonon mode by measurements of the layer spacing dependence and the dependence on total density, and through comparison of these measurements with existing theoretical calculations. Finally, we present the results for a very remotely spaced layer sample and discuss several features which cannot be explained by current theory. \section{Phonon drag: background and technique} Interactions between electrons are known to have dramatic effects on the properties of 2DEG systems. While direct measurement of these interactions is generally difficult, it was proposed by Pogrebinskii \cite{pogrebinskii} and later by Price\cite{price} that the e-e interactions might be measurable in a double layer, independent, 2D electron system due to the presence of interactions between layers. Such measurements, termed electron drag, have been achieved in the double layer 2DEG systems\cite{gramila}, as well as in a 2D/3D electron system\cite{solomon} and the electron-hole double layer system\cite{sivan}. In electron drag, a current, $I$, is driven through one of two closely spaced but electrically isolated 2DEG layers. Scattering between electrons in opposing layers results in a transfer of momentum from the current carrying layer to the other, effectively dragging the electrons in the second layer. If current is not permitted to flow from the second layer, charge accumulates at one end resulting in a voltage, $V_{D}$. The voltage increases until the force of the electric field due to charge accumulation balances the effective drag force resulting from interlayer interactions. Measurements of the drag resistivity, $\rho_{D}$, the ratio of the induced voltage to the drive current per square, has been shown\cite{gramila} to be directly related to an interlayer electron-electron scattering rate, $\tau_{D}^{-1}$, through a simple Drude style relation\cite{gramila} $\rho_{D}=m^{*}/ne^{2}\tau_{D}$, where $m^{*}$ is the effective mass and $n$ is the electron density. This technique thus provides quantitative evaluation of e-e scattering. Prior studies have established two distinct contributions to the interlayer e-e interactions: direct Coulomb scattering\cite{gramila,coulomb} and phonon exchange scattering\cite{bonsager,nara,gramila2,rubel,tso,zhang}. For very closely spaced layers, the Coulomb contribution dominates, resulting in $\rho_{D}$ that varies as $T^{2}$ at low temperatures. This contribution to interlayer scattering has a strong layer spacing dependence\cite{gramila,coulomb}, $d^{-4}$. This results from a cutoff in the maximum scattering wavevector at the inverse of the layer spacing and modifications to screening efficiency as the layer spacing is changed. The strong layer spacing dependence permits phonon scattering to dominate $\rho_{D}$ for remotely spaced layers. However, the contribution of phonon scattering is still evident even for closely spaced layers through a peak near 2 K when $\rho_{D}$ is scaled by $T^{2}$, emphasizing deviations from the quadratic dependence of Coulomb scattering. While previous studies clearly established the existence of a phonon mediated e-e scattering process and the important role of $2k_{F}$ scattering, questions regarding details of the underlying mechanism, especially in regard to the magnitude of $\rho_{D}$, are still not conclusively understood. Gramila {\it et al.}\cite{gramila} showed that real phonon exchange, while generating approximately the right temperature dependence, could not explain the magnitude of $\rho_{D}$ and proposed a virtual phonon exchange\cite{nara}. Tso {\it et al.}\cite{tso} included virtual phonon exchange in a calculation of $\rho_{D}$ and obtained reasonable agreement in both magnitude and temperature dependence, but questions have been raised\cite{bonsager} about the electron-phonon coupling used in their work. Zhang {\it et al.}\cite{zhang} used a first-principles approach but found a phonon mediated drag which has the same layer spacing dependence as the Coulomb interaction, inconsistent with prior experimental results\cite{nara,gramila2}. Virtual phonons have also been considered by Badalyan {\it et al.}\cite{badalyan}. The existence of the coupled electron-phonon collective mode was proposed by B\o nsager {\it et al.}\cite{bonsager} as a possible basis for the strong scattering if the phonon mean free path is large. The aim of this work is to firmly establish the essential physical elements of phonon drag and to test for the existence of the coupled electron-phonon collective mode. The samples used in the study are GaAs/Al$_{0.3}$Ga$_{0.7}$As double quantum well structures grown by Molecular Beam Epitaxy. Two 2DEG layers formed in 200 \AA\ wide quantum wells are separated by a potential barrier. Samples with barrier thicknesses of 225, 500, 2400, and 5000 \AA\ were used, corresponding to well center-to-center spacings, $d$, of 425, 700, 2600, and 5200 \AA, respectively. The density of each 2DEG layer as grown is approximately $1.5 \times 10^{11}$ cm$^{-2}$ with mobilities $\sim 2 \times 10^{6}$ cm$^{2}$/V s. An active region, in which interactions are probed, was defined through standard photolithographic techniques in a mesa approximately $440\ \mu$m long and $40\ \mu$m wide. Electrical connection was provided through Au/Ge/Ni ohmic contacts. The ability to separately contact the individual 2DEG's was provided through a gating technique\cite{contact} which uses both front and backside aluminum Schottky gates. The use of an interlayer bias and application of a voltage to an overall top gate allowed the density of the layers to be altered, as calibrated by Shubnikov-de Haas measurements. In the drag measurement itself, low-frequency, low-excitation currents of $\sim$ 100 nA were used, with the nV or smaller drag signals detected using established drag techniques\cite{gramila}. \section{General Properties of Phonon Drag} This section presents new measurements of the temperature and density dependences of phonon drag in a sample with $d=$ 2600 \AA. General features of the phonon scattering process are discussed. These include a temperature dependence which is characteristic of an electron-phonon scattering processes and the dominant role of $2k_{F}$ scattering. \subsection{Temperature dependence} Measured drag resistivity, $\rho_{D}$, scaled by $T^{2}$ as a function of temperature for the 2600 \AA\ spacing sample is shown in Fig.~\ref{1}. For this layer spacing, the strong $d$ dependence of the Coulomb interaction makes its contribution to $\rho_{D}$ completely negligible. The magnitude of Coulomb scattering can be estimated based on its $d^{-4}$ spacing dependence\cite{gramila,coulomb} and its magnitude as determined in other samples\cite{nara,gramila2}. The estimate for the Coulomb component of $\rho_{D}/T^{2}$ ($\sim 0.5\ \mu\Omega / \Box$K$^{2}$) is three orders of magnitude less than the phonon contribution. The most striking feature in these data is a clear transition which occurs near 2 K from a strong temperature dependence at lower temperatures to a dependence weaker than quadratic at higher temperatures. The general character of the transition is identical to that observed in other samples, for both closely spaced layers\cite{gramila,nara} where Coulomb scattering is present and for a sample even more remotely spaced\cite{nara,gramila2}. This transition is a general feature of 2DEG phonon scattering and was first observed in the Bloch-Gr\"{u}neisen transition of acoustic-phonon-limited mobility in 2DEG's\cite{stormer}. There are two essential elements which contribute to this change in temperature dependence. The first is a sharp cutoff for the wavevectors of phonons which can be scattered by the electron system. At low temperatures, only phonons with wavevector $q < 2k_{F}$ can be thermally excited. All such phonons can participate in the scattering process. As the temperature increases, progressively larger wavevector phonon states become occupied. The increase in wavevector provides both a larger momentum transfer and access to more of the electron phase space, leading to a strong temperature dependence for phonon scattering. This increase continues until phonons with $q \sim 2k_{F}$ become thermally excited. Low energy excitations of the electron system are absent for changes in $q > 2k_{F}$, so phonons with these $q$'s cannot scatter and conserve both energy and momentum. The transition to a weaker temperature dependence resulting from this cutoff will scale with the size of the Fermi surface. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig1.ps}} \end{center} \caption{Drag resistivity divided by $T^{2}$ versus temperature for $d=$ 2600 \AA. The densities of both layers are matched at $1.53 \times 10^{11}$ cm$^{-2}$. Coulomb scattering is completely negligible in this data. A transition from a strong to a weak temperature dependence occurs in $\rho_{D}/T^{2}$ near 2 K. The behavior reflects the important role of $2k_{F}$ scattering, which is characteristic of electron-phonon scattering.} \label{1} \end{figure} A second element which contributes to the behavior is the enhanced phase space available for large angle scattering in the electron system, which diverges\cite{smith} at zero temperature for a scattering wavevector equal to $2k_{F}$. The divergence allows $2k_{F}$ phonons to dominate scattering, even at temperatures below the energy of such phonons. The important role of $2k_{F}$ scattering is illustrated in Fig.~\ref{2}. It shows the relative net momentum $P_{Q}$ transfered to the phonon system per unit time calculated for a single current carrying electron layer as a function of wavevector parallel to the 2DEG, $Q$. The calculation is directly related to earlier approaches for calculating phonon scattering\cite{gramila2,stormer}. At 7 K, the momentum transfer rate is dominated by phonons in the vicinity of $Q=2k_{F}$ for both deformation potential and piezo-electric coupling. As the temperature is reduced to $\sim$ 3 K, this peak sharpens, with $\sim 2k_{F}$ scattering continuing to dominate the momentum transfer process. Even at 1.02 K, a fraction of the energy of a $2k_{F}$ phonon, $2k_{F}$ phonons continue to represent a significant portion of the total momentum, although low $Q$ phonons also become important. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3.5in \epsffile{fig2.ps}} \end{center} \caption{Calculations of the net momentum transfer rate from a single-layer electron system to phonons for both deformation potential (DP) and piezo-electric (PE) coupling at various temperatures as a function of Q/k$_{F}$. Rates are plotted relative to the DP coupling results at 7.0 K. The cutoff in phonon scattering at $2k_F$ is evident with a strong contribution from $2k_F$ scattering for all temperatures.} \label{2} \end{figure} This predominance of $2k_{F}$ scattering is critical for determining the temperature which characterizes the change in temperature dependence. If there were simply a cutoff at $2k_{F}$, the transition would be expected to occur when the characteristic phonon wavevector equals $2k_{F}$; at a temperature of $T_{c}=2k_{F} \hbar s/k_{B}$, where $s$ is the sound velocity. By using the longitudinal acoustic-phonon velocity for $s$, the transition temperature $T_{c} \sim$ 7.8 K is obtained; the transverse sound velocity yields $T_{c} \sim$ 4.6 K. However, the strong contribution of $2k_{F}$ scattering still dominates well below this temperature. This dominance permits the scattering rate to be determined primarily by the thermal occupancy of $2k_{F}$ phonons, even at temperatures well below the expected transition. Indeed, the characteristic shape for the drag momentum transfer of Fig. \ref{1} can be reasonably well represented by simply the thermal occupancy of the $2k_{F}$ phonons. At the lowest temperatures, where the contribution of $2k_{F}$ phonons is lost, a characteristic strong power law dependence is recovered. \subsection{Density dependence} Evidence for the dominance of $2k_{F}$ phonons in the drag process is available in the dependence of $\rho_{D}$ on density. The dependence of $\rho_{D}$ on the relative densities of the two layers is shown in Fig. \ref{3}. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig3.ps}} \end{center} \caption{Measured drag resistivity versus layer density ratio for $d=$ 2600 \AA\ at four different temperatures. The top layer density, n$_{top}$, is varied while the bottom layer density n$_{bot}$ is fixed at $1.53 \times 10^{11}$ cm$^{-2}$. The peak at matched densities confirm the importance of $2k_F$ scattering, which persists even to the lowest temperatures. Coulomb scattering is entirely negligible in this sample.} \label{3} \end{figure} In this measurement, the density of one layer is fixed while the density of the other is changed using an overall top gate. Measurements at four different temperatures for the 2600 \AA\ spacing sample are shown. The key aspect of this measurement is a clear maximum in the drag resistivity when the densities of both layers are matched. At matched densities, each electron system has an identical Fermi surface size so the dominant emission of $2k_{F}$ phonons in one layer matches those phonons most likely to be absorbed by the other. When the densities of two layers are mismatched, $2k_{F}$ phonons of the higher density layer have $q$'s too large to permit absorption in the other layer. The substantial reduction in phonon exchange at $q$'s other than $2k_{F}$ cause a significant reduction in $\rho_{D}$ as compared to matched densities. While the peak at matched density has been observed earlier in a $d=$ 425 \AA\ sample\cite{gramila2} at 2.3 K and in a $d=$ 500 \AA\ sample\cite{rubel} at 4.2 K, the behavior for both of those measurements included a significant contribution of Coulomb scattering. The data presented here can be directly compared to calculations of phonon drag without the significant added uncertainty arising from a subtraction of Coulomb scattering. The elimination of Coulomb scattering also permits the observation of a peak in $\rho_{D}$ down to previously unmeasured temperatures, where the signal is well below a nanovolt. This observation of a maximum in $\rho_{D}$ at matched density for this low temperature clearly supports the dominant role of $2k_{F}$ scattering, even for a temperature below the peak in $\rho_{D}/T^{2}$ of Fig.~\ref{1}, and well below the temperature corresponding to the energy of a $2k_{F}$ phonon. Further evidence for phonon scattering across the Fermi surface can be obtained from the temperature dependence measured for various matched layer densities. Measurements of $\rho_{D}/T^{2}$ are shown in Fig. \ref{4} for three different densities with $n$ changing by a factor of 2. This measurement directly tests the assumption that the observed transition temperature is determined by the size of the Fermi surface. Since the size of $2k_{F}$ scales as $\sqrt n$, decreasing the density should move this transition to lower temperatures. Both the observed direction and magnitude of this change are in agreement with a dominance of phonon drag by $2k_{F}$ phonon exchange. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig4.ps}} \end{center} \caption{The drag resistivities divided by $T^{2}$ versus temperature with different matched densities for $d=$ 2600 \AA. The number beside each plot is the matched density (in $10^{11}$ cm$^{-2}$). The change in peak position reflects changes in the size of the Fermi surface.} \label{4} \end{figure} \section{Coupled Electron-phonon mode} In this section, measurements of layer spacing and matched density dependence are presented and compared to theoretical predictions. Fits to the phonon mean free path determined by this comparison, detailed density dependence measurements, and the overall magnitude of $\rho_{D}$ are considered as tests of the existence of a coupled electron phonon collective mode. \subsection{Layer spacing dependence} It has been experimentally established that phonon drag has a weak layer spacing dependence\cite{nara,gramila2}. A specific form of the dependence has not been studied in detail, however, until the recent calculation by B\o nsager {\it et al.}\cite{bonsager}. Their work predicts that in the limit of small phonon mean free path, drag varies as $ln(d_{a}/d)$ until $d$ reaches $d_{a}$, which equals $l_{ph}/2k_{F}L$. For $d > d_{a}$, $\rho_{D}$ decreases more abruptly as $(d_{a}/d)exp(-d/d_{a})$. In the coupled mode regime, which requires large phonon mean free paths, drag was also expected to decrease logarithmically as long as $d$ is less than $d_{B}=(1+q_{TF}/2k_{F})/16k_{F}C_{DP}$, where $q_{TF}$ is the Thomas-Fermi wavevector and $C_{DP}$ is the dimensionless deformation potential coupling constant as defined in eq.~(30) of their paper. For the parameters typical of our samples, $d_{B}$ is approximately 5000 \AA. For $d$ greater than $d_{B}$, it was found that $\rho_{D}$ has a local maximum at $d \sim \sqrt{l_{ph}/k_{F}}$, beyond which the electron-phonon coupled mode, which involves both electron layers, begins to separate into two independent modes, substantially reducing interlayer drag. Measurements of the layer spacing dependence of $\rho_{D}$ provides a test of both this general theoretical approach, as a logarithmic spacing dependences should be observed, and as an indication of the existence of the electron-phonon collective mode. Measurements of $\rho_{D}$ for five different values of $d$ are shown in Fig.~\ref{5}. Data for one layer spacing, $d=$ 375 \AA, is reproduced \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig5.ps}} \end{center} \caption{Drag resistivities divided by $T^{2}$ versus temperature for different layer spacings. Data for d=375~\AA\ is from Ref.~2. Coulomb scattering, which can be represented by a horizontal line, is present for the three smallest layer spacings; phonon scattering is evident for all layer spacings. } \label{5} \end{figure} from the work of Ref.~2; current measurements for $d=$ 425, 700, and 5200 \AA\ generally confirm prior experimental results. The 2600 \AA\ spacing has not been previously measured. The data is plotted as $\rho_{D}/T^{2}$, where Coulomb contributions can be represented by a horizontal line. Visual inspection shows that the phonon contribution to $\rho_{D}$, which can be estimated by the deviation from a quadratic temperature dependence, shows little variation for layer spacing 700 \AA\ or less, but reduces significantly for larger layer spacings. This variation can be quantified after subtraction of the Coulomb contribution of $\rho_{D}$. This contribution for $d=$ 700 \AA\ is that determined in Ref.~3 and 4; a $d^{-4}$ spacing dependence was used to infer the value for smaller layer spacings. No subtraction is necessary for $d=$ 2600 \AA\ and 5200 \AA. The resultant phonon scattering for the $d=$ 700 \AA\ sample was used to fit the Coulomb adjusted data sets; a single multiplicative constant was applied to obtain a best fit to the other layer spacings. Nearly identical results for the 375 \AA\ and 425 \AA\ spacing samples are obtained if their Coulomb contribution is permitted to be a free fitting parameter. The overall relative magnitude of phonon drag determined through this fitting is shown in Fig.~\ref{6}. A logarithmic layer spacing dependence is evident for the data with $d \le$ 2600 \AA. This behavior does not extend to the 5200 \AA\ spacing sample, the behavior of which will be postponed for a later section; we focus here on the smaller layer spacings. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig6.ps}} \end{center} \caption{Dependence of the relative magnitude of phonon drag on layer spacing, determined by fits to the data of Fig.~5 as described in the text. The errors for the smallest layer spacings are dominated by potential inaccuracies in the subtraction of Coulomb scattering contributions. A logarithmic spacing dependence is observed for spacings below 3000 \AA. This agrees with the theoretically predicted dependence, and determines a phonon mean free path of 15 $\mu$m (dashed line: see text). The magnitude for the 5200 \AA\ spacing sample is not consistent with an overall logarithmic dependence.} \label{6} \end{figure} The logarithmic dependence seen in Fig. \ref{6} can be used to determine a value for the phonon mean free path, assuming that the drag corresponds to the independent system regime where $\rho_{D} \propto ln(d_{a}/d)$. The corresponding spacing dependence is shown in the figure as a dashed line. The value determined for the phonon mean free path from this fit is 15 $\mu$m. This value is substantially less than the critical value which determines the onset of the dominance of the coupled mode regime; that value was shown\cite{bonsager} to be $\sim$ 200 $\mu$m for the parameters of our sample. This small value for $l_{ph}$ appears to indicate that the coupled electron-phonon collective mode is absent. A difficulty with the value obtained for $l_{ph}$ is that it differs considerably from the phonon mean free path determined through thermal conductivity measurements on similar samples. Measurements by Eisenstein {\it et al.}\cite{eisenstein} found $l_{ph}$ to be nearly two orders of magnitude larger. It is possible that this difference can be accounted for by the fact that the phonons which scatter from the 2DEG are predominantly $2k_{F}$ phonons. Interactions with the electron system could result in a $l_{ph}$ for $2k_{F}$ phonons substantially shorter than for phonons with other wavevectors. The removal of $2k_{F}$ phonons from a thermal conductivity measurement could have a small effect, as a broad range of phonon wavevectors contribute as determined by the Bose distribution. The scale of the difference in $l_{ph}$ weakens the use of its value as determined in Fig.~\ref{6} as a sole indicator for the absence of an electron-phonon collective mode. \subsection{Matched density dependence} An important additional test for the existence of the electron-phonon collective mode is the dependence on total matched density at various temperatures. This dependence has been calculated\cite{bonsager} to show distinctly different behavior in the two regimes of $l_{ph}$. At 1 K in both regimes, $\rho_{D}$ should decrease monotonically as the density increases. At higher temperatures, a peak in the density dependence emerges for $l_{ph} < l_{c}$, i.e. for the case where no electron-phonon collective mode contributes to drag. The emergence of this peak at 2 to 4 K is accompanied by a change in the background density dependence. The decrease with increasing density seen at 1 K is smaller at 2 K. For 3 K and 4 K, there is only a small overall change with density in the magnitude of $\rho_{D}$. This behavior contrasts with that expected for $l_{ph} > l_{c}$, where the coupled electron-phonon collective mode dominates. In this regime, $\rho_{D}$'s behavior changes from a decrease with increasing density at 1 K to a substantial increase with density for higher temperatures. By 4 K the dependence is nearly proportional to density. These general differences in behavior, which are related to differences in screening and the role of the coupled electron-phonon collective mode, should be readily discernible in experiments. Measurements of the density dependence of $\rho_{D}/T^{2}$ for the 2600 \AA\ spacing sample are shown in Fig. \ref{7}. The general dependence seen at 1.2, 2.0, 3.0 and 4.0 K is strikingly similar to that predicted in the calculation of Ref.~1 for the small $l_{ph}$ regime. For each temperature measured, both the overall dependence on density and the emergence of the peak at high temperatures is in good agreement with predictions based on weak coupling and display clear differences from expectations for the strong coupling regime. These measurements clearly support the assertion that phonon drag in this sample corresponds to the short phonon mean free path regime, that is, the regime for which the coupled electron-phonon collective mode is absent. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig7.ps}} \end{center} \caption{Dependence of drag resistivity for $d=$ 2600 \AA\ at various temperatures on the individual layer density with layer densities matched. The changes in behavior as the temperature increases match theoretical predictions for the small $l_{ph}$ regime, providing evidence against an electron-phonon collective mode.} \label{7} \end{figure} A key element which distinguishes these measurements is the confidence that Coulomb scattering can be neglected. Earlier measurements\cite{rubel} which confirmed the existence of a peak in the relative density dependence at various matched densities, as had been seen in Ref.~4, also explored the dependence of $\rho_{D}$ at 4.2 K on total matched density. Those measurements, however, explore a density regime generally above the theoretical investigations of Ref.~1, and required a subtraction of Coulomb scattering. The relatively small increase of $\rho_D$ observed ($\sim$ 25 \%) for densities between 2 and $3.8 \times 10^{11}$ cm$^{-2}$ appear inconsistent with extrapolation of the current 4 K measurements; but this small increase assumes subtraction of a Coulomb contribution with a density dependence that has not been established in experiment. The measurements presented here can be compared to theory without the additional uncertainty introduced by such subtraction. \subsection{Magnitude of the drag} Although both the layer spacing and the matched density dependence show results clearly consistent with the short mean free path regime, there remains a contradiction as regards the magnitude of $\rho_{D}$. The measured magnitude is inconsistent with calculations assuming a small phonon mean free path, and is in much better agreement with the assumption of a large $l_{ph}$. It could be argued that the magnitude of $\rho_{D}$ alone could be considered evidence for the existence of a coupled electron-phonon collective mode. It is this mode which provides the increase in $\rho_{D}$ in Ref.~1 beyond that expected for ordinary phonon exchange, which has been shown to be too small to account for the strength of the phonon based interactions\cite{bonsager,gramila,nara,gramila2,tso,badalyan}. This argument cannot be considered conclusive, however, as there is considerable uncertainty in the overall magnitude calculated for $\rho_{D}$. This uncertainty arises from the use of the random-phase approximation (RPA) for screening. It is possible that the use of a more complete description of screening would generate a substantially larger magnitude for $\rho_{D}$ where $l_{ph}<l_c$, which must increase by an order of magnitude to match the observed $\rho_{D}$. The contradiction between evidence for a short phonon mean free path, as seen in the layer spacing and density dependences, and that for a long phonon mean free path, as seen in the magnitude of $\rho_{D}$, remains. The question of the existence of the coupled electron-phonon collective mode has thus not been conclusively addressed. \section{Remotely Spaced Layer Sample} The discussion above is restricted to those samples for which a logarithmic spacing dependence is observed. Measurements made on a sample with $d=$ 5200 \AA\ has a magnitude for $\rho_{D}$ which lies well below that expected from the logarithmic dependence, consistent with prior measurements. A deviation from a logarithmic dependence is expected in the short $l_{ph}$ regime\cite{bonsager} for layer spacings larger than $d_{a}$ where $\rho_D$ decreases exponentially with $d$. Use of the value of $l_{ph}$ derived from the spacing dependence measurements yields a value for $d_{a}$ of 3.8~$\mu$m. Significant deviations from a logarithmic dependence should not occur in the small $l_{ph}$ limit until the spacing is an order of magnitude larger than 5200 \AA. The unexpected reduction in $\rho_{D}$ requires consideration of alternate origins. A possible explanation is that drag occurs in the large $l_{ph}$ collective mode limit despite the evidence for a small $l_{ph}$. In this regime, a change in characteristic behavior occurs at a length scale much smaller than for the small $l_{ph}$ regime. This characteristic layer spacing in our samples corresponds to a length of approximately 5000 \AA. If the large $l_{ph}$ limit applies, a substantial deviation below the observed logarithmic dependence for $d=$ 5200 \AA\ could then be consistent with expectations for a electron-phonon collective mode. The absence of a clear indication of which regime is appropriate provides motivation for re-examining the dependence on matched densities for the 5200 \AA\ sample. Such a measurement is extremely difficult, requiring accurate detection of signals as small as 250 pV. The central difficulty in measuring such small signals lies in ensuring that potential spurious signals are absent; this was verified here to a level of 30 pV. An additional complication for the measurement is that large layer spacings require a substantial interlayer bias to be applied, limiting the range of densities measured to between 1.1 and $1.7 \times 10^{11}$ cm$^{-2}$ per layer. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig8.ps}} \end{center} \caption{Dependence of drag resistivity for $d=$ 5200 \AA\ at various temperatures on individual layer density with layer densities matched. The behavior is markedly different from that of the 2600 \AA\ spacing sample, and disagrees with theoretical predictions for both small and large $l_{ph}$ regimes.} \label{8} \end{figure} The results of these measurements are shown in Fig.~\ref{8}. A reduction in $\rho_{D}$ with increasing density at 1.2 K is consistent with the behavior of the 2600 \AA\ spacing sample. For higher temperatures, however, there are distinct differences from the earlier measurements. A clear reduction in magnitude with increasing density is observed for all measured temperatures. There is no sign of a peak in the density dependence, as was clearly evident for the 2600~\AA\ spacing sample. These data are inconsistent with both the small $l_{ph}$ limit, which well describes the density dependence for the 2600~\AA\ spacing sample, and with the large $l_{ph}$ limit. In either case a peak in the density dependence can be discerned, but no peak is apparent in the 5200~\AA\ spacing measurements. In both cases, the theoretical dependence at high temperatures makes a clear departure from the decreasing dependence typical at $\sim$ 1 K: to a weak density dependence for small $l_{ph}$ and to a strongly increasing density dependence for large $l_{ph}$. The 5200~\AA\ spacing sample, by contrast, retains a decreasing density dependance as the temperature is increased, clearly contradicting either theoretical prediction. \begin{figure}[!t] \begin{center} \leavevmode \hbox{% \epsfysize=3in \epsffile{fig9.ps}} \end{center} \caption{Drag resistivities divided by $T^{2}$ versus temperature for $d=$ 5200 \AA\ with two different matched densities. The role of $2k_{F}$ scattering remains evident in the change in temperature dependence in this sample, reflecting the characteristic wavevector cutoff that scales with the size of the Fermi surface.} \label{9} \end{figure} The unusual behavior seen for this sample is not reflected in measurements exploring more fundamental elements of the electron-phonon scattering process, such as the presense of a cutoff at $2k_{F}$. This can be seen in the data of Fig.~\ref{9} which shows the temperature dependence of $\rho_{D}/T^{2}$ for two densities. The data show the change in $T$ dependence characteristic of this cutoff, and a shift in the transition temperature corresponding to the change in the size of the Fermi surface. Quantitative information about this transition temperature is somewhat uncertain, as it was not possible to verify the lack of spurious signals to a level below 30~pV. Adding a small offset to the measured signal can change the apparent peak position. The general confirmation of the properties related to the $2k_{F}$ wavevector cutoff in the 5200~\AA\ spacing sample make its magnitude and its dependence on matched densities more puzzling. The only free parameter in the theoretical calculations is $l_{ph}$, which determines whether the electron and phonon systems are weakly or strongly coupled. We would expect this parameter not to vary between samples, as they are all essentially indentical apart from the size of the barrier. Yet the 5200 \AA\ spacing sample is clearly distinct in behavior from the more closely spaced layer samples, and from theoretical predictions for either $l_{ph}$ regime. The measurements for this sample strongly suggest that understanding of interlayer phonon mediated electron-electron scattering is not yet complete. \section{Conclusions} Extensive measurements of phonon mediated electron-electron scattering have been performed using electron drag on double layer 2DEG samples. Measurements for a sample in which Coulomb scattering is absent explored fundamental elements of this scattering process. These include the presence of a cutoff for scattering of phonons having wavevectors greater than twice the Fermi wavevector of the electron system and the dominance of $2k_{F}$ scattering even at relatively low temperatures. These properties of phonon drag, which are directly related to elements of electron-phonon scattering in single layers, have been verified in temperature and density dependence measurements. Additional measurements explore the existence of a coupled electron-phonon collective mode, as has been recently proposed\cite{bonsager} for 2DEG systems. Contradictory evidence concerning the existence of this mode has been observed. Both the dependence of the scattering on layer spacing below 3000 \AA\ and its dependence on matched layer densities at various temperatures show behavior consistent with the absence of this mode. The overall magnitude of the scattering, however, suggests the presense of the collective mode. The phonon mean free path, as determined by fits to the spacing dependence assuming the absense of the mode, also differs by orders of magnitude with previous measurements of this length via thermal conductivity. Phonon drag measurements for a sample with a 5200 \AA\ layer spacing raise further questions about details of the scattering process. Although the fundamental elements of scattering related to the size of the Fermi surface are confirmed in this sample, its scattering strength is well below the logarithmic spacing dependence found for more closely spaced layers. The sample further shows a decrease in scattering strength as the density of both layers is increased, a behavior generally inconsistent with theoretical predictions assuming either the absence or presense of the collective mode. In general, while basic properties of phonon drag are well established, these measurements show that a full understanding of the process, especially with respect to the magnitude of the scattering and the presense of a coupled electron-phonon collective mode, requires further investigations. \acknowledgements Discussions with A. H. MacDonald and M. C. B\o nsager are gratefully acknowledged. We are indebted to Jim Eisenstein for his contributions to prior investigations of phonon drag and for his considerable efforts in the growth of the 2600 \AA\ spacing sample. This work was supported by the NSF through grant DMR-9503080 and DMR-9802109, by the Alfred P. Sloan Foundation, and by the Research Corporations Cottrell Scholar program. \vspace{-0.125in}
\section*{Figure Captions} \begin{description} \item[Figure 1] (a) Acoustic bands for a square lattice of air-cylinders for $\beta = 10^{-3}$. A complete band gap lies between the two horizontal dashed lines. (b) The dependence of the complete band gap on the cylinder concentration $\beta$. \item[Figure 2] The left column: The phase diagram for the two dimensional phase vectors defined in the text. Right column: The spatial distribution of acoustic energy. For brevity, 200 cylinders are considered in the computation. In the plot of the energy distribution, the geometrical spreading factor has been removed. Note that in the case of $ka = 0.01$ there are a few phase vectors which do not point to the same direction as the others because of the effect of the finite boundary; when the cylinder number goes to infinity, this effect disappears. \end{description} \vspace{24 pt} \noindent {\bf Acknowledgments.} The work received support from the National Central University and the National Science Council. \vspace{24pt} \noindent Correspondence and request for materials should be addressed to Z.Y. \\ Email : <EMAIL> \end{document}
\section{INTRODUCTION} A full quantum theory of gravity is still out of reach. However, in situations where the spacetime curvature is well below Planck's curvature, it is possible to compute some quantum gravity effects. Indeed, metric fluctuations can be quantized using standard methods. The non renormalizability of the resulting quantum field theory is not an impediment for making meaningful quantum corrections. The key point is to consider general relativity as an effective field theory \cite{dono}. Although the leading long distance quantum corrections are expected to be too small in realistic situations, the analysis of general relativity as an effective field theory is of conceptual interest. Moreover, tiny but measurable quantum gravity effects could show up when measuring the decoherence of wavepackets of a non relativistic particle subjected to the gravitational potential \cite{perci}. On the other hand, recent speculations raise the length scale relevant for quantum gravity effects from Planck length to a TeV scale \cite{cern}. In this situation, the effects of metric fluctuations could be easier to observe. In the context of effective field theories, it is in principle possible to compute an effective action and effective field equations for the mean value of the spacetime metric. The effective field equations (known as semiclassical Einstein equations or backreaction equations) include the backreaction of quantum matter fields and of the metric fluctuations on the spacetime metric. These equations should be the starting point to investigate interesting physical problems like, for example, the dynamical evolution of a black hole geometry taking into account the evaporation process. The backreaction equations have been investigated by several authors in the last twenty years or so \cite {huetal}. However, due to the complexity of the problem (and also to the non renormalizability of the theory) most works considered scalar or spinor quantum matter fields, but the graviton contribution was simply omitted. It is in general stressed that the graviton effects should be similar to those of a couple of massless, minimally coupled scalar fields. While this is true at the level of the backreaction equations, there is an important physical difference that has been pointed only recently \cite{nos}. When metric fluctuations are taken into account, the background geometry (i.e. the metric that solves the backreaction equations), turns out to be non physical. The reason is the following: any classical or quantum device used to measure the spacetime geometry will also feel the graviton fluctuations. As the coupling between the device and the metric is non linear, the device will not measure the background geometry, which therefore is not the relevant physical quantity to compute. In particular, in Ref. \cite{nos} we have shown that, working in the Newtonian approximation, the trajectory of a classical test particle is not a geodesic of the background metric. Instead its motion is determined by a quantum corrected equation that takes into account its coupling to the gravitons. Moreover, while the backreaction equations and their solutions depend on the gauge fixing of the gravitons, this dependence cancels out in the quantum corrected equation of motion for the test particle. The aim of this paper is to analyze the effect of the gravitons on the motion of a test particle beyond the Newtonian approximation. In order to avoid technical complications, we will assume we know a solution to the backreaction equations, and will focus only on the departure of the test particle's equation of motion from the geodesic equation of the background metric. Moreover, we will consider models where it is easy to fix completely the gauge of the gravitons and quantize the theory by taking into account the remaining degrees of freedom. The paper is organized as follows. In Section II we prove that the effective action that governs the motion of the test particle is the mean value of the classical action. In Section III we consider Robertson Walker universes. We first briefly describe how to quantize the metric fluctuations in terms of massless scalar fields. Then we compute the quantum corrections to the geodesic equation and solve the quantum corrected equations of motion perturbatively. In particular, we find the graviton corrections to the cosmological redshift. In Section IV we consider three dimensional gravity coupled to a Maxwell field. Following Ref. \cite{ash1}, we first show that this model is exactly soluble: one can fully fix the gauge and show that the degrees of freedom reside in the Maxwell field. Then we compute the quantum corrected equation of motion for the test particle. We show that, even in regions where the background metric is locally flat, the trajectory of the test particle is not a straight line. Section V contains our final remarks. \section{EFFECTIVE ACTION FOR A TEST PARTICLE} In this Section we will show that, when quantum metric fluctuations are taken into account, the effective action for the test particle is the mean value of its classical action. This result is summarized in Eq.(\ref{seef}) below (the reader may want to accept this as a reasonable assumption and skip this section). Consider pure gravity described by Einstein-Hilbert action \footnote{ Our metric has signature $(-+++)$ and the curvature tensor is defined as ${R}^{\mu}_{\, \cdot \, \nu\alpha\beta} = \partial_{\alpha} \Gamma^{\mu}_{\nu\beta} - \ldots$, ${R}_{\alpha\beta} = {R}^{\mu}_{\, \cdot \, \alpha\mu\beta}$ and ${R}= {g}^{\alpha\beta} {R}_{\alpha\beta}$. We use units $\hbar=c=1$.} $S_{\rm G}=(2/\kappa^2) \int d^4x \sqrt{-g} R$, where $\kappa^2=32 \pi G$, and imagine that in addition we have some type of matter content described by an action $S_{\rm M}$. The effect of quantum metric fluctuations can be analyzed with the background field method, expanding the whole action $S_{\rm G}+S_{\rm M}$ around a background metric as $g_{\mu\nu} \rightarrow g_{\mu\nu} + \kappa h_{\mu\nu}$, and integrating over the graviton field $h_{\mu\nu}$ to get an effective action for the background metric. In order to fix the gauge one chooses a gauge-fixing function $\chi^{\mu}[g,h]$, a gauge-fixing action $S_{\rm gf}[g,h]= -(1/2) \int d^4x \sqrt{-g} \chi^{\mu} g_{\mu\nu} \chi^{\nu}$, and the corresponding ghost action $S_{\rm gh}$. Imagine that in addition we have a classical test particle that moves in the above background metric and we wish to study the effects of metric fluctuations on it. We couple gravity to the particle by means of the standard action $S_m[x]=-m \int \sqrt{-g_{\mu\nu}} dx^{\mu} dx^{\nu}$, where $x^{\mu}$ denotes the path of this test particle. The complete effective action $S_{\rm eff}$ for the background metric $g_{\mu\nu}$ and for the test particle $m$ is obtained by integrating the whole action $S\equiv S_{\rm G}+S_{\rm M}+S_{\rm gf}+S_m +S_{\rm gh}$ over the graviton and ghost fields. To evaluate it in the one loop approximation we first expand $S$ up to second order in gravitons. The second order term reads \begin{equation} S^{(2)}= \int d^4y \sqrt{-g} h_{\mu\nu} F^{\mu\nu\rho\sigma} h_{\rho\sigma} - \int d^4y \sqrt{-g} h_{\mu\nu} m^{\mu\nu\rho\sigma} h_{\rho\sigma} \end{equation} where ${\hat F}\equiv F^{\mu\nu\rho\sigma}$ is a second order differential operator that depends on the background metric, and $m^{\mu\nu\rho\sigma}$ is a tensor depending on the position and velocity of the test particle, \begin{equation} m^{\mu\nu\rho\sigma}(y)=\frac{m \kappa^2}{8} \int d\tau \delta^4(y-x(\tau)) \dot{x}^{\mu}\dot{x}^{\nu}\dot{x}^{\rho}\dot{x}^{\sigma} \end{equation} There is also a second order term in ghost fields, that for gauge-fixing functions linear in the metric fluctuations decouple from the gravitons, and couple only to the background metric. The result of the path integral is the classical action $S_{\rm clas}=S_{\rm G}+S_{\rm M} + S_{m}$ plus the sum of two functional determinants, \begin{equation} S_{\rm eff}=S_{\rm clas} + \frac{i}{2} {\rm Tr} \ln ({\hat F} - {\hat m}) - i {\rm Tr} \ln {\hat G} \label{efe} \end{equation} where ${\hat G}$ is also a second order differential operator that arises from integrating over ghosts. Once the effective action is evaluated, one can derive the equations of motion for the background metric $g_{\mu\nu}$, the so called semiclassical Einstein equations, i.e. $\delta S_{\rm eff}/\delta g_{\mu\nu}=0$. To solve these equations one can discard all contributions coming from the test particle, as they are vanishingly small. As they stand, these equations (obtained from the standard {\it in-out} effective action) are neither real nor causal. In order to get real and causal equations of motion for the background metric, the {\it in-in} effective action must be evaluated \cite{ctp}. Alternatively, one can take twice the real and causal part of the propagators in the {\it in-out} field equations. In both ways one gets semiclassical Einstein equations suitable for initial value problems. >From the effective action given above one can also derive the quantum corrected equation of motion for the test particle, i.e. $\delta S_{\rm eff}/\delta x^{\rho}=0$, which will be our main concern in what follows. The same comments about reality and causality apply to this equation of motion. In this paper we will work with the usual {\it in-out} effective action and use the adequate propagators in the quantum corrected equations. In general it is extremely complicated, if not impossible, to work out the functional traces in Eq.(\ref{efe}), so several approximation methods have been developed to deal with them. However, in this paper we will only focus on the quantum effects of the coupling between the test particle and gravitons. We can make use of the fact that the test particle has a small mass, so we can expand Eq.(\ref{efe}) in powers of $m$ and just keep the leading contribution. In this way we find that the whole effective action reads \begin{equation} S_{\rm eff}[g_{\mu\nu},x]= S_{\rm clas} + \frac{i}{2} {\rm Tr} {\ln {\hat F}} - i {\rm Tr} \ln {\hat G} - \int d^4x \sqrt{-g} \langle h_{\mu\nu} m^{\mu\nu\rho\sigma} h_{\rho\sigma} \rangle \end{equation} The expectation value is taken with respect to the graviton state. The effective action for the test particle will be the sum of the classical term $S_m[x]$ and this last term, so that we conclude that in fact that effective action is the expectation value of the classical one \begin{equation} S_{\rm eff}[x] = \langle S_m[x] \rangle \label{seef} \end{equation} It is important to stress that due to the non linear nature of the coupling between gravity and test particle, the effective lagrangian is not the same as the classical lagrangian evaluated in the expectation value for the particle's path. The calculation described so far preserves the covariance in the background metric $g_{\mu\nu}$. Alternatively, one can fully fix the gauge of the quantum fluctuations of the geometry and quantize the remaining degrees of freedom. As can be easily proved, the argument leading to Eq. (\ref{seef}) remains unchanged, since it relies only on the fact that the test particle mass is small. \section{QUANTUM CORRECTIONS TO GEODESICS FOR FLAT ROBERTSON-WALKER METRICS} \subsection{Non covariant quantization} In this subsection we briefly review the non-covariant method of quantization for flat Robertson-Walker universes. The metrics we are dealing with are therefore of the form $ds^2=- dt^2 + a^2(t) d{\bf x}^2$, where $a(t)$ is the expansion coefficient . The action for the matter content in RW metrics has the form \begin{equation} S_{\rm M} = \int d^4x \sqrt{-g} \left[ \frac{1}{2} (\rho + p) u^{\mu} u^{\nu} g_{\mu\nu} + \frac{1}{2} (\rho + 3 p) \right] \end{equation} where $u^{\mu}$, $\rho$ and $p$ and the fluid's four-velocity, density and pressure respectively. The associated classical Einstein equations are \begin{equation} R_{\mu\nu}=-\frac{1}{2} \left( T_{\mu\nu} - \frac{1}{2} g_{\mu\nu} T_{\lambda}^{\lambda} \right) \end{equation} where the classical energy-momentum tensor is $T_{\mu\nu}=(\rho+p) u_{\mu} u_{\nu} - p g_{\mu\nu}$. There are different ways to quantize the theory. One is based on the background field method, which was described above. Here we follow another quantization procedure that starts from the classical theory of perturbations in RW, developed in \cite{Lif}. One considers perturbations such that $\delta \rho=\delta p=\delta u^{\mu}=0$, and metric perturbations $h_{\mu\nu}$ that satisfy $u^{\mu} h_{\mu\nu}=0$, and further imposes the gauge conditions $h^{\mu\nu}_{~~;\nu}=0$. Finally one ends up with only two independent components of the metric, $h_{+}$ and $h_{\times}$, which can be expressed in terms of the original components of $h_{\mu\nu}$, and that correspond to the two polarizations of a gravitational wave. The above conditions on the metric imply that $h_{0\mu}=0$ and a transversality condition ${\tilde \nabla}_j h^{ij}=0$, where ${\tilde \nabla}_j$ denotes the covariant derivative with respect to the spatial part of the metric. Both components $h_+$ and $h_{\times}$, and also $h_i^j$, verify the field equation for a minimally coupled massless scalar field in RW \begin{equation} \Box \phi = - a^{-3} \frac{\partial}{\partial t} \left( a^3 \frac{\partial}{\partial t} \phi \right) + \nabla^2 \phi =0 \label{scalar} \end{equation} To quantize we use the non-covariant quantization procedure of \cite{Ford77,BLH78}. First one writes the second order term of the action $S_{\rm G}+S_{\rm M}$ in terms of the two independent degrees of freedom of the field, $h_+$ and $h_{\times}$ \begin{equation} S^{(2)}_{\rm G+M} = \frac{1}{2} \int d^4x \sqrt{-g} [ \partial_{\mu} h_+(x) \partial^{\mu} h_+(x) + \partial_{\mu} h_{\times}(x) \partial^{\mu} h_{\times}(x) ] \end{equation} and then imposes equal-time canonical conmutation relations for the two scalar fields $[h_a({\bf x},t),\Pi_b({\bf x'},t)]= i \delta_{ab} \delta({\bf x}-{\bf x'})$, where $a,b=+,\times$ and $\Pi_a$ is the canonical momentum conjugate to $h_a$. This quantization procedure is equivalent to that for the individual modes $h_i^j$. Instead of using canonical quantization, one can also do path integrals. One expands the action in terms of the individual modes $h_i^j$ (or in terms of $h_+$ and $h_{\times}$) and integrates over them in order to get an effective action for the background metric. For the one loop effective action one needs the second order term of the expansion of the action in terms of metric perturbations, namely $S^{(2)}_{\rm G+M}=1/2 \int d^4y \sqrt{-g} h^i_j \Box h_i^j$, where $\Box$ denotes the scalar D'Alambertian operator. Finally one has to evaluate the functional determinant of this differential operator. \subsection{Quantum corrected geodesic equation} Having summed up how to quantize metric perturbations in RW, let us see how such quantum metric fluctuations affect the motion of a classical test particle. As described in the previous section, the effective action for the test particle is the expectation value of the classical action, namely \begin{equation} S_{\rm eff}[x]= - m \int \sqrt{-g_{\mu\nu}(x) dx^{\mu} dx^{\nu}} - \frac{m \kappa^2}{8} \int d\tau \langle h_{ij}(x) h_{lm}(x) \rangle \dot{x}^i \dot{x}^j \dot{x}^l \dot{x}^m \end{equation} where the dot denotes the derivative with respect to $\tau$. The graviton two-point function can be expressed in terms of the scalar two-point function $\langle \phi(x) \phi(x') \rangle$ as \begin{equation} \langle h_{ij}(x) h_{lm}(x') \rangle = - \frac{1}{3} a^2(t) a^2(t') \left( \delta_{ij} \delta_{lm} - \frac{3}{2} \delta_{il} \delta_{jm} - \frac{3}{2} \delta_{im} \delta_{jl} \right) \langle \phi(x) \phi(x') \rangle \end{equation} We recall that in these expressions the metric $g_{\mu\nu}$ is the solution to the semiclassical Einstein equations that follow from quantizing gravity in a RW universe. In the following we will assume that these equations have been solved and that the quantum corrected expansion factor $a(t)$ has been found. The geodesic equation for the test particle follows from $\delta S_{\rm eff}[x]/\delta x^{\rho}=0$. For the temporal component we get \begin{equation} \frac{d^2t}{d \tau^2} + a(t) a'(t) \left( \frac{d \bf{x}}{d \tau} \right)^2 - \frac{\kappa^2}{8} \dot{x}^i \dot{x}^j \dot{x}^l \dot{x}^m \frac{\partial}{\partial t} G_{ijlm}[x(t)]=0 \label{geotemp} \end{equation} where $a'(t)\equiv da/dt$ and $G_{ijlm}[x(t)]$ is the coincident limit of the graviton two-point function, evaluated along the trayectory of the particle. For the $n$-th spatial component ($n=1,2,3$) we obtain \begin{equation} \frac{d}{d \tau} \left( a^2(t) \frac{d x^n}{d \tau} - \frac{\kappa^2}{2} G_{ijkl}[x(t)] \delta^{in} \dot{x}^j \dot{x}^k \dot{x}^l \right)=0 \label{geoesp} \end{equation} Now let us solve Eqs.(\ref{geotemp},\ref{geoesp}) for $d {\bf x}/d \tau$ and $dt/d \tau$. From Eq.(\ref{geoesp}) we see that the expression in parenthesis is conserved. These conserved three quantities reflect the spatial translational invariance of RW metric, which is preserved upon the quantization procedure. Therefore \begin{equation} a^2(t) \frac{d x^n}{d \tau} - \frac{\kappa^2}{2} G_{ijkl}[x(t)] \delta^{in} \dot{x}^j \dot{x}^k \dot{x}^l = \alpha \label{consesp} \end{equation} where $\alpha$ is a dimensionless constant that depends on the initial velocity of the particle. Plugging this identity into Eq.(\ref{geotemp}) we find \footnote{ This equation also follows from the very definition of the proper time. Indeed, from $1=(dt/d\tau)^2 - a^2(t) (d{\bf x}/d\tau)^2$ we easily get Eq.(\ref{constem}). } \begin{equation} \frac{dt}{d \tau} = \sqrt{ 1 + a^{-2}(t) \sum_{n=1}^{3} \left( \alpha + \frac{\kappa^2}{2} G_{ijkl}[x(t)] \delta^{in} \dot{x}^j \dot{x}^k \dot{x}^l \right)^2 } \label{constem} \end{equation} Now we solve Eqs.(\ref{consesp},\ref{constem}) perturbatively in terms of the coupling between the test particle and gravitons. Let us assume that the initial velocity of the test particle is in the $x=x^1$ direction. The zeroth order approximation corresponds to neglecting the coupling between the particle and gravitons, which results in \begin{eqnarray} \frac{d x}{d \tau} &=& \frac{\alpha}{a^2(t)} \\ \frac{dt}{d \tau} &=& \sqrt{1 + \alpha^2 a^{-2}(t)} \label{geoclas} \end{eqnarray} Note that the limiting case of a light ray (null limit) $dx/dt=1$ is obtained when $\alpha \rightarrow \infty$. When the coupling is taken into account, we see that the particle still moves in the same $x$ direction, and we get \begin{eqnarray} \frac{dx}{d\tau} &=& \frac{\alpha}{a^2(t)} \left( 1 + \frac{\alpha^2 \kappa^2}{3 a^2(t)} \langle \phi^2(t) \rangle \right) \\ \frac{dt}{d\tau} &=& \sqrt{1+\alpha^2 a^{-2}(t)} \left(1 + \frac{\alpha^2 \kappa^2}{3 a^4(t)} \frac{\langle \phi^2(t) \rangle}{1+\alpha^2 a^{-2}(t)} \right) \end{eqnarray} where we expressed the graviton two point function in terms of the scalar two-point function as $G_{xxxx}(t)=(2/3) a^4(t) \langle \phi^2(t) \rangle$. The speed of the particle results \begin{equation} \frac{dx}{dt}= \frac{\alpha a^{-2}(t)}{\sqrt{1+\alpha^2 a^{-2}(t)}} \left( 1 + \frac{32}{3} \pi G \langle \phi^2(t) \rangle \frac{\alpha^2 a^{-2}(t)}{1+\alpha^2 a^{-2}(t)} \right) \label{qvel} \end{equation} This is the main result of this section. It expresses the quantum corrections to the velocity of a test particle that moves in a flat Robertson-Walker quantum background. In the null limit \begin{equation} dx/dt \approx a^{-1}(t) [1+(32/3) \pi G \langle \phi^2(t) \rangle] \end{equation} describes the graviton correction to the cosmological redshift. To estimate the effect of this quantum correction on the classical trayectory of the test particle, we first have to evaluate the two-point function in the coincident limit, $\langle \phi^2(t) \rangle$. As is well known, this coincident limit is divergent, so a renormalization procedure is compelling. In the following we will calculate $\langle \phi^2(t) \rangle$ for particular RW metrics, namely $a(t)=a_0 e^{H t}$ (de Sitter) and $a(t)=a_0 t^c$ For de Sitter spacetime, the two-point function not only has UV problems but also IR ones. However, in the late time limit $t=t' \gg H^{-1}$ it is possible to give an approximate form for the renormalized function. It was shown by several authors \cite{Ford1,Ford2,Linde,Staro} that the coincident limit grows linearly with the coordinate time, $\langle \phi^2(t) \rangle \approx H^3 t/2 \pi^2$. Using that $\kappa^2 \propto R^{-1}_{\rm Planck}$ and that for de Sitter the curvature is constant $R \propto H^2$, we conclude that the quantum correction is proportional to $(R/R_{\rm Planck}) F(t)$, where the function $F(t)=\alpha^2 H t a_0^{-2} e^{-2 H t}/(1+\alpha^2 a_0^{-2} \exp^{-2 H t})$ decreases exponentially for late times. The velocity of the test particle in the late time limit is therefore given by \begin{equation} \frac{dx}{dt} = \frac{\alpha a_0^{-2} e^{-2H t}} {\sqrt{1+\alpha^2 a_0^{-2} e^{-2 H t}}} \left( 1+ \frac{16 G \hbar H^2 F(t)}{3 \pi c^5} \right) \label{vel1} \end{equation} where we have restored units $\hbar$ and $c$. As we pointed out before, the scale factor $a(t)$ should be a solution to the semiclassical Einstein equations. A perturbative solution will be of the form $a(t)=a_{\rm clas}(t) + \delta a (t)$, $a_{\rm clas} $ being the classical scale factor and $\delta a \ll a_{\rm clas}$. It is well known that the semiclassical Einstein equations admit de Sitter solutions \cite{wada} $a(t)=a_0 e^{Ht}$ with $H=H_{\rm clas}(1+\gamma {H_{\rm clas}^2\over R_{\rm Planck}}), \,\,\gamma = O(1)$. Therefore, as long as ${H_{\rm clas}^3 t\over R_{\rm Planck}}\ll 1$ the correction to the scale factor is given by ${\delta a\over a_{\rm clas}} \simeq \gamma {H_{\rm clas}^3 t\over R_{\rm Planck}}$. Replacing $a(t)=a_{\rm clas}(t) + \delta a (t)$ in Eq. (\ref{vel1}) we obtain, to first order in all quantum corrections \begin{equation} \frac{dx}{dt} = \frac{\alpha a_0^{-2} e^{-2H_{\rm clas} t}} {\sqrt{1+\alpha^2 a_0^{-2} e^{-2 H_{\rm clas} t}}} \left( 1+ \frac{16 G \hbar H^2 F(t)}{3 \pi c^5} -\gamma {2+\alpha^2a^{-2}_{\rm clas}(t)\over 1+\alpha^2a^{-2}_{\rm clas}(t)} {H_{\rm clas}^3 t\over R_{\rm Planck}}\right) \label{vel2} \end{equation} where $F(t)$ is to be evaluated with the classical value for the Hubble parameter. This shows that the quantum correction to the geodesics coming from the graviton coupling (second term in Eq. (\ref{vel2})) and the one coming from the semiclassical Einstein equations (third term) are of the same order of magnitude. Consider now metrics with $a(t)=a_0 t^c$. Although these are not solutions to the semiclassical Einstein equation, they are useful to illustrate the corrections to the geodesics. In this case there are no infrared divergencies. In the Appendix we give some details as to how to evaluate the renormalized two-point function. The result is $\langle \phi^2(t) \rangle \propto t^{-2} \log(t^2 \mu^2)$, where $\mu$ is an (arbitrary) renormalization scale. Since for these metrics the curvature is $R \propto t^{-2}$, we obtain that the quantum correction also has the form $(R/R_{\rm Planck}) F(t)$, where now $F(t)=\alpha^2 a_0^{-2} t^{-2c} \log(t^2 \mu^2)/(1+\alpha^2 a_0^{-2} t^{-2c})$, which also decreases for long times. The velocity of the test particle is \begin{equation} \frac{dx}{dt} = \frac{\alpha a_0^{-2} t^{-2c}} {\sqrt{1+\alpha^2 a_0^{-2} t^{-2c}}} \left( 1+ \frac{2 c(2c-1) G \hbar F(t) }{3 \pi c^5 t^2} \right) \end{equation} \section{QUANTUM CORRECTIONS TO GEODESICS IN THREE DIMENSIONAL GRAVITY} \subsection{Three dimensional General Relativity} In this section we will consider 2+1 gravity coupled to Maxwell fields. Under the assumption of rotational symmetry, this model is exactly soluble. Moreover, it is possible to associate a well defined quantum operator to the spacetime metric. Therefore, it is particularly useful to analyze the effective action for a test particle and the corrections to the geodesics. In this subsection we will follow closely Refs. \cite{ash1,ash2}. At the classical level, the theory is governed by the Einstein-Maxwell equations, which read \begin{eqnarray} && R_{ab}= 8\pi G\nabla_a\phi\nabla_b\phi\\ &&g^{ab}\nabla_a\nabla_b\phi = 0 \label{mee} \end{eqnarray} where the electromagnetic field has been written in terms of a scalar field as $F_{ab}=\epsilon_{abc}\nabla^c\phi$. Assuming rotational symmetry, the above equations can be easily solved. The metric can be written as \begin{equation} g_{ab}dx^a dx^b = e^{G\Gamma (r,t)}[-dt^2+dr^2]+r^2 d\theta^2 \end{equation} Moreover, the scalar field decouples from the metric \begin{equation} g^{ab}\nabla_a\nabla_b\phi = 0 \rightarrow (-\partial_t^2+\partial_r^2)\phi = 0 \end{equation} Therefore, one can solve the $1+1$ Klein Gordon equation for $\phi$ and then determine $\Gamma$ from the Einstein equation. The result is \begin{equation} \Gamma (r,t)={1\over 2}\int_0^r dr'~~r'~~[(\partial_t\phi)^2 + (\partial_{r'}\phi)^2] \label{ene} \end{equation} Note that, as $r\rightarrow\infty$, $\Gamma$ tends to a constant value $\Gamma (\infty ,t) =H_0$. The metric becomes locally flat with a deficit angle $2\pi (1- e^{-G H_0 /2})$. To quantize the theory, one can promote $\phi$ to an operator $\hat\phi$ describing a free quantum scalar field in $1+1$ dimensions. The spacetime metric is a secondary operator that can be expressed in terms of $\hat\phi$ as \begin{equation} \hat{g}_{rr}=-\hat{g}_{tt}=e^{G\hat\Gamma} \end{equation} where $\hat\Gamma$ is the operator defined by Eq.(\ref{ene}) with $\phi\rightarrow\hat\phi$. For simplicity in what follows we will consider the metric operator in the asymptotic region $r\rightarrow\infty$, where the operator $\hat\Gamma$ is time independent. For a given coherent state of the scalar field (denoted by $|F\rangle$ and peaked around a classical configuration $F(r,t)$), it is easy to show that \begin{eqnarray} \langle F|\hat\phi |F \rangle &=& F(r,t)\nonumber\\ \langle F|\hat g_{rr} |F \rangle &=& \exp \left[ {1\over \hbar} \int_0^{\infty} dw~~|F(w)|^2 (e^{G\hbar w} - 1) \right] \end{eqnarray} For sufficiently low frequencies (i.e. when the Fourier transform of the classical configuration is peaked around a low frequency), the mean value of the metric operator can be approximated by \begin{equation} <F|\hat g_{rr} |F> = g_{rr} \left( 1 + \hbar {G^2\over 2} \int_0^{\infty} dw~w^2~|F(w)|^2 \right) \end{equation} The first term is the value of the metric we would obtain from the classical field equations for a classical scalar field configuration given by $F(r,t)$. The second term represents a small quantum correction. As in the classical case, for $r\rightarrow\infty$ the mean value of the metric describes a locally flat spacetime, but with a quantum corrected deficit angle. \bigskip \subsection{Effective action for a test particle} According to our general discussion in Section 2, the effective action for a test particle moving in the $2+1$ dimensional spacetime is given by \begin{equation} S_{\rm eff}[x]=<S_m[x]>=-m <\int dt \sqrt{e^{G\hat\Gamma}(1-\dot r^2)-r^2\dot \theta ^2} > \end{equation} where the mean value is taken with respect to the coherent state $|F \rangle$. Here a dot denotes derivative with respect to $t$. As in the previous section we will consider only the asymptotic region where the metric operator is time independent. We write the metric operator as $e^{G\hat\Gamma}= \langle e^{G\hat\Gamma} \rangle +\hat\Delta$. The effective lagrangian then becomes \begin{equation} L_{\rm eff}= -m \bar L \langle \sqrt{ \left[ 1+{\hat\Delta (1-\dot r^2)\over {\bar L}^2} \right] } \rangle \end{equation} where $\bar L$ is proportional to the classical lagrangian evaluated in the mean value of the metric \begin{equation} \bar L = \sqrt { \langle e^{G\hat\Gamma} \rangle (1-\dot r^2)-r^2\dot \theta ^2 } \end{equation} Note that, after a redefinition of the angular variable $\theta\rightarrow \sqrt{\langle e^{G\hat\Gamma} \rangle} \theta$, $\bar L$ becomes proportional to the lagrangian of the test particle in a locally flat spacetime. The deficit angle is given by $2\pi (1- \sqrt{\langle e^{G\hat\Gamma} \rangle})$. Assuming that the quantum fluctuations around the mean value are small \footnote {This is not always the case. See Ref. \cite{ash2}.} we get \begin{equation} L_{\rm eff}= -m\bar L \left[ 1-{1\over 8}{(1-\dot r^2)^2\Delta^2\over \bar L^4} \right] \label{leff} \end{equation} where $\Delta^2=\langle \hat\Delta^2 \rangle = \langle (e^{G\hat\Gamma}- \langle e^{G\hat\Gamma} \rangle )^2\rangle$. The above equation is the starting point to describe the quantum corrections to the trajectory of the test particle. Let us first consider a non relativistic motion of the particle. In this situation we have \begin{equation} \bar L\simeq \sqrt{ \langle e^{G\hat\Gamma} \rangle} \left[ 1-{\dot r^2\over 2}-{r^2\dot\theta^2 \over 2 \langle e^{G\hat\Gamma} \rangle } \right] \end{equation} Therefore, the effective lagrangian can be approximated by \begin{equation} L_{\rm eff}\simeq -m \sqrt{ \langle e^{G\hat\Gamma} \rangle } \left[ 1- {1\over 8}\left ({\Delta g\over g}\right )^2\right ] \left [ 1-{\dot r^2\over 2}-{r^2\dot\theta^2 \over 2 \langle e^{G\hat\Gamma} \rangle } \left( 1+ {1\over 2}\left ({\Delta g\over g}\right ) ^2 \right) \right ] \label{lefnr} \end{equation} where $\left ({\Delta g\over g}\right )^2={\Delta^2\over \langle e^{G\hat\Gamma} \rangle^2}$. We can see from the Eq. (\ref{lefnr}) that in this nonrelativistic limit the effective lagrangian has, up to an irrelevant constant factor, the same form that $\bar L$, but with a different deficit angle. Indeed, after the redefinition of the angular variable $\theta \rightarrow \sqrt{\langle e^{G\hat\Gamma}\rangle} \left( 1- {1\over 4}\left ({\Delta g\over g}\right )^2 \right)\theta$, the effective lagrangian becomes proportional to the flat spacetime lagrangian. Therefore the trajectories will be straight lines in a locally flat spacetime. However, the global properties of the trajectories will be different from the ones obtained with the mean value of the metric $\langle e^{G\hat\Gamma} \rangle$, since the deficit angle for the effective lagrangian is now given by $2\pi (1- \sqrt{\langle e^{G\hat\Gamma}\rangle} \left( 1- {1\over 4}\left ({\Delta g\over g}\right )^2 \right) )$. In the general case (a relativistic particle), the situation is different. Indeed, one can prove that it is not possible to redefine $\theta$ in order to bring $L_{\rm eff}$ (Eq. (\ref{leff})) to a flat spacetime form. As a consequence, although the mean value of the metric is locally flat, the test particle ``sees'' a much more complex geometry. The conclusion of this section is that, again, the trajectories of the test particle do not coincide with the geodesics of the mean value of the metric. \section{FINAL REMARKS} Let us summarize the new results contained in this paper. We have computed the quantum corrections to the trajectory of a test particle by taking into account the quantum fluctuations of the spacetime metric. We have analyzed two particular models where it is easy to fix completely the gauge of the quantum fluctuations and quantize the remaining degrees of freedom. For a Robertson Walker spacetime, the fluctuations of the metric can be described by two massless, minimally coupled scalar fields. The quantum corrected trajectory has the same symmetries as the classical trajectory. However, it contains a quantum correction proportional to the graviton two point function and to the initial velocity of the test particle. This additional term produces, in particular, a quantum correction to the gravitational redshift. Let us assume that we solve the backreaction equations perturbatively and find a solution $a(t)=a_c(t)+\delta a(t)$, where $a_c(t)$ is the classical scale factor. Had we neglected the coupling between gravitons and test particle, we would have concluded that the test particle's trajectories coincide with the geodesics of the metric $a(t)=a_c(t)+\delta a(t)$. However, this coupling induces an additional correction to the equation of motion that is of the same order of magnitude as the one produced by $\delta a(t)$. (we have shown this in the particular case of a de Sitter solution and, in a previous paper \cite{nos}, in the Newtonian approximation). As a consequence, it is meaningless to compute $\delta a(t)$ and neglect the graviton effects on the motion of the particle, which is the physical observable. An interesting feature of our result is that the quantum corrections to the geodesic depend on the velocity of the test particle in such a way that one cannot define an "effective metric" for the trajectory, i.e. a metric such that its geodesics coincide with {\it all} the quantum corrected trajectories. It is worth to note that, if one tries to define observationally an "effective spacetime curvature" through a geodesic deviation equation, this effective curvature will be dependent on the initial four velocity of the geodesics under consideration. In the case of three dimensional general relativity, there are no propagating degrees of freedom associated to the geometry. At the classical level one can make the degrees of freedom to reside in the matter field. At the quantum level, the operator associated to the metric can be written in terms of the matter field operator. In this model, given a quantum state of the matter fields, it is easy to compute the mean value of the metric and of any function of it. In particular, we computed the mean value of the lagrangian for a test particle. We have shown that, even in the asymptotic region, where both the classical metric and the mean value of the quantum metric operator describe locally flat spacetimes, the test particle "feels" the quantum fluctuations and the trajectory is not a straight line. Now we would like to comment about related works. To our knowledge, the fact that the mean value of the metric is not enough to describe the spacetime geometry when the graviton contribution is taken into account, was first pointed out in Ref. \cite{vilkocqg}. It was stressed there that one can assign an effective metric to a given observable ${\cal O} (g_{\mu\nu})$, through the identity \begin{equation} g^{\rm eff}_{\mu\nu}= {\cal O}^{-1}\langle {\cal O} (g_{\mu\nu})\rangle \end{equation} The effective metric obviously depends on $\cal O$. We agree with this point of view. Indeed, from our results it is easy to illustrate this fact. Consider for example the quantum corrected velocity of the test particle given in Eq. (\ref{qvel}). Taking into account the classical result for the velocity, one can introduce an "effective scale factor" through the identity \begin{equation} \frac{\alpha a^{-2}_{\rm eff}(t)}{\sqrt{1+\alpha^2 a^{-2}_{\rm eff}(t)}} = \frac{\alpha a^{-2}(t)}{\sqrt{1+\alpha^2 a^{-2}(t)}} \left( 1 + \frac{32}{3} \pi G \langle \phi^2(t) \rangle \frac{\alpha^2 a^{-2}(t)}{1+\alpha^2 a^{-2}(t)} \right) \label{qvel2} \end{equation} This gives ${a_{\rm eff}\over a}\simeq 1-{\alpha^2+a^2\over \alpha^2+2 a^2} \frac{32}{3} \pi G \langle \phi^2(t) \rangle \frac{\alpha^2 a^{-2}(t)}{1+\alpha^2 a^{-2}(t)}$. The "effective scale factor" depends on the initial velocity of the particle. In Ref.\cite{fordsvai} the authors analyzed the graviton induced fluctuations of horizons in Robertson Walker and Schwarszchild spacetimes. The analysis was based on the study of the effects of gravitons on (nearly) null geodesics. They pointed out that, due to the interaction with the fluctuations of the metric, there are two effects on the trajectories of photons: the mean geodesic will deviate from the classical geodesic, and there will be stochastic fluctuations around the mean value. They studied the stochastic fluctuations and neglected the deviation of the mean value. In this sense, our work is complementary to Ref. \cite{fordsvai}, since we computed the mean value corrections. In our framework, the stochastic fluctuations could be analyzed by using the CTP formalism to compute the effective action for the test particle. It can be shown that the imaginary part of this CTP effective action introduces a noise term in the equation of motion (similar ideas have been applied to the semiclassical Einstein equations, see for example \cite{varios}). In this paper we fixed completely the gauge of the metric fluctuations before quantization. Alternatively, one could use the covariant method described in Section II. We showed in a previous work \cite{nos} that the solution to the backreaction equation and the quantum corrections to the geodesics are both dependent on the gauge fixing procedure. In the Newtonian approximation, this dependence cancels when computing the trajectory of the test particle. Whether this is true or not beyond the Newtonian approximation is an open question, that will be addressed in a forthcoming paper. \section{Acknowledgments} We acknowledge the support from Universidad de Buenos Aires, Fundaci\'on Antorchas and CONICET (Argentina). We would like to thank M. Banados and D. Tiglio for pointing out Refs. \cite{ash1,ash2} to us. \newpage
\section*{} Superkamiokande~\cite{REF:SKAM,REF:Totsuka,REF:Inou} has recently observed an excess of solar-neutrino events at electron energies higher than $13~$MeV. This excess cannot be interpreted as a distortion of the boron neutrino spectrum due to neutrino oscillations~\cite{REF:SKAM,REF:Totsuka,REF:Inou,REF:BKS}, if one restricts oneself to those oscillation solutions that explain the observed gallium, chlorine and water-cerenkov neutrino rates. It is tempting to think that this excess is the result of low statistics or small systematic errors at the end of the boron neutrino spectrum. For example, because of very steep end of the electron spectrum, even small systematic error in electron energy ({\em e.g.}, due to calibration) could enhance the number of events in the highest energy bins. One should wait for future Superkamiokande data, where such possible systematic effects will be further elaborated. The data from the SNO detector, which will come in the operation soon, {\em e.g.}, see Ref.~\cite{REF:SNO}, can shed light on this excess. Another possible explanation of this excess~\cite{escri,REF:BaKr} is that the Hep neutrino flux might be significantly larger (about a factor 10--20) than the SSM prediction. The Hep flux depends on solar properties, such as the $^3$He abundance and the temperature, and on $S_{13}$, the zero-energy astrophysical $S$-factor of the \mbox{$p+{}^3He \rightarrow {}^4He + e^+ + \nu$} reaction. Both SSM based~\cite{REF:BaKr} and model-independent~\cite{REF:BDFR} approaches give a robust prediction for the ratio $\Phi_{\nu}(Hep)/S_{13}$. Therefore, this scenario implies a cross-section larger by a factor 10--20 than the present calculations (for reviews see~\cite{REF:BaKr,INT}). Such a large correction to the calculation does not seem likely, though it is not excluded. A large Hep neutrino flux remains a possible explanation of the excess. The signature of Hep neutrinos, the presence of electrons above the maximum boron neutrino energy, can be tested by the SNO experiment. The Superkamiokande collaboration noticed~\cite{REF:SKAM} that vacuum oscillations with large $\Delta m^2$ explain the observed high-energy excess. However, the same Ref.~\cite{REF:SKAM} emphasizes that those oscillation parameters that reproduce the excess do not solve the Solar Neutrino Problem (SNP), {\em i.e.}, they do not explain the global rates observed by the four solar neutrino experiments. In Ref.~\cite{REF:dvo} it was demonstrated that if the SSM prediction for the boron neutrino flux is reduced by factor $f_B= 0.8$ and the chlorine experimental signal is arbitrarily assumed to be larger by a factor $f_{Cl} =1.3$, the vacuum oscillation solution to SNP (global rates) corresponds to $\Delta m^2 = 4.2\cdot 10^{-10}$~eV$^2$ and $\sin^2 2\theta=0.93$: this choice of parameters reproduces also the excess of high-energy recoil electrons in the Superkamiokande spectrum~\cite{REF:SKAM}. In this paper we shall further elaborate upon this specific vacuum oscillation solution. For the sake of conciseness, we shall refer to this solution ($\Delta m^2 = 4.2\cdot 10^{-10}$~eV$^2$ and $\sin^2 2\theta=0.93$) as HEE (High-Energy Excess) VO. We shall start with a short description of relevant features of those vacuum oscillation solutions, whose parameters fit all global rates; these solutions will be indicated altogether as VO. Vacuum oscillations can reconcile the SSM with the observed rates of all three kinds of solar neutrino experiments (for reviews see~\cite{REF:Bahc,REF:BiPe,REF:Turck,REF:Hax}). A recent detailed study~\cite{REF:HaLa95,REF:BK96,REF:KP96,REF:HaLa97,REF:Lisi} of VO solutions shows that global fits to the data result in oscillation parameters within the ranges $5\cdot 10^{-11}$~eV$^2\leq \Delta m^2 \leq 1\cdot 10^{-10}$~eV$^2$ and $0.7\leq\sin^2 2\theta\leq 1$ for oscillations between active neutrinos. The large range of $\sin^2 2\theta$ is mainly caused by uncertainties in the B-neutrino flux, though other uncertainties contribute too; $\Delta m^2$ is much less sensitive to changes of the B-neutrino flux. These effects have been explicitly investigated in Refs.~\cite{REF:KrSm,REF:KP96}. In the SSM the B-neutrino flux uncertainties ($+19\%,-14\%, $~\cite{REF:BP98}) are mainly caused by the uncertainties in $S_{17}$ (the $p$-Be cross section is poorly known) and by the strong temperature dependence of this flux. The above uncertainties are only $1\sigma$ errors and the actual discrepancy could be larger, especially due to the $S_{17}$ factor. This large uncertainty of the B-neutrino flux has motivated several authors to consider the boron flux as $\Phi_B=f_B\Phi_B^{SSM}$ with $f_B$ as a free parameter~\cite{REF:BKS,REF:KrSm,REF:HaLa95,REF:KP96}. A signature of vacuum oscillation is the anomalous seasonal variation of the neutrino flux at low energies~\cite{REF:Pom,REF:BiPe}. The distance between the Sun and the Earth varies during the year by about 3\% affecting the detected flux both because of the $1/r^2$ geometrical factor and because of the dependence of the survival probability $P(\nu_e \to \nu_e)$ on the distance. The second effect is absent for MSW solutions. Nevertheless, the MSW solution also predicts seasonal variations of neutrino flux, which are connected with the day/night effect and are caused by the longer winter nights (for recent calculations see~\cite{Valle,BKS1} and references to earlier works therein). The MSW seasonal variations are weaker than the VO ones at low energies. In the case of VO, the monochromatic Be-neutrinos are expected to show the strongest seasonal variations~\cite{REF:BiPo,REF:GK,REF:KrPe95,REF:Lisi}; on the contrary, Be-neutrinos should show very small seasonal variations in the case of MSW oscillations. Since Be-neutrinos are monochromatic, their flux shows the entire seasonal variation predicted by VO; the effect is reduced for the other fluxes due to the averaging over the different phases of neutrinos with different energies within the interval of observation, $\Delta E$. Seasonal variations for $\Delta m^2$ larger than the values allowed by VO solutions were recently analyzed in Ref.~\cite{REF:GKK}. The authors found some significant consequences such as energy dependence and correlation with distortion of the spectrum. The latter effect was also discussed earlier in Ref.~\cite{REF:MS}. In relation to the chlorine signal, seasonal variations were analysed in early work~\cite{REF:EHRLICH}. A clear discussion of the seasonal variation effect has been presented in Refs.~\cite{REF:Rosen}. To explain the excess in electron spectrum observed by Superkamiokande we allow a boron neutrino flux 15--20\% smaller than the SSM prediction, and we allow that the chlorine signal be about 30\% larger than the Homestake observation. This assumed $3.4\sigma$ increase could have a combined statistical and systematic origin though we do not have any concrete argument in favor of such systematic error in the Homestake experiment. In our calculation, we shall use neutrino fluxes from the BP98 model~\cite{REF:BP98} with the B-neutrino flux rescaled as $\Phi_B= f_B \Phi_B^{SSM}$. For the chlorine rate we assume $R_{Cl} = 2.56 f_{Cl}$~SNU (the Homestake experiment gives the rate~\cite{REF:Hom} $2.56 \pm 0.16 \pm 0.16$~SNU). It is easy to see that for $f_{Cl}=1.3$ the assumed signal 3.33~SNU is $3.4\sigma$ higher than one given by Homestake (systematic and statistical error are incoherently combined). For the gallium rate we use the average of the GALLEX~\cite{REF:Ki} and SAGE~\cite{REF:Ga} results: $72.5 \pm 5.7$~SNU. Finally, we take the Superkamiokande result~\cite{REF:SKAM}: $(2.46 \pm 0.09)\cdot 10^6$~cm$^{-2}$s$^{-1}$. For each pair $f_B$ and $f_{Cl}$ we find the VO solution, {\em i.e.}, the parameters ($\Delta m^2, \sin^22\theta$), that explain the observed rates, and then we calculate the corresponding boron neutrino spectrum. For example, for $f_B=0.8$ and $f_{Cl}=1.3$ the oscillation parameters ($\Delta m^2=4.2 \cdot 10^{-10}$~eV$^2, \sin^22\theta=0.93$) give a good fit to all rates ($\chi^2$/d.o.f. = 3.0/3): This is not the best fit point, which has $\chi^2\approx 0$, therefore the 3 d.o.f. are the three experimental rates. On the other hand, the spectrum with these oscillation parameters reproduces~\cite{REF:SKAM} the excess of high-energy events observed in the Superkamiokande spectrum. More generally, this choice of oscillation parameters gives rates in agreement with the experiments at the $2\sigma$ level for $0.77 \leq f_B \leq 0.83$ and $1.3 \leq f_{Cl} \leq 1.55$. In Fig.~1 we present the neutrino-induced electron spectra for the vacuum oscillation solutions as the ratio to the SSM unmodified spectrum~\cite{REF:BP98}. The dotted and dashed curves show two spectra corresponding to the VO solutions of Ref.~\cite{REF:BKS} and Ref.~\cite{REF:HaLa97}, respectively. The solid line shows the VO oscillation solution that is discussed in this paper (HEE VO) corresponding to $\Delta m^2=4.2\cdot 10^{-10}$~eV$^2$ and $\sin^2 2\theta=0.93$ ($f_B=0.8$ and $f_{Cl}=1.3$). The role of the two parameters, $f_B$ and $f_{Cl}$, for the best fit of the spectrum is different: while $f_B$ mostly changes $\sin^2 2\theta$, $f_{Cl}$ affects $\Delta m^2$ and, therefore, the spectrum. Values of $f_{Cl}$ as low as 1.2 already give a bad fit to the observed spectrum. The anomalous seasonal variations of Be-neutrino flux and of the gallium signal are shown in the Fig.~2 (see also \cite{REF:dvo}). Anomalous seasonal variation is described by the survival probability of the electron neutrino $P(\nu_e \to \nu_e)$. For Be-neutrinos with energy $E=0.862$~MeV the survival probability (the suppression factor for electron neutrinos) is given by \begin{equation} \label{EQ1} P(\nu_e \to \nu_e)= 1-\sin^2 2\theta \sin^2\left( \frac{\Delta m^2 a}{4E}\, (1+e\cos \frac{2\pi t}{T} ) \, \right) \, , \end{equation} where $a=1.496\cdot 10^{13}$~cm is the semimajor axis, $e=0.01675$ is the eccentricity of the Earth's orbit, and $T=1$~yr is the orbital period. The phase in Eq.~(\ref{EQ1}) is such that $t=0$ corresponds to the aphelion. In Fig.~2 the solid and dashed curves show the variation of the Be-neutrino flux for the HEE VO and VO~\cite{REF:BKS} cases, respectively. The case of the HEE VO (solid curve) is dramatically different from the VO case: there are two maxima and minima during one year and the survival probability oscillates between $1-\sin^2 2\theta \approx 0.14$ and 1. The explanation is obvious: the HEE VO solution has a large $\Delta m^2$, which results in a phase $\Delta m^2 a/(4E) \approx 93$, large enough to produce two full harmonics during one year, when the phase changes by about 3\% due to the factor $(1+e\cos 2\pi t/T)$. The flat central maximum with a shallow local minimum has a trivial origin: the extrema of $P(\nu_e \to \nu_e)$ in Eq.~(\ref{EQ1}) correspond to phases $k\pi/2$, where k are integers, and to the phases with $\cos 2\pi t/T = \pm 1$. The accidental proximity of these phases can result in three nearby extrema. The shallow minimum in Fig.~2 disappears with small changes in $\Delta m^2$. The phases of maxima and minima in terms of t/T are not fixed in the HEE VO solution, because tiny changes of $\Delta m^2$ shift their positions: {\em e.g.}, 1\% change in $\Delta m^2$ shifts the position of an extremum by more than one month (see Eq.~(\ref{EQ1})). As one can see from Fig.~2, the HEE VO solution predicts that the beryllium electron neutrinos should arrive almost unsuppressed during about four months in a year! According to the SSM, beryllium neutrinos contribute 34.4~SNU out of the total gallium signal of 129~SNU. Therefore, the strong $^7$Be neutrino oscillation predicted by the HEE VO solution also implies an appreciable variation of total gallium signal. In Fig.~2 the dotted curve shows this variation corresponding to the HEE VO solution, which can be compared with the weaker variation corresponding to the best-fit VO solution (dashed-dotted curve). It is possible that the HEE VO variation could already be partially testable by the existing gallium data, and this possibility will significantly increase when the results from GNO with its larger statistics are available. In Fig.~3 the predicted time variation of the gallium signal is compared with GALLEX data (see also \cite{REF:dvo}). GALLEX data have been analysed according to the time of the year of the exposures and grouped in six two-month bins (M.~Cribier cited in \cite{REF:Ki}): the data points with error bars in Fig.~3 reproduce the result of this analysis. The data give the rates averaged for the same two months every year of observations. The theoretical prediction (solid curve) is plotted with the same averaging. The $7\%$ geometrical variation is included. Both the phase of the time-variation and the average flux have been taken to fit the data. The fit by the theoretical curve has $\chi^2$/d.o.f.= 0.85/4; the fit by a nonoscillating signal is also good: $\chi^2$/d.o.f.=1.36/5. Because of the limited statistics, we do not interpret the good visual agreement in Fig.~3 as a proof of HEE VO solution, though it is certainly suggestive. The comparison of the predicted time variation with preliminary data \cite{REF:Ga} of the other gallium detector SAGE is shown in Fig.~4. Note that this time we can not choose the phase arbitrary: it is already fixed by the fit to the GALLEX data. Because of the larger fluctuations of the SAGE data (compare Fig.~3 and Fig.~4) the agreement with the predicted variation is worse. In Fig.~5 the predicted variation is compared with the Homestake data (see Ref.~\cite{REF:EHRLICH} for an earlier analysis of indications for seasonal oscillations in the Homestake data). The phase of the theoretical dependence is kept fixed at the value fitted to the GALLEX data. As for the GALLEX data we find that the HEE VO theoretical curve gives a better fit ($\chi^2$/d.o.f.=1.4/5) than the time-independent signal, which however cannot be excluded ($\chi^2$/d.o.f.=3.1/5). This agreement is further strengthened by the fact that the phase of time dependence was not chosen to fit the Homestake data, since it was already fixed by the GALLEX data. One should consider this agreement as additional indication for the HEE VO solution. Finally, in Fig.~6 (Fig.~7) we compare the time variation of the Superkamiokande signal for recoil electrons with energies higher than 10~MeV (11.5~MeV) with the HEE VO predictions. The fit of the data is good: $\chi^2$/d.o.f.=2.7/7 ($\chi^2$/d.o.f.=5.1/7). Similar calculations were done by the Superkamiokande collaboration~\cite{REF:Suzuki}, by A.~Smirnov (private communication) and by M.~Maris and S.~Petcov~\cite{REF:Petcov}. While the agreement between the HEE VO solution and each single observational datum on seasonal variations might appear accidental and not statistically significant, the combined agreement with all data on seasonal variations, as shown in Figs.~3--7 (total $\chi^2/d.o.f = 34.7/41$), appears to be quite a suggestion in favour of the HEE VO solution. As in our previous work~\cite{REF:dvo}, we prefer not to make a global fit in terms of $\chi^2/d.o.f.$ to all available data (rates, spectrum and time variations). The large number of degrees of freedom can hide a discrepancy with some particular data, especially if it corresponds to only one degree of freedom, like the chlorine rate in our case. A small $\chi^2$ is only a necessary condition for the correct model. One can find such a global fit in the paper by Barger and Whisnant~\cite{BarWhi}, which appeared after this work was completed. The authors study the VO solution with $\Delta m^2=4.42\cdot 10^{-10}$~eV$^2$ and $\sin^2 2\theta=0.93$, {\em i.e.}, parameters close the ones we consider. They find this solution as the global best fit to the rates, spectrum and time dependence of SuperKamiokande signal ($\chi^2=39$ for 26 degrees of freedom). In conclusion, the combination of a B-neutrino flux 20\% lower than in the SSM (easily allowed by the present uncertainties) and of the assumption that the chlorine signal be $30\%$ ($3.4\sigma$) higher than the one presently observed by Homestake results in a vacuum oscillation solution (HEE VO) that fits the electron spectrum recently observed by Superkamiokande. This solution predicts strong seasonal variation of $^7$Be-neutrino flux: some indication to such a variation is already seen in the GALLEX and Homestake data. Seasonal dependence of the Superkamiokande data for electron energies higher than 10~MeV and 11.5~MeV provide further indication in favour of the HEE VO solution. The anomalous seasonal variation of Be-neutrino flux predicted by the HEE VO solution can be reliably observed by the future BOREXINO~\cite{REF:BOREXINO} and LENS~\cite{REF:LENS} detectors. Additionally, LENS, which should measure the flux and spectrum of $pp$ neutrinos, will be able to observe the suppression of $pp$ neutrino flux, $P(\nu_e \to \nu_e)= 1- (1/2)\sin^2 2\theta=0.53$, which is another signature of VO solutions. \section*{Acknowledgements} We are grateful to A.~Bettini, T.Kirsten and A.~Yu.~Smirnov for useful discussions. The participation of F.~Villante at some stage of work is gratefully acknowledged.
\section{Introduction} The element lithium has provided many clues concerning stellar evolution and nucleosynthesis, as well as cosmology. However, its origin is still far from clear. Nowadays there is an increasing belief that this element has a multi-source nature: primordial nucleosynthesis (see the review by Wallerstein et al. 1997), galactic cosmic ray spallation in the interstellar medium and/or in the neighbourhood of supernova remnants (Meneguzzi et al. 1971; Feltzing \& Gustafsson 1994), late type stars and novae (Abia et al. 1993a; Hernanz et al. 1996), supernovae explosions (Woosley \& Weaver 1995) and even spallation reactions around compact objects (Guessoum \& Kazanas 1998). However, except perhaps for the Li production during the Big Bang there is no consensus about the contribution of these sources to the present cosmic abundance, log $\epsilon$(Li)$\approx 3.3$\footnote{The abundance of a given element X is noted as log $\epsilon$(X)$\equiv 12 +$log(N(X)/N(H)) where N(X)/N(H) is the abundance by number of the element X.} The discovery of an unusually strong Li line at $\lambda6708$ {\AA} by McKellar (1940) in the carbon star (C-star) WZ Cas strongly suggested that AGB stars might be an important source of Li in the galaxy. The work of McKellar was followed by similar discoveries in other AGB stars: the C-stars WX Cyg and IY Hya (Sanford 1950; Abia et al. 1991), the S-stars T Ara and T Sgr (Feast 1974; Boesgaard 1970) and the SC-stars Henize 166 and VX Aql (Catchpole \& Feast 1971; Warner \& Dean 1970). The same figure has also been reported in AGB stars of the Magellanic Clouds (Plez et al. 1993; Smith et al. 1995). The pioneering analysis by Cohen (1974) showed that the measured equivalent width of the $\lambda6708$ {\AA} Li I line (W$_{\lambda}\sim 1-10$ {\AA}) in these stars is mainly the consequence of a strong enhancement of the lithium abundance in the atmosphere. Quantitative determinations by spectral synthesis (Denn et al. 1991; Abia et al. 1993b; Plez et al. 1993; Smith et al. 1995) have shown that the abundance of Li in these stars is 1-2 orders of magnitude higher than the present cosmic Li abundance. Therefore, these stars have been named super lithium-rich (SLiR) stars and might well constitute the main source of Li in the galaxy. The production of Li in AGB stars has received special attention in recent years (e.g. Sackmann \& Boothroyd 1992). Basically, Li is produced by the reaction $\rm{^3He(^4He,\gamma)^7Be}$, followed by $\rm{^7Be(e^-,\nu)^7Li}$ in a hot convective region that brings the $\rm{^7Be}$ or $\rm{^7Li}$ to cooler regions before the $^7$Li is destroyed by $\rm{^7Li(p,\alpha)^4He}$. This is the so-called $^7$Be-transport mechanism (Cameron \& Fowler 1971). Results of these theoretical studies quantitatively agree with the Li abundances derived in AGB stars of the galaxy and the Magellanic Clouds, although there are still many open questions related to the Li production in these stars. Among others we find questions about the minimum initial stellar mass which can eventually become a SLiR star during the AGB phase, the duration of the SLiR phase or the actual Li yield into the interstellar medium. Our purpose in this work is related to the reality of the Li abundances in SLiR stars and its consequences on the Li yield. At present, the uncertainty in the derivation of the Li abundances in AGB stars is not lower than 0.4-0.5 dex (see references above), mainly due to uncertainties in the stellar parameters and the fit to the observed spectra. Note, that in many situations no single set of stellar parameters is found to fit the observations. However, there are also systematic errors that are not usually taken into account that might dramatically change the Li abundance derived: uncertainties in the atmosphere models, the existence of velocity stratifications (most SLiR stars are actually variable), sphericity and NLTE effects are not currently considered as possible systematic sources of error. Of course, the consequences of each of these phenomena on Li abundance merit individual study and are beyond the scope of this work (see, however, Scholtz 1992; J\o rgensen et al. 1992). Here, we will focus our attention on the effects of departures from LTE in the formation of the Li lines in C-stars. The study of NLTE effects in the formation of lithium lines in stellar atmospheres was begun by the work of M\"uller et al. (1975). They performed NLTE analysis of the very weak lithium resonance line $\lambda 6708$ {\AA} in the spectrum of the Sun. Later, Luck (1977) investigated the statistical balance of lithium in the atmospheres of G-K giants for a 4-level atom model of Li. A similar atom model was used by de la Reza \& Querci (1978) and de la Reza et al. (1981). Historically, these papers considered for the first time the impact of a stellar chromosphere on the lithium lines. A new stage of research was begun with the work of Steenbock \& Holweger (1984), who used the technique of complete linearization for an 8-level atom model. They studied NLTE effects in lithium lines in atmospheres of dwarfs and giants. Pavlenko (1991) considered in detail the effects of deviation from LTE in the atmosphere of red giants. Later, Magazz\'u et al. (1992), Martin et al. (1994) and Pavlenko (1994), continued the NLTE studies in T-Tau stars, G-K giants, subgiants and dwarfs. The main results of these studies were confirmed by the independent work of Carlsson et al. (1994), who made a similar investigation using atmosphere models with various effective temperatures, luminosities and metallicities. Finally, Houdebine et al. (1995), Pavlenko et al. (1995), Pavlenko \& Magazz\'u (1996) and Martin et al. (1997) have continued the studies of different aspects of NLTE formation of lithium lines in stellar atmospheres. As far as we know, the sole study on this subject for AGB stars is that by de la Reza \& Querci (1978) who performed kinetic equilibrium calculations of neutral lithium lines in C-stars and determined the influence of the possible chromospheric radiation into the photosphere. In the present work we revise this study using up to date atomic data, collisional and radiative rates, an extended Li atom model and more reliable model atmospheres for C-stars in a wider range of effective temperatures (T$\rm{_{eff}}=2500-3100$ K) and C/O ratios (1.0-1.35) (see below). We explicitly apply our results deriving Li abundances from synthetic spectra, both in LTE and NLTE, in three well known SLiR stars (WX Cyg, WZ Cas and IY Hya) from four accessible Li I lines: the resonance line at $\lambda6708$ {\AA} and the subordinate transitions at $\lambda4603$, $\lambda6104$ and $\lambda8126$ {\AA}, respectively, benefiting from the high signal-to-noise ratio and high resolution spectra of these stars. The consequences on the real Li abundances in AGB stars and on their net Li yield into the interstellar medium is then reexamined. \section{Observations} The observations were made during 1997 and 1998 in two different observatories. We used the 4.2 m WHT at the Observatory of El Roque de los Muchachos with the Utrecht Echelle Spectrograph as the main instrument and a 2048$\times$2048 CCD with 24 $\mu$m pixel size. We used the 79.0 lines/mm grating which provides less wavelength coverage, but more space between orders (20-30 arcsec). The projected size of the slit on the chip was around two pixels which gave a resolving power of 50000, the effective resolution ranging between 0.05-0.19 {\AA} from the blue orders to the red ones. The total number of orders on the chip were 30 covering the wavelength range 0.4-1.0 $\mu$m with some gaps between orders. WZ Cas and WX Cyg were also observed by the 2.2 m telescope at the Calar Alto Observatory. For this observational run a fibre optics cassegrain echelle spectrograph (FOCES) (Pfeiffer et al. 1998) was used. This time the chip was a 1024$\times$1024 Tektronik CCD with 24 $\mu$m pixel size. The FOCES image covers the visible spectral region from 0.38 to 0.96 $\mu$m in about 80 orders with full spectral coverage. Spectral orders are separated by 20 pixels in the blue and 10 in the red. The maximum resolving power is 40000 with a two pixel resolution element. \begin{table*} \caption[]{Log of the observations and stellar parameters} {!}{\includegraphics{tabnlte1.eps}} \end{table*} The reduction of the spectra was made following the standard procedures using the ECHELLE task of the IRAF software package: bias subtraction, division by flat-field images, removal of the scattered-light, extraction of the orders and wavelength calibration with Th-Ar lamps. Typically the rms in the residuals of the calibrations were better than 10 m\AA. Finally, the spectra were divided by the spectrum of a hot rapidly rotating star to remove telluric absorptions, although they are only important in the $\lambda8126$ {\AA} range. The signal-to-noise ratio of the spectra vary along the wavelength, from S/N$\sim 20$ in the bluest orders to S/N$\sim 700$ around $\lambda8000$ {\AA}. Since our stars are relatively bright objects (V$\sim 7-11$), this strong variation in the S/N ratio achieved is mainly due to the very low emissivity of the stars below $\lambda\sim 4500$ \AA. This is particularly evident in IY Hya, which is an extremely red object. For this reason, its spectrum in the $\lambda4603$ {\AA} range is of too low quality (S/N$\sim 30$) and has not been considered. \section{The abundance analysis} \subsection{Atomic and molecular line lists} The line list described in L\`ebre et al. (1999) completed with C$_2$ lines has been used for the analysis of the $\lambda6708$~\AA\ region. Lines from the $^{12}$C$_2$, $^{12}$C$^{13}$C and $^{13}$C$_2$ Swan system (Kurucz 1998) were considered and from the Phillips red system were predicted as in de Laverny \& Gustafsson (1998). The line lists for the $\lambda4603, \lambda6104$ and $\lambda8126$ {\AA} spectral domains were built in a similar way for the C$_2$ and CN molecules, and atomic line data were found in VALD (Piskunov et al. 1995). The broadening by radiation and van der Waals damping were calculated as in de Laverny \& Gustafsson (1998). Line data were adjusted as described in L\`ebre et al. (1998) by comparison with the solar spectrum using the Holweger \& M\"uller (1974) model atmosphere for the Sun with element abundances from Anders \& Grevesse (1989). A good fit to the solar spectrum was obtained in the $\lambda6708$ and $\lambda8126$ {\AA} spectral domains in $\sim 50$ {\AA} around the corresponding Li line. The fit, however, was not as good in the other two spectral ranges. Indeed, in the $\lambda4603$ and $\lambda6104$ {\AA} regions there are a number of unidentified features in the solar spectrum with a modest absorption ($\sim 5-25$ m{\AA}). Most of these features could be atomic in nature because no such intense molecular absorptions are expected in the Sun in these spectral ranges. However, we did not find any atomic lines in the data base of VALD or of Kurucz (1998) at the wavelengths of the missing features. Nevertheless, we were able to obtain a good fit to the solar spectrum in at least 10 {\AA} around the $\lambda4603$ {\AA} and $\lambda6104$ {\AA} Li I features. Thus, we do not believe this problem will introduce an additional source of error into the abundance of Li derived from these two spectral domains. \subsection{Atmospheric parameters} Stellar parameters for the stars studied were taken from the literature when available. For WX Cyg the effective temperature derived by Ohnaka \& Tsuji (1996) from the infrared flux method (IRFM) was adopted. For WZ Cas, effective temperature derived from the IRFM and angular radii measurements (Dyck et al. 1996) agree quite well ($\pm 30$ K). However, the value derived from these two methods (3150 K) contrasts with that derived from infrared photometry ($\sim 2800$ K; see Noguchi et al. 1981; Frogel et al. 1972). We adopted here the value obtained by Dyck et al. (1996). For IY Hya there is no estimate of its effective temperature nor its photometry is available in the literature. We estimated the effective temperature by comparing its spectrum with the temperature sequence spectral atlas for C-stars created by Barnbaum et al. (1996). From that comparison we believe that IY Hya is a cool N-star and consequently, we adopted the lower effective temperature in our grid of atmosphere models i.e.: T$_{\rm{eff}}=2500$ K. The uncertainty in T$_{\rm{eff}}$ is, however, no less than $\pm 200$ K; it might well be larger for IY Hya. Note also that these stars are variable and we suspect T$_{\rm eff}$~variations as large as $\sim 300$ K during their cycle (see Richichi et al. 1995). The consideration of an unique effective temperature for these stars is therefore a strong assumption. A solar metallicity ([Fe/H]=0.0) was considered for the three stars. Most galactic C-stars are of Population I with near solar metallicity although typically they show strong enhancements of heavy elements (Zr,Ba,La etc...)(Utsumi 1985; Dominy 1985). We checked that there is no intense heavy element line close enough to the four Li absorptions in our line lists that may affect the Li abundance derivation. A study of the heavy element enhancements in the SLiR stars and other normal C-stars will be presented in a separate work. A gravity of log g$=0.0$ and a microturbulence parameter of $\xi=2.5$ kms$^{-1}$ were adopted. These are typical values for C-stars (see Lambert et al. 1986). Concerning gravity, note that studies by Pavlenko (1990) and Magazz\'u et al. (1992) on O-rich dwarfs, subgiants and giant stars showed no important sensitivity in the ratio of LTE/NLTE Li abundances caused by changes in the gravity of $\Delta$log g$=\pm 1$, especially in the case of strong lines. Thus, uncertainties in gravity play a minor role in the formation of the lithium lines in C-rich atmospheres. Radiative damping (which does not depend on log g) plays the main role in the formation of the wings of the saturated lithium lines. More serious seem to be the consequences of changes in the temperature structure of the atmosphere with decreasing gravity due to the increase in the effectiveness of sphericity effects. However, analysis of these effects lies beyond the scope of this paper. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte1.eps}} \caption{Theoretical LTE (solid) and NLTE (dotted) curves of growth for the four Li I absorptions for different~T$_{\rm eff}$. The pairs of curves (solid-dotted) from top to bottom in each plot are for~T$_{\rm eff}$=2500, 2800 and 3000 K, respectively.} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte2.eps}} \caption{Theoretical LTE (solid) and NLTE (dotted) curves of growth for the four Li I absorptions for different C/O ratios. The pairs of curves (solid-dotted) from top to bottom in each plot are for C/O=1.007, 1.02, 1.1 and 1.35, respectively.} \end{figure*} Finally, a model atmosphere was interpolated in T$_{\rm{eff}}$ and the C/O ratio for each star from an unpublished grid of models for C-stars (Eriksson et al., private communication). These models are constructed under the basic hypotheses of hydrostatic equilibrium and plane-parallel approximation. The models include more complete opacity data of HCN and C$_2$H$_2$ (in addition to diatomic molecules) with the sampling treated by an opacity distribution function (see Eriksson et al. 1984, for details). The final C/O ratio of the model atmosphere used in the spectral synthesis was obtained by an iterative procedure comparing synthetic and observed spectra until a good fit was obtained. Li lines were not included in this procedure. However, the same C/O ratio was not found to give the best fit in the four spectral domains for a given star, although differences in the C/O ratio never exceeded a few hundredths. For instance, in the $\lambda4603$ and $\lambda6104$ {\AA} regions we systematically derived a lower C/O ratio by $\sim 0.03$ hundredths. Table 1 shows the log of the observations and the final stellar parameters used in the analysis. The quoted CNO abundances are the mean values of those obtained from the best fit for each spectral domain (N abundance plays a minor role). \subsection{LTE abundances} Lithium abundances were derived by synthetic spectra {\bf generated by using a modified version of code written at the Uppsala Astronomical Observatory. The theoretical spectra were} convolved with the corresponding FWMH to match the instrumental profile in the four spectral domains following the procedure used in Abia et al. (1993b). The carbon isotopic ratios ($^{12}$C/$^{13}$C) derived in Abia \& Isern (1997) from spectral synthesis to the $^{12}$CN and $^{13}$CN lines in the $\lambda7990-8030$ {\AA} spectral region were adopted (see also Abia \& Isern (1996) and de Laverny \& Gustafsson (1998), for details on the analysis and the choice of the CN parameters, respectively). Table 1 shows the carbon isotopic ratio adopted in the stars. They are in fair agreement with those derived by Ohnaka \& Tsuji (1996) and Lambert et al. (1986) in WX Cyg and WZ Cas. The total error due to stellar atmosphere parameter uncertainties, continuum placement and the fit itself amount to $\pm 0.4-0.5$ dex for Li (LTE) abundances and $\pm 6-13$ for the $^{12}$C/$^{13}$C ratio (see references above). Figures 1 and 2 (solid lines) show the behaviour with T$_{\rm{eff}}$ and the C/O ratio of curves of growth calculated in LTE for the four Li absorptions. As can be clearly seen, subordinate lines are saturated for high Li abundances (log $\epsilon$(Li)$> 3$). For the resonance line, however, the Li absorption increases more steeply with increasing Li abundance. For very high Li abundances (log $\epsilon$(Li)$>4$), the subordinate lines are not very sensitive to the effective temperature, the C/O ratio or gravity (as mentioned before). This is a consequence of their saturated behaviour. The opposite occurs with the resonance line. This might lead us to consider the resonance line the better tool for abundance determinations in C-stars. However, the resonance line forms in the outermost layers (see below) where the atmosphere structure is quite uncertain in C-stars and the microturbulence may even become supersonic. We will also see that NLTE effects are much larger for this transition. Figures 3 to 6 show synthetic fits in LTE to the different spectral domains in WZ Cas. Synthetic spectra agree quite well with the observations for the $\lambda6708$ and $\lambda8126$ {\AA} regions but this is not the case for the $\lambda4603$ and $\lambda6104$ {\AA} ones. Some of the discrepancy is certainly due to the missing features in our line lists as mentioned in 3.1. Moroever, in the $\lambda4603$ and $\lambda6104$ {\AA} regions we obtain systematically larger residual fluxes than observed. This, by the way, means that the Li abundances derived from the $\lambda4603$ and $\lambda6104$ {\AA} Li lines are systematically lower (see Table 3) than those derived from the red Li lines. The same figure was found by de la Reza \& da Silva (1995) studying the formation of Li lines in K-giant stars. The reason for this is unknown. Bad gf-values for C$_2$ lines, which are numerous in these bluest spectral regions, might partially explain this problem. Note that in these two spectral domains most of the C$_2$ lines were taken from the data compilation by Kurucz (1998), which is probably not very accurate. Note also that we were not able to check their gf values with the solar spectrum as mentioned above. The discrepancy between theoretical and observed fluxes is particularly evident in the $\lambda$4603 {\AA} region. We confirmed that most of the strong absorptions in this spectral range are due to atomic lines while molecular lines contribute most to the background absorption. However, no reasonable reduction of the metallic abundances in the star can solve this discrepancy. Therefore, we are led to believe that an incorrect figure for continuous opacity might be the main cause of this. The problem of the missing opacity in the blue has been widely discussed in the literature. Recently there have been some new results in this sense. Yakovina \& Pavlenko (1998) showed that to fit the head of the strong NH band $A^3\Pi-X^3\Sigma^-$~ at $\lambda$336 nm in the spectrum of the Sun, one should increase the continuum absorption coefficient in the region by a factor $\kappa_\nu\sim 1.7 - 1.9$. Recently, Bell \& Balachandran (1998) increased $\kappa_\nu\sim 1.6$ to fit Be II lines at $\lambda$313.0 nm in the solar spectrum. We tried to artificially simulate this missing opacity by increasing the H$^-$ opacity by a free factor (namely, increasing the electronic pressure by a factor $\sim 3 $). Indeed, in this case theoretical residual fluxes are lower (except for the very strong (saturated) lines) and the agreement between observed and theoretical spectra improve. We preferred, nevertheless, to be coherent in our analysis in the four spectral domains and did not artificially change our set of conventional opacity sources. Thus, we used the same model atmosphere for all the spectral ranges and decided, for that reason, not to consider Li abundances derived from the $\lambda$4603 {\AA} range. We will have to wait for the next generation of atmosphere models for C-stars to study this problem better (Plez et al. 1999). Note that in principle sphericity effects should be more pronounced in the blue part of the spectrum because here the sensitivity to the temperature structure is higher even for strong spectroscopic features formed in the outer atmosphere. On the other hand, this discrepancy may be caused (at least partially) by the impact of dust opacity on the star's spectrum. Indeed, the formation of dust in the atmosphere of AGB stars and their envelopes has been known for a long time (see references in Wallerstein \& Knapp 1998). Dust particles can contribute to total opacity via scattering processes, i.e. with cross-sections $\sim 1/\lambda^n$. In that case the contribution of the dust opacity would increase to the blue. Li-rich K-giants also show an red excess, probably caused by a dusty envelope (although of a different nature; de la Reza \& da Silva 1995). One may propose other explanations to explain this effect. For instance, the procedure used to place the continuum level in the observed spectrum still seems very subjective. To provide more balanced conclusions regarding this problem we suggest that a more detailed study of these regions should be carried out in several SLiR stars. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte3.eps}} \caption{Synthetic spectral fit to the $\lambda4603$ {\AA} spectral domain in WZ Cas for log $\epsilon$(Li)=3.0. Note the strong discrepancy between observed (dots) and synthetic spectra in LTE (solid) and NLTE (dashed) (LTE and NLTE fits to the Li line almost coincide). Most of this discrepancy is probably due to an incorrect continuous opacity in this spectral range. The Li abundances derived from this line were not considered.} \label{} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte4.eps}} \caption{Synthetic spectral fit to the $\lambda6104$ {\AA} spectral domain of WZ Cas for log $\epsilon$(Li)=3.0 LTE (solid line), NLTE (dashed line), observed spectrum (dots).} \label{} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte5.eps}} \caption{As Figure 4 in the $\lambda6708$ {\AA} spectral domain of WZ Cas for log $\epsilon$(Li)=5.0} \label{} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte6.eps}} \caption{As Figure 4 in the $\lambda8126$ {\AA} spectral domain of WZ Cas for log $\epsilon$(Li)=4.5} \label{} \end{figure*} \subsection{NLTE procedure} \begin{table*} \caption[]{Data for JOLA opacities used in this work} {!}{\includegraphics{tabnlte2.eps}} \end{table*} To carry out the NLTE analysis for a 20-level Li atom model, we followed the procedure described in Pavlenko (1994) and Pavlenko \& Magazz\'u (1996). A few specific items were taken into account here: \begin{itemize} \item ionization-dissociation equilibria were computed for the carbon-rich case, i.e. considering C-contained molecules formation: CO, CH, CN, C$_2$, CS, HCN, H$_2$C, H$_3$C, H$_4$C, HC$_2$, H$_2$C$_2$ and HC$_3$. \item in the frequencies of bound-free transitions of Li I the opacity of diatomic molecules was computed in the framework of just-overlapping-line approximation (JOLA, see Table 2). \item in the frequencies of bound-bound linearized transitions of Li I the atomic absorption was included. Furthermore, in the frequencies of the four Li lines we considered the absorption of CN and C$_2$ for a given $^{12}$C/$^{13}$C ratio (see Table 1). \end{itemize} The work most relevant to the present study was done by de la Reza \& Querci (1978; thereafter RQ). Using a divergence flux method these authors carried out an extensive modelling of NLTE effects of lithium lines in C-stars. They used a 4 level lithium atom model, atmosphere models by Querci et al. (1974) and Johnson (1974) with effective temperatures of 3800, 3600 and 3500 K (warmer than the T$\rm{_{eff}}$ values usually derived in C-N stars), gravity log g$=1.0$, C/O$=1.3$ and modelled three Li I lines, namely the $\lambda6104$, $\lambda6708$ and $\lambda8126$ {\AA} lines. They also tried to model the impact of a chromosphere on the formation of lithium lines using the approach of {\it radiative temperature} (T$_{\rm{rad}}$) on the bound-free transitions of lithium. Several aspects in their NLTE calculation differ from the present study: \begin{enumerate} \item {\it The atom model}. We used a 20-level atom model, which allowed us to consider the interlocking of lithium lines (transitions) more appropriately. The ionization equilibrium of lithium is formed by the whole system of the bound-free transitions. Hovewer, as noted by RQ (see also Pavlenko 1991), transitions from/on the second level play the main role. \item{\it Bound-free transitions.} The radiation field in bound-free transition computations is very important in the modelling of lithium lines without LTE. Indeed, the effectiveness of the overionization of lithium depends directly on the mean intensities of the radiation field in the blue part of the spectrum (Pavlenko 1991). The approximation $\rm{T_{rad}=T_{e}}$ ({\it electronic temperature}) used by RQ does not produce any overionization of lithium. Only in the case of $\rm{T_{rad}>T_{e}}$ does the ionization of lithium increase with respect to the LTE case. Note that $\rm{T_{rad}>T_{e}}$ was considered by RQ only when a chromospheric radiation field was included. As noted by these authors, this approximation seems to be rather crude for the radiation field. \item{\it bound-bound transitions}. Lithium lines were treated by RQ as singlets. Basically, radiation transfer in the frequencies of several multiplet lines should differ from the case one single (strongest) line. Moreover, we include molecular line absorption in the continuous (background) opacity. In principle, this should reduce the probability of photon losses from the atmosphere, i.e. it directly affects the processes of radiative transfer. On the other hand, the radiative transfer in the continuum differs substantially from the case of continuum+lines in an atmosphere with chromosphere, as here the impact of the hot chromospheric layers on the photosphere might be weaker. \end{enumerate} \subsection{NLTE curves of growth, synthetic spectra and abundances} {\bf NLTE computations were performed by using a modified version of the code WITA2 (Pavlenko et al. 1995) described in Pavlenko (1999)}\footnote{\bf {Comparison between LTE abundances derived with this code and those obtained with the Uppsala's code showed an excellent agreement ($\sim\pm 0.01$ dex).}}. The dependence of the departure coefficients $b_i = \rm{N^i_{NLTE}/N^i_{\rm LTE}}$, where $\rm{N^i_{\rm NLTE}}$ and N$^i_{\rm LTE}$ are the NLTE and LTE populations of the $i$ level, upon the depth in a typical C-star atmosphere is shown in Fig. 7. The behaviour of the departure coefficients of lower and upper lithium levels differ substantially. The populations of the lower levels are reduced by the overionization processes. On the contrary, upper levels are overpopulated relative to the LTE case because they are more closely linked with the continuum (see also Pavlenko 1994, Carlsson et al. 1994, Pavlenko \& Magazzu 1996). Indeed, for log $\epsilon$(Li)$>$ 4.0 the departure coefficients of the first level become very low in the outermost layers ($b_1<10^{-4}$ at P$_g< 0.01$), where the NLTE core of the Li resonance line forms. On the other hand, the depth of formation of lithium resonance and subordinate lines is quite different. For large lithium abundances (log $\epsilon$(Li)$> 4.0$) the formation region for the resonance line shifts toward the outermost layers, where the temperature drops below 1000 K. In fact, we have to extrapolate the model atmosphere structure to log P$_{gas}\approx -5$ to take the line formation consistently into account. This produces additional problems in the computation of collisional and radiative rates. Here we are approaching the interstellar medium regime, where our approximations for collisional and radiative rates may not be valid. Moroever, in this regime the splitting of Li terms into sublevels may also become inportant although, since we are dealing with saturated lines, we believe that this effect is not critical. Lithium subordinate lines are formed in deeper layers. For instance, even for log $\epsilon$(Li)$> 4.5$ the core of the $\lambda6104$ {\AA} doublet forms in a region with temperature $\sim 2000$ K. As a consequence, the NLTE formation regime of this Li line changes dramatically: the interlocking processes of radiative transitions (see Magazz\'u et al. 1992; Carlsson et al. 1994) become more important than overionization. In this situation, line profiles depend on the behaviour of the source function which is affected by the whole set of radiative transitions. In brief, considering the formation of lines in the outermost layers we note: i) the opacity here drops sharply due to the low molecular and H$^{-}$ density, ii) the electron density is low and its effect on collision rates is small and iii) the electron temperature is also low, so that the effectiveness of the overionization $\propto\rm{e^{(-(T_e-T_{rad})/kT)}}$ would be high. Figures 1 and 2 (dashed lines) also show theoretical NLTE curves of growth for the four Li lines. Departures from LTE are very important for the resonance $\lambda6708$ {\AA} line. NLTE corrections for this line can amount to 0.6 dex! However, NLTE effects for the subordinate lines are weak ($\leq 0.2$ dex) and even decrease for strong Li absorptions (high Li abundances, log $\epsilon$(Li)$\geq 4$). We also see from these figures that NLTE effects decrease for decreasing T$_{\rm eff}$~and increasing C/O ratio in the model atmosphere. When T$_{\rm eff}$~drops the opacity of the carbon-contained molecules increases so that the difference between T$_{\rm{rad}}$ (of the radiation field in the bound-free frequencies of Li) and T$_e$ decreases. The same holds when the C/O ratio increases in the atmosphere. It is interesting to note that LTE/NLTE abundances converge for high Li abundances for the subordinate lines whatever the T$_{\rm eff}$~and/or the C/O ratio in the atmosphere. This gives us a chance to use subordinate lines as a tool for deriving Li abundances in Li-rich AGB stars even in the LTE approach, despite their saturated nature. Figures 3 to 6 show synthetic fits in NLTE (only for Li lines; i.e. dashed lines) in our stars. The difference between the LTE/NLTE fits to the resonance line is remarkable. For the subordinate lines, LTE and NLTE fits differ only slightly. Table 3 shows the final NLTE Li abundances derived in our stars. NLTE corrections are always positive, ranging from 0.1 to 0.5 dex. As was shown by Magazz\'u et al. (1992) for the case of G-M stars with solar abundances, the significance of NLTE effects depends upon the line strength. Li lines of moderate intensity (W$_\lambda >0.2$ {\AA}) form in a region where S$_\nu(\tau_{\rm{NLTE}}\approx 1) < \rm{B}_\nu(\tau_{\rm {LTE}}\approx 1)$. The NLTE cores of the resonance doublet become stronger than the LTE ones and, as a result, NLTE abundance corrections $\Delta$log $\epsilon$(Li)$=$log $\epsilon$(Li)$_{\rm {NLTE}}-$log $\epsilon$(Li)$_{\rm{LTE}}$ are negative. However, in the case of the strongest (saturated) Li resonance doublets the difference between LTE and NLTE cores cannot be large (because both approach to zero) then, NLTE corrections again become positive as in the case of weak lines. Note that there is a region in the atmosphere where $S_\nu<B_{\nu}$ holds for all except the $\lambda4603$ {\AA} line (Fig. 8) in the case of intermediate strong lines (log $\epsilon$(Li)$\sim 3$). In these regions the interlocking processes of Li radiative transitions play the main role. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte7.eps}} \caption{(a) Departure coefficients of five Li levels in the atmosphere of a T$_{\rm eff}$/log g/(C/O)$=3000/0.0/1.007$ C-star computed with a lithium abundance of log $\epsilon$(Li)=3.0. (b) The same for log $\epsilon$(Li)=4.5.} \label{} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{fnlte8.eps}} \caption{S$_{\nu}$/B$_{\nu}$ ratio for the four Li lines in the atmosphere 3000/0.0/1.007 of a C-star for different lithium abundances. Curves correspond to the $\lambda4603$ (solid), $\lambda6104$ (dotted), $\lambda6708$ (long dashed) and $\lambda8126$ {\AA} (short dashed) Li lines, respectively. For each pair of curves the upper line is computed with log $\epsilon$(Li)=3.0 and the lower one with log $\epsilon$(Li)=4.5.} \label{} \end{figure*} \section{Discussion and conclusions} We now compare our LTE/NLTE results with the literature. Abia et al. (1991) derived Li abundances in the same stars using the $\lambda6708$ Li line. The abundances obtained here are considerably lower. In WZ Cas and WX Cyg the differences can be ascribed to the different choice of T$_{\rm eff}$~and the C/O ratio in the model atmosphere but also to the more complete molecular line list used in this work, which makes the background absorption more intense in the $\lambda6708$ {\AA} region. Note that Abia et al. (1991) did not include the C$_2$ molecule (in any isotopic form) or the $^{13}$CN molecule in their line list. In IY Hya, moroever, the spectrum now analyzed in the $\lambda6708$ {\AA} range shows a much less intense Li line than that from which the Li abundance was derived by these authors: W$_{6708}\sim 2.5$ {\AA} here against 7.5 {\AA} in Abia et al. (1991). A similar variation in time might be present in the spectrum of WX Cyg analyzed here (4.5 vs. 6.5 {\AA}). Whether this variation is an effect of a varying T$_{\rm eff}$~along the pulsational phase or just evidence of the time scale for Li formation/depletion in AGB stars is a question which merits further studies. Note that Asplund et al. (1999) found the same figure in a post-AGB star and they concluded that they were observing the characteristic time scale of the mechanism responsible for the lithium production. Qualitatively our NLTE results agree with RQ but quantitatively some discrepancies are found. {\bf For instance, inspection to their Fig. 7 shows a NLTE correction by $\sim 0.7$ dex at log $\epsilon$(Li)$=4$ for the $\lambda 6708$ {\AA} line. However, our computations give a correction of $\sim 0.4$ dex (see Fig. 2)}. This can be easily explained by differences in the NLTE procedure, model atmospheres etc. {\bf Namely}, in this paper we fully computed the opacities in the bound-free and bound-bound lithium transitions. This allowed us to establish the dependence of NLTE in lithium lines on C/O and the $^{12}$C/$^{13}$C isotopic ratio. Moreover, the analysis of RQ was based on the comparison between the LTE and NLTE curves of growths. Our LTE/NLTE spectral synthesis approach to determine Li abundances in the atmosphere of Li-rich AGB stars seems to be better. Only in this way is the lithium abundance determination possible, since the conventional definition of the equivalent width becomes meaningless due to the severe blending of Li lines. In this sense our results for subordinate Li lines are of great interest. Since we underestimate the opacity of the bound-free and bound-bound transitions of lithium we get an upper limit to the NLTE effects in these lines, but comparatively, we obtain low NLTE corrections on Li abundances. Using a more complete opacity source list, NLTE effects should be even less pronounced. That would comprise a good chance to obtain accurate abundance determinations based on the subordinate Li lines in the frame of the LTE approach, which is simpler and therefore more useful. {\bf On the other hand, we can only qualitatively compare our NLTE results with those by Pavlenko \& Magazz\'u (1996) in M (O-rich) stars due to the obvious differences in opacities, model atmospheres etc, although in both cases, strong Li resonance lines are formed in the outermost layers of the atmosphere. Their curves of growth (see their Fig. 5) show a similar behaviour than ours (Figs. 1-2): for the $\lambda 6708$ {\AA} line W$_\lambda^{\rm{LTE}}>$ W$_\lambda^{\rm{NLTE}}$ for log $\epsilon$(Li)$>2$; i.e., in the case of strong lines formed high enough in the atmosphere NLTE effects reduce the equivalent width due to the dominant role of the overionization. For the $\lambda 6104$ {\AA} Li line they found W$_\lambda^{\rm{LTE}}<$W$_\lambda^{\rm{NLTE}}$ for log $\epsilon$(Li)$=4.0$ (see their Fig. 6). However, we obtain the contrary. This can be explained since in our case (C-rich atmospheres) the overionization should increase in the line-forming regions due to the drop of the density and pressure.} \begin{table*} \caption[]{LTE/NLTE lithium abundances in the stars studied} {!}{\includegraphics{tabnlte3.eps}} \end{table*} Abia et al. (1993a) computed an empirical Li yield from a homogeneous sample of galactic C-stars. They derived Li abundances from the $\lambda6708$ {\AA} line in $\sim 220$ C-stars and by considering the mass-loss rate estimates for the stars in their study, obtained a Li yield into the interstellar medium of $\sim 2\times 10^{-9}$ M$_\odot$pc$^{-2}$Gyr$^{-1}$. These authors showed that this yield is extremely dependent on the stellar mass-loss rate assumed and on the real Li abundance in the star. Since approximately 90\% of this value is determined by the actual yield from the SLiR C-stars, it is straightforward to revise this figure on the basis of the more accurate Li abundances obtained in this work. Now, considering the NLTE abundances obtained from the $\lambda8126$ {\AA} Li line as the best estimate of the lithium abundance, we obtain the revised yield of $\sim 8\times 10^{-10}$ M$_\odot$pc$^{-2}$Gyr$^{-1}$, a factor $\sim 3$ lower than the previous value. This is mainly because the revised Li abundances in WX Cyg and IY Hya are considerably lower. As in Abia et al. (1993a), we can roughly estimate the contribution of C-stars to the galactic Li. Assume that the above production rate has been constant during the lifetime of the galaxy ($\sim 13$ Gyr) and that the surface density of C-stars has been uniform and also constant during this time within a galactocentric radius of $\sim 15$ kpc ($\sim 50$ C-stars kpc$^{-2}$, see Claussen et al. 1987). In this case, the total contribution to galactic Li abundance by C-stars is M$_{\rm{Li}}\sim 3$ M$_\odot$, i.e. $\sim 10\%$ of the total Li in the galaxy. This contribution might be higher if evolutionary effects are taken into account, since the star formation rate was certainly higher in the past. Correspondingly, the number of C-stars that eventually became a SLiR C-star at a given time was also higher (see Abia et al 1993a for details). In summary: 1. The formation of the lithium resonance line in C-stars is severely affected by NLTE effects. NLTE Li abundances are higher and can differ from the LTE ones up to 0.6 dex. The core of this strong Li line is formed in the outermost layers where we approach the interstellar medium regime. Furthermore, these layers are affected by different phenomena such as the stellar chromosphere, inhomogeneities, dusty shells, outflow/infall velocities etc. Our computations also show that lithium is severely overionized. Therefore, the use of this saturated lithium line for abundance determinations in AGB stars seems quite impossible. In that sense, we confirm the results of RQ. 2. Subordinate lithium lines are formed in the inner parts of the atmosphere. We have shown that NLTE effects are rather weak even for strong lines (0.1-0.3 dex). Thus, they may be used for lithium abundance determinations in AGB stars even in the framework of the LTE approach. Of the three subordinate lines, the $\lambda8126$ {\AA} Li I line is probably the best one for Li abundance determinations because it is less blended, because NLTE effects are weak, because the molecular and atomic line list appears to be complete and because continuous opacity seems rather well reproduced in this spectral domain. 3. NLTE effects in the lithium lines show a complicated dependence on the input parameters T$_{\rm eff}$~and the C/O ratio. We found that NLTE effects increase with increasing T$_{\rm eff}$~and a decreasing C/O ratio. For a given T$_{\rm eff}$, NLTE effects decrease with increasing Li abundance for the subordinate lines but increase for the resonance line. Furthermore, for high Li abundances (log $\epsilon$(Li)$>4$) LTE/NLTE abundances from the subordinate lines are not very sensitive to T$_{\rm eff}$~or the C/O ratio. For the resonance line, however, the opposite happens. 4. The possibility of using the lithium subordinate doublet at $\lambda$4603 {\AA} was considered. Theoretical spectra, however, show a bad agreement with the observations, probably due to a wrong estimate of the continuous opacity in this spectral range and to an incomplete line list in this region. Furthermore, this lithium line forms a strong blend with iron lines. Due to the very weak NLTE corrections found for it, the use of this Li doublet would be promising in the case of metal deficient stars if using very high resolution spectra. Since we are confident of the atomic and molecular line list used beyond $\sim 6000$ {\AA}, we suggest that the C$_2$ lines of the Swan system observed in the blue part of the spectrum are the largest source of error. More accurate line lists for these transitions are therefore needed. Such an improvement, together with better atmosphere modelling, solving the problem of the missing continuous opacity in the blue part of the spectrum, should improve the agreement between observed and synthetic spectra in the future. 5. We have used observational data of high quality to determine lithium abundances in three C-stars: WX Cyg, WZ Cas and IY Hya. We have shown, that they are indeed SLiR stars with Li abundances in the range log $\epsilon$(Li)$\approx 3$ to 5. However, up to now we have still failed to accurately determine Li abundances in AGB stars, mainly due to uncertainties in modelling their atmosphere. This leads to an important uncertainty concerning the estimate of the Li yield by AGB stars. These stars might account for up to 30$\%$ of the currently observed Li or merely be a secondary source of Li in the galaxy. \begin{acknowledgements} Patrick de Laverny acknowledges support from the {\it Soci\'et\'e de Secours des Amis des Sciences}. Data from the VALD data base at Vienna, Austria, were used for the preparation of this paper. K. Eriksson and the stellar atmosphere group of the Uppsala Observatory are thanked for providing the grid of model atmospheres. The 4.2 m WHT is operated on the island of La Palma by the RGO in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof\'\i sica de Canarias. Based in part on observations collected at the German-Spanish Astronomical Center, Calar Alto, Spain. This work was partially supported by grant PB96-1428. \end{acknowledgements}
\section{Introduction} A well-studied problem in non-equilibrium statistical mechanics is the growth of domains in quenching processes \cite{Gunton}. Namely, when a system is suddenly quenched from a disordered initial state into a thermodynamic region where different phases coexist, macroscopic domains can be observed, usually characterized by a single time-dependent length scale, which grows as a power law $L(t) \sim t^{z}$. The spatial patterns of the domains at two different times are related by a global change of this length scale. A signature of this dynamical scaling is the fact that the structure factor $C(\vec{k},t)$ can be cast in the form \begin{equation} C(\vec{k},t) = L^{d}(t) C_{o} (\vec{k} L(t)) \label{prima} \end{equation} where $C_{o}$ is the scaling function. In the case of a binary mixture, the evolution of the system is described by a scalar order parameter $\phi$ representing the difference of concentrations of the two liquids in the mixture. When hydrodynamic effects are not considered, the quenching process can be described by an equation of the form $\partial_{t} \phi= \!- \vec{\nabla}\cdot \vec{j}$, where $\vec{j}=\! -\Gamma \vec{\nabla}(\delta H/\delta \phi)$, $H$ is a free-energy functional describing the ordered phases and $\Gamma$ is a transport coefficient called mobility, usually taken constant. It has been argued \cite{Langer} that the dynamics is more accurately mimicked considering a field-dependent mobility of the form \begin{equation} \Gamma(\phi) = (1 - a(T) \phi^{2}) \label{mobility} \end{equation} where $a(T) \rightarrow 1$ for temperature $T \rightarrow 0$, $a(T) \rightarrow 0$ for $T \rightarrow T_{c}$. In this way it is possible to take into account the different nature of the two mechanisms which operate during the phase separation: surface diffusion and bulk diffusion. The first has a growth exponent $z=1/4$ \cite{Furukawa} and is due to the diffusion of molecules of the two species along the interfaces. The second, which is also called Lifshitz-Slyozov mechanism, is due to the diffusion of molecules of one species from more curved interfaces, where they evaporate, to less curved ones through the bulk of the other phase. The corresponding growth exponent is $z=1/3$ \cite{Lifshitz}. At high temperatures (but still less than the critical temperature) the bulk diffusion is the only observed because it is faster than the other. Since bulk diffusion is a thermal activated process \cite{Corberi}, one expects that, lowering the temperature, a regime with $z=1/4$ can be observed. The proposed form (\ref{mobility}) for the mobility is able to catch these features. For shallow quenches ($a \ll 1$) the mobility remains constant in the whole system, while for very deep quenches ($a=1$) $\Gamma (\phi)$ vanishes in the bulk phases where $\phi^{2}=1$, suppressing the bulk diffusion. The diffusion along the interfaces is unaffected because $\Gamma (\phi) \simeq 1$ on domain boundaries. The effect of an order parameter-dependent mobility (\ref{mobility}) on systems with scalar order parameter has been studied by simulations \cite{Lacasta,Puri}. It has been found that for $a=1$ the length scale $L(t)$ grows as $t^{1/4}$ and for $0<a<1$ there is a crossover between $L(t) \sim t^{1/4}$ and $L(t) \sim t^{1/3}$. Recently, a non constant mobility has been also used to study the phase-ordering dynamics in systems with a vectorial order parameter \cite{Corberi,Bray}. In \cite{Corberi} the expression (\ref{mobility}) for the mobility has been adopted. The limit $N \rightarrow \infty$ , where $N$ is the number of vectorial components, is analytically solvable and corresponds to the self-consistent approximation of the scalar case. It is characterized by a growth exponent $1/6$ for the case $a=1$. For $0<a<1$ the usual value $1/4$ of the growth exponent for vectorial systems is recovered in the asymptotic regime. A more general form of the mobility has been introduced in \cite{Bray}, being given by \begin{equation} \Gamma(\phi) = (1 - \phi^{2})^{\alpha} \label{newmob} \end{equation} where $\alpha$ is a positive real number. In this paper we are concerned with the phase separation of a binary mixture subject to an uniform shear flow when the mobility is given by Eq.~ (\ref{newmob}). When a shear flow is applied to a quenched binary mixture, the pattern of the phase-separating domains as well as the time evolution are strongly modified by the flow \cite{Onuki}. Domains greatly elongated in the flow direction have been observed in simulations \cite{Ohta,Rothman} and in experiments \cite{Hashimoto}. A difference $\Delta z = 0.8 \div 1$ between growth exponents has been measured in some experiments, with the larger exponent in the flow direction \cite{Lauger,Chan}. The deformations of domains affects the rheological properties of the system, giving rise to excess stresses and to an increase $\Delta \eta$ of the viscosity \cite{Onuki2,Krall}. This behavior has its explanation in the fact that mechanical energy is expended to deform the domains against the interfacial tension. When the domains are stretched to such an extent that they start to burst, the stored mechanical energy is dissipated and the excess viscosity decreases. The phase separation in steady shear for binary mixtures with constant mobility has been recently studied in \cite{Noi}, where the existence of an anisotropic dynamical scaling theory with different growth exponents in flow and in other directions has been shown. It is found that the excess viscosity, after reaching a maximum, relaxes to zero, exhibiting log-time periodic oscillations. Also other physical observables are modulated by such oscillations, which can be related to a cyclical mechanism of storing and dissipation of elastic energy. Here we want to see the effects of a non constant mobility on this {\it scenario}. The outline of the paper is as follows. In section 2 we present the model and derive the equation of time evolution for the structure factor in a self-consistent approximation. In section 3 we report the asymptotic behavior of the model. We found two different growth exponents for the flow and the shear directions, given in the leading scaling regime by $z_{x}=(5+2 \alpha)/2(2+\alpha)$ and $z_{y}=1/2(2+\alpha)$, respectively. The asymptotic behavior of the rheological quantities is also calculated. Finally, in section 4 we integrate numerically the time evolution equation of the structure factor and calculate the whole evolution of the physical observables as moments of the structure factor. In the asymptotic regime they are modulated by dumped log-time oscillations. Our conclusions complete the article. \section{The model} We consider a binary mixture whose evolution is described by the diffusion-convection equation \begin{equation} {\label{eqM01}} \frac{\partial \phi(\vec{r},t)}{\partial t}+ \vec{\nabla} \cdot (\phi(\vec{r},t) \vec{v})= \vec{\nabla} \cdot \left [ \Gamma (\phi) \vec{\nabla} \left ( \frac{\partial H[\phi(\vec{r},t)]}{\partial \phi}\right ) \right] \end{equation} where the field $\phi(\vec{r},t)$ describes the concentration difference between the two components of the mixture, $\Gamma (\phi)$ is the mobility, which depends on the order parameter as in Eq.~(\ref{newmob}). The order parameter $\phi$ is convected by an external velocity field \cite{Onuki}. We choose a planar shear flow with \begin{equation} \vec{v}=\gamma y \vec{e}_x, \end{equation} where $\gamma$ is the shear rate, assumed constant, and $\vec{e}_x$ is a unit vector in the flow direction. The free energy functional is chosen to be of the standard $\phi^4$ form: \begin{equation} {\label{eqM02}} H[ \phi]= \!\!\int \!\!d \vec{r} \left [ -\frac{1}{2} \phi^2(\vec{r},t)+ \frac{1}{4} \phi^4(\vec{r},t)+\frac{1}{2} |\vec{\nabla} \phi(\vec{r},t) |^2 \right] \end{equation} where we assume that all parameters have been rescaled into dimensionless units \cite{Sagui} and the system is below the critical temperature. The two minima of the polynomial part of $H[ \phi]$ describe the pure states of the mixture. In this article we deal with the non-linear term of Eq.~ (\ref{eqM01}) in a self-consistent approximation; hence the term $\phi^3$ appearing in the functional derivative $\delta H/\delta \phi$ is linearized as $< \phi^2 > \phi$, where $<... >$ stands for the average over the system. In the same way the term $\phi^2$ in the mobility is substituted by $< \phi^2 >$. In the Fourier space Eq.~ (\ref{eqM01}) becomes: \begin{eqnarray} {\label{eqM03}} \frac{\partial \hat{\phi}(\vec{k},t)}{\partial t}&=& \gamma k_x \frac{\partial}{\partial k_y} \hat{\phi}(\vec{k},t) \nonumber \\ &-& (1-S(t))^\alpha k^2 \left[ S(t)-1+k^2 \right] \hat{\phi}(\vec{k},t) \end{eqnarray} where $S(t)=< \phi^{2}(\vec{r},t) >$ and $\hat{\phi}(\vec{k},t)$ is the Fourier transform of $\phi(\vec{r},t)$. The statistical quantity of experimental interest is the time-dependent structure factor $C(\vec{k},t)$ which is defined as $<\hat{\phi}(\vec{k},t) \hat{\phi}(-\vec{k},t)>$. It obeys to the evolution equation: \begin{eqnarray} {\label{eqM04}} \frac{\partial C(\vec{k},t)}{\partial t}&=& \gamma k_x \frac{\partial}{\partial k_y} C(\vec{k},t) \nonumber \\ &-& 2 (1-S(t))^\alpha k^2 \left [ S(t)-1+ k^2 \right ] C(\vec{k},t) \end{eqnarray} which is closed by the self-consistency condition \begin{equation} S(t)=\int_{|\vec{k}|<q} \frac {d \vec{k}}{(2 \pi)^{d}} \;\: C(\vec{k},t) \end{equation} where $q$ is a phenomenological cutoff. Rheological quantities of interest can be calculated as momentum integrals of the structure factor. Since we restrict the solution of the model to the two-dimensional case, we consider the excess viscosity $\Delta \eta$ and the first normal stress $\Delta N_{1}$ defined by \cite{Dawson} \begin{equation} \Delta \eta = - \gamma^{-1} \int_{|\vec{k}|<q} \frac{d \vec{k}}{(2 \pi)^{d}} \:k_{x}\: k_{y}\: C(\vec{k},t) \label{eqn5b} \end{equation} \begin{equation} \Delta N_{1} = \int_{|\vec{k}|<q} \frac{d \vec{k}}{(2 \pi)^{d}}\; \big [ k_{y}^{2}- k_{x}^{2} \big ] \;C(\vec{k},t) \label{stress} \end{equation} \section{The scaling behavior} Assuming simple scaling for the structure factor, we write, for arbitrary space dimensionality $d$, \begin{equation} \label{scaling} C(\vec{k},t) = \prod_{i=1}^{d} R_i(t) F (\vec{X}, \tau(\gamma t)) \end{equation} where the subscript $i$ labels the space directions with $i=1$ along the flow, $R_{i}$ is the average size of domains in the $i$-th direction, $\vec X$ is a vector of components $X_i=k_i R_i(t)$ and $F$ is a scaling function \footnote{Our self-consistent approximation is equivalent to that made in the large-$N$ limit for vectorial systems \cite{Ma}. In this case simple scaling is not verified for $\gamma=0$ when $\alpha=0$ \cite{Coniglio} and $\alpha \neq 0$ \cite{Bray} so that $C(\vec{k},t)$ has not the form (\ref{prima}) but can be written as $L^{d \xi(k/k_{m})}$, where $k_{m}$ is the position of the maximum in the structure factor and $\xi$ is a function which depends continuously on $k$ (multiscaling). In the case with shear the exact solution is still lacking, so we cannot, in principle, rule out multiscaling. However, since simple scaling is the leading approximation in the regions of the maxima of the structure factor, we can use it to obtain the correct value of the growth exponents (apart from logarithmic corrections) because the momentum integrals which define the observables are dominated by the maxima of $C(\vec{k},t)$.}. We also allow an explicit time dependence of the structure factor through $\tau(\gamma t)$; notice that since $C(\vec k,t)$ scales as the domains volume below the critical temperature, $\tau $ must not introduce any further algebraic time dependence in $C(\vec k,t)$. From the numerical results of the next section, we will see that $F$ is a dumped periodic function of $\tau$. Inserting the form (\ref{scaling}) of $C(\vec k,t)$ into Eq.~(\ref{eqM04}) we obtain : \begin{eqnarray} &&\gamma X_1F_2 = R_{1} R_{2}^{-1}\Bigg \{ \dot \tau \;\frac{\partial F}{\partial \tau} + \sum_{i=1}^{d} \bigg [ R_{i}^{-1} \dot R_{i} (F+X_{i} F_{i})\nonumber\\ &+&\! 2 \Big [ 1-S(t)\Big ]^{\alpha} \!\!R_{i}^{-2} X_{i}^{2} \bigg ( \sum_{k=1}^{d} R_{k}^{-2}X_{k}^{2}-\!1 \! +\!S(t) \bigg) F \bigg]\Bigg\} \label{ciccio} \end{eqnarray} where $F_i=\partial F/\partial X_i$ and a dot means a time derivative. Under the assumptions that $R_{1} \gg R_{i} \;\; (i=2,d)$ and $R_{i} \simeq \widetilde{R} \;\; (i=2,d)$ we can write \begin{eqnarray} \gamma X_1F_2 &=& \widetilde{R}^{-1} R_{1} \Bigg \{ \dot \tau \;\frac{\partial F}{\partial \tau} + R_{1}^{-1} \dot R_{1}(F+X_{1} F_{1}) \nonumber \\ &+&\widetilde{R}^{-1} \dot {\widetilde{R}} \sum_{i=2}^{d} \bigg [ (F+X_{i} F_{i})+ 2 \Big [ 1-S(t)\Big ]^{\alpha} \nonumber \\ &\times&\widetilde{R}^{-4} X_{i}^{2} \bigg ( \sum_{k=2}^{d} X_{k}^{2}-(1 -S(t))\widetilde{R}^{2} \bigg) F \bigg]\Bigg\} \label{ciccio2} \end{eqnarray} Since the l.h.s. of Eq.(\ref{ciccio2}) has no explicit algebraic time dependence, one has the asymptotic solutions \begin{eqnarray} R_1(t)&\sim& \gamma \: t^{(5+2\alpha)/2(2+\alpha)}\nonumber \\ \widetilde{R}(t)&\sim& t^{1/2(2+\alpha)}\nonumber \\ \big(1-S(t)\big)&\sim& t^{1/(2+\alpha)}\\ \tau (\gamma t)&\sim& \log \gamma t \nonumber \end{eqnarray} The growth exponents in the flow and in the shear directions are $z_{x}=(5+2\alpha)/2(2+\alpha)$ and $z_{y}=1/2(2+\alpha)$. We observe that $\displaystyle z_{x}-z_{y}=\frac{4+2 \alpha}{2(2+\alpha)}=1$. The value $z_{y}$ is the same found in \cite{Bray}, in the leading approximation, in a model having a field-dependent mobility with vectorial conserved order parameter without shear. Increasing the values of $\alpha$, one obtains values smaller with respect to the case with constant mobility, when $z_{x}=5/4$ and $z_{y}=1/4$. The shear affects only the growth exponent $z_{x}$ which remains greater than $1$ for every real and positive value of $\alpha$. The previous arguments can be used to establish the scaling properties of the rheological coefficients. Inserting the form~(\ref{scaling}) into Eq.~(\ref{eqn5b}) we obtain \begin{eqnarray} \Delta \eta (t)&\sim& (\gamma t) ^{-(3+\alpha)/(2+\alpha)} \gamma ^{-(1+\alpha)/(2+\alpha)}\nonumber \\ &&\times \int X_1 X_2 F \left [ \vec X,\tau (t)\right ]d\vec X \end{eqnarray} Therefore, in the scaling regime, for each value of $\gamma t$, the functions $\Delta \eta$ corresponding to different values of $\gamma$ collapse each on the others if rescaled as $\Delta \eta \!\rightarrow \!\gamma ^{(1+\alpha)/(2+\alpha)} \!\Delta \eta$. A similar analysis can be done for the normal stress. It is straightforward to show that in the asymptotic regime \begin{equation} \Delta N_{1} \sim t^{-1/(2+\alpha)} \int X_2^2 F \left [ \vec X,\tau (t)\right ] d\vec {X} \end{equation} Setting $\alpha=0$, we recover the previous results for the case with constant mobility\cite{Noi}. \section{Results and discussion} In this section we consider the numerical solution of Eq.~ (\ref{eqM03}). We will present the results for the calculation of the average size of domains and of the rheological indicators $\Delta \eta$ and $\Delta N_{1}$, stressing the effects of a non-costant mobility. \begin{figure*} \resizebox{0.80\textwidth}{!}{% \includegraphics{epjfig1.eps} } \caption{The structure factor at consecutive times for $\gamma=0.01$ and $\alpha=1$. The $k_{x}$ coordinate assumes positive values on the right of the picture, while the $k_{y}$ is positive towards the upper part of the plane. The maximum value of $|k_{y}|$ is $0.8$. In the other direction, $k_{x}$ varies in the range $[-0.8,0.8]$, $[-0.4,0.4]$ and $[-0.2,0.2]$, respectively. The positions of the peaks on the left foil of the structure factor is $(-0.22,0.37)$ for the highest and $(-0.08,-0.24)$ for the other at $\gamma t=4$. At $\gamma t=9$ the highest peak is located at $(-0.05,-0.23)$ and the other, at $(-0.10,0.36)$.} \label{fig:1} \end{figure*} We solved the equation of time evolution of the structure factor numerically in two dimensions, implementing a first-order Euler scheme with an adaptive mesh. The starting configuration for the structure factor is a constant value, which corresponds to a disordered state at very high temperature. Eq.~ (\ref{eqM03}) has been solved for different values of $\gamma$. In the following, results will be shown for the case $\gamma=0.01$. Similar results have been obtained in the other cases. The two values $\alpha=1$ and $\alpha=2$ for the parameter $\alpha$ appearing in (\ref{newmob}) have been considered. At the beginning the function $C(\vec{k},t)$ develops a circular volcano shaped structure. This is then deformed, as consequence of shear, into an elliptic structure. The sizes of the axes of the ellipse decrease in time at a different rate, this being larger in the $k_{x}$-direction. During this evolution two dips start to develop in the volcano edge until $C(\vec{k},t)$ is made of two foils. In Fig.~ 1 at $\gamma t=1$, the shape of $C(\vec{k},t)$ representative of this stage of the evolution is plotted. Later, on each foil, two well-formed peaks can be seen. At $\gamma t \simeq 4$ the peaks characterized by the larger values of $|k_{y}|$ prevail. Observe that the structure factor is symmetric with respect to the change $\vec{k} \rightarrow -\vec{k}$. The peaks with the smaller $|k_{y}|$ corresponding to a more isotropic configuration of domains, grow faster than the others until they prevail as it can be seen in Fig.~ 1 at $\gamma t=9$. These peaks continue to prevail along all the time evolution. What we observed is that the peaks continue always to grow and that the difference of their heights as a function of the shear strain $\gamma t$ is an increasing function being modulated by dumped oscillations. In order to have information about the growth of domains, we computed the typical domain size as \begin{equation} R_{x}(t)=\left(\frac{\int d \vec{k} C(\vec{k},t)} {\int d \vec{k} k_{x}^{2} C(\vec{k},t)} \right)^{1/2} \end{equation} and the same for the other direction. The values of $R_{x}$ and $R_{y}$ are plotted in Fig.~ 2 as function of the shear strain $\gamma t$ for $\alpha=1$ and $\alpha=2$. We plotted also the values for the case with constant mobility ($\alpha=0$). Some comments are in order here. The asymptotic behavior is the one expected through the previous scaling analysis: for $\alpha=1$ one has $R_{x} \sim t^{7/6}$ and $R_{y} \sim t^{1/6}$; for $\alpha=2$, $R_{x} \sim t^{9/8}$ and $R_{y} \sim t^{1/8}$. The growth exponents in both the directions are decreasing functions of the mobility exponent $\alpha$. This is reasonable since the mobility (\ref{newmob}) becomes smaller and smaller when $\alpha$ increases, if $(1-\phi^{2})<1$. Another consideration is about the superimposed log-time periodic oscillations. In the case with constant mobility these oscillations have an apparently constant amplitude. For non-zero values of $\alpha$ these oscillations are dumped. We will see that this feature is common to all the observables and will be discussed later. \begin{figure} \resizebox{0.49\textwidth}{!}{% \includegraphics{epjfig2.eps} } \caption{The typical size of domains as function of the strain $\gamma t$ in the flow and in the shear directions for different values of $\alpha$: $\alpha=0$ (full line), $\alpha=1$ (dashed line), $\alpha=2$ (dotted line).} \label{fig:2} \end{figure} We turn now to the study of the rheological behavior of the system. The external velocity field causes additional stresses on the mixture. Important indicators are the excess viscosity and the first normal stress. We calculated numerically $\Delta \eta$ and $\Delta N_{1}$ using their definitions (\ref{eqn5b}) and (\ref{stress}), respectively. The results are shown in Fig.~ 3 and Fig.~ 4. \begin{figure} \resizebox{0.49\textwidth}{!}{% \includegraphics{epjfig3.eps} } \caption{Plots of the excess viscosity vs the strain $\gamma t$ for different values of $\alpha$: $\alpha=0$ (full line), $\alpha=1$ (dashed line), $\alpha=2$ (dotted line). The inset shows the maxima of $\Delta \eta$ as function of $\gamma$ for the case $\alpha=1$. The slope of the straight line is 0.8.} \label{fig:3} \end{figure} \begin{figure} \resizebox{0.49\textwidth}{!}{% \includegraphics{epjfig4.eps} } \caption{The first normal stress as function of the strain $\gamma t$ for different values of $\alpha$: $\alpha=0$ (full line), $\alpha=1$ (dashed line), $\alpha=2$ (dotted line).} \label{fig:4} \end{figure} The excess viscosity reaches its maximum at the onset of the scaling when the domains are expected to be maximally stretched in the flow direction and the structure of $C(\vec{k},t)$ is the one shown at $\gamma t=4$ in Fig.~ 1. According to our analysis of the scaling behavior, we expect the excess viscosity to scale with $\gamma$ as $\Delta \eta \sim \gamma^{-(1+\alpha)/(2+\alpha)}$ for fixed value of $\gamma t$. In the inset of Fig.~ 3 we report for the case $\alpha=1$ the dependance of $\Delta \eta_{M}$ on $\gamma$. We find an exponent $0.8$ slightly larger than the expected $2/3$. The reason can be related to the fact that the excess viscosity reaches the maximum before the asymptotic regime is fully realized. Then $\Delta \eta$ decreases as a consequence of the dissipation of the elastic energy stored by domains which start to burst when they are stretched furtherly. Therefore more isotropic patterns form and the typical structure of the function $C(\vec{k},t)$ is the one at $\gamma t=9$ in Fig.~ 1. The excess viscosity decreases with a power law behavior which is consistent with the predicted exponent $-(3+\alpha)/(2+\alpha)$. The first normal stress reported in Fig.~ 4 decreases in time according to the exponent $-1/(2+\alpha)$ after reaching a maximum. The amplitudes of all quantites plotted in Figures 2, 3, 4 are modulated by dumped log-time oscillations \cite{Sornette}. We believe that the physical explanation for the dumping of oscillations may be found in the vanishing value of the mobility at equilibrium. In the case with constant mobility the origin of the oscillations is related to a cyclical mechanism of elongation and bursting of domains, which allows to store and dissipate elastic energy in the system \cite{Noi}. In the present case, during the time evolution, the decreasing values of $\Gamma (\phi)$ suppress diffusion in the bulk phase and inhibit the growth of small bubbles coming from bursting. Therefore, they cannot be stretched too much by the flow. In this way it is more difficult to store elastic energy in the system and the excess viscosity can increase only of a small amount. This mechanism of growth inihibition becomes stronger and stronger in the course of evolution causing the observed dumping of oscillations. In conclusion, we have studied the phase separation of a binary mixture with field-dependent mobility in shear flow. We proved that dynamical scaling holds for this system. There are different growth exponents in the flow and in the shear directions which depend on the mobility exponent $\alpha$. The difference in growth exponents is always $1$. All the physical observables have amplitudes decorated by dumped oscillations which are periodic in logarithmic time. We made a guess about the origin of this behavior. It would be an important endeavour to study this system in direct simulation of Eq.~ (\ref{eqM01}) to deeply understand this phenomenon. \begin{acknowledgement} We thank Federico Corberi for valuable discussions about the subject of this work. \end{acknowledgement}
\section{Introduction} \label{intro} It is well known that not all discontinuous solutions of hyperbolic conservation laws are admissible. Some of these can be excluded on physical grounds. For example, expansion shocks in gas dynamics must be discarded since they do not satisfy the second law of thermodynamics. Others can be excluded for purely mathematical reasons such as the fact that they do not satisfy uniqueness and existence conditions or are structurally unstable with respect to small perturbations of the initial data. These mathematical conditions are usually called evolutionary conditions. For example, intermediate shocks in magnetohydrodynamics (MHD) satisfy the second law but are not evolutionary. This subject was extensively studied between the late 1940's and early 1960's (e.g. Courant \& Friedrichs 1948, Lax 1957, Akhiezer et al. 1959, Germain 1960, Gel'fand 1963, Polovin 1961) and a full account can be found in numerous textbooks (e.g. Jeffrey \& Taniuti 1964, Cabannes 1970, Somov 1994). Until recently there was general agreement that admissible shocks must both satisfy the evolutionary condition and possess a steady dissipative shock structure, although the relation between these conditions was not entirely clear. There the matter rested until time-dependent numerical solutions of the dissipative MHD equations showed that certain types of intermediate shocks can arise from smooth initial data (Wu 1987). Shortly thereafter, Brio \& Wu (1988) found intermediate shocks in their numerical solution for a particular MHD Riemann problem. More recently, intermediate shocks have been also been found in two-dimensional simulations (De Sterck et al. 1998). Furthermore, Chao et al. (1993) have reported a detection of an interplanetary intermediate shock in the Voyager 1 data. All this has caused some authors to reject the classical theory and to suggest that the evolutionary condition is not relevant to dissipative MHD (Wu 1987, 1988a,b, 1990; Kennel, Blandford \& Wu 1990; Hada 1994, Myong \& Roe 1997a,b) and has led to a reexamination of the whole question of the existence, or otherwise, of non-classical shocks (see Glimm 1988, Freistuhler \& Liu 1993, Myong \& Roe 1997a and references therein). There are, however, others who argue that there is nothing wrong with the classical theory (e.g. Barmin, Kulikovsky \& Pogorelov 1996; Falle \& Komissarov 1997). The matter clearly needs to be resolved, particularly since the existence, or otherwise of intermediate shocks is of crucial importance not only for fundamental MHD processes such as reconnection (Wu 1995), but is also relevant to many other astrophysical applications. The purpose of this paper is to try and clear the matter up by showing that there is neither a real conflict between the classical shock theory and the results of numerical calculations nor any incompatibility between ideal and dissipative MHD. In order to make the discussion complete, we have put together and extended a number of results from the literature that have tended be ignored or misunderstood. This paper is organised as follows. In \S\ref{sgeneral} we briefly review the classical shock theory and the evolutionary conditions. In \S\ref{dis-str} we study the relationship between these conditions and the uniqueness and existence of steady dissipative shock structures for systems with a concave entropy function. In \S\ref{apple} we apply these results to the full system of MHD equations and to the reduced system of planar MHD. In \S\ref{numerics} we present the results of numerical calculations which show that, for both these systems, the behaviour of the shocks is entirely consistent with the predictions of the classical shock theory. In \S\ref{dissc} we consider various aspects of the problem of intermediate shocks and discuss ways in which to avoid their appearance in MHD simulations with planar symmetry. In particular, we present the results of one dimensional simulations using a modified Glimm scheme (Glimm 1965) in which these shocks do not appear. \section{General Theory of Shocks} \label{sgeneral} In this section we give a brief review of the classical theory of discontinuous solutions of hyperbolic conservation laws. For our purposes it is sufficient to consider only the dimensional equations of the form \begin{equation} {{\partial {\bf u}} \over {\partial t}} + {{\partial {\bf f}} \over {\partial x}} = 0, \label{a1} \end{equation} where ${\bf u} \in {\cal R}^n$ is a vector of conserved variables and ${\bf f}({\bf u}) \in {\cal R}^n$ is a vector of the corresponding fluxes. As is well known, the system (\ref{a1}) is called hyperbolic if the Jacobian matrix \[ {\bf A} = {{\partial {\bf f}} \over {\partial {\bf u}}}. \] has $n$ real eigenvalues, $\lambda_k$ ($k = 1 \ldots n$) corresponding to $n$ linearly independent right eigenvectors, ${\bf r}_k$ and is called strictly hyperbolic if all the $\lambda_k$ are different. The physical significance of the $\lambda_k$ is that they are the speeds of small amplitude waves. Waves are classified as linear or nonlinear according to the behaviour of \[ C_k({\bf u}) \equiv {\bf r}_k({\bf u}) \cdot \nabla_u \lambda_k({\bf u}). \] If $C_k({\bf u})=0$ for all ${\bf u}$, then the k-wave is called linear, whereas if the dimension of the surface defined by $C_k({\bf u})=0$ is less then $n$, then it is called nonlinear or genuinely nonlinear. The states, ${\bf u}_l$, ${\bf u}_r$ on either side of a discontinuity travelling with speed $s$ must satisfy the shock equations \begin{equation} s({\bf u}_l - {\bf u}_r) = {\bf f}_l - {\bf f}_r , \label{a4} \end{equation} The number $n_s$ of independent shock equations can be less then $n$. For example, a contact discontinuity in gas dynamics has $n_s=3$ whereas $n=5$. Since ${\bf A}$ is the Jacobian, we clearly have $s \rightarrow \lambda_k$ for some $k$ as ${\bf u}_l \rightarrow {\bf u}_r$, which means that one can associate each discontinuity that allows this limit with one of the waves of the system. A discontinuity is called linear if the corresponding characteristic speed does not change across it, otherwise it is called nonlinear. There mere fact that a discontinuity satisfies (\ref{a4}) it does not necessarily imply that it is either stable or that it can arise from continuous initial data. For some hyperbolic systems equations (\ref{a4}) allow nonlinear shocks that propagate with a characteristic speed associated with a nonlinear wave, which means that they can be attached to such a wave to form compound waves. Systems with such shock solutions are called non-convex. Compound waves may arise from continuous initial data if the system allows single simple waves in which $C_k({\bf u})$ changes sign along the phase curve of a simple wave. This condition is therefore often used as an alternative definition of non-convexity. Although these definitions are equivalent for a single conservation law, they are not necessarily so for systems. The evolutionary condition is directly related to the question of existence and uniqueness of discontinuous solutions. It is well known that for hyperbolic equations there is a general way of deciding this question, which is to use the compatibility conditions that must be satisfied along the characteristics (Friedrichs 1955). If a characteristic with wave speed $\lambda_k$ enters one side of a discontinuity, then the state on that side must satisfy the compatibility relation associated with that characteristic, \[ {\bf l}_{k}({\bf u}) \cdot {\bf du}=0, \] where ${\bf l}_{k}(\bf u)$ is the left eigenvector of ${\bf A}$ corresponding to that characteristic. These equations are independent provided the ${\bf l}_k$ are linearly independent i.e. for all hyperbolic systems. If the wave speeds on either side of the discontinuity are such that $m_i$ compatibility relations have to be satisfied, then there are $n_s+m_{i}$ equations relating the $2n+1$ unknowns associated with the discontinuity, ${\bf u}_l$, ${\bf u}_r$ and the shock speed, $s$. A discontinuous solution can therefore only exist and be unique if \begin{equation} m_{i} = 2n-n_s+1. \label{a5} \end{equation} Obviously, when $n_s=n$, (\ref{a5}) reduces to \begin{equation} m_{i} = n+1. \label{a5a} \end{equation} It is clear from this that if a characteristic is parallel to the shock curve, then it is counted as incoming since the corresponding compatibility relation must be satisfied (Gelfand 1963). If $m_{i}>2n-n_s+1$ then the system is overdetermined and there is no solution except for certain special initial conditions. There will therefore always be arbitrarily small perturbations of this data that will destroy such a discontinuity by splitting it into a number of waves, just as an arbitrary initial dicontinuity splits in a Riemann problem. If $m_{i}<2n-n_s+1$, then the solution exists, but is not unique and one might hope that this nonuniqueness can be removed by including dissipative terms. In the following we will call condition (\ref{a5}) the strong evolutionary condition and call the condition \[ m_{i} \le 2n-n_s+1 , \] which allows nonunique solutions a relaxed evolutionary condition. An equivalent way of obtaining (\ref{a5}) is by a linear structural stability analysis of shock solutions (e.g. Landau \& Lifshitz 1959, Jeffrey \& Taniuti 1964). A discontinuity that is exposed to a small amplitude incident wave will only survive if it can respond by changing its speed and emitting small amplitude waves. Each such wave is described by one parameter and we also have the perturbation in the shock speed, which means that there are $m_{o}+1$ unknowns in this problem, where $m_{o}$ is the number of outgoing characteristics. Since these are related to the amplitude of the incoming wave by the $n_s$ shock relations, the discontinuity can only have a unique response if \begin{equation} m_{o} = n_s-1. \label{a6} \end{equation} It is worth pointing out that, contrary to what is claimed in Myong \& Roe, 1997a, this analysis does not assume that the discontinuity is weak. This suggests that non-unique discontinuous solutions should spontaneously self-destruct by emitting waves even if they are not perturbed (Anderson 1963). Although the conditions (\ref{a5}) and (\ref{a6}) appear to be different, the fact that $m_{o}+m_{i}=2n$ means that they are entirely equivalent (Gel'fand 1963). Note that, if the system of shock and compatibility equations splits into independent subsets, then the discontinuity is only evolutionary if each of these subsets has the same number of equations as variables (Jeffrey \& Tanuiti 1964). Finally, as far as the evolutionary conditions are concerned it does not matter whether, or not, the system (\ref{a1}) is strictly hyperbolic and convex since these properties are not used in the derivation of (\ref{a5},\ref{a6}). However, it is only in the case of strictly hyperbolic systems that these conditions reduce to the Lax conditions (Lax 1957) \[ \begin{array}{rcl} \lambda_{k-1}({\bf u}_l) < & s & < \lambda_{k}({\bf u}_l) \\ \lambda_{k}({\bf u}_r) < & s & < \lambda_{k+1}({\bf u}_r) \end{array} , \] for a nonlinear discontinuity associated with the $k$th characteristic (here we have assumed that $\lambda_1 < \lambda_2 < ... < \lambda_n$). \section{Evolutionary conditions and dissipative shock structure} \label{dis-str} In order to assess recent claims that nonevolutionary shocks become admissible if dissipative terms are included, we need to look at the general relationship between the evolutionary conditions and the uniqueness and existence of steady dissipative shock structures. Godunov (1961) has shown that it is much easier to explore this question if the equations can be transformed to a symmetric form. Although this is not possible for arbitrary hyperbolic systems of conservation laws, it can certainly be done for gasdynamics, MHD, and the shallow water equations and probably for any system that can arise in nature. \subsection{Symmetric Form of the Ideal Equations} \label{sb-symeq} We start by summarizing some of the results described by Friedrichs (1954), Friedrichs \& Lax (1971) and Boillat (1974, 1982). As before, it is only necessary to consider the one dimensional case. Consider a dissipation-free system of conservation laws described by the equations (\ref{a1}). Suppose now that there exists a quantity, $h({\bf u})$, which is also conserved as long as the solution to this system is continuous. For example, $h({\bf u})$ is the entropy in gasdynamics or MHD, whereas it is the total energy for the shallow water equations. If such a quantity exists, then there must exist a flux function, $g({\bf u})$, such that \begin{equation} \frac{\partial h}{\partial t} + \frac{\partial g}{\partial x} = 0, \label{9} \end{equation} (\ref{a1}) and (\ref{9}) can only be consistent if \begin{equation} \frac{\partial h}{\partial u_i} \frac{\partial f_i}{\partial u_j} = \frac{\partial g}{\partial u_j}, \label{10} \end{equation} (summation convention assumed), since then \[ \frac{\partial h}{\partial t} + \frac{\partial g}{\partial x} = \frac{\partial h}{\partial u_i} \left( \frac{\partial u_i}{\partial t} + \frac{\partial f_i}{\partial x} \right) = 0 . \] for any $C^1$ solution satisfying (\ref{a1}). If we now use $h$ to define the Legendre transformation \begin{eqnarray} \label{12} u^{\prime}_i & = & -\frac{\partial h}{\partial u_i}, \\ \label{13} u_i & = & \frac{\partial h^\prime}{\partial u^\prime_i}, \\ \label{14} h^\prime &=& h + {u^\prime}_i {u_i}, \end{eqnarray} then (\ref{10}) allows us to write the fluxes as \[ f_i = {{\partial g^\prime} \over {\partial u^\prime_i}} \] where \[ g^\prime = g + {u^\prime}_i f_i. \] In terms of the variables ${\bf u^\prime}$, (\ref{a1}) becomes a symmetric system \begin{equation} {\bf P} \frac{\partial {\bf u}^\prime}{\partial t} + {\bf Q}\frac{\partial {\bf u}^\prime}{\partial x} = 0 \label{17} \end{equation} where the symmetric matrices ${\bf P}$ and ${\bf Q}$ are given by \begin{equation} \begin{array}{ccccccc} P_{ij} & = & \displaystyle{\frac{\partial u_i}{\partial u^{\prime}_j}} & = & \displaystyle{\frac{\partial^2 h^\prime} {\partial {u^\prime}_i \partial u^{\prime}_j}} & = & \displaystyle{- \frac{\partial^2 h} {\partial u_i \partial u_j}}, \\ & & & & & & \\ Q_{ij} & = & \displaystyle{\frac{\partial f_i}{\partial {u^\prime}_j}} & = & \displaystyle{\frac{\partial^2 g^\prime} {\partial {u^\prime}_i \partial {u^\prime}_j}}.\\ \end{array} \label{18} \end{equation} Note that $h$ is usually a strictly concave function, in which case (\ref{18}) ensures that ${\bf P}$ is positive definite and the transformation is non-singular. In ordinary gasdynamics or MHD, $h$ is the entropy per unit volume and is therefore guaranteed to be concave by the second law of thermodynamics. For the shallow water equations $h = -e$, where $e$ is the sum of the kinetic and potential energy and dissipation ensures that this is also concave. \subsection{Dissipative Equations} \label{sb-diss} If we now assume that the dissipative fluxes are proportional to the spatial gradients of the dependent variables, then the dissipative version of (\ref{17}) is \begin{equation} \frac{\partial {\bf u}}{\partial t} + \frac{\partial {\bf f}}{\partial x} = {\bf P} \frac{\partial {\bf u}^\prime}{\partial t} + {\bf Q}\frac{\partial {\bf u}^\prime}{\partial x} = \frac{\partial} {\partial x} {\bf D} \frac{\partial {\bf u^\prime}}{\partial x} \label{20} \end{equation} where ${\bf D}$ is a matrix of dissipation coefficients. Multiplying this on the left by ${\bf u}^{\prime t}$ (the superfix t denotes the transpose) and using (\ref{9}--\ref{12}) gives the evolution equation for $h$ \[ \frac{\partial h}{\partial t} + \frac{\partial g}{\partial x} = -{\bf u}^{\prime t} \frac{\partial} {\partial x} {\bf D} \frac{\partial {\bf u}^\prime}{\partial x} , \] Integrating this over an arbitrary fixed interval $[a,b]$ and integrating the dissipative term by parts gives \[ \frac{d}{d t} \int\limits_a^b h d x + \left[ g +{\bf u}^{\prime t} {\bf D} \frac{\partial {\bf u}^\prime}{\partial x} \right]^b_a = \int\limits_a^b \frac{\partial {\bf u}^{\prime t}}{\partial x} {\bf D} \frac{\partial {\bf u}^\prime}{\partial x} d x . \] Since the term on the RHS of this equation represents a source term for $h$ and the second law of thermodynamic requires that this be positive if $h$ is the entropy per unit volume, the matrix ${\bf D}$ must be positive definite for gasdynamics and MHD. The dissipative shallow water equations must also satisfy this condition if we set $h = -e$, where $e$ is the total energy. One can also show that all linear waves decay if ${\bf D}$ is positive definite and $h$ is a strictly concave. The linear version of (\ref{20}) is simply \[ {\bf P} \frac{\partial {\bf u}^\prime}{\partial t} + {\bf Q} \frac{\partial {\bf u}^\prime}{\partial x} = {\bf D} {{\partial^2 {\bf u}^\prime} \over {\partial x^2}} \] where ${\bf P}$, ${\bf Q}$, and ${\bf D}$ are now constant matrices. Multiplying this by ${\bf u}^{\prime t}$ and integrating over $[a,b]$ gives \[ {d \over {d t}} \int_a^b {{\bf u}^{\prime t} {\bf P} {\bf u}^\prime d x} + \left[{\bf u}^{\prime t} {\bf Q} {\bf u}^\prime - 2 {\bf u}^{\prime t} {\bf D} \frac{\partial {\bf u}^\prime}{\partial x} \right]^b_a = - 2 \int_a^b {{{\partial {\bf u}^{\prime t}} \over {\partial x}} {\bf D} {{\partial {\bf u}^\prime} \over {\partial x}} d x}, \] after integrating the dissipative term by parts. Since ${\bf P}$ is positive definite if $h$ is strictly concave, the term on the RHS ensures that all linear waves decay if ${\bf D}$ is positive definite, \subsection{Steady Shock Structures} \label{sb-shst} Now consider a solution of the steady version of (\ref{20}) \begin{equation} {d \over {d x}} {\bf f} = {d \over {d x}} {\bf D} {d \over {d x}} {\bf u^\prime} \label{27} \end{equation} with the boundary conditions \begin{equation} \begin{array}{l} {\bf u}^\prime \rightarrow \left\{ { \begin{array}{l} {\bf u}^\prime_l~~x \rightarrow -\infty \\ {\bf u}^\prime_r~~x \rightarrow +\infty. \end{array} } \right. \end{array} \label{28} \end{equation} If this represents a shock structure, then ${\bf u}^\prime_l$ and ${\bf u}^\prime_r$ must satisfy the shock relations in the shock frame \begin{equation} {\bf f}({\bf u}^\prime_l) = {\bf f}({\bf u}^\prime_r). \label{29} \end{equation} Integrating (\ref{27}) and applying the boundary conditions (\ref{28}) gives \begin{equation} {\bf D} {{d {\bf u}^\prime} \over {d x} } = {\bf f}({\bf u}^\prime) - {\bf f}({\bf u}^\prime_l) = {\bf f}({\bf u}^\prime) - {\bf f}({\bf u}^\prime_r) \label{30} \end{equation} A steady shock structure therefore corresponds to a solution of (\ref{30}) that connects the equilibrium points ${\bf u}^\prime_l$ and ${\bf u}^\prime_r$. We now show that there is no guarantee that this solution is unique and structurally stable unless the corresponding discontinuous solution of the ideal system are satisfies the evolutionary conditions (\ref{a5}). Let $L_u$ be the unstable manifold of the point ${\bf u}^\prime_l$ and $R_s$ the stable manifold of the point ${\bf u}^\prime_r$. Then the trajectories in $L_u$ and $R_s$ are described by $dim(L_u) - 1$ and $dim(R_s) - 1$ parameters respectively. Since any trajectory which lies in both has to satisfy $n - 1$ matching conditions, this means that, in general, there will only be a unique trajectory connecting ${\bf u}^\prime_l$ and ${\bf u}_r$ if $dim(L_u) + dim(R_s) = n + 1$. If $dim(L_u) + dim(R_s) > n + 1$, then the trajectory may not be unique, whereas if $dim(L_u) + dim(R_s) < n + 1$, then any trajectory that does exist can be destroyed by perturbations of ${\bf u}^\prime_l$, ${\bf u}^\prime_r$ i.e. it is not structurally stable. The following theorem relates $dim(L_u)$ and $dim(R_s)$ to the number of characteristics entering the shock: \begin{theorem} If ${\bf u}^\prime_e$ is an equilibrium point of the dissipative shock equations (\ref{30}) at which none of the characteristic speeds vanish, then the equilibrium point is hyperbolic and the dimension of its stable (unstable) manifold is given by the number of positive (negative) characteristic speeds in the state ${\bf u}^\prime_e$. \end{theorem} \begin{proof} Suppose that ${\bf u}^\prime_e = {\bf u}^\prime_l$ (the proof for ${\bf u}^\prime_r$ is identical). Then linearizing (\ref{30}) in the neighbourhood of ${\bf u}^\prime_l$ gives \[ {\bf D}_l {{d {\bf v}} \over {d x}} = {\bf Q}_l{\bf v}, \] where ${\bf v} = {\bf u}^\prime - {\bf u}^\prime_l$, ${\bf Q}_l = {\bf Q}({\bf u}^\prime_l)$ and ${\bf D}_l = {\bf D}({\bf u}^\prime_l)$. If this equilibrium point is hyperbolic, then the dimension of its stable (unstable) manifold are given by the numbers of eigenvalues, $\mu_k$, satisfying \begin{equation} |{\bf Q}_l -\mu {\bf D}_l| =0 . \label{32} \end{equation} with positive (negative) real parts. On the other hand, the characteristic speeds for the system (\ref{17}), $\lambda_k$, in the state ${\bf u}^\prime_l$ are given by \begin{equation} \left| {\bf Q}_l - \lambda {\bf P}_l \right| = 0 . \label{33} \end{equation} A standard result (e.g. Gantmacher 1959) tells us that, since ${\bf P}_l$, ${\bf Q}_l$ are symmetric and ${\bf P}_l$ is positive definite, ${\bf Q}_l$ has the same number of positive, negative and zero eigenvalues as the set $\lambda_k$. If, like Godunov (1961), we assume that ${\bf D}_l$ is symmetric as well as positive definite, then the theorem would follow immediately from (\ref{32}) and (\ref{33}). However, the following lemma shows that this is an unnecessary restriction. \begin{lemma} Let ${\bf Q}$ be a non-singular symmetric matrix, ${\bf D}$ a positive definite matrix and $\mu_k$ the solutions of \[ |{\bf Q} -\mu {\bf D}| =0. \] Then the number of $\mu_k$ with positive (negative) real part is the same as the number of positive (negative) eigenvalues of ${\bf Q}$. \label{lem1} \end{lemma} \begin{proof} Define \[ {\bf D}_\epsilon = {\bf D}_s + \epsilon {\bf D}_a, \] where $\epsilon \in [0,1]$ and \[ {\bf D}_s = \frac{1}{2}({\bf D}+{\bf D}^t) ,\quad {\bf D}_a = \frac{1}{2}({\bf D}-{\bf D}^t) . \] It easy to see that ${\bf D}_\epsilon$ is also positive definite. Now consider the eigenvalue problem \[ |{\bf Q} -\mu(\epsilon) {\bf D}_\epsilon| =0 . \] The conclusion of the lemma is certainly true for $\epsilon = 0$, since then ${\bf D}_\epsilon$ is symmetric. If we can show that the $\mu_k(\epsilon)$ are continuous functions of $\epsilon$ and that $\Re\{\mu_k(\epsilon)\} \not = 0 ~\forall k$ for $\epsilon \in [0,1]$, then it will also be true for $\epsilon = 1$. The $\mu_k(\epsilon)$ are the roots of a polynomial of degree $n$ whose coefficients are polynomials in $\epsilon$. A root can therefore only change discontinuously by going to infinity, which can only occur if the coefficient, $|D_\epsilon|$, of the highest power of $\mu$ vanishes. However, this cannot happen since $D_\epsilon$ is positive definite for $\epsilon \in [0,1]$. The $\mu_k(\epsilon)$ must therefore be continuous functions of $\epsilon$ for $\epsilon \in [0,1]$. In order to prove that the $\mu_k$ cannot cross the imaginary axis, suppose that for some $k$, $\mu_k(\epsilon) = i \eta$, where $\eta$ is real. If ${\bf a}+i{\bf b}$ is the corresponding eigenvector, we have \[ \begin{array}{rcl} {\bf Qa}+\eta{\bf D}_\epsilon {\bf b} &=& 0 , \\ {\bf Qb}-\eta{\bf D}_\epsilon {\bf a} &=& 0 . \end{array} \] Multiplying the first of these by ${\bf b}^t$. the second by ${\bf a}^t$ and substracting gives \[ \eta ( {\bf b}^t{\bf D}_\epsilon{\bf b} + {\bf a}^t{\bf D}_\epsilon{\bf a} ) = 0 . \] Since ${\bf D}_\epsilon$ is positive definite this requires $\eta =0$ and hence $\mu_k = 0$, which cannot be true if the eigenvalues of $\bf Q$ are non-zero. This completes the proof of the lemma. \end{proof} (\ref{32}), (\ref{33}) and lemma (\ref{lem1}) show that the theorem is true even if ${\bf D}$ is not symmetric. \end{proof} This is a somewhat more direct proof of a result which has also been obtained by Kulikovsky \& Lyubimov (1965). In their analysis of viscous shock structures, Myong \& Roe (1997a) assumed that Theorem 3.1 holds for MHD, but did not give a proof. This analysis tells us that if the shock relations (\ref{29}) have a solution such that none of the characteristic speeds given by (\ref{33}) vanish in both the left and the right state and $m_i$ is the number of characteristics entering the shock, then \begin{enumerate} \item for $m_i = n+1$ the shock can have a unique structurally stable dissipative structure; \item for $m_i > n+1$ the dissipative structure is not guaranteed to be unique. \item for $m_i < n+1$ there might be a unique dissipative structure but it cannot be structurally stable. \end{enumerate} These conditions are not only compatible with the evolutionary conditions, they are complementary to them. Shocks for which $m_i > n + 1$ have a dissipative shock structure and could therefore be regarded as admissible on these grounds. However, the left and right states of such shocks must be carefully tuned since they cannot adjust themselves to an arbitrary small perturbation of their left and right states. Shocks that satisfy the relaxed evolutionary condition, $m_i < n + 1$, are apparently permitted by the ideal equations, but cannot establish a dissipative structure and must spontaneously self-destruct. It is therefore clear that the only physically admissible shocks are those those that satisfy the strong evolutionary conditions (\ref{a5}) or (\ref{a5a}). Theorem 3.1 gives us no information in those cases for which the shock speed coincides with at least one of the characteristic speeds. The corresponding critical point is then no longer hyperbolic and its type depends on the details of the particular system. \section{Application to Magnetohydrodynamics} \label{apple} As we shall see, the mathematical properties of the full system of MHD and the reduced planar system of MHD are somewhat different and this has to be clearly understood when the evolutionary conditions are applied. We therefore discuss these systems separately. \subsection{Full System of MHD} \label{sb-full} It is well known that the one dimensional equations of MHD can be written in the form (\ref{a1}) (e.g. Brio \& Wu 1988). The conserved quantities ${\bf u}$ and the corresponding fluxes ${\bf f}$ are \[ \begin{array}{ll} {\bf u} = \left[{ \begin{array}{c} \rho \\ \rho v_x \\ \rho v_y \\ \rho v_z \\ e \\ B_y \\ B_z \end{array} }\right] & {\bf f} = \left[{ \begin{array}{c} \rho v_x \\ \rho v_x^2 + p_g + B^2/2 - B_x^2 \\ \rho v_x v_y - B_x B_y \\ \rho v_x v_z - B_x B_z \\ \{e + p_g + B^2/2 \}v_x - B_x({\bf v}.{\bf B}) \\ v_x B_y - v_y B_x \\ v_x B_z - v_z B_x \end{array} }\right] \end{array}. \] Here $p_g$ is the gas pressure, \[ e = i + {1 \over 2} B^2 + {1 \over 2} \rho v^2 \] is the total energy per unit volume and $i$ is the enthalpy per unit volume. Here we use units such that the velocity of light and the factor $4\pi$ do not appear. As we have already discussed, ideal MHD has a supplementary conservation law representing the conservation of thermodynamic entropy. The second law of thermodynamics guarantees that the function $h = \rho S$, where $S$ is the entropy per unit mass, is strictly concave (e.g. TerHaar \& Wergeland 1966) and hence that the matrix ${\bf P}$ defined by (\ref{18}) is positive definite. The system of MHD equations can therefore be written in the symmetric form (\ref{17}) and is hyperbolic. Although this has been demonstrated for relativistic MHD by Ruggeri \& Strumia (1981), we have been unable to find an account of the corresponding analysis for classical MHD in the literature. However, since the derivations are similar to those for the relativistic case, we shall simply give the symmetric variables. They are \[ \begin{array}{rclrclrclrcl} u_1^\prime & = & \displaystyle{\frac{1}{T}\left({w\over \rho} - {1 \over 2}v^2 \right)}, & u_2^\prime & = & \displaystyle{\frac{v_x}{T}}, & u_3^\prime & = & \displaystyle{\frac{v_y}{T}}, & u_4^\prime & = & \displaystyle{\frac{v_z}{T}}, \\ & & & & & & & & & & & \\ u_5^\prime & = & \displaystyle{-\frac{1}{T}}, & u_6^\prime & = & \displaystyle{\frac{B_y}{T}}, & u_7^\prime & = & \displaystyle{\frac{B_z}{T}}. \end{array} \] There is no need to verify that the matrix, ${\bf D}$, of dissipation coefficients is positive definite, since this must be true for any system that obeys the second law of thermodynamics. Indeed, this condition is used to derive the dissipative equations in the first place (e.g. Landau \& Lifshitz 1960). The exact form of symmetrized equations is also of no importance for our purposes. Their existence, does, however, allow us to apply the conclusions of the general theory described in Sections 2 and 3 to dissipative MHD. \subsubsection{Characteristic Wave Speeds} Since there are seven variables in this system, there are seven waves whose speeds are \begin{tabular}{lrcl} & & & \\ Fast Waves & $\lambda_{f\mp}$ & $=$ & $v_x \mp c_f$, \\ Alfv\'en Waves & $\lambda_{a\mp}$ & $=$ & $v_x \mp c_a$, \\ Slow Waves & $\lambda_{s\mp}$ & $=$ & $v_x \mp c_s$, \\ Entropy Wave & $\lambda_e$ & $=$ & $v_x$, \\ & & & \end{tabular} \noindent where the alfv\'en speed, $c_a$, and the slow and fast speeds, $c_s$, $c_f$ are given by \[ c_a = |B_x| / \surd \rho, \] \[ c_{s,f}^2 = {1 \over 2} \left[ {a^2 + {B^2 \over \rho} \mp {\left\{ {{\left({a^2 + {B^2 \over \rho}}\right)}^2 - {{4a^2B_x^2} \over \rho}} \right\} }^{1/2}}\right], \] where $a$ is the adiabatic sound speed. Note that $0 \leq c_s \leq c_a \leq c_f$. If $B_x=0$ then $c_s=c_a=0$, whereas if the transverse component of the magnetic field, ${\bf B}_t$, vanishes then $c_f=c_a$ if $c_a>a$, $c_s=c_a$ if $c_a<a$ and $c_s=c_f=c_a$ if $c_a=a$. The MHD equations are therefore not strictly hyperbolic. Brio \& Wu (1988) also argued that they are non-convex, but we shall postpone discussion of this until later. \subsubsection{Shock Types} The MHD shock equations allow two linear solutions and several distinct types of nonlinear solutions which satisfy the entropy principle that the entropy of a fluid element always increases. A convenient way of classifying these is to use the jump in the transverse component of magnetic field, ${\bf B}_t$. From the shock equations one finds (Jeffrey \& Taniuti 1964) \begin{equation} [{\bf B}_t(c_a^2 - v_x^2)]_l = [{\bf B}_t(c_a^2 - v_x^2)]_r, \label{7} \end{equation} where $v_x$ is the velocity in the shock frame. Note that if $c_a^2-v_x^2$ does not vanish, then ${\bf B}_t$ on one side of the discontinuity must be either parallel or anti-parallel to that on the other. The nonlinear solutions are \begin{enumerate} \item Slow/Fast shocks, which have non-zero ${\bf B}_t$ in the same direction on both sides. (\ref{7}) then implies that there is no change in sign of $(c_a^2 - v_x^2)$. The magnitude of magnetic field is larger on the downstream side for fast shocks and smaller downstream for slow shocks. \item Intermediate shocks, which also have non-zero ${\bf B}_t$ but in opposite directions on either side of the shock (Anderson 1963, Cabannes 1970). (\ref{7}) then implies that $(c_a^2 - v_x^2)$ changes sign. \item Switch-on shocks, which have vanishing ${\bf B}_t$ upstream. (\ref{7}) then implies that $v_x^2=c_a^2$ on the downstream side. \item Switch-off shocks, which have vanishing ${\bf B}_t$ downstream. (\ref{7}) then implies that $v_x^2=c_a^2$ on the upstream side. \end{enumerate} The linear discontinuities are \begin{enumerate} \item Alfv\'en discontinuities, which have $v_x^2=c_a^2$ on both sides. case (\ref{7}) then allows an arbitrary change in the direction of ${\bf B}_t$. However, the magnitude of ${\bf B}_t$ remains unchanged, which is why these are sometimes called rotational discontinuities. \item Contact discontinuities, which have the same value of $v_x$ on both sides, but $v_x^2 \not= c_a^2$. (\ref{7}) then requires that ${\bf B}_t$ be continuous unless $B_x = 0$ and the other shock conditions require all other variables, except for the density, to be continuous. \end{enumerate} We shall also find occasion to use the following classification of nonlinear MHD shocks, which is due to Germain (1960). The states in the shock frame are divided into four types \[ \begin{array}{ll} {\rm 1)}~~|v_x| > c_f, & {\rm 2)}~~c_f > |v_x| > c_a, \\ {\rm 3)}~~ c_a > |v_x| > c_s, & {\rm 4)}~ c_s > |v_x|, \end{array} \] and a shock is defined to be of type $m \rightarrow n$ if the upstream and downstream states are of types $m$ and $n$ respectively. From the MHD shock equations, one finds that pressure and specific volume, $\tau$ ($\tau = 1/\rho$), on each side of a nonlinear shock satisfy the following equations \[ p + G^2 \tau + {1 \over 2} \frac{F_y^2}{(\tau-\tau_a)^2} = F_x , \] \[ w\tau + {1 \over 2} G^2 \tau^2 + {\tau \over 2 \tau_a } \frac{F_y^2}{(\tau-\tau_a)^2} = H , \] where $G$ is the mass flux, $F_x$, $F_y$, $H$ are shock invariants and $\tau_a= B_x^2/G^2$. The analysis in Anderson (1963) can be used to show that the function $H(\tau)$ is as shown in Figure 1. $\tau-\tau_i$ has the same sign as $v_x^2-c_i^2$, where $i=s,a,f$. One can see that there are six different types of compressive shocks: fast shocks ($1 \rightarrow 2$), slow shocks ($3 \rightarrow 4$), and four intermediate shocks: $1 \rightarrow 3$, $1 \rightarrow 4$, $2 \rightarrow 3$, and $2 \rightarrow 4$. Depending on the relative position of the maxima of $H$, there are also limit shocks which propagate with the fast speed relative to the upstream state and/or the slow speed relative to the downstream state (see figures 1{\it b,c}). We shall denote such such shocks by $f \rightarrow n$ and $n \rightarrow s$ respectively. These shocks turn out not to be evolutionary, but if they were, then MHD would be a non-convex system. \begin{figure} \begin{center} \leavevmode \epsffile[24 26 378 137]{curve.ps} \caption{The shock invariant, $H$, as a function of specific volume, $\tau$, for three different cases.} \end{center} \label{curve-figure} \end{figure} \subsubsection{Evolutionary conditions} When we apply the evolutionary conditions to MHD discontinuities we have to take into account the fact that the system of shock and compatibility equations split into two independent subsets for all types of discontinuities, except the alfv\'en discontinuity,. If we choose a reference frame such that on one side of a discontinuity $B_z=0$ and $v_z=0$ then the system of shock equations contains two equations involving $B_z$ and $v_z$. These are \[ {B_z}_l = {B_z}_r , \] \[ {v_z}_l = {v_z}_r . \] The compatibility relations along the alfv\'en characteristics only involve $B_z$ and $v_z$ and they also the only ones that do so. An evolutionary discontinuity that is not an alfv\'en discontinuity must therefore not only satisfy the general condition (\ref{a5}), but also have exactly two incoming, and hence two outgoing alfv\'en characteristics. These conditions also follow from the linear stability analysis (Syrovatskii 1959, Jeffrey \& Tanuiti 1964). In the rest of this section we simply state the well known results on the evolutionary properties of MHD discontinuities. We do, however, pay particular attention to those cases in which there are characteristics travelling with the same speed as the discontinuity. As we have pointed out in Section 2, such characteristics must be counted as incoming. There is no dispute about the fact that fast and slow shocks are evolutionary because they have eight incoming characteristics, two of which are alfv\'en waves. Furthermore, since their speed can never be equal to a characteristic speed, Theorem 3.1 tells us that they also have a unique structurally stable dissipative structure. All intermediate shocks are super-alfv\'enic with respect to the upstream state and sub-alfv\'enic with respect to the downstream state, which means that they have too many ($>2$) incoming alfv\'en characteristics. They are therefore nonevolutionary and can be destroyed by interactions alfv\'en waves. The same argument applies to switch-on and switch-off shocks which also have too many (9) incoming characteristics, 3 of which are alfv\'en characteristics. However, these solutions are clearly limits of fast and slow shocks and therefore have evolutionary solutions in their immediate neighbourhood, which is why Jeffrey \& Taniuti (1964) call them weakly evolutionary. That they are not strictly evolutionary can also be understood from the following example. Consider a switch-on shock overtaking weak switch-off fast rarefaction travelling in the same direction. Once these have merged, the shock is no longer propagating into a state with zero transverse magnetic field. Since the shock is superfast, it has no way of modifying its upstream state and therefore cannot remain a switch-on shock. Instead, such an interaction leads to the appearance of a neighbouring fast shock solution, together with some other waves, at least one of which must, in general, be an alfv\'en wave. If we count the two entropy characteristics as incoming on the grounds that they have the same speed as the discontinuity, then contact discontinuities have eight incoming characteristics, two of which are alfv\'en characteristics. They are therefore evolutionary. Alfv\'en discontinuities also have eight incoming characteristics if we include the two alfv\'en characteristics that have the same speed as the discontinuity. The total number of incoming alfv\'en characteristics is three, but this is allowed since the fact that the shock equations for these discontinuities couple the $y$ and $z$ components of velocity and magnetic field means that this is the one case for which the shock equations do not decompose into two sets. Theorem 3.1 cannot be applied to contact and alfv\'en discontinuities since they propagate with a characteristic speed. However, they would in any case not possess a steady dissipative structure simply because they are linear and therefore have no nonlinear steepening to balance the spreading due to dissipation. For this reason, Wu (1988b) considers them to be inadmissible, but since their width grows like $t^{1/2}$, whereas the separation between the waves in a Riemann problem grows like $t$, they must be regarded as admissible components of the solution for large times. \subsection{Reduced system of planar MHD} \label{sb-planar} In this section we discuss the system of equations which describes MHD in a world in which the plane defined by the velocity and the magnetic field is invariant. There are several reasons for doing this. Firstly, it has some interesting properties. Secondly, we want to show that the general classical theory of shocks is as valid for this system as it is for the full system. Finally, the numerical simulations that gave rise to the current conroversy surrounding intermediate shocks reflect the properties of this system. When the z components of the magnetic field and velocity vanish, the equations reduce to a system of $5$ variables with the following vectors of conserved quantities and fluxes \[ \begin{array}{ll} {\bf u} = \left[{ \begin{array}{c} \rho \\ \rho v_x \\ \rho v_y \\ e \\ B_y \\ \end{array} }\right] & {\bf f} = \left[{ \begin{array}{c} \rho v_x \\ \rho v_x^2 + p_g + B^2/2 - B_x^2 \\ \rho v_x v_y - B_x B_y \\ \{e + p_g + B^2/2 \}v_x - B_x({\bf v}.{\bf B}) \\ v_x B_y - v_y B_x \\ \end{array} }\right] \end{array}. \] This is still a hyperbolic system but it is fundamentally different from the full system of MHD because it does not have alfv\'en waves. However, the other characteristic fields are still present with the same eigenvalues and with eigenvectors that are the same apart from the reduced number of components. Moreover, it has the same solutions of the shock equations including the alfv\'en discontinuity, except that these are now only allowed to change the direction of the transverse magnetic field by $\pi$. This follows from the remarkable property of the full system of MHD that there exists an inertial frame in which the variation of the transverse components of the magnetic field and velocity induced by all characteristic waves and shocks, except for alfv\'en waves, are confined to single plane. Note that the alfv\'en discontinuity still propagates with the alfve\'n speed, but this is no longer on of the characteristic speeds. The Riemann problem for this system has been analysed in considerable detail by Myong \& Roe (1997b) who came to the conclusion that the classical evolutionary conditions are inadequate for this system. However, we intend to show that this claim is based on a failure to recognise the essential difference between the reduced system and full MHD. \subsubsection{Evolutionary conditions} Since the number of equation is reduced by two and it is the alfv\'en waves that are lost, we can conclude that all evolutionary discontinuities that have two incoming alfv\'en characteristics in the full system remain evolutionary in the planar system. This implies that fast, slow and contact discontinuities are evolutionary. On the other hand, discontinuities that are evolutionary in the full system, but which do not have exactly two incoming alfv\'en characteristics must be non-evolutionary in the planar system. There is only one such discontinuity, the alfv\'en discontinuity, which now only has $5$ incoming characteristics and should therefore spontaneously self-destruct even if it is not perturbed. Another interesting feature is that some of the shocks that are non-evolutionary in the full system become evolutionary in the reduced system. $1\rightarrow 3$ shocks now satisfy the strong evolutionary condition, in fact they have the same incoming and outgoing characteristics as fast and switch-on shocks. As far as the characteristic count is concerned these three shocks are therefore indistinguishable so that one can use a single name, {\it plane fast shock}, say, for all of them. Similarly, $2 \rightarrow 4$ shocks, switch-off shocks and slow shocks become slightly different versions of evolutionary {\it plane slow shocks}. However, $1 \rightarrow 4$ shocks remain non-evolutionary even in the plane system since they have $7$ incoming characteristics. Such shocks, which have too many incoming characteristics, are often called {\it overcompressive} in the literature. As we have shown, although they do have a steady dissipative structure, it is not unique and it does not help them to survive interactions with external perturbations. $2 \rightarrow 3$ shocks have only $5$ incoming characteristics and are therefore non-evolutionary. Such shocks, which have too few incoming characteristics, are often called {\it undercompressive}. Since they do not have a structurally stable steady dissipative structure they should disintegrate spontaneously even without any external perturbation. Now consider shocks that propagate at one of the characteristic speeds in either the upstream or downstream state. $1 \rightarrow s$, $f \rightarrow 4$, and $f \rightarrow s$ shocks are non-evolutionary since they have $7$ incoming characteristics. On the other hand, $2 \rightarrow s$ and $f \rightarrow 3$ shocks have $6$ incoming characteristics and are therefore evolutionary. The planar system of MHD is therefore genuinely non-convex and admits the two evolutionary compound waves: a {\it slow compound wave} consisting of a $2 \rightarrow s$ shock with an attached slow rarefaction and a {\it fast compound wave} consisting of a fast rarefaction with an attached $f \rightarrow 3$ shock. Finally, we list the evolutionary shocks and compound waves of the planar system along with the notation used in Myong \& Roe (1997b): \begin{description} \item{Slow planar shock} (S1); \item{Fast planar shock} (S2); \item{Slow compound wave} (C1); \item{Fast compound wave} (C2); \item{Contact discontinuity} (not considered). \end{description} Myong \& Roe (1997b) found that some Riemann problems only have a solution if non-evolutionary shocks are permitted. However, as we discuss in \S\ref{dissc}, these Riemann problems are confined to regions of parameter space with zero volume, which is exactly what is meant by the statement that non-evolutionary shocks are structurally unstable. In the next section we show that the results of numerical calculations are entirely consistent with these conclusions. \section{Numerical Calculations} \label{numerics} The numerical calculations were carried out using the scheme described in Falle, Komissarov \& Joarder (1998). This is an upwind shock capturing scheme which is capable of dealing with shocks of arbitrary strength even without the inclusion of any dissipation other than that introduced by the truncation errors. Careful test simulations have shown that this scheme provides accurate solutions for all types of MHD waves in all regimes. One can argue that if a numerical scheme works well then its numerical dissipation must have the same qualitative properties as the physical dissipation. However, in order to remove any doubts, we modified our scheme so that it can now handle dissipative MHD and all the calculations described here have a fully resolved dissipative shock structures (about 15 mesh points wide). For this we used a simple scalar form for the dissipation for which equations (\ref{a1}) become \[ {{\partial {\bf u}} \over {\partial t}} + {{\partial {\bf f}} \over {\partial x}} = {\partial \over {\partial x}} {\bf g}, \] where the diffusive fluxes are \[ {\bf g} = \left({ \begin{array}{c} 0 \\ \\ \displaystyle{{{4 \mu} \over 3}{{\partial v_x} \over {\partial x}}} \\ \\ \displaystyle{\mu {{\partial v_y} \over {\partial x}}} \\ \\ \displaystyle{\mu {{\partial v_z} \over {\partial x}}} \\ \\ \displaystyle{{{4 \mu v_x} \over 3} {{\partial v_x} \over {\partial x}} + \mu v_y{{\partial v_y} \over {\partial x}} + \mu v_z{{\partial v_z} \over {\partial x}} + \nu_m \left[ {B_y {{\partial B_y} \over {\partial x}} + B_z {{\partial B_z} \over {\partial x}}} \right]} \\ \\ \displaystyle{\nu_m {{\partial B_y} \over {\partial x}}}\\ \\ \displaystyle{\nu_m {{\partial B_z} \over {\partial x}}}\\ \end{array} } \right) \] where $\mu$ is the dynamic viscosity, $\kappa$ the thermal conductivity and $\nu_m$ the resistivity. As expected, the outcomes of all the simulations presented here did not not depend on the size of dissipation and were the same even when only numerical and/or artificial dissipation was present. The only effect of changing the dissipation was to alter the form and width of the shock structures. \begin{table} \begin{center} \begin{tabular}{l} $2 \rightarrow 3$ Intermediate Shock (figure 2a)\\ ~~Left state: $\rho = 1,~p_g = 1,~{\bf v} = (-0.95,0,0),~{\bf B} = (1,0.5,0)$ \\ ~~Right state: $\rho = 0.837,~p_g = 0.705,~{\bf v} = (-1.135,1.266,0),~{\bf B} = (1,-0.7,0)$ \\ \\ Alfv\'en Shock (figure 2b)\\ ~~Left state: $\rho = 1,~p_g = 1,~{\bf v} = (-1,1,0),~{\bf B} = (1,1,0)$ \\ ~~Right state: $\rho = 1,~p_g = 1,~{\bf v} = (-1,3,0),~{\bf B} = (1,-1,0)$ \\ \\ $1 \rightarrow 3$ Intermediate Shock (figures 3a, 5 and 7a) \\ ~~Left state: $\rho = 1,~p_g = 1,~{\bf v} = (-0.925,0,0),~{\bf B} = (1,0.5,0)$ \\ ~~Right state: $\rho = 0.498,~p_g = 0.258,~{\bf v} = (-1.857,0.648,0), ~{\bf B} = (1,-0.1,0)$ \\ \\ $2 \rightarrow 4$ Intermediate Shock (figures 3b and 7b )\\ ~~Left state: $\rho = 1,~p_g = 1,~{\bf v} = (-0.4,0,0),~{\bf B} = (0.5,0.5,0)$ \\ ~~Right state: $\rho = 0.561,~p_g = 0.155,~{\bf v} = (-0.714,2.252,0),~{\bf B} = (0.5,-1.3,0)$ \\ \\ $1 \rightarrow 4$ Intermediate Shock (figure 4)\\ ~~Left state: $\rho = 1,~p_g = 1.2,~{\bf v} = (-0.842,0.0,0.0),~{\bf B} = (1.0,0.4,0)$ \\ ~~Right state: $\rho = 0.390,~p_g = 0.161,~{\bf v} = (-2.16,0.644,0),~{\bf B} = (1.0,-0.142,0)$ \\ \\ Brio \& Wu Problem (figure 8)\\ ~~Left state: $\rho = 1,~p_g = 1,~{\bf v} = (0,0,0),~{\bf B} = (0.75,1, 0)$ \\ ~~Right state: $\rho = 0.125,~p_g = 0.1,~{\bf v} = (0,0,0), ~{\bf B} = (0.75,-1,0)$ \\ \\ \end{tabular} \end{center} \caption{Riemann problems for the numerical calculations.} \end{table} \begin{table} \begin{center} \begin{tabular}{llllll} Problem & Domain & $n$ & $\mu /\rho $ & $\kappa/\rho$ & $\nu_m$ \\ \hline \hline Figure 2a & $[-4,1]$ & 250 & 0.02 & 0.01 & 0.01 \\ Figure 2b & $[-2,1]$ & 150 & 0.02 & 0.01 & 0.01 \\ Figure 3a,b & $[-4,1]$ & 250 & 0.02 & 0.01 & 0.01 \\ Figure 4 & $[-4,1]$ & 250 & 0.02 & 0.01 & 0.01 \\ Figure 5 & $[-1,1]$ & 200 & 0.01 & 0.005 & 0.005 \\ Figure 6 & $[-2,1]$ & 300 & 0.01 & 0.005 & 0.005 \\ Figure 7a & $[-8,2]$ & 500 & 0.02 & 0.01 & 0.01 \\ Figure 7a & $[-14,1]$ & 750 & 0.02 & 0.01 & 0.01 \\ Figure 8a,b & $[2.5,4.5]$ & 200 & 0.0 & 0.0 & 0.0 \\ \hline \end{tabular} \end{center} \caption{Other parameters for the numerical calculations. $n$ is the number of mesh points, $\mu$ is the kinematic viscosity, $\kappa$ is the thermal conductivity, $\nu_m$ is the resistivity.} \end{table} First of all, we need to establish whether the behaviour of numerical MHD shocks agrees with the predictions of the evolutionary theory. In order to do this, we adopt the following procedure. First we test whether a shock has a steady dissipative structure by setting up the relevant Riemann problem and running the calculation until a well resolved steady dissipative shock structure is established, as expected for evolutionary and overdetermined shocks, or a completely different solution emerges, as expected for underdetermined shocks. If a steady structure exists, then we test to see whether it can survive small perturbations. This can be accomplished by considering a slightly different Riemann problem, as in Barmin et al. (1996) or, like Wu (1988a), allowing a small amplitude wave to interact with the shock. \begin{figure} \begin{center} \leavevmode \hbox \epsffile[15 13 380 290]{noshst.ps}} \caption{Planar simulations of shocks that should not have a steady dissipative structure in planar MHD: (a) $2 \rightarrow 3$ shock, (b) Alfv\'en shock. In both cases the outcome is a slow compound wave (SCW). The dashed lines show the corresponding initial solutions. The continuous lines show the final solutions.} \end{center} \label{noshst} \end{figure} \begin{figure} \begin{center} \leavevmode \hbox \epsffile[15 13 380 432]{evsh.ps}} \caption{Planar simulations of the interaction between evolutionary shocks and small amplitude fast rarefactions ($\delta B_t = 10\%$). (a) fast ($1 \rightarrow 3$) shock, (b) slow ($2 \rightarrow 4$) shock. In both cases the outcome is a shock of the same type, together with some other waves. Here FR denotes a fast rarefaction and $V_x$ is the x-component of velocity as measured in the shock frame. The dashed lines show the initial solutions. The continuous and dotted lines show the final solutions.} \end{center} \label{evsh} \end{figure} \subsection{Planar MHD} We start by discussing the results of the planar simulations. They show that if the initial discontinuity corresponds to a slow planar shock then a smooth steady shock structure connecting the initial left and right states finally develops and it does not matter whether the shock is $3 \rightarrow 4$ or $2\rightarrow 4$. The same thing happens for the fast planar shock and the overdetermined (overcompressive) $1 \rightarrow 4$ shock. In contrast, figure 2 shows that $2 \rightarrow 3$ shocks and alfv\'en shocks always turn into a slow compound wave. All this is exactly as perdicted by the theory described in \S\ref{dis-str} and \S\ref{apple}. Our simulations cannot be used to determine whether limit shocks, such $1 \rightarrow s$ and $f \rightarrow 3$), have a steady dissipative shock structure, simply because it is impossible to set up a shock whose speed is exactly equal to a characteristic speed. However, if we compute a Riemann problem that corresponds to a compound wave of any of the types discussed above, the wave that is expected, or strictly speaking a solution close to such a wave, always emerges. This is hardly surprising because all of them have neighbouring solutions containing shocks with a steady dissipative structure. \begin{figure} \begin{center} \leavevmode \hbox \epsffile[15 13 380 432]{is14.ps}} \caption{Planar simulations of $1 \rightarrow 4$ shock subjected to a small variation of pressure ($\pm 10\%$) in the left state. This shock is non-evolutionary even in planar MHD and splits as the result of the perturbation into two evolutionary shocks plus other small amplitude waves The outcome is (a) $1 \rightarrow 3$ and $ 3 \rightarrow 4$ shocks if $\delta p = -10\%$ and (b) $1 \rightarrow 2$ and $2 \rightarrow 4$ shocks if $\delta p = +10\%$. The dashed lines show the initial solutions. The continuous and dotted lines show the final solutions. $V_x$ is the x-component of velocity as measured in the frame of the emerged intemediate shock.} \end{center} \label{is14} \end{figure} As is shown in figure~3, evolutionary shocks always survive interactions with small amplitude waves and persist if the Riemann problem is perturbed. Figure~4 shows how a small variation of the initial data forces an overdetermined $1 \rightarrow 4$ shock to split into two evolutionary shocks. Depending on the form of the perturbation, the shock either splits into a $1 \rightarrow 2$ shock followed by a $2 \rightarrow 4$ shock or a $1 \rightarrow 3$ shock followed by a $3 \rightarrow 4$ shock. This is to be expected because, as one can see from figure~1, a $1 \rightarrow 4$ shock is exactly equivalent to one or other of these shock pairs propagating with the same speed. In fact, this result is in complete agreement with the analysis of the Riemann problem for planar MHD in Myong and Roe (1997b). $1 \rightarrow 4$ shocks, {\it O-shocks} in their notation, are only required on the boundary between the two domains of parameter space in which their solution involves a combination of fast and slow planar shocks (S2 and S1). The results for compound waves involving non-evolutionary shocks are similar. Figure~1 shows that the non-evolutionary $1 \rightarrow s$ limit shock can be understood as a {\it double-layer} shock composed of two evolutionary shocks, a $1 \rightarrow 2$ and a $2 \rightarrow s$. Indeed, if the Riemann problem corresponding to a compound wave containing such a shock is perturbed, then in some cases the outcome is a $1 \rightarrow 2$ shock and a slow compound wave and in other cases it is a $1 \rightarrow 3$ shock and a detached slow rarefaction. {\it All this can be summed up by saying that in planar MHD the behavior of shocks in our numerical simulations is entirely consistent with the classical evolutionary theory of shocks and the theory of dissipative shock structures as described in \S\ref{sgeneral} and \S\ref{dis-str}.} \subsection{Full MHD} Since both fast ($1 \rightarrow 2$) and slow ($3 \rightarrow 4$) shocks satisfy the strong evolutionary condition in full MHD they are expected to have unique dissipative structure and be stable with respect to small perturbations of any kind. This is precisely what we find from our simulations. $1 \rightarrow 3$ and $2 \rightarrow 4$ shocks are overdetermined in full MHD and it is therefore possible that they might have a nonunique steady dissipative structure, indeed it turns out that they do. These shocks, as well as $1 \rightarrow 4$ shocks, can now have a nonvanishing z-component of magnetic field inside the shock layer even if $B_z=0$ outside. For given the dissipative coefficients their stucture can be parameterised by the value of the following integral \[ I_z = \int_{-\infty}^{+\infty} B_z dx . \] We can gradually increase or decrease the value of $I_z$ by sending from the downstream side of the shock an alfv\'en wave that first rotates the magnetic field by a small angle and then restores the original state. This wave is absorbed by the shock which develops a new steady structure (see Figure~5{\it a}). However, like Kennel et al. (1990) we found that there is a maximum value of $|I_z|$ that the shock can manage. If this limit is exceeded, then the shock disintegrates (see Figure~{\it b}). This does not occur in the case of fast and slow shocks because the alfv\'en waves do not get trapped inside the shocks, but instead pass straight through. $2 \rightarrow 3$ shocks have the right number of incoming characteristics and may therefore have a unique dissipative structure in full MHD. Since such a structure does not exist in planar MHD, we can only expect to find them in our simulations by allowing a non-zero $B_z$. In order to do this, we modified the initial data by inserting a layer in which the transverse field rotates smoothly from that in the original left state to that in the original right state. We found that the solution never relaxed to a smooth steady $2 \rightarrow 3$ transition and were about to conclude that no steady structure exists until we realised that the solution shown in figure~5{\it b} actually contains a $2 \rightarrow 3$ shock, which was produced by the disintegration of the $1 \rightarrow 3$ shock. We therefore we studied the reaction of a $1 \rightarrow 3$ shock to an increase in $I_z$. After absorbng another alfv\'en wave the shock splits and one of the emerging waves is again a $2 \rightarrow 3$ shock but of smaller amplitude (figure~6). This behaviour is consistent with the existence of a unique dissipative structure for $2 \rightarrow 3$ shocks. In fact, what happens is that, as $I_z$ increases, the shock tends to an alfv\'en shock that rotates the transverse field by $\pi$ Finally, we have also verified that all intermediate shocks and compound waves disintegrate when exposed to perturbations that render the left and right states non-coplanar. For example, figure~7 shows how $1 \rightarrow 3$ and $2 \rightarrow 4$ shocks split into evolutionary waves after interaction with a small amplitude alfv\'en wave. After the alfv\'en wave has been absorbed the transverse fields on either side of the shock are no longer parallel or antiparallel as required by the shock equations. The shock can only become coplanar by emitting alfv\'en waves, which, for an intermediate shock, can only be done in the downstream direction. However, since there is no downstream travelling alfv\'en wave that can restore the original post-shock state, the shock must split. This argument is not new, in fact it was used by Kantrowitz \& Petschek (1966) to prove that intermediate shocks are unphysical. The wave designated as AW in figure~7 can be called a dissipative alfv\'en wave but it could also be described as an evolving $2 \rightarrow 3$ shock with a gradually increasing value of $I_z$. {\it We therefore conclude that for full MHD the behaviour of shocks in our numerical simulations is also entirely consistent with the classical evolutionary theory of shocks and the theory of dissipative shock structures as described in \S\ref{sgeneral} and \S\ref{dis-str}.} \begin{figure} \begin{center} \leavevmode \hbox \epsffile[15 13 380 289]{is13.ps}} \caption{Dissipative structure of a $1 \rightarrow 3$ shock in full MHD for different values of $I_z$. This shock has a nonunique steady dissipative structure that depends upon $I_z$. For relatively small values of $I_z$ this structure is steady (left panel, $I_z=-0.085$) but for larger values it splits into $1 \rightarrow 2$ and $1 \rightarrow 2$ shocks (right panel, $I_z=-0.20$.} \end{center} \label{is13} \end{figure} \begin{figure} \begin{center} \leavevmode \hbox \epsffile[15 13 380 195]{is23.ps}} \caption{Dissipative structure of a $2 \rightarrow 3$ shock in full MHD. This shock has a unique steady dissipative structure and therefore reacts to a change in $I_z$ by emitting some waves and turning into a different $2 \rightarrow 3$ shock. The continuous lines show the solution for $I_z=-0.20$ and the dashed lines for $I_z=-0.52$.} \end{center} \label{is23} \end{figure} \begin{figure} \begin{center} \leavevmode \hbox \epsffile[15 13 380 445]{sinta.ps}} \caption{MHD shocks interacting with small amplitude alfv\'en waves. The evolutionary shocks survive, but the non-evolutionary ones split. (a) A $1 \rightarrow 3$ shock splits into a fast shock (FS), an alfv\'en wave (AW) and a slow shock (SS); b) A $2 \rightarrow 4$ shock splits into an alfv\'en wave and a slow shock. Other small amplitude waves are also emitted. The dashed line shows the exact ideal solution of the Riemann problem for the initial state formed by the collision of the intermediate and alfv\'en shocks.} \end{center} \label{sinta} \end{figure} \begin{figure} \begin{center} \leavevmode \hbox \epsffile[0 0 390 357]{brio.ps}} \caption{Brio \& Wu problem (Brio \& Wu (1988). (a) Numerical solution found using a Godunov type scheme. This is a proper solution of the reduced system of planar MHD but is inadmissible in full MHD. (b) Numerical solution found using Glimm's scheme to track alfv\'en discontinuties (markers) and the exact solution involving only evolutionary shocks (lines). This is a proper solution for full MHD and is the only physically admissible solution for this problem.} \end{center} \label{brio} \end{figure} \section{Discussion} \label{dissc} The results described in the previous sections have clarified many aspects of the shock theory in general and MHD shocks in particular and provide a basis upon which we can discuss other important, related, issues. \subsection{Riemann problems and evolutionary conditions} One of the arguments in favour of non-evolutionary shocks used in current literature is that some Riemann problems do not have a solution unless non-evolutionary shocks are admitted (e.g. Glimm 1988, Myong \& Roe 1997a,b). This is presumably based on the belief that any Riemann problem must have a physically admissible solution. Although this is certainly true for gas dynamics, there is surely no reason why this has to hold for any system. It all comes down to the notions of bifurcations and structural stability. One has to ask the following question: is it, or is it not, possible to carry out the relevant experiment in a laboratory? If the qualitative result of the experiment does not change when the initial conditions are slightly changed, then the problem is structurally stable and the experiment is possible, at least in principle. However, if this is not true, then the problem is structurally unstable and no appropriate experiment is possible. It therefore follows that the set of structurally unstable Riemann problem are confined to regions of parameter space whose total volume is zero. Now suppose there is an MHD Riemann problem that has no other solutions than those containing non-evolutionary shocks. Since there are arbitrary small perturbations of the parameters that cause these shocks to split into evolutionary shocks, this Riemann problem must be structurally unstable. In full MHD the only known case for which a non-evolutionary shock, a $1 \rightarrow 4$ shock, is required is a piston problem in which the piston velocity is parallel to the magnetic field (Jeffrey \& Taniuti 1964). If this condition is not exactly satisfied then the non-evolutionary shock does not arise. Close inspection of the solution of the Riemann problem for planar MHD presented by Myong \& Roe (1997b) shows that non-evolutionary shocks are required only on the boundaries between domains in parameter space that contain only evolutionary shocks. \subsection{Steepening of continuous waves} Another argument that appears to justify the existence of intermediate shocks is based on the results of numerical simulations by Wu (1987), which suggest that intermediate shocks can be formed by nonlinear steepening of simple magnetosonic waves. Since the transverse component of magnetic field changes sign across an intermediate shock the simple wave must have the same property, which means that the transverse component of magnetic field must vanish somewhere within the wave. However, at this point the magnetosonic speed is equal to the alfv\'en speed and it is impossible to assign a unique eigenvector to the simple wave. As the result, the direction of the tangential component of the field can rotate by an arbitrary angle at this point so that simple wave really consists of two distinct parts, which are disconnected as far as the direction of the magnetic field is concerned. This can be put in a slightly different way. Alfv\'en waves propagating in the same direction as such a simple wave cannot pass throught the alfv\'en point. During the steepening they will accumulate near this point giving rise to a net field rotation so that the discontinuity that forms has non-coplanar left and right states and can therefore not be a single shock. Instead, it must split into evolutionary shocks, one of which must be an alfv\'en shock. Incidentally, this seems to be the only way of generating alfv\'en shocks. However, in planar MHD the transition through the alfv\'enic point is unique and as we have seen some of the intermediate shocks are in fact evolutionary. This is the explanation for the outcome of the planar simulations performed by Wu (1987). He also found that the results were not very different if the initial data was perturbed so that it was no longer exactly coplanar. However, because of the periodic boundary conditions used in this simulation, there was no net rotation in the perturbed problem, which makes it rather artificial. The reason why this perturbation did not destroy the intermediate shock is that these boundary conditions, together with the initial data, only allowed a small value of $I_z$ per shock. It is therefore hardly surprising that an intermediate shock appeared since, as we have shown, these shocks can survive if $I_z$ is small enough. \subsection{Timescale for disintegration} Let us suppose that an intermediate shock has somehow been formed and then interacts with an alfv\'en wave that rotates the magnetic field by a small angle $\delta\phi$. It is clearly of some importance to know how long it takes for the shock to split. Our simulations show that it splits when the the value of $I_z$ associated with the shock structure becomes comparable with $lB_y$, where $l$ is the shock thickness. If the incident alfv\'en wave has a small amplitude, $\delta \phi$, then this gives us the following estimate for the disintegration time, $t_s$ \begin{equation} t_s \approx {l \over {c_a \delta\phi}}, \label{46} \end{equation} \noindent where we have used the alfv\'en speed as a characteristic fluid velocity in the shock frame. This also tells us that the shock will only propagate for a distance $\approx l/\delta\phi$ before it falls apart. From this we conclude that in all cases for which the dissipative scale is much smaller then the characteristic length scale of the flow, intermediate shocks can only appear as very short lived time-dependent phenomena. It is instructive to apply equation (\ref{46}) to the interplanetary intermediate shock for which Chao et al. (1993) claim to have found evidence in the Voyager 1 data. In this case $c_a = 40 {\rm ~km~s^{-1}}$ and $l = 5 \times 10^4 {\rm ~ km}$, which gives $t_s = 1.2 10^3 {\delta\phi}^{-1}{\rm~s}$. The flow time for the solar wind at this distance ($\simeq 9 {\rm ~AU}$) is $\simeq 3 \times 10^7 {\rm ~s}$. It is therefore clear that $\delta \phi$ would have to be ridiculously small for the shock to survive for a significant fraction of a flow time. This is most unlikely since the flow of the solar wind is sufficiently complex to contain plenty of alfv\'en waves for which $\delta \phi \sim 1$ and indeed Chao et al. find plenty of evidence for strong alfv\'en waves in the data. Actually, the evidence for an intermediate shock is not really very convincing. The uncertainties are such that it could just as well be a slow shock. Exactly the same arguments can be applied to the magnetohydrodynamic shocks in the interstellar medium. Not only does the theory of collisionless shocks (see e.g. Tidman \& Krall 1971) predict that, in these conditions, such shocks are extremely thin compared to the scale of the flow but there are numerous observations that confirm that this is indeed true (see e.g. Draine \& McKee 1993). \subsection{Convexity of MHD} From the above discussion it is quite clear that a hyperbolic system is genuinely non-convex if it allows structurally stable compound waves that only contain evolutionary shocks. Planar MHD is therefore genuinely non-convex whereas full MHD is convex. \subsection{Non-evolutionary shocks in numerical simulations} The appearance of non-evolutionary shocks in numerical calculations is not something that is unique to MHD since it is well known that, even in gas dynamics, some numerical schemes can generate expansion shocks in certain circumstances. However, this phenomenon is both more subtle and more interesting in the case of MHD. The essential point is that, unlike gas dynamics, planar MHD is is very different from full MHD in the sense that there are shocks that are non-evolutionary in full MHD, but evolutionary in planar MHD and vice-versa. Unfortunately, this property means that the results of planar MHD simulations can be very misleading because, although most upwind schemes seem to give perfectly good solutions for planar MHD, these are of no relevance to the real universe with its three spatial dimensions. This is not at all unusual, indeed it may very well be the rule rather than the exception. For example, the properties of fluid turbulence are very different in two and three dimensions as are those of magnetohydrodynamic dynamos. The other properties of non-evolutionary MHD shocks, that are not shared by gas dynamical expansion shocks, are that all of them satisfy the second law of thermodynamics and most of them also possess a steady dissipative structure. This, together with the fact that the ratio of the thickness of numerical shock structures to the overall scale of the flow is almost always many orders of magnitude greater than in the corresponding physical system, means that they can persist for a significant time even in nonplanar problems. For example, if the piston problem discussed by Jeffrey \& Taniuti (1964, p. 256--258) is slightly modified so that it has a small transverse component of the field, then the evolutionary solution contains fast, slow and alfv\'en shocks all propagating with very similar speeds. In a numerical simulation this complex would remain unresolved for some time, during which it would be classified as a $1 \rightarrow 4$ shock. The only truly satisfactory solution to this difficulty is to devise schemes that only allow evolutionary shocks. Figure~8 shows that there are schemes that will do this. Here we have a numerical solution to the Brio \& Wu problem obtained with our MHD version of Glimm's scheme (Glimm 1965). This method requires a nonlinear Riemann solver and we employ the one described in Falle et al. (1998), which specifically excludes intermediate shocks. In fact we do not use Glimm's scheme everywhere, but only to track the alfv\'en shock. One can see that in this way we can avoid the appearance of intermediate shocks even in planar problems. Unfortunately, it is not a simple matter to generalise this to more than one dimension. The only viable option, that we can think of, is to subject all numerical calculations to a careful analysis using the theory described in this paper. As an example of this, it is instructive look at some recent calculations of steady MHD flow past a cylinder. \subsection{2D bow shock simulations} Recently, De Sterck et al.(1998) have carried out numerical MHD calculations of the flow past an infinite, perfectly conducting cylinder. These are planar simulations and must therefore be interpreted in the light of the theory of planar MHD. The parameters are chosen in such way that the usual convex bow shock is impossible. Instead, the analysis given in Steinolfson \& Hundhausen (1990) suggests that the shock has a dimple. They assumed that there is only a single shock, in which case a cosistent solution requires the shock type to change from $1 \rightarrow 2$ to $1 \rightarrow 3$ and then to $1 \rightarrow 4$ as the distance from the symmetry axis decreases. Although the $1 \rightarrow 4$ shock is non-evolutionary even in planar MHD, in this case it seems that such a shock must occur on the symmetry axis for the same reason that it occurs when a piston moves parallel to the magnetic field. However, one would expect it to split into $1 \rightarrow 2$ and $2 \rightarrow 4$ or $1 \rightarrow 3$ and $3 \rightarrow 4$ shocks further away from the the axis. Indeed, De Sterck et al.(1998) find that not far from the axis the $1 \rightarrow 4$ shock splits and the leading shock (ED in their notation) is a $1 \rightarrow 2$. At some distance from this branching point the other shock (EG) is identified by them as $f \rightarrow s$, but this is unlikely to be true everywhere for such an inhomogeneous flow. One would also expect another branching at the point where Steinolfson \& Hundhausen (1990) predict a transition from $1 \rightarrow 3$ to $1 \rightarrow 4$. The results of De Sterck et al.(1998) do, indeed, show this branching (DE and DG), with the trailing shock being clearly identifiable as a $2 \rightarrow 4$ shock. \section{Conclusions} \label{conc} Both our analysis and numerical results show that the evolutionary conditions for existence and uniqueness of discontinuous solutions of the equations of ideal MHD are not only compatible with the conditions for existence and uniqueness of steady dissipative shock structures, they are actually complementary to them. The general theory suggests that this will be true for all nonlinear hyperbolic systems that can arise in nature. Non-evolutionary shock can have a nonunique dissipative structure and may, perhaps, appear under some exeptional curcumstances as transient phenomena. However, they are not persistent and are bound to split when subjected to small perturbations. In the case of MHD, alfv\'en waves are the most effective killers since not only our calculations but also those described by Wu (1988a) show that intermediate MHD shocks are destroyed by interactions with alfv\'en waves. It is true that it takes a finite time for this interaction to take place, but in any physical system that we know of, this time is so short that it is most unlikely that such shocks can be detected. The occurrence of intermediate MHD shocks in planar numerical simulations is consistent with the mathematical properties of planar MHD, in which $1 \rightarrow 3$ and $2 \rightarrow 4$ shocks become evolutionary but the alfv\'en shock becomes non-evolutionary. However, the planar limit is a singular limit of full MHD and we suggest that planar numerical simulations should be avoided, especially since they are hardly any cheaper than for full MHD. Intermediate shocks may even pollute full MHD simulations because numerical shock structures are usually not very thin compared to the length scale of the flow. It is therefore essential that the results of such simulations be subjected to a careful analysis in order to make sure that they do not contain any intermediate shocks. If they do, then additional work is required to determine the extent to which they are corrupted. The results of our calculations with Glimm's scheme show that this problem can be eliminated in numerical schemes that treat shocks especially alfv\'en shocks, as discontinuities.
\section{Introduction} Studies of the intracluster medium (ICM) address many important cosmological and astrophysical questions. Precision measurements of individual ICM spectra and metallicities can be used to constrain models of massive star formation. The slope and evolution of cluster scaling relations (e.g. the luminosity vs. temperature or ICM mass vs. temperature) can provide insight into the nature and importance of galactic winds, as well as the background cosmology in which the ICM develops. In addition, current models of structure formation suggest that the mass components of clusters are representative of the universe as a whole; for this reason there has been much effort expended towards measuring the contributions of dark matter, ICM, and galaxies in the population. The mean ICM mass fraction is therefore a useful lower limit on the global baryon fraction, and can also be used to constrain $\Omega_0$ if combined with primordial nucleosynthesis calculations. Many studies have been carried out recently which attempt to measure the form of the ICM density profile using X--ray images (e.g. Mohr, Mathiesen, \& Evrard 1999 (MME); White, Jones, \& Forman 1997; Loewenstein \& Mushotzky 1996; David, Jones, \& Forman 1995, White \& Fabian 1995). Following the lead of early work in the field (Forman \& Jones 1982), such studies typically create an analytic model for the azimuthally averaged ICM density profile, and use the X--ray surface brightness to constrain its parameters. These models generally assume that the ICM is uniform at a given radius, that the density profile decreases monotonically from the center, and that the cluster is spherically symmetric. These assumptions break down in real clusters; even apparently relaxed clusters tend to exhibit small asphericities. The prevalence of accretion events and major mergers in the local population is also well established, which calls into question the validity of a monotonic density profile. The assumption of a uniform ICM is a reasonable approximation, but it too must fail on sufficiently small scales. A certain level of random density fluctuations is to be expected, and even minor accretion events are found in simulations to produce persistent mild shocks and acoustic disturbances in the ICM. The small scatter in observed scaling relations, for example the ICM mass and mass fraction vs. temperature (MME), implies that these assumptions produce results with at least 20\% precision; this fact inspires confidence. It is important, however, to directly measure the systematic errors incurred by these inaccuracies. In accordance with the increasing quality of X--ray observations, the scientific community has recently begun to address these issues. A number of groups are investigating the possibility of a multiphase medium (Thomas 1998, Lima Neto et al. 1997, Waxman \& Miralde-Escude 1995), although most of the work to date focuses on cooling flows rather than global ICM properties. A notable exception is the work of Gunn \& Thomas (1996), who examine the effect of a multiphase ICM on baryon fraction measurements. There have also been attempts to get around the assumption of spherical symmetry by using the isophotal area (Mohr \& Evrard 1997) or elliptical isophotes (Knopp \& Henry 1996, Buote \& Canizares 1996) to constrain a density model. All of these methods, however, still fall prey to one or more of the assumptions mentioned above. Hydrodynamic simulations of cluster evolution are ideally suited to the task of estimating the systematic errors incurred by these assumptions. In this study, we produce realistic X--ray surface brightness images from a set of simulations and measure ICM masses by fitting the emission profile to a spherically symmetric beta model for the density. Our independent knowledge of the three-dimensional structure allows us to observe the effects of gross substructure and small-scale density fluctuations on the derived ICM mass directly, and correlate these errors with the kind of substructure present. In \S 2 of this paper, we present some detail on the nature of the simulations and how we create our X--ray images. In \S 3, we describe how we analyze these images and extract ICM density profiles. In \S 4, we compare the ``observed'' masses to the simulations and correlate the errors with structural properties of the ICM. Finally, in \S 5 we restate our results and remark on how they can be applied to real observations. Our results are phrased throughout in a manner indpendent of the Hubble constant $H_0$. \section{Data} We use an ensemble of 48 hydrodynamical cluster simulations, divided among four different cold dark matter (CDM) cosmological models. These models are (i) SCDM ($\Omega_0 = 1$, $\sigma_8 = 0.6$, $h = 0.5$, $\Gamma = 0.5$); (ii) $\tau$CDM ($\Omega_0 = 1$, $\sigma_8 = 0.6$, $h = 0.5$, $\Gamma = 0.24$); (iii) OCDM ($\Omega_0 = 0.3$, $\sigma_8 = 1.0$, $h = 0.8$, $\Gamma = 0.24$); and (iv) $\Lambda$CDM ($\Omega_0 = 0.3$, $\lambda_0 = 0.7$, $\sigma_8 = 1.0$, $h = 0.8$, $\Gamma = 0.24$). Here the Hubble constant is $100h$ km s$^{-1}$ Mpc$^{-1}$, and $\sigma_8$ is the power spectrum normalization on $8h^{-1}$ Mpc scales. The initial conditions are Gaussian random fields consistent with a CDM transfer function with the specified $\Gamma$ (Davis et al. 1985). The baryon density is set in each case to a fixed fraction of the total density ($\Omega_b = 0.2\Omega_0$). The simulation scheme is P3MSPH; first a P$^3$M (DM only) simulation is used to find cluster formation sites in a large volume, then a hydrodynamic simulation is performed on individual clusters to resolve their DM halo and ICM structure in detail. The resulting cluster sample covers a little more than a decade in total mass, ranging from about $10^{14}$ to $2 \times 10^{15} M_{\odot}$. We work with X--ray surface brightness maps derived from the simulations according to procedures described by Evrard (1990a,b). The particles representing pieces of the ICM are treated as bremsstrahlung emitters at the local temperature and density, and this emission is distributed over the image pixels according to a two-dimensional Gaussian with a width equal to that of the particle's SPH smoothing kernel. Emission is collected in the energy band [0.1,2.4] keV, and the resulting surface brightness in each pixel is converted to counts per second according to an approximate ROSAT PSPC energy conversion factor. We then ``observe'' this model image using an effective area map of the PSPC and an exposure sufficient to yield $10^4$ cluster photons, comparable to ROSAT exposures of clusters in the Edge sample (Edge et al. 1990). Finally, each pixel is given a Poisson uncertainty based on the number of photons and the whole image is smoothed on a scale of $24.4\arcsec$ to improve the signal-to-noise ratio. Angular distances are calculated as $d_A = cz/H_0$ with $z = 0.06$, so the physical smoothing scale is either 42.6 or 26.6 kpc. Note that beause we are trying to isolate the systematic errors due to substructure, we do not add Poisson noise to the image; the error bars are only used to assign appropriate weights to the data in the analysis. It should be noted that the simulations do not include certain processes known to be present in real clusters, such as radiative cooling and the injection of gas and energy by galactic supernovae. The cooling time for at least 99\% of the SPH particles is much longer than the Hubble time, and would have little effect on the structure of the ICM; nevertheless, our simulations cannot develop cooling flows. Detailed studies of real cooling flows find mass deposition rates no larger than hundreds of solar masses per year, suggesting that they comprise only a small fraction of the ICM mass (White et al. 1998). Simulations which include galactic winds create clusters with more realistic density profiles, but the winds are not found to greatly affect the temperature structure of the gas (Metzler \& Evrard 1994). It is conceivable that the presence of galactic winds would also create more clumping in the ICM, but this effect is probably negligible next to the variations caused by accretion events. The clusters display great morphological diversity, ranging from examples which appear almost perfectly spherical and relaxed to others which are undergoing a three-way merger event. Each cluster is imaged in three orthogonal projections, allowing us to compare the biases caused by substructure along the line-of-sight and substructure in the plane of the sky. We use all three projections in general, but confine ourselves to one projection per cluster when making statistical comparisons to insure that the probabilities are derived from statistically independent data. Further details on the nature of these simulations can be found in both MME and Mohr \& Evrard (1997). \section{Image Analysis} We fit the emission to a beta model, with three-dimensional density profile $\rho(r) = \rho_0[1+(r/r_c)^2]^{-3\beta/2}$ (Cavaliere \& Fusco-Femiano, 1978). In this model $r_c$ and $\beta$ are free parameters to be constrained by the X--ray emission profile, and $\rho_0$ is found by normalizing the emission integral to the bolometric luminosity of the cluster. Producing a cluster emission profile appropriate for fitting requires three steps. First, the emission center is found by sliding a circular aperture of radius 10 pixels over the image and minimizing the distance between the center of the aperture and the centroid of the X-ray photon distribution within the aperture. This approach converges even in cases where the cluster emission is significantly skew (Mohr, Fabricant, \& Geller 1993). Second, we calculate the azimuthally averaged radial profile. Finally, each point in the profile is assigned a Poisson uncertainty based on the number of photons in the annulus; this allows the fitting program to assign approprate relative weights to the datapoints. The PSPC point spread function (PSF) reduces a cluster's central intensity and increases the apparent value of $r_c$. This effect changes the best-fit values of our model parameters and biases the derived ICM mass if left untreated. The incurred error is potentially significant for clusters with small angular diameter or pronounced cooling flows, and is simulated here by the image smoothing mentioned in \S 2. Thus, rather than fitting the beta model directly to our profile, we first convolve it with a PSF appropriate for our smoothing kernel. Details on the mathematics of this technique can be found in MME and Saglia et al. (1993). The fitting was performed from the center of the cluster out to $r_{500}$, the radius at which the mean interior density is 500 times the critical density. This measure was chosen to probe the easily visible region of each cluster given typical backgrounds and sensitivities, as well as for physical reasons which will become apparent in the next section. By fitting out to a fixed fraction of the virial radius rather than a fixed metric radius, we insure that we are probing regions with similar dynamical gravitational time scales. The ratio $\Upsilon$ between cluster baryon fractions and the universal baryon fraction has also been calibrated by simulations; a bias factor of $\Upsilon = 0.9 \pm 0.1$ was found at $r_{500}$ (Evrard 1997). Recent work with an independent simulation has corroborated this result, finding $\Upsilon = 0.92 \pm 0.06$ at $r_{200}$ (Frenk et al. 1999). The radius $r_{500}$ is found in our simulations to scale as $1.4 h^{-1} (T/10\mathrm{keV})^{1/2}$ Mpc, regardless of cosmology. Finally, we checked our process by fitting 100 monte carlo images of true beta models with similar smoothing and individual realizations of Poisson noise, and found a chi-squared distribution consistent with a perfect match between the fitting function and the underlying model. This confirms that the high chi-squared values obtained in some images result from real physical deviations from a spherical beta model. A more complete discussion of our method, as applied to real PSPC data, is given in MME. \section{Results} We measure the variance of density fluctuations in spherical shells around the ICM particle with the lowest gravitational potential. This variance is expressed in terms of a clumping factor $C$, \begin{equation} C \equiv \langle\rho^2\rangle/\langle\rho\rangle^2. \end{equation} Each ICM particle in the simulation has mass $m_{\mathrm{gas}}$. For a given shell lying between radii $r_1$ and $r_2$ with $N_{12}$ gas particles, $\langle\rho\rangle$ is defined as the total mass in gas particles divided by the volume of the shell, $m_{\mathrm{gas}}N_{12}/V$. The numerator is calculated as \begin{equation} \langle\rho^2\rangle \equiv \frac{1}{V_{12}}\int^{r_2}_{r_1}d^3r\rho^2 \rightarrow\frac{m_{\rm gas}}{V_{12}}\sum_i^{N_{12}}\rho_i, \end{equation} with the right-hand term being the Lagrangian limit of the integral and $\rho_i$ the SPH gas density of particle $i$ (Evrard 1988). The results of this analysis are displayed in Figure~1. Although more complicated structure is visible in some of our clusters, the average magnitude of density fluctuations interior to $r_{500}$ seems to be a slowly increasing function of radius. The radial bins are chosen such that they represent fixed logarithmic intervals in the density constrast $\delta_c$. Dramatic increases in clumping are occasionally seen between $r_{500}$ and the more traditional virial radius; this is another reason that for choosing to work within a density constrast of 500. Extending our analysis to the virial radius would return essentially the same results with lower statistical significance. The presence of these fluctuations enhances the cluster luminosity over what would be expected for a smooth, single-phase ICM. The emissivity in a given shell is given by \begin{equation} \varepsilon(r) = r^2dr\int d\Omega \frac{\rho^2(r,\Omega)}{\mu_e\mu_Hm_p^2} \Lambda(T,\Omega). \end{equation} If the density fluctuations are modest, the temperature and ionization structure of the gas in a given shell will be approximately uniform. Taking these factors outside the integral and using equation (1) to incorporate our clumping factor, we have \begin{equation} \varepsilon(r) = 4\pi r^2dr\frac{\Lambda(T)}{\mu_e\mu_Hm_p^2} \langle\rho(r)\rangle^2 C(r). \end{equation} $C$, therefore, represents the factor by which a shell's emissivity is enhanced over the single-phase case. This formalism can be invoked to describe a true multiphase medium under arbitrary constraints by choosing an appropriate function $C(r)$. Nagai, Evrard, \& Sulkanen (1999) look at such distributions in more detail, investigating the observable consequences of applying isobaric multiphase models. For each cluster we calculate the mass-weighted mean $\bar{C}$ over all shells as an approximation to the overall bias on its total luminosity and inferred core density $\rho_0^2$. These values, averaged again over all clusters within a cosmological model, are as follows: $\bar{C}_o = 1.40$, $\bar{C}_s = 1.32$, $\bar{C}_{\Lambda} = 1.29$, and $\bar{C}_{\tau} = 1.38$. The mean values of $M_{\beta}/M_{\mathrm{true}}$ for the four cosmologies are 1.34 (OCDM), 1.13 (SCDM), 1.11 ($\Lambda$CDM), and 1.14 ($\tau$CDM). The mass error distributions for the four cosmologies are mutually consistent, so we combine them to determine a mean ICM bias for the entire ensemble of $1.18 \pm 0.02$. We also calculate the ensemble mean of $\bar{C}$, arriving at $1.34 \pm 0.02$. Since the total luminosity is proportional to the square of the central gas density, we expect an overall bias in our ICM mass estimates of $\bar{C}^{1/2}$, or $1.16 \pm 0.01$. The error bars given here are one standard deviation of the mean, so this is excellent agreement. We do not, however, find a strong correspondence between the level of clumping and the ICM mass error in individual clusters, however. Variations in the shape of the mass profile and large-scale asymmetries, which contribute a random component to the error, mask this relationship. It is difficult to judge what weighting scheme is most appropriate for this calculation, since the pivotal regions of the emission profile vary from cluster to cluster. The central points are very accurate and do much to constrain the fit, but are few in number; the outer regions have large error bars but many more data points. We began our anaysis with a mass-weighted clumping factor under the simple reasoning that it would greatly favor neither region. We also performed the exercise of weighting the shells by their luminosity (which gives more importance to the core) and volume (which emphasizes the outskirts). The resulting ensemble mean clumping factors do not vary greatly under the different weighting schemes: we found averages of $\bar{C} = 1.27$ for luminosity-weighted shells and $\bar{C} = 1.39$ for volume-weighted shells. We also attempt to identify correlations of mass error with large-scale substructure signatures. A surprising result is that we find no correlation of $M_{\beta}/M_{\rm true}$ with centroid shift, which has been found to be a good indicator of a cluster's dynamical state (Mohr, Evrard, Fabricant, \& Geller 1995). Even those clusters with the highest centroid shifts don't show a significantly different gas mass bias from the rest of the sample. Apparently, azimuthally averaging the cluster profile compensates for mild asphericities very well. We also looked for correlations of mass error with various bulk properties of the ICM such as mass, emission-weighted temperature, and luminosity, finding none. There is evidence for a weak correlation of mass error with the model parameters $\beta$ and $r_c$, but this can be entirely attributed to the tendency for strongly bimodal clusters to have profiles which are not well fit by the model. Although our clusters typically have steeper profiles than are observed in reality (c.f.), the lack of shape dependence indicates that this is probably not a problem. In none of the above tests were there discernible differences among the four cosmologies. In fact, we found only one ICM property to be correlated with $\delta M$. Defining a ``secondary peak'' as a local maximum in the surface brightness at least 1\% of the global maximum, we find that the most important property is the existence of secondary peaks within $r_{500}$. We attempted to further quantify the degree of asymmetry caused by subclumping, but the mass error turned out to be uncorrelated with the number and strength of the subpeaks. We therefore divide our ensemble into two subsets on this basis, hereafter referred to as ``regular'' and ``bimodal'' clusters for ease of language. Note that the subset of ``regular'' clusters contains examples with large centroid shifts or asymmetries, and the ``bimodal'' subset includes a few trimodal clusters as well. The two subsets comprise 62\% and 38\% of the total, respectively. Figure~2 displays a histogram of ICM mass errors for all the images in our ensemble. The shaded region corresponds to the regular clusters, while the remaining subset represents the bimodal sample. The bimodal population has a distribution of mass errors with mean $\delta M = 0.188$ ($\delta M = 1 - M_{\beta}/M_{\rm true}$) and standard deviation $\sigma = 0.084$. The remaining population has a distribution with $\bar{\delta M} = 0.093$ and $\sigma = 0.041$. Applying the K--S test, we find that the two populations are drawn from different distributions with very high confidence (greater than 5$\sigma$) and that the subset of non-bimodal images has an error distribution consistent with the Gaussian form. In applying the K--S tests we worked with a subset of the data consisting of just one image from each simulation, in order to maintain statistical independence. The means and standard deviations just quoted, however, reflect the distribution of the entire image ensemble. The smaller samples in each category have distributions entirely consistent with the complete sample, but we feel that the stated values better reflect the size of real uncertainties by taking advantage of multiple projections. Within our sample there are also 17 (out of 144) images which contain a major subclump in line-of-sight with the primary cluster and indistinguishable as a secondary peak. (A major subclump is one which produces a secondary peak when the cluster is viewed from a different angle.) These images, which occur in both the regular and bimodal populations, have a distribution in $\delta M$ consistent with the remaining set of bimodal images. This is reasonable given the similarity of their three-dimensional physical structures. Their partial membership in the set of regular images, however, does not change the shape or mean of that distribution significantly. Such occurences seem to be sufficiently uncommon that our heuristic classification system remains useful. Since strongly bimodal clusters are more uncommon than our simulations indicate, real samples will probably suffer even less from this contamination. \section{Conclusions} Assuming a density profile which is uniform at a given radius introduces a significant bias into measurements of the ICM mass, because it ignores the presence of luminosity enhancements from overdense regions. Our simulations indicate that these small-scale fluctuations produce a mean overestimate of $\sim 10\%$ when we confine our analysis to regular clusters. The advent of spatially resolved X--ray spectral imaging should allow us to test for irregularity in ICM structures and constrain the level of these fluctuations with direct observations; in the meantime it seems prudent to apply a correction of this scale to current measurements, as is done in MME. The relationship between these fluctuations and the cluster accretion history remains an open question, but this analysis suggests that there is no strong connection. Applying our spherically symmetric model to a cluster's azimuthally averaged surface brightness profile returns reasonably accurate ICM mass measurements even in clusters exhibiting significant asymmetries, so long as there are no significant secondary peaks in the image. The assumption of spherical symmetry, while formally invalid in most cases, seems to be borne out in practice in that the process of azimuthal averaging returns a mean density profile which is unbiased by the presence of moderate substructure. Such deviations from spherical symmetry contribute a random error of $\sim 5\%$ to ICM mass measurments under our method. The subset of bimodal (or multi-peaked) clusters has a much larger bias and dispersion, and is perhaps best excluded from population studies of the ICM. \acknowledgments This research was supported by NASA grants NAG5-2790, NAG5-3401, and NAG5-7108, as well as NSF grant AST-9803199. JJM is supported through Chandra Fellowship grant PF8-1003, awarded through the Chandra Science Center. The Chandra Science Center is operated by the Smithsonian Astrophysical Observatory for NASA under contract NAS8-39073.
\section{$\kat{smooth affine geometry}~\text{\bf @}_{\text{\bf n}}$.} \begin{paragraaf} Let $\kat{cat}$ be a category of associative $\mathbb{C}$-algebras with unit. An algebra $A \in Ob(\kat{cat})$ is said to be {\it $\kat{cat}$-smooth} iff for every {\it test-object} $(B,I)$ in $\kat{cat}$ (that is, $B \in Ob(\kat{cat})$, $I \triangleleft B$ a nilpotent ideal such that $\tfrac{B}{I} \in Ob(\kat{cat})$) and every $A \rTo^{\phi} \tfrac{B}{I} \in Mor(\kat{cat})$, there exists a {\it lifting morphism} $A \rTo^{\tilde{\phi}} B \in Mor(\kat{cat})$ making the diagram below commutative \[ \begin{diagram} B & \rOnto & \dfrac{B}{I} \\ & \luDotsto_{ \exists \tilde{\phi}} & \uTo_{\phi} \\ & & A \end{diagram} \] When $\kat{cat} = \kat{commalg}$, the category of all commutative $\mathbb{C}$-algebras, we recover Grothendieck's formulation of smooth (regular) commutative algebras. For this reason we call a $\kat{commalg}$-smooth algebra {\it g-smooth}. {\par \noindent} When $\kat{cat} = \kat{alg}$, the category of all $\mathbb{C}$-algebras, we recover Quillen's notion of {\it quasi-free} or {\it formally smooth} algebras. For this reason we call a $\kat{alg}$-smooth algebra {\it q-smooth}. {\par \noindent} Usually, we will assume that $\kat{cat}$-smooth algebras are {\it affine} algebras in $\kat{cat}$. {\par \noindent} Note however, that a commutative q-smooth algebra need not be g-smooth. For example, consider the polynomial algebra $\mathbb{C}[x_1,\hdots,x_d]$ and the $4$-dimensional noncommutative local algebra \[ B = \dfrac{\mathbb{C} \langle x,y \rangle}{(x^2,y^2,xy+yx)} = \mathbb{C} \oplus \mathbb{C} x \oplus \mathbb{C} y \oplus \mathbb{C} xy \] Consider the one-dimensional nilpotent ideal $I = \mathbb{C} (xy-yx)$ of $B$, then the $3$-dimensional quotient $\tfrac{B}{I}$ is commutative and we have a morphism $\mathbb{C}[x_1,\hdots,x_d] \rTo^{\phi} \tfrac{B}{I}$ by $x_1 \mapsto x, x_2 \mapsto y$ and $x_i \mapsto 0$ for $i \geq 2$. This morphism admits no lift to $B$ as for any potential lift $[\tilde{\phi}(x),\tilde{\phi}(y)] \not= 0$ in $B$. Therefore, $\mathbb{C}[x_1,\hdots,x_d]$ can only be q-smooth if $d=1$. \end{paragraaf} \begin{paragraaf} In fact, W. Schelter proved in \cite{Schelter} that $A$ is q-smooth if and only if the $A$-bimodule $\Omega^1_A = Ker~A \otimes A \rTo^{m_A} A$ is projective. If $A$ is an affine commutative $q$-smooth algebra, $A$ must be the coordinate ring of a finite set of points or of a smooth affine curve. Apart from semisimple algebras and some algebras which are finite modules over commutative $q$-smooth algebras, noncommutative q-smooth algebras are rather exotic objects such as free algebras and algebras arising from universal constructions (some of which we will encounter below). {\par \noindent} A fairly innocent class of q-smooth algebras are the {\it path algebras}. Let $Q$ be a {\it quiver}, that is a directed graph on a finite set $Q_v = \{ v_i,\hdots,v_k \}$ of vertices, having a finite set $Q_a = \{ a_1, \hdots, a_l \}$ of arrows. The path algebra $\mathbb{C}~Q$ has as $\mathbb{C}$-basis the oriented paths in $Q$ (including those of length zero corresponding to the vertices $v_i$) and multiplication induced by concatenation, that is, $1 = v_1 + \hdots + v_k$ is a decomposition into orthogonal idempotents and we have \begin{itemize} \item{$v_j.a$ is zero unless $\underset{v_j}{\bullet} \lTo^a \bullet$, } \item{$a.v_i$ is zero unless $\bullet \lTo^a \underset{v_i}{\bullet}$, } \item{$a_i.a_j$ is zero unless $\bullet \lTo^{a_i} \bullet \lTo^{a_j} \bullet$.} \end{itemize} To prove that $\mathbb{C}~Q$ is q-smooth, take a test-object $(B,I)$ in $\kat{alg}$ and an algebra map $\mathbb{C}~Q \rTo^{\phi} \tfrac{B}{I}$. The decomposition $1 = \phi(v_1) + \hdots + \phi(v_k)$ into orthogonal idempotents can be lifted modulo the nilpotent ideal to a decomposition $1 = \tilde{\phi}(v_1) + \hdots + \tilde{\phi}(v_k)$ into orthogonal idempotents. But then, taking for every arrow \[ \underset{v_j}{\bullet} \lTo^a \underset{v_i}{\bullet} \quad \quad \tilde{\phi}(a) \in \tilde{\phi}(v_j) (\phi(a) + I) \tilde{\phi}(v_i) \] gives a required lift. \end{paragraaf} \begin{paragraaf} J. Cuntz and D. Quillen argue in \cite{CQ1,CQ2} that q-smooth algebras behave (for example with respect to deRham cohomology) as commutative affine smooth algebras. In \cite[\S 9]{Kont1} M. Kontsevich gave a (somewhat cryptic) sketch how one might go about to develop an affine noncommutative geometry for q-smooth algebras. He suggests that one should {\it approximate} the noncommutative geometry of $A$ {\it at level $n$} by the {\it representation space} $\wis{rep}_n~A$ which is the affine scheme representing the functor \[ \kat{commalg} \rTo^{Hom_{\kat{alg}}(A,M_n(-))} \kat{sets} . \] When $A$ is q-smooth, it follows from the lifting property for q- and g-smooth algebras that $\wis{rep}_n~A$ is a smooth affine variety. We will denote this {\it approximation at level $n$} by \[ \wis{space}~A \text{\bf @}_{\text{\bf n}} = \wis{rep}_n~A . \] Recently, Kontsevich and A. Rosenberg \cite{KontRos} made this proposal more explicit. They argue that in order to extend a structure $\wis{struct}$ in commutative geometry to the noncommutative affine smooth variety $\wis{spec}~A$ we must be able to define it at every level \[ \wis{struct}(\wis{space}~A) \Rightarrow \ \forall n \ : \ \wis{struct} \text{\bf @}_{\text{\bf n}} = \wis{struct}(\wis{rep}_n~A) \] That is, a noncommutative structure of some kind on $\wis{spec}~A$ should induce analogous commutative structures on all the representation spaces $\wis{rep}_n~A$. These structures $\wis{struct}$ can either be \begin{itemize} \item{{\it classical }, that is, ordinary commutative gadgets such as functions, vector fields and so on (see \cite[\S 1.3.2]{KontRos} for more examples), or } \item{{\it non-classical}, that is, new structures on commutative schemes coming from noncommutative algebra such as the {\it formal structures} of Kapranov \cite{Kapranov} to be defined below.} \end{itemize} \end{paragraaf} \begin{paragraaf} In this talk i want to indicate how one can extend this approximation proposal from q-smooth algebras to arbitrary algebras and obtain in this way a rich \wis{affine geometry} $\text{\bf @}_{\text{\bf n}}$. Then i will briefly indicate how one can construct global objects at level $n$ and how one might build a noncommutative geometry from these approximate objects. \end{paragraaf} \section{$\kat{affine geometry}~\text{\bf @}_{\text{\bf n}}$.} \setcounter{Prob}{0} \begin{paragraaf} First we will specify $\kat{alg} \text{\bf @}_{\text{\bf n}}$, that is, the algebras that are level $n$ approximations of associative algebras. A {\it trace map} on an associative $\mathbb{C}$-algebra $A$ is a $\mathbb{C}$-linear map $A \rTo^{tr} A$ such that for all $a,b \in A$ we have $tr(ab) = tr(ba)$, $tr(a)b = b tr(a)$ and $tr(tr(a)b) = tr(a)tr(b)$. Algebras with trace are the objects of a category $\kat{alg}^{\kat{tr}}$ with morphisms the trace preserving $\mathbb{C}$-algebra maps. The forgetful functor $\kat{alg}^{\kat{tr}} \rTo \kat{alg}$ has a left adjoint \[ \kat{alg} \rTo^{\tau} \kat{alg}^{\kat{tr}} \] that is, given an algebra $A$ we can construct an algebra $A^{\tau}$ with trace in a universal way by adding formally the traces. {\par \noindent} Fix a number $n$ and express $\prod_{i=1}^n (t-\lambda_i)$ as a polynomial in $t$ with coefficients polynomials in the Newton functions $\nu_k = \sum_{i=1}^n \lambda_i^k$. Replacing $\nu_k$ by $tr(x^k)$ we get a formal {\it Cayley-Hamilton polynomial of degree $n$} : $\chi^{(n)}_x(t)$. Let $A$ be an algebra with trace $tr$, we say that $A$ is a {\it Cayley-Hamilton algebra of degree $n$} if $tr(1) = n$ and $\chi^{(n)}_a(a) = 0$ in $A$ for all $a \in A$. Cayley-Hamilton algebras of degree $n$ are the objects of a category $\kat{alg}^{\kat{tr}}_{\kat{n}}$ with trace preserving algebra maps as morphisms. There is a natural functor \[ \kat{alg}^{\kat{tr}} \rTo^n \kat{alg}^{\kat{tr}}_{\kat{n}} \] by sending an algebra $A$ with trace $tr$ to the quotient $A_n$ by the twosided ideal generated by the elements $tr(1)-n$ and $\chi_a^{(n)}(a)$ for all $a \in A$. Starting with an arbitrary algebra $A$ we propose to take $A~\text{\bf @}_{\text{\bf n}} = A^{\tau}_n$. That is, \[ \begin{diagram} \kat{alg} & & \\ \dTo^{\tau} \uInto & \rdDotsto^{\text{\bf @}_{\text{\bf n}}} & \\ \kat{alg}^{\kat{tr}} & \rTo_n & \kat{alg}^{\kat{tr}}_{\kat{n}} = \kat{alg} \text{\bf @}_{\text{\bf n}} \end{diagram} \quad \quad \quad \begin{diagram} A & & \\ \dTo^{\tau} & \rdDotsto^{\text{\bf @}_{\text{\bf n}}} & \\ A^{\tau} & \rTo_n & A^{\tau}_n = A \text{\bf @}_{\text{\bf n}} \end{diagram} \] Of course it may happen that $A~\text{\bf @}_{\text{\bf n}} = 0$ (for example if $A = A_m(\mathbb{C})$ the $m$-th Weyl algebra). For more details on algebras with trace and their properties (some of which we will recall below) we refer to the paper \cite{ProcCH} of C. Procesi. \end{paragraaf} \begin{paragraaf} For $A$ an arbitrary associative $\mathbb{C}$-algebra we define $\wis{space}~A \text{\bf @}_{\text{\bf n}}$ as before to be the representation space $\wis{rep}_n~A$ which is the affine scheme representing the functor \[ \kat{commalg} \rTo^{Hom_{\kat{alg}}(A,M_n(-))} \kat{sets} \] That is, there is a universal representation $A \rTo^{j_A} M_n(\mathbb{C}[\wis{rep}_n~A])$ such that for every $\mathbb{C}$-algebra map $A \rTo^{\phi} M_n(B)$ with $B$ a commutative algebra, we have a uniquely determined morphism $\mathbb{C}[\wis{rep}_n~A] \rTo^{\psi} B$ such that the diagram below is commutative \[ \begin{diagram} A & & \rTo^{\phi} & & M_n(B) \\ & \rdTo_{j_A} & & \ruTo_{M_n(\psi)} \\ & & M_n(\mathbb{C}[\wis{rep}_n~A]) & & \end{diagram} \] $GL_n$ acts by conjugation on $M_n(B)$ in a functorial way making $\wis{rep}_n~A$ into an affine $GL_n$-scheme and so there are actions of $GL_n$ by automorphisms on $\mathbb{C}[\wis{rep}_n~A]$ and also on $M_n(\mathbb{C}[\wis{rep}_n~A]) = \mathbb{C}[\wis{rep}_n~A] \otimes M_n(\mathbb{C})$ (tensor product action). The image of $A$ under $j_A$ is contained in the ring of invariants $M_n(\mathbb{C}[\wis{rep}_n~A])^{GL_n}$ which is the ring of $GL_n$-equivariant maps $\wis{rep}_n~A \rTo M_n(\mathbb{C})$ with algebra structure induced by the one on the target space. The main results on this $GL_n$-setting are due to Procesi \cite{ProcCH} and assert that we have a commutative functorial diagram \[ \begin{diagram} A & \rTo^{j_A} & M_n(\mathbb{C}[\wis{rep}_n~A]) \\ \dTo^{\text{\bf @}_{\text{\bf n}}} & \rdTo^{j_A} & \uInto \\ A \text{\bf @}_{\text{\bf n}} & \cong & M_n(\mathbb{C}[\wis{rep}_n~A])^{GL_n} \end{diagram} \] and we recover the algebra with trace $A \text{\bf @}_{\text{\bf n}} \in Ob(\kat{alg}^{\kat{tr}}_{\kat{n}})$ from the $GL_n$-affine scheme $\wis{space}~A~\text{\bf @}_{\text{\bf n}} = \wis{rep}_n~A$. Clearly, this scheme is in general not smooth nor even reduced. Procesi studied $\kat{alg}^{\kat{tr}}_{\kat{n}}$-smooth algebras in \cite{ProcCH} which we will call p-smooth from now on. In fact, he proved that $A \text{\bf @}_{\text{\bf n}}$ is p-smooth if and only if $\wis{rep}_n~A$ is a smooth $GL_n$-scheme. In particular, we have \[ A \text{ q-smooth \ } \quad \Rightarrow \ \forall n \ : \quad A \text{\bf @}_{\text{\bf n}} \text{ p-smooth} \] consistent with our approximation strategy. \end{paragraaf} \begin{paragraaf} We now come to classical structures on $\wis{space}~A~\text{\bf @}_{\text{\bf n}} = \wis{rep}_n~A$. Let us consider functions as proposed in \cite[\S 1.3.2]{KontRos}. Let $a \in A$ and $A \rTo^x M_n(\mathbb{C})$ a geometric point $x \in \wis{rep}_n~A$, then we can take the ordinary trace $tr(x(a))$. This gives us a linear map \[ \dfrac{A}{[A,A]} \rTo \mathbb{C}[\wis{rep}_n~A] \] Kontsevich and Rosenberg propose to take for $\wis{funct}~A~\text{\bf @}_{\text{\bf n}}$ the subalgebra generated by the image of this map. Observe that these functions are $GL_n$-invariant and recall that Procesi proved that the ring of invariant polynomial functions \[ \mathbb{C}[\wis{rep}_n~A]^{GL_n} = tr~A~\text{\bf @}_{\text{\bf n}} \] where $tr$ is the trace map on $A~\text{\bf @}_{\text{\bf n}}$. That is, we have \[ \wis{funct}~A~\text{\bf @}_{\text{\bf n}} = tr~A~\text{\bf @}_{\text{\bf n}} . \] This ring has the following representation theoretic interpretation. The $GL_n$-orbits in $\wis{rep}_n~A$ correspond to isomorphism classes of $n$-dimensional representations of $A$ and the {\it closed} orbits correspond to $n$-dimensional {\it semi-simple} representations. Invariant theory tells us that the closed orbits are parametrized by the maximal ideals of the ring of polynomial invariants. The inclusion $tr~A~\text{\bf @}_{\text{\bf n}} \rInto \mathbb{C}[\wis{rep}_n~A]$ induces a morphism of schemes \[ \wis{rep}_n~A \rOnto^{\pi} \wis{fac}_n~A \] and sends an $n$-dimensional representation of $A$ to the isomorphism class of the semi-simple $n$-dimensional representation which is the direct sum of the Jordan-H\"older factors. Moreover, the algebra with trace $A~\text{\bf @}_{\text{\bf n}}$ is a finitely generated module over the subalgebra $tr~A~\text{\bf @}_{\text{\bf n}} = \mathbb{C}[\wis{fac}_n~A]$ and hence we can associate to $A~\text{\bf @}_{\text{\bf n}}$ its classical structure sheaf $\Oscr_{A~\text{\bf @}_{\text{\bf n}}}$ which is a sheaf of noncommutative algebras over $\wis{fac}_n~A$. {\par \noindent} We propose that the classical structures on $\wis{space}~A~\text{\bf @}_{\text{\bf n}}$ are given by $GL_n$-equivariant structures associated with the $GL_n$-invariant theoretic setting \[ \begin{diagram} & & \Oscr_{A \text{\bf @}_{\text{\bf n}}} \\ & & \dDots \\ \wis{rep}_n~A & \rOnto^{\pi} & \wis{fac}_n~A \end{diagram} \] All the classical structures proposed in \cite[\S 1.3.2]{KontRos} are of this general form. \end{paragraaf} \begin{paragraaf} Let us compute all of this in the case of path algebras. Consider the semisimple subalgebra $V=\underbrace{\mathbb{C} \times \hdots \times \mathbb{C}}_k$ generated by the vertex-idempotents $\{ v_1,\hdots,v_k \}$. Every $n$-dimensional representation of $V$ is semi-simple and determined by the multiplicities by which the factors occur. That is, we have a decomposition \[ \wis{rep}_n~V = \bigsqcup_{\sum a_i = n} GL_n / (GL_{a_1} \times \hdots \times GL_{a_k}) = \bigsqcup_{\alpha} \wis{rep}_{\alpha}~V \] into homogeneous spaces where $\alpha$ runs over the {\it dimension vectors} $\alpha = (a_1,\hdots, a_k)$ such that $\sum_i a_i = n$. The inclusion $V \rInto \mathbb{C}~Q$ induces a map $\wis{rep}_n~\mathbb{C}~Q \rTo^{\psi} \wis{rep}_n~V$ and we have the decomposition of $\wis{rep}_n~\mathbb{C}~Q$ into associated fiber bundles \[ \psi^{-1}(\wis{rep}_{\alpha}~V) = GL_n \times^{GL(\alpha)} \wis{rep}_{\alpha}~Q . \] Here, $GL(\alpha) = GL_{a_1} \times \hdots \times GL_{a_k}$ embedded along the diagonal in $GL_n$ and $\wis{rep}_{\alpha}~Q$ is the affine space of $\alpha$-dimensional representations of the quiver $Q$. That is, \[ \wis{rep}_{\alpha}~Q = \bigoplus_{\underset{v_j}{\bullet} \lTo^a \underset{v_i}{\bullet}} M_{a_j \times a_i}(\mathbb{C}) \] and $GL(\alpha)$ acts on this space via base-change in the vertex-spaces. That is, $\wis{rep}_n~\mathbb{C}~Q$ is the disjoint union of smooth affine components depending on the dimension vectors $\alpha = (a_1, \hdots,a_k)$ such that $\sum a_i = n$. This decomposition also translates to the algebra at level $n$ \[ \mathbb{C}~Q~\text{\bf @}_{\text{\bf n}} = \bigoplus_{\alpha}~T^Q_{\alpha} \] where the trace map on the component $T^Q_{\alpha}$ is determined by $tr(v_i) = a_i$. For the functions at level $n$ we have the decomposition \[ tr~\mathbb{C}~Q~\text{\bf @}_{\text{\bf n}} = \bigoplus_{\alpha}~t^Q_{\alpha} \quad \text{where} \quad t^Q_{\alpha} = tr~T^Q_{\alpha} \] and is the ring of $GL(\alpha)$-invariants $\mathbb{C}[\wis{rep}_{\alpha}~Q]^{GL(\alpha)}$. In \cite{LBProc} it was proved that this ring of invariants is generated by traces along oriented cycles in the quiver $Q$ of length $\leq n^2$. That is, for fixed $\alpha$ replace any arrow $\underset{v_j}{\bullet} \lTo^a \underset{v_i}{\bullet}$ by the $a_j \times a_i$ matrix of coordinate functions on $\wis{rep}_{\alpha}~Q$ corresponding to the arrow $a$. Then, multiplying these matrices along a cycle produces a square matrix whose trace is a $GL(\alpha)$-invariant. {\par \noindent} In \cite{LBProc} we also gave a computational description of the algebra with trace $T^Q_{\alpha}$. It has a block decomposition \[ T^Q_{\alpha} = \begin{bmatrix} T_{11} & \hdots & T_{1k} \\ \vdots & & \vdots \\ T_{k1} & \hdots & T_{kk} \end{bmatrix} \] where $T_{ij}$ is the $t^Q_{\alpha}$-module spanned by the paths in the quiver $Q$ starting at vertex $v_j$ and ending in vertex $v_i$ (again, the length of the required paths can be bounded by $n^2$). {\par \noindent} Observe that $t^Q_{\alpha}$ is positively graded and e will denote the $\mathfrak{m}$-adic completion with respect to the graded maximal ideal $\mathfrak{m} = \oplus_{i \geq 1} t^Q_{\alpha}(i)$ by $\hat{t}^Q_{\alpha}$. Analogously we denote $\hat{T}^Q_{\alpha} = T^Q_{\alpha} \otimes \hat{t}^Q_{\alpha}$. \end{paragraaf} \begin{paragraaf} In order for the approximation strategy for a q-smooth algebra $A$ to succeed we need to control the local structure of $A~\text{\bf @}_{\text{\bf n}}$ and $\wis{fac}_n~A$. In fact a much more general result holds. Let $A$ be an associative algebra and $x$ a geometric point of $\wis{fac}_n~A$. We say that $A~\text{\bf @}_{\text{\bf n}}$ is {\it locally smooth} in $x$ provided $\wis{rep}_n~A$ is smooth along the closed orbit determined by $x$. Observe that if $A$ is q-smooth (or even p-smooth), then $A~\text{\bf @}_{\text{\bf n}}$ is locally smooth in {\it all} $x$. {\par \noindent} The point $x$ determines an $n$-dimensional semi-simple representation $M_x$ of $A$, say \[ M_x = S_1^{\oplus m_1} \oplus \hdots \oplus S_k^{\oplus m_k} \] where $S_i$ is a simple $A$-representation of dimension $d_i$ occurring in $M$ with multiplicity $m_i$, that is $n = \sum_i d_im_i$. {\par \noindent} We associate a {\it local quiver} $Q(x)$ to the point $x$. It has $k$ vertices $v_i$ (the number of distinct simple components $S_i$ of $M_x$) and the number of arrows from $v_i$ to $v_j$ is given by \[ \#~\underset{v_j}{\bullet} \lTo \underset{v_i}{\bullet} = \text{dim}_{\mathbb{C}}~Ext^1_A(S_i,S_j) . \] Further, we define a {\it local dimension vector} $\alpha(x) = (m_1,\hdots,m_k)$ determined by the multiplicities of the simple components in $M_x$. {\par \noindent} Denote the completion of the stalk of the structure sheaf $\Oscr_{\wis{fac}_n~A}$ in $x$ by $\hat{\Oscr}_x$, then with notation as before we have an isomorphism of local algebras \[ \hat{\Oscr}_x \cong \hat{t}^{Q(x)}_{\alpha(x)} . \] This is an application of the {\it Luna slice theorem} in invariant theory, see \cite{LBsmooth}. Further, if $\hat{\Oscr}^{A~\text{\bf @}_{\text{\bf n}}}_x$ denotes the algebra $A~\text{\bf @}_{\text{\bf n}} \otimes \hat{\Oscr}_x$, then it is also proved in \cite{LBsmooth} that \[ \hat{\Oscr}^{A~\text{\bf @}_{\text{\bf n}}}_x \underset{Morita}{\sim} \hat{T}^{Q(x)}_{\alpha(x)} \] where $\underset{Morita}{\sim}$ stands for {\it Morita-equivalence}, that is, equivalence of the module categories. The precise form of the Morita equivalence is determined by the embedding of $GL(\alpha(x)) = Stab_{GL_n}(M_x) \rInto GL_n$. {\par \noindent} Hence, if $A~\text{\bf @}_{\text{\bf n}}$ is locally smooth in $x$, the \'etale local structure of $\Oscr_{A~\text{\bf @}_{\text{\bf n}}}$ and $\wis{fac}_n~A$ near $x$ is fully determined by the combinatorial data $Q(x)$ and $\alpha(x)$. A lot more can be said about these combinatorics. For example, $(Q(x),\alpha(x))$ determines the local quiver-data in neighbouring points, one can compute the dimensions of the strata in $\wis{fac}_n~A$ consisting of points with the same local quiver-data and given the local dimension of $\wis{fac}_n~A$ in $x$ one can even classify all the possible quiver-data. For some of these we refer to \cite{LBetale} and \cite{LBsmooth}. {\par \noindent} In particular, this applies to a q-smooth algebra $A$ in any point $x$ of $\wis{fac}_n~A$. That is, the the approximation $A~\text{\bf @}_{\text{\bf n}}$ of {\it any} q-smooth algebra $A$ is locally (in the \'etale topology) isomorphic to the approximation of the subclass of path algebras providing a handle on the exotic class of q-smooth affine algebras\footnote{A. Rosenberg suggests it would be nice to obtain the quiver-data from the non-commutative space as defined in \cite[\S 2]{KontRos}. I will look into this.} \end{paragraaf} \begin{paragraaf} Now, we turn to non-classical structures on $\wis{space}~A~\text{\bf @}_{\text{\bf n}}$ such as Kapranov's {\it formal noncommutative structure}, see \cite{Kapranov}. Let $R$ be an associative $\mathbb{C}$-algebra, $R^{Lie}$ its Lie structure and $R^{Lie}_m$ the subspace spanned by the expressions $[r_1,[r_2,\hdots,[r_{m-1},r_m]\hdots ]$ containing $m-1$ instances of Lie brackets. The {\it commutator filtration} of $R$ is the (increasing) filtration by ideals $(F^d~R)_{d \in \mathbb{Z}}$ with $F^d~R = R$ for $d \in \mathbb{N}$ and \[ F^{-d}~R = \underset{m}{\sum}~\underset{i_1 + \hdots + i_m = d}{\sum} R R^{Lie}_{i_1} R \hdots R R^{Lie}_{i_m} R \] The associated graded $gr_F~R$ is a (negatively) graded commutative Poisson algebra with part of degree zero $R_{ab} = \tfrac{R}{[R,R]}$. {\par \noindent} Denote with $\kat{nil}_d$ the category of associative $\mathbb{C}$-algebras $R$ such that $F^{-d-1}R = 0$ (note that an algebra map is filtration preserving). Kapranov studied in \cite{Kapranov} $\kat{nil}_d$-smooth algebras which we will call $k_d$-smooth algebras from now on. {\par \noindent} Kapranov proves \cite[Thm 1.6.1]{Kapranov} that any affine commutative smooth algebra $C$ has a unique $k_d$-smooth {\it thickening} $R$ with $R_{ab} \simeq C$ (up to isomorphisms identical on $C$). The approach of \cite[\S 1]{Kapranov} may be compared to that of M. Artin in \cite{ArtinDef}. {\par \noindent} For $R \in Ob(\kat{nil}_d)$, Kapranov introduces a sheaf $\Oscr_R$ of noncommutative rings on the commutative scheme $X_{ab} = \wis{spec}~R_{ab}$. Observing that the commutator filtration for such $R$ is {\it Zariskian} as in \cite{LiFVO} the approach of \cite[\S 2]{Kapranov} may be compared to that of F. Van Oystaeyen in \cite[Chpt II]{FVObook}. {\par \noindent} If $X$ is an affine smooth commutative variety and $R_d$ the canonical $k_d$-smooth thickening of $\mathbb{C}[X]$, then the sheaf of noncommutative algebras on $X$ \[ \Oscr^{f} = \underset{\leftarrow}{\text{lim}}~\Oscr_{R_d} \] is Kapranov's noncommutative formal structure on $X$. {\par \noindent} If $A$ is q-smooth and affine, the representation space $\wis{rep}_n~A$ is a smooth affine variety and hence is equipped with such a formal noncommutative structure sheaf $\Oscr^f$. Part of the Kontsevich-Rosenberg proposal is that the level $n$ approximation $\wis{spec}~A~\text{\bf @}_{\text{\bf n}}$ should carry such a formal structure. \end{paragraaf} \begin{paragraaf} At first sight we face a serious problem as Kapranov's inductive construction of $k_d$-smooth thickenings of a commutative $C$ only works when $C$ is g-smooth (in that case one can define at each stage a {\it universal} central extension). Moreover, the construction of such a formal structure must be sufficiently functorial to suit our purposes. I do not know of such a structure for all commutative affine $\mathbb{C}$-algebras (or even only those having a $GL_n$-action). {\par \noindent} Fortunately, we only need it for coordinate rings of representation spaces $\mathbb{C}[\wis{rep}_n~A]$ and there we can apply some ringtheory of the early 70ties, in particular G. Bergman's coproduct theorems \cite{Bergman} (see also the lecture notes of A. Schofield \cite[Chp. 2]{Schofield} for more details). The starting point is that for every associative algebra $A$ the functor \[ \kat{alg} \rTo^{Hom_{\kat{alg}}(A,M_n(-))} \kat{sets} \] is {\it representable} in $\kat{alg}$. That is, there exists an associative $\mathbb{C}$-algebra $\sqrt[n]{A}$ such that there is a natural equivalence between the functors \[ Hom_{\kat{alg}}(A,M_n(-)) \underset{n.e.}{\sim} Hom_{\kat{alg}}(\sqrt[n]{A},-) . \] In other words, for every associative $\mathbb{C}$-algebra $B$, there is a functorial one-to-one correspondence between the sets \[ \begin{cases} \text{algebra maps} \quad A \rTo M_n(B) \\ \text{algebra maps} \quad \sqrt[n]{A} \rTo B \end{cases} \] To define $\sqrt[n]{A}$ consider the free algebra product $A \ast M_n(\mathbb{C})$ and consider the subalgebra \[ \sqrt[n]{A} = A \ast M_n(\mathbb{C})^{M_n(\mathbb{C})} = \{ p \in A \ast M_n(\mathbb{C}) \mid p.(1 \ast m) = (1 \ast m).p \ \forall m \in M_n(\mathbb{C}) \} \] Before we can prove the universal property of $\sqrt[n]{A}$ we need to recall a property that $M_n(\mathbb{C})$ shares with any Azumaya algebra : if $M_n(\mathbb{C}) \rTo^{\phi} R$ is an algebra morphism and if $R^{M_n(\mathbb{C})} = \{ r \in R \mid r.\phi(m) = \phi(m).r \ \forall m \in M_n(\mathbb{C}) \}$, then we have $R \simeq M_n(\mathbb{C}) \otimes_{\mathbb{C}} R^{M_n(\mathbb{C})}$. In particular, if we apply this to $R = A \ast M_n(\mathbb{C})$ and the canonical map $M_n(\mathbb{C}) \rTo^{\phi} A \ast M_n(\mathbb{C})$ where $\phi(m) = 1 \ast m$ we obtain that $M_n(\sqrt[n]{A}) = M_n(\mathbb{C}) \otimes_{\mathbb{C}} \sqrt[n]{A} = A \ast M_n(\mathbb{C})$. {\par \noindent} Hence, if $\sqrt[n]{A} \rTo^{f} B$ is an algebra map we can consider the composition \[ A \rTo^{id_A \ast 1} A \ast M_n(\mathbb{C}) \simeq M_n(\sqrt[n]{A}) \rTo^{M_n(f)} M_n(B) \] to obtain an algebra map $A \rTo M_n(B)$. Conversely, consider an algebra map $A \rTo^g M_n(B)$ and the canonical map $M_n(\mathbb{C}) \rTo^i M_n(B)$ which centralizes $B$ in $M_n(B)$. Then, by the universal property of free algebra products we have an algebra map $A \ast M_n(\mathbb{C}) \rTo^{g \ast i} M_n(B)$ and restricting to $\sqrt[n]{A}$ we see that this maps factors \[ \begin{diagram} A \ast M_n(\mathbb{C}) & \rTo^{g \ast i} & M_n(B) \\ \uInto & & \uInto \\ \sqrt[n]{A} & \rDotsto & B \end{diagram} \] and one verifies that these two operations are each others inverses. The algebra $\sqrt[n]{A}$ has other surprising properties. For example, no matter how bad $A$ is, $\sqrt[n]{A}$ is a {\it domain} having $\mathbb{C}^*$ as its group of invertible elements by \cite[Thm. 2.19]{Schofield}. Now, equip $\sqrt[n]{A}$ with the commutator filtration \[ \begin{diagram} \hdots & & -2 & & -1 & & 0 & & 1 & & \hdots \\ \hdots & \subset & F_{-2} & \subset & F_{-1} & \subset & \sqrt[n]{A} & = & \sqrt[n]{A} & = & \hdots \end{diagram} \] By the universal property of $\sqrt[n]{A}$ we see that $\tfrac{A}{F_{-d-1}} \in Ob(\kat{nil}_d)$ is the object representing the functor \[ \kat{nil}_d \rTo^{Hom_{\kat{alg}}(A,M_n(-))} \kat{sets} . \] In particular, as $\kat{nil}_0 = \kat{commalg}$ we deduce that \[ gr_0~\sqrt[n]{A} = \dfrac{\sqrt[n]{A}}{F_{-1}} = \dfrac{\sqrt[n]{A}}{[\sqrt[n]{A},\sqrt[n]{A}]} \simeq \mathbb{C}[\wis{rep}_n~A] \] as both algebras represent the same functor. The construction of the formal sheaf of noncommutative algebras is similar to that of \cite{Kapranov}. For fixed $d$, the induced filtration on the quotient $\sqrt[n]{A}_d = \tfrac{\sqrt[n]{A}}{F_{-d-1}}$ is Zariskian (even discrete). Hence, taking the {\it saturation} of a multiplicatively closed subset of the associated graded $gr~\sqrt[n]{A}$ is an Ore-set in $\sqrt[n]{A}_d$. Hence, we can construct a sheaf of algebras $\Oscr_{\sqrt[n]{A}_d}$ on $\wis{rep}_n~A$ as in \cite{FVObook}. The formal noncommutative structure sheaf on $\wis{rep}_n~A$ is then the inverse limit \[ \Oscr_{\sqrt[n]{A}}^f = \underset{\leftarrow}{lim}~\Oscr_{\sqrt[n]{A}_d} \] We now claim that the approximation at level $n$ of $\wis{spec}~A$ is given by the {\it vier-span} \[ \wis{spec}~A~\text{\bf @}_{\text{\bf n}} \quad = \quad \quad \begin{diagram} \Oscr_{\sqrt[n]{A}}^f & & \Oscr_{A \text{\bf @}_{\text{\bf n}}} \\ \dDots & & \dDots \\ \wis{rep}_n~A & \rOnto & \wis{fac}_n~A \end{diagram} \] There remains to prove that the formal structure $\Oscr_{\sqrt[n]{A}}^f$ defined above coincides with Kapranov's formal structure in case $A$ is q-smooth. We have seen that \[ M_n(\sqrt[n]{A}) \simeq A \ast M_n(\mathbb{C}) \] hence if $A$ is q-smooth, so is $\sqrt[n]{A}$ by the coproduct theorems, see for example \cite[Thm. 2.20]{Schofield} for a strong version. But then clearly $\sqrt[n]{A}_d$ is $k_d$-smooth for all $d$ (or use \cite[Prop. 1.4.6]{Kapranov}). By the unicity of $k_d$-smooth thickenings \cite[\S 1.6]{Kapranov} it is then immediate that Kapranov's formal structure $\Oscr^f$ on $\wis{rep}_n~A$ coincides with our $\Oscr_{\sqrt[n]{A}}^f$. {\par \noindent} Note also the perhaps surprising fact that for $\wis{rep}_n~A$ for $A$ a q-smooth algebra one does not need the perturbative approach of Kapranov to describe a formal neighborhood of $\wis{rep}_n~A$ into a noncommutative smooth space. By the above, $\wis{rep}_n~A$ is a closed subvariety of the noncommutative smooth space $\wis{spec}~\sqrt[n]{A}$. {\par \noindent} Our description also clarifies the importance of this formal noncommutative structure. In order to understand the (commutative) scheme structure of $\wis{rep}_n~A$ one needs to consider also representation $A \rTo M_n(F)$ where $F$ is a finite dimensional commutative algebra (rather than restrict to $\mathbb{C}$). Similarly, if one wants to understand the noncommutative scheme structure of $\wis{rep}_n~A$ one ought to consider representations $A \rTo M_n(F)$ where $F$ is a finite dimensional non-commutative algebra. If $F$ is {\it basic} (that is, all simple $F$-representations are one-dimensional over $\mathbb{C}$), then $F \in Ob(\kat{nil}_d)$ for some $d$, and then this representation is controlled by the formal structure. In general, any finite dimensional noncommutative algebra $F$ is Morita equivalent to a basic algebra, so certainly if we vary $n$ all the formal structures $\Oscr_{\sqrt[n]{A}}$ control representation $A \rTo M_m(F)$. Expressed differently, one can view the formal structures $\Oscr_{\sqrt[n]{A}}$ as sewing-machines to stitch the different $\wis{rep}_n~A$ together. \end{paragraaf} \begin{paragraaf} I cannot resist the temptation to add an infinite family of formal structures on $\wis{space}~A~\text{\bf @}_{\text{\bf n}}$. If at layer $1$ the commutator filtration is defined using the Lie bracket $[r_1,r_2]$, one can similarly define at layer $m$ the $m$-commutator filtration based on the expressions \[ S_{2m}(r_1,r_2,\hdots,r_{2m}) = \sum_{\sigma \in S_{2m}} (-1)^{sgn(\sigma)} r_{\sigma(1)} r_{\sigma(2)} \hdots r_{\sigma(2m)} . \] The associated graded algebra with respect to the $m$-commutator filtration on $\sqrt[n]{A}$ is no longer commutative but is a {\it polynomial identity algebra} of degree $m$ (that is, it basically lives at level $\text{\bf @}_{\text{\bf m}}$). As these algebras are close to commutative algebras, they have plenty of Ore sets and again the Zariskian argument provides us with new formal structure sheaves. \end{paragraaf} \section{$\kat{geometry}~\text{\bf @}_{\text{\bf n}}$.} \setcounter{Prob}{0} \begin{paragraaf} In this section we will briefly indicate how one can define global objects in the approximate geometry $\text{\bf @}_{\text{\bf n}}$ and afterwards how one can put approximate objects together to define a noncommutative geometry. More details will have to await another occasion. {\par \noindent} The strategy to define a $\kat{geometry}~\text{\bf @}_{\text{\bf n}}$ is simple : use $GL_n$-equivariant (commutative) geometry as inspiration and try to define all concepts in terms of $\kat{alg}~\text{\bf @}_{\text{\bf n}}$. The latter is essential in order to define formal structures as we have seen above. {\par \noindent} In view of our local combinatorial description of p-smooth algebras, it is clear that the natural topology of $\kat{affine geometry}~\text{\bf @}_{\text{\bf n}}$ is the \'etale topology. Therefore, in order to define global objects we choose for the approach via {\it algebraic spaces}\footnote{V. Hinich suggests it might be more natural to extend the notion of algebraic stack in order to maintain compatibility with the Kontsevich-Rosenberg proposal of noncommutative spaces in \cite[\S 2]{KontRos}. I agree and will try to work this out.} as developed by M. Artin \cite{ArtinAS}. Algebraic spaces are defined by \'etale equivalence relations, so we need \begin{itemize} \item{a product in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$, and} \item{\'etale morphisms in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$.} \end{itemize} As a product, the tensor-product $\otimes_{\mathbb{C}}$ is not suitable as it takes us out of $\kat{alg}~\text{\bf @}_{\text{\bf n}}$ (think of tensorproducts of matrixrings). However, free algebra products provide us with a suitable definition \[ A \underset{n}{\boxtimes} B \overset{def}{=} A \ast B~\text{\bf @}_{\text{\bf n}} \] In order to define \'etale morphisms in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$ we extend the notions of formally \'etale (resp. formally unramified, formally smooth) from $\kat{commalg}$ verbatim. That is, let $A$ be an associative algebra, then a morphism $A~\text{\bf @}_{\text{\bf n}} \rTo^{\phi} B$ in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$ is said to be {\it formally \'etale} iff for every test-object $(T,I)$ in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$, we have a unique lift, where all the morphisms below are $A~\text{\bf @}_{\text{\bf n}}$-algebra maps \[ \begin{diagram} T & \rOnto & \dfrac{T}{I} \\ \uTo & \luDotsto_{\exists ! \tilde{\psi}} & \uTo^{\psi} \\ A~\text{\bf @}_{\text{\bf n}} & \rTo_{\phi} & B \end{diagram} \] If one replaces unicity by existence $\phi$ is said to be {\it formally smooth} and if unicity is replaced by the existence of at most one lift, $\phi$ is said to be {\it formally unramified}. These notions also have a geometric interpretation : consider an algebra map $A \rTo^{\phi} B$, then $A~\text{\bf @}_{\text{\bf n}} \rTo^{\phi~\text{\bf @}_{\text{\bf n}}} B~\text{\bf @}_{\text{\bf n}}$ is \'etale if and only if the induced morphism $\wis{rep}_n~B \rTo \wis{rep}_n~A$ is an \'etale morphism of commutative schemes. {\par \noindent} In order to show that this {\it \'etale topology} on $\wis{spec}~A~\text{\bf @}_{\text{\bf n}}$ is rich enough, let us give tow classes of \'etale maps. The first is classical, coming from commutative \'etale maps. Consider the diagram \[ \begin{diagram} A~\text{\bf @}_{\text{\bf n}} & \rTo^{id \otimes f} & A~\text{\bf @}_{\text{\bf n}} \otimes_{tr~A~\text{\bf @}_{\text{\bf n}}} S \\ \uInto & & \uTo \\ tr~A~\text{\bf @}_{\text{\bf n}} & \rTo^f & S \end{diagram} \] where $f$ is an \'etale morphism in $\kat{commalg}$, then $id \otimes f$ is \'etale in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$. The second class is more exotic, though it is the natural substitute for localizations, {\it universal localizations} in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$. Let $\kat{projmod}~A$ denote the category of finitely generated (left) modules over an associative algebra $A$ and let $\Sigma$ be some class of maps in this category (that is some $A$-module morphisms between certain projective modules). In \cite[Chp. 4]{Schofield} it is shown that there exists an algebra map $A \rTo^{j_{\Sigma}} A_{\Sigma}$ with the universal property that the maps $A_{\Sigma} \otimes_A \sigma$ have an inverse for all $\sigma \in \Sigma$. Using the above geometric characterization of \'etale maps it follows that the induced maps \[ A~\text{\bf @}_{\text{\bf n}} \rTo^{j_{\Sigma}~\text{\bf @}_{\text{\bf n}}} A_{\Sigma}~\text{\bf @}_{\text{\bf n}} \] are \'etale in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$. In particular, it may happen that a finite dimensional algebra has hugely infinite dimensional \'etale extensions in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$. For example consider the $m+3$-dimensional algebra \[ A = \begin{bmatrix} \mathbb{C} & V \\ 0 & \mathbb{C} \end{bmatrix} \] where $V$ is a $\mathbb{C}$-vectorspace of dimension $m+1$. Then, there is a universal localization of $A$ isomorphic to $M_2(\mathbb{C} \langle x_1,\hdots,x_m \rangle)$. We refer to the book \cite{Schofield} of A. Schofield for more details. {\par \noindent} Using these \'etale maps and the product $\underset{n}{\boxtimes}$ one can define $\kat{algebraic spaces}~\text{\bf @}_{\text{\bf n}}$ as in \cite{ArtinAS}. \end{paragraaf} \begin{paragraaf} It may be interesting to reconsider the notion of {\it orbit topos} introduced by R.W. Thomason in \cite{Thomason} in this setting. For example, one wonders how much of the orbit topos of $\wis{rep}_n~A$ can be described using algebraic spaces in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$. {\par \noindent} In general it is not true that any $GL_n$-equivariant technique can be performed in $\kat{algebraic spaces}~\text{\bf @}_{\text{\bf n}}$. A noteworthy exception is $GL_n$-equivariant desingularization. In \cite{LBetale} the following problem was handled. Let $X$ be a smooth (commutative) surface and $\Delta$ a central simple algebra in $\kat{alg}~\text{\bf @}_{\text{\bf n}}$ over the function field $\mathbb{C}(X)$. Recall that $\Delta$ is determined by \begin{itemize} \item{a divisor $D \rInto X$ and a list of its irreducible components $C$,} \item{the list of singular points $p_j$ of $D$, and} \item{for each branch $B_k$ of $D$ at $p_i$ a number $n_{i,k} \in \mathbb{Z}/n\mathbb{Z}$ such that $\sum_k n_{i,k} = 0$.} \end{itemize} One might ask whether for each $\Delta$ there exists a smooth object in $\kat{algebraic spaces}~\text{\bf @}_{\text{\bf n}}$ having $\Delta$ as its noncommutative function algebra. An idea might be to start with a {\it maximal order} $\mathcal{A}$ in $D$ over $X$, consider $\wis{rep}_n~\mathcal{A}$ and construct its $GL_n$-equivariant desingularization. In \cite[Chp. 6]{LBetale} it was shown that the above problem has a positive solution if and only if \[ \sum_{p_i \in C} \sum_{B_k \in C} n_{i,k} = 0 \quad \text{ for all irreducible components $C$ of $D$} \] In general, one can construct an object in $\kat{algebraic spaces}~\text{\bf @}_{\text{\bf n}}$ having $D$ as function algebra and having only a finite number of points where it is not locally smooth. All of these singularities are locally (in the \'etale topology) Morita equivalent to that of the quantum plane $\mathbb{C}_q[x,y]$ with $xy = q yx$ for $q$ an $m$-th root of unity where $m \mid n$. {\par \noindent} Based on this example, our strategy to develop $\kat{geometry}~\text{\bf @}_{\text{\bf n}}$ is to use $GL_n$-equivariant theory as far as it can serve us, but then study the remaining cases (which will lead to interesting algebras) rather than solving the remaining problems by extending our algebraic framework. \end{paragraaf} \begin{paragraaf} Finally, how can one use $\kat{geometry}~\text{\bf @}_{\text{\bf n}}$ to develop a noncommutative geometry. Again the strategy is simple : formalize the setting $A \Rightarrow~\forall n~A~\text{\bf @}_{\text{\bf n}}$ and make a list of relations holding naturally among the $\wis{spec}~A~\text{\bf @}_{\text{\bf n}}$. Then, define an object in $\kat{geometry}$ by a list of objects $X~\text{\bf @}_{\text{\bf n}}$ in $\kat{geometry}~\text{\bf @}_{\text{\bf n}}$ satisfying this list of relations. At this moment i do not have an elegant set of axioms. Let me conclude by proposing one axiom which should illustrate the principle. {\par \noindent} If $n = \sum_i m_i$ and let $V_i$ be an $m_i$-dimensional representation of $A$, then the direct sum $\oplus V_i$ is an $n$-dimensional representation of $A$. Hence there are morphisms \[ \underset{i}{\times}~\wis{spec}~A~\text{\bf @}_{{\text{\bf m}}_i} \rTo \wis{spec}~A~\text{\bf @}_{\text{\bf n}} \] satisfying obvious compatibility relations. {\par \noindent} To formalize this condition for objects $X = (X~\text{\bf @}_{\text{\bf n}})_n$ in $\kat{geometry}$, let us stratify the geometric points of the underlying $GL_n$-algebraic space $X~\text{\bf @}_{\text{\bf n}}$ by \[ X~\text{\bf @}_{\text{\bf n}}(r) = \{ x \in X~\text{\bf @}_{\text{\bf n}} \mid Stab_{GL_n}(x) \ \text{has a maximal torus of dimension $r$} \ \} \] Then, for each integral solution $n = \sum_i m_i$ one must have connecting morphisms \[ \underset{i}{\times} X~\text{\bf @}_{{\text{\bf m}}_i} \rTo^{c_{(m_i)}} X~\text{\bf @}_{\text{\bf n}} \] with the additional condition that \[ X~\text{\bf @}_{\text{\bf n}}(r) \subset \underset{m_1+\hdots+m_r=n}{\cup} GL_n.Im(c_{(m_1,\hdots,m_r)}). \] It is clear that a lot of additional work needs to be done. \end{paragraaf}
\section*{Appendix \thesection\protect\indent \parbox[t]{11.715cm} {#1}} \addcontentsline{toc}{section}{Appendix \thesection\ \ \ #1} } \renewcommand{\theequation}{\thesection.\arabic{equation}} \renewcommand{\thefootnote}{\fnsymbol{footnote}} \newcommand{\newsection}{ \setcounter{equation}{0} \section} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \newcommand{\tr}[1]{\:{\rm tr}\,#1} \newcommand{\Tr}[1]{\:{\rm Tr}\,#1} \def{\rm const}{{\rm const}} \def{\,\rm e}\,{{\,\rm e}\,} \def\partial{\partial} \def\delta{\delta} \def^{\dagger}{^{\dagger}} \newcommand{\br}[1]{\left( #1 \right)} \newcommand{\rf}[1]{(\ref{#1})} \newcommand{\nonumber \\*}{\nonumber \\*} \hyphenation{di-men-sion-al} \hyphenation{di-men-sion-al-ly} \def{\cal N}=4{{\cal N}=4} \def\alpha'{\alpha'} \def\alpha{\alpha} \def\epsilon{\epsilon} \def\theta{\theta} \def\phi{\phi} \def{\cal A}{{\cal A}} \def\left\langle W^\dagger(C_1)W(C_2)\right\rangle_{{\rm conn}}{\left\langle W^\dagger(C_1)W(C_2)\right\rangle_{{\rm conn}}} \begin{document} \begin{titlepage} \begin{flushright} ITEP--TH--17/99 \end{flushright} \vspace{1.5cm} \begin{center} {\LARGE Wilson Loop Correlator} \\[.5cm] {\LARGE in the AdS/CFT Correspondence}\\ \vspace{1.9cm} {\large K.~Zarembo}\footnote{Permanent address: {\it Department of Physics and Astronomy,} {\it University of British Columbia,} {\it 6224 Agricultural Road, Vancouver, B.C. Canada V6T 1Z1} and {\it Institute of Theoretical and Experimental Physics,} {\it B. Cheremushkinskaya 25, 117259 Moscow, Russia}.\\ E-mail: {\tt <EMAIL>}} \\ \vspace{24pt} {\it The Niels Bohr Institute\\ Blegdamsvej 17\\ DK-2100 Copenhagen 0\\ Denmark} \end{center} \vskip 2 cm \begin{abstract} The AdS/CFT correspondence predicts a phase transition in Wilson loop correlators in the strong coupling ${\cal N}=4$, $D=4$ SYM theory which arises due to instability of the classical string stretched between the loops. We study this transition in detail by solving equations of motion for the string in the particular case of two circular Wilson loops. The transition is argued to be smoothened at finite `t~Hooft coupling by fluctuations of the string world sheet and to be promoted to a sharp crossover. Some general comments about Wilson loop correlators in gauge theories are made. \end{abstract} \end{titlepage} \setcounter{page}{2} \newsection{Introduction} According to \cite{Mal98,Rej98}, calculation of Wilson loop correlators in $D=4$, ${\cal N}=4$ supersymmetric $SU(N)$ Yang-Mills theory at large $N$ and large 't~Hooft coupling amounts in evaluation of the classical string action in anti-de-Sitter space. The string propagates in the bulk of $AdS_5\times S^5$ with its ends attached to the Wilson loops lying on the boundary. This prescription is a consequence of the AdS/CFT correspondence \cite{Mal97,Gub98,Wit98}. Gross and Ooguri pointed out that it implies a kind of phase transition in the correlation function of two Wilson loops \cite{Gro98a}. The reason for the Gross-Ooguri phase transition is that the string action, the area of a minimal surface bounded by the loops, generically has two competing saddle points. The minimal surface can have a topology of annulus or can consist of two disconnected pieces spanning individual loops. The annulus evidently has smaller area when the loops are closed enough to one another. But the area of the annulus increases with separation between the loops and eventually disconnected surface becomes energetically more favourable. At large distances, the string behaves classically only in the vicinity of the loops and the connected correlator is saturated by perturbative exchange of lightest supergravity modes between disconnected pieces of the classical world sheet \cite{Gro98a,Ber98}. A jump from one saddle point to the other should lead to the phase transition in the Wilson loop correlator considered as a function of the separation. Two Wilson loop correlators in the AdS/CFT correspondence were studied at large distances when they are saturated by supergraviton exchange \cite{Ber98}. To our knowledge, the connected solution for the minimal surface has not yet been considered. The arguments of \cite{Gro98a} relied upon the pattern of topology change for the surface stretched between two concentric circles in the flat space, fig.~\ref{cat}. For further references, we sketch these arguments here. When the distance $L$ between the circles is small, the minimal surface is the catenoid: $r(x)=R_m\cosh(x/R_m)$. The area of the catenoid is given by \begin{equation}\label{am} A=\pi R_mL+\pi R_m^2\sinh\left(\frac{L}{R_m}\right), \end{equation} where $R_m$, the radius of its narrowest section, is determined by boundary conditions: \begin{equation}\label{rm} R_m\cosh\left(\frac{L}{2R_m}\right)=R. \end{equation} The left hand side of this equation, as a function of $R_m$, has a minimum at some $R_m={\rm const}\cdot L$, so for large enough $L$ this equation ceases to have any solutions. At the critical distance $L_*$, the connected minimal surface becomes unstable. Yet before its area starts to exceed the area of two discs, another minimal surface with the same boundary. \begin{figure}[t] \hspace*{5cm} \epsfxsize=7cm \epsfbox{cat.eps} \caption[x]{Minimal surface spanned by two concentric discs.} \label{cat} \end{figure} In this paper, we address the issue of how the Gross-Ooguri phase transition is affected by the geometry of $AdS_5$ and by quantum corrections due to fluctuations of the string world sheet. The former problem is much simpler. The solution for two circular Wilson loops is found in sec.~\ref{ads}. The surface with the topology of annulus becomes unstable at certain value of $L$. As a consequence, the Wilson loop correlator in ${\cal N}=4$ SYM undergoes the phase transition at infinite 't~Hooft coupling. The influence of world-sheet fluctuations is more difficult to estimate. It is still not known how to calculate superstring amplitudes in $AdS_5\times S^5$. Presence of the background Ramond-Ramond flux compels one to use the Green-Schwarz formalism \cite{Met98,Kal98a,Pes98,Kal98b,Kal98c,Raj99}, the covariant quantisation of the superstring in which is notoriously complicated \cite{GSW}. In view of the difficulties with passing to finite $\alpha'$ in $AdS$ geometry, which is equivalent to finite 't~Hooft coupling on the SYM side, we pursue the above flat space example to study the world-sheet fluctuations in sec.~\ref{flat}. The string amplitude between two loops can be calculated non-perturbatively in $\alpha'$ in this case \cite{Coh86,Mar89}. \newsection{Two Loop Correlator}\label{ads} In the supergravity (strong coupling) limit, the Wilson loop correlator is expressed in terms of an area of the classical string world sheet stretched between the loops \cite{Mal98,Rej98}: \begin{equation}\label{mal1} \left\langle W^\dagger(C_1)W(C_2)\right\rangle_{{\rm conn}} =\exp\left(-\frac{S}{2\pi\alpha'}\right), \end{equation} \begin{equation}\label{mal2} S=\int d\sigma d\tau\,\sqrt{\det_{ab}g_{\mu\nu}\partial_a x^\mu\partial_b x^\nu}, \end{equation} \begin{equation}\label{mal3} \frac{\delta S}{\delta x^\mu}=0. \end{equation} We take $C_1$ and $C_2$ to be concentric circles of radius $R$ separated by distance $L$ (fig.~\ref{cat}). This configuration suggests the use of the cylindric coordinates in ${\bf R}^4$. The $AdS_5$ metric is then \begin{equation} ds^2=\frac{1}{z^2}\left(dz^2+dt^2+dx^2+dr^2+r^2d\varphi^2\right). \end{equation} The boundary of $AdS$ is at $z=0$. We chose the units in which the radius of $AdS$ is unity and the string tension is proportional to the Yang-Mills coupling: $$\alpha'=(2g_{\rm YM}^2N)^{-1/2}.$$ \subsection{Minimal Surface} Using symmetries of the problem we take the following ansatz for the minimal surface: \begin{equation} t=0,~~\varphi=\sigma,~~r=r(\tau),~~x=x(\tau),~~z=z(\tau). \end{equation} Then the equations $\delta S/\delta t=0$ and $\delta S/\delta\varphi=0$ are satisfied identically. The equations for $r$, $x$, and $z$ follow from the action \begin{equation} S=2\pi\int d\tau\,\frac{r}{z^2}\,\sqrt{(r')^2+(x')^2+(z')^2}. \end{equation} The variation with respect to $r$ and $z$ yields: \begin{eqnarray} \left(\frac{r}{z^2}\,\frac{r'}{\sqrt{(r')^2+(x')^2+(z')^2}}\right)' -\frac{1}{z^2}\,\sqrt{(r')^2+(x')^2+(z')^2}&=&0, \\* \left(\frac{r}{z^2}\,\frac{z'}{\sqrt{(r')^2+(x')^2+(z')^2}}\right)' +\frac{2r}{z^3}\,\sqrt{(r')^2+(x')^2+(z')^2}&=&0. \end{eqnarray} The equation which follows from the variation with respect to $x$ can be integrated once to give \begin{equation} \frac{r}{z^2}\,\frac{x'}{\sqrt{(r')^2+(x')^2+(z')^2}}=k, \end{equation} where $k$ is an integration constant. If $k=0$, the string does not propagate along the $x$ direction and sweeps out a surface with the disc topology. The minimal surface in $AdS_5$ bounding the single circular Wilson loop was found in \cite{Gro98b,Ber98}: \begin{equation}\label{disc} r^2+z^2=R^2. \end{equation} If $k\neq 0$ (without loss of generality we assume that $k>0$), $x'$ is always positive and we can choose $x$ to be one of the coordinates on the string world sheet: $\tau=x$. With this gauge choice, the equations of motions take the form: \begin{eqnarray} r''-\frac{r}{k^2z^4}&=&0, \label{one} \\* z''+\frac{2r^2}{k^2z^5}&=&0, \label{two} \\* (z')^2+(r')^2+1-\frac{r^2}{k^2z^4}&=&0. \label{three} \end{eqnarray} Boundary conditions for these equations are \begin{equation} r(-L/2)=r(L/2)=R, \end{equation} \begin{equation} z(-L/2)=z(L/2)=0. \end{equation} In principle, it is necessary to regularise the problem by shifting the Wilson loops from the boundary of $AdS$ in order to get the solution with a finite area. Strictly speaking, more appropriate boundary conditions for $z$ are $z(\pm L)=\epsilon$, but the solution itself is not singular in the limit $\epsilon\rightarrow 0$ and, since a regularisation does not influence the critical behaviour, we shall first find unregularised solution and shall return to the issue of regularisation later. Qualitative structure of the solution is clear from eqs.~\rf{one}, \rf{two}. Closed string emitted by the Wilson loop at $x=-L/2$ falls into the interior of the $AdS$ space, then bounces back, and is absorbed by the Wilson loop at $x=L/2$. The radius of the string decreases till the bouncing point and then starts to increase (fig.~\ref{sol}), like in the flat space. \begin{figure}[t] \hspace*{5cm} \epsfxsize=7cm \epsfbox{sol.eps} \caption[x]{Schematic form of the solution for the minimal surface.} \label{sol} \end{figure} It is possible to integrate equations \rf{one}--\rf{three}. Adding first two ones multiplied by $r$ and by $u$, respectively, to the third equation, we get: \begin{equation} (r^2+z^2)''+2=0, \end{equation} or, taking into account boundary conditions: \begin{equation}\label{sph} r^2+z^2+x^2=R^2+\frac{L^2}{4}\equiv a^2. \end{equation} This equality suggests the substitution: \begin{eqnarray} r&=&\sqrt{a^2-x^2}\cos\theta, \nonumber \\* z&=&\sqrt{a^2-x^2}\sin\theta. \label{trig} \end{eqnarray} The equation~\rf{three} considerably simplifies in the new variables: \begin{equation} (a^2-x^2)(\theta ')^2+\frac{x^2}{a^2-x^2}+1 -\frac{\cos^2\theta}{k^2(a^2-x^2)\sin^4\theta}=0. \end{equation} The variables separate: \begin{equation}\label{tetash} \theta '=\pm\frac{a}{a^2-x^2}\,\sqrt{\frac{\cos^2\theta}{k^2a^2\sin^4\theta}-1}, \end{equation} where the upper sign is to be taken for $x\in [-L/2,0]$ and the lower sign for $x\in [0,L/2]$. The solution on the interval $[-L/2,0]$ is determined by equation \begin{equation}\label{tx} ka\int_0^\theta \,\frac{d\phi\,\sin^2\phi}{\sqrt{\cos^2\phi-k^2a^2\sin^4\phi}} =\frac12\,\ln\left(\frac{a+\frac{L}{2}}{a-\frac{L}{2}}\right) -\frac12\,\ln\left(\frac{a-x}{a+x}\right) \end{equation} and should be continued to positive $x$ by the symmetry $\theta(x)=\theta(-x)$. The parameter $k$ is fixed by requirement of continuity of the solution at $x=0$. Since $r$ and $z$ are even functions of $x$, so is $\theta(x)$. Hence, the derivative $\theta'(0)$ vanishes. Equation \rf{tetash} then fixes the value of $\theta$ at $x=0$: \begin{equation}\label{to} \theta_0\equiv \theta(0)=\arccos\left(\frac{\sqrt{4k^2a^2+1}-1}{2ka}\right). \end{equation} Substituting $x=0$ in eq.~\rf{tx}, we get an equation for $k$: \begin{equation}\label{fka} F(ka)=\frac12\,\ln\left(\frac{a+\frac{L}{2}}{a-\frac{L}{2}}\right) =\ln\left(\frac{\sqrt{R^2+\frac{L^2}{4}}+\frac{L}{2}}{R}\right), \end{equation} where \begin{equation}\label{defk} F(ka)\equiv ka\int_0^{\theta_0} \, \frac{d\phi\,\sin^2\phi}{\sqrt{\cos^2\phi-k^2a^2\sin^4\phi}}. \end{equation} The right hand side is the complete elliptic integral, since the upper bound of integration is at the square root branching point: $\cos^2\theta_0-k^2a^2\sin^4\theta_0=0$. The equations \rf{trig}, \rf{tx}, and \rf{fka} express the solution for the minimal surface in terms of elliptic functions. \subsection{Critical behaviour} The existence of the phase transition in the Wilson loop correlator can be inferred from eq.~\rf{fka}. The function $F(\xi)$ turns to zero at $\xi=0$ and at $\xi=\infty$: \begin{eqnarray} F(\xi)&\simeq& -\xi\ln\xi~~~(\xi\rightarrow 0), \nonumber \\*\label{fxi} F(\xi)&\simeq&\frac{\sqrt{2\pi^3}}{\Gamma^2\left(\frac14\right)} \,\frac{1}{\sqrt{\xi}}~~~(\xi\rightarrow\infty). \end{eqnarray} Consequently, it has a maximum at some $\xi=\xi_*$. At the same time, the right hand side of equation \rf{fka} varies from zero to infinity as the separation between loops increases. Continuity arguments show that the branch with larger $ka$ should be chosen from the two of the solutions of equation \rf{fka}. The two branches meet at $ka=\xi_*$. If $L$ exceeds the critical value $$L_*=2R\sinh F(\xi_*),$$ the equation \rf{fka} does not have any solutions and the minimal surface with the topology of annulus ceases to exist. Numerically, $\xi_*=0.58$ and \begin{equation} L_*=1.04 R. \end{equation} The connected minimal surface becomes unstable when Wilson loops are separated by this distance. In fact, the transition from the annulus to the disc topology is of the first order. The critical point is determined by the equality of the free energies in the two phases, the areas of the disconnected and the connected minimal surfaces. The area of regularised disconnected solution is \cite{Gro98b,Ber98} \begin{equation} S_{\rm disc}=2\left(\frac{2\pi\sqrt{R^2+\epsilon^2}}{\epsilon}-2\pi\right) =\frac{4\pi R}{\epsilon}-4\pi+O(\epsilon). \end{equation} The area of the connected surface also requires regularisation. Shift in the boundary conditions for $z$, $z(\pm L/2)=\epsilon$ instead of $z(\pm L/2)=0$, renders the area finite. The boundary conditions for $\theta(x)$ change, accordingly. From \rf{trig} we find: \begin{equation} \theta(\pm L/2)=\arctan\left(\frac{\epsilon}{R}\right)\simeq\frac{\epsilon}{R}. \end{equation} Using equations of motion, we get for the regularised area: \begin{eqnarray} S&=&2\pi\int_{-L/2}^{L/2}dx\,\frac{r}{z^2}\,\sqrt{1+(r')^2+(z')^2} =\frac{2\pi}{k}\int_{-L/2}^{L/2}dx\,\frac{r^2}{z^4} =\frac{4\pi}{k}\int_{-L/2}^{0}\frac{dx}{a^2-x^2}\,\frac{\cos^2\theta}{\sin^4\theta} \nonumber \\* &=&4\pi\int_{\epsilon/R}^{\theta_0}\frac{d\theta\,\cot^2\theta}{\sqrt{\cos^2\theta-k^2a^2 \sin^4\theta}}. \end{eqnarray} This integral can be simplified by the change of variables $$\tan\theta=\left(\frac{\sqrt{4k^2a^2+1}-1}{2}\right)^{-1/2}\sin\psi.$$ After some algebra, we obtain: \begin{equation}\label{ar} S=\frac{4\pi R}{\epsilon}-4\pi\,\frac{\alpha}{\sqrt{\alpha-1}} \int_0^{\pi/2}\frac{d\psi}{1+\alpha\sin^2\psi+\sqrt{1+\alpha\sin^2\psi}}, \end{equation} where \begin{equation}\label{al} \alpha=\frac{1+2k^2a^2+\sqrt{1+4k^2a^2}}{2k^2a^2}. \end{equation} The equations \rf{mal1}, \rf{fka}, \rf{defk}, \rf{ar}, and \rf{al} determine the correlator of two Wilson loops in the strong-coupling ${\cal N}=4$ SYM theory. The divergent part of the area is the same for the connected and the disconnected solutions. Its origin can be attributed to perimeter divergency of a Wilson loop due to self-energy contributions \cite{Gro98b,Ber98}. The renormalised area is always negative. Its short-distance asymptotics can be found with the help of eq.~\rf{fxi}: \begin{equation}\label{asym} S\simeq\frac{4\pi R}{\epsilon}-4\pi\,\frac{\sqrt{2\pi^3}} {\Gamma^2\left(\frac14\right)} \,ka\simeq \frac{4\pi R}{\epsilon}-\frac{16\pi^4}{\Gamma^4\left(\frac14\right)}\, \frac{R}{L}. \end{equation} The area grows with separation between loops and at the point of the phase transition meets the area of the disconnected surface. Solving equations for the critical point numerically, we find: \begin{equation} L_c=0.91 R. \end{equation} In spite of evident differences between the classical string propagation in $AdS_5$ and in the flat space, the qualitative pattern of the Gross-Ooguri phase transition appears to be not much influenced by geometry of the target space. In both cases, the transition is of the first order and takes place at $L\sim R$. We expect that an influence of the string fluctuations on the phase transition is also universal, at least in some respects. This is our motivation to study the free superstring propagator between two loops in the next section. \newsection{Two Loop Amplitude in the Free String Theory} \label{flat} The equations \rf{mal1}--\rf{mal3} are expected to be valid at infinite 't~Hooft coupling and to be replaced by a full sum over random surfaces in $AdS_5\times S^5$ with boundary conditions set by Wilson loops, when the coupling is finite. Presently, fluctuations of the string world sheet in $AdS$ geometry can be accounted for only to the first order in $\alpha'$ \cite{Gre99,For99}. The superstring amplitude in the flat space is much easier to calculate. It is known exactly and provides a possibility to trace qualitative changes brought in by non-perturbative $\alpha'$ corrections. The amplitude for the free type IIB superstring to propagate between two concentric circular loops is a particular case of the boundary state amplitude calculated for the bosonic string in \cite{Coh86} and for NSR string in \cite{Mar89}. After GSO projection, the amplitude takes the form: \begin{eqnarray}\label{ampl} {\cal A}&=&{\rm const}\,\int_0^\infty\frac{ds}{s^5}\, \frac{\Theta_2^4(0|is)+\Theta_3^4(0|is)-\Theta_4^4(0|is)}{2\eta^{12}(is)}\, \exp\left(-\frac{S(s)}{2\pi\alpha'}\right) \nonumber \\* &=&{\rm const}\,\int_0^\infty\frac{ds}{s^5}\, \frac{\Theta_2^4(0|is)}{\eta^{12}(is)}\, \exp\left(-\frac{S(s)}{2\pi\alpha'}\right), \end{eqnarray} where \begin{equation}\label{strac} S(s)=\frac{L^2}{s}+2\pi R^2 \tanh\left(\frac{\pi s}{2}\right). \end{equation} Suppose that $L^2\gg\alpha'$ and $R^2\gg\alpha'$, then $S$ is large and the proper time integral \rf{ampl} is saturated by a saddle point, if the latter exists. The saddle point of the action \rf{strac} is determined by equation \begin{equation}\label{smsp} \frac{\cosh\left(\frac{\pi s_m}{2}\right)}{\pi s_m}=\frac{R}{L}. \end{equation} By the substitution $\pi s_m=L/ R_m$ we recover the boundary condition \rf{rm} for the catenoid from this equation. The area of the catenoid \rf{am} is also reproduced: \begin{equation} S(s_m)=\frac{L^2}{s_m}+\frac{L^2}{\pi s_m^2}\,\sinh \pi s_m. \end{equation} The saddle point disappears at sufficiently large $L$, because eq.~\rf{smsp} has no solutions for $L>L_*$. Beyond the critical point, the main contribution to the integral comes from $s\rightarrow\infty$. The action at $s=\infty$ is $2\pi R^2$, the area of two discs. At large distances, the amplitude is saturated by an exchange of massless string modes: \begin{equation} {\cal A}\simeq {\rm const}\,\frac{96}{L^8}\,{\,\rm e}\,^{-R^2/\alpha'}. \end{equation} The semiclassical amplitude has a discontinuous first derivative in $L$ at the point of the transition between the two saddle points. The discontinuity persists to any finite order in the $\alpha'$ expansion. The integral \rf{ampl}, however, defines an analytic function of $L$. The phase transition is thus replaced by a smooth crossover in the exact amplitude. The larger $R^2$ is, the sharper this crossover will be. The abrupt change in the amplitude is clearly seen in fig.~\ref{amp} which displays the result of numerical integration of eq.~\rf{ampl} at $R^2=40\pi\alpha'$. Note that retaining only massless string modes is a good approximation everywhere above the transition, since $L^2$ is much larger than $\alpha'$, whereas below the transition all modes give comparable contribution. \begin{figure}[t] \hspace*{3cm} \epsfxsize=10cm \epsfbox{amp.eps} \caption[x]{The superstring amplitude between two circular loops versus $L^2$ at $R^2=20$ (the units are such that $2\pi\alpha'=1$). The thick solid line is $-\ln{\cal A}$. Two other curves represent approximations for the amplitude valid in different phases. The thin solid line is the area of the catenoid with $O(\alpha'^0)$ correction added. The line terminates at the point of instability, $L=L_*^2$. The dashed line is $2\pi R^2-\ln(\mbox{supergraviton~exchange})$. The estimate for the crossover point based on matching of the classical actions in the two phases is $L_c\simeq 22$ for the chosen value of $R$. Account of the supergraviton exchange and of the first order in the semiclassical expansion shift the crossover point by about $20\%$. The reason for numerically large deviation from the classical estimate is that the first order corrections are logarithmic in $L^2$.} \label{amp} \end{figure} The conclusion from the above consideration is that the world-sheet fluctuations smoothen the Gross-Ooguri phase transition. However, if $\alpha'=(2g_{\rm YM}^2N)^{-1/2}$ is sufficiently small, the transition should show up as a sharp crossover in the Wilson loop correlator. \newsection{Discussion} Calculations carried out in the previous sections demonstrate that the Wilson loop two point function in ${\cal N}=4$ SYM undergoes the first order phase transition as the distance between loops changes, if the 't~Hooft coupling is infinite. If the coupling is large but finite, tunnelling between the saddle points of the string action smoothens the dependence of the correlator on the distance and the phase transition is replaced by a sharp crossover. The transition seems to be completely washed out in the weak coupling, perturbative regime. In conclusion, we would like to comment on Wilson loop correlation functions in asymptotically free, confining gauge theories, which are also believed to have some kind of approximate, or even exact string representation. At short distances, $\Lambda_{QCD}^{-1}\ll L\ll R$, the Wilson loop correlator should be dominated by an open string stretched between the loops. This representation suggests that the large-radius limit of the correlator should determine an interaction potential between probe static charges: \begin{equation} V(L)=-\lim_{R\rightarrow\infty}\frac{\ln\left\langle W^\dagger(C_1)W(C_2)\right\rangle_{{\rm conn}}}{2\pi R}. \end{equation} The comparison of eq.~\rf{asym} with the potential found in \cite{Mal98} shows that this relation is satisfied by the two loop correlator calculated using the AdS/CFT correspondence. At large distances between loops, a more appropriate picture is that of the closed string exchange. It is not quite clear whether the open and the closed string regimes are separated by the Gross-Ooguri crossover. General arguments seem to indicate that this is the case. If geometric characteristics of the loops are large in the units of the string tension, a confining string has a large action and hopefully can be described within the semiclassical approximation. The instability of the classical string world sheet will then lead to the phase transition in the semiclassical amplitude. This kind of behaviour certainly holds for the free string. However, the free string amplitude is rather loose substitute for Wilson loop correlation functions in a gauge theory, basically because of the lack of zig zag invariance \cite{Pol97,Pol98}. One of the consequences of the violation of zig zag invariance is the Hagedorn transition in the free string case: The amplitude \rf{ampl} diverges at $L^2\leq L_H^2=2\pi\alpha'$ due to exponential growth of the closed string density of states. Such kind of divergency of correlation functions at short distances is expected in quantum gravity \cite{Aha99}, but not in gauge theories. In fact, intermediate states in the spectral representation for the Wilson loop correlator, \begin{equation}\label{sr} \left\langle W^\dagger(C_1)W(C_2)\right\rangle_{{\rm conn}}=\int_0^\infty dE\,\rho_C(E){\,\rm e}\,^{-EL}, \end{equation} \begin{equation} \rho_C(E)=\sum_{n\neq 0}\delta(E-E_n)\left|\langle 0|W(C)| n\rangle\right|^2, \end{equation} which generically are glueballs, are expected to have exponentially growing spectrum. But zig zag symmetry suppresses coupling of Wilson loops to glueballs of very high spin, which can be seen expanding the Wilson loop in local operators \cite{Shi80}. Spin $2n$ operators come with coefficients of order $1/n!$. For example, the operator with a minimal number of derivatives is $$\frac{1}{n!}\,\tr\left(\oint dx_\mu x_\nu\,F_{\mu\nu}(x_0)\right)^{n}.$$ Hence, Wilson loop effectively couples only to states with spin $J<J_0$, where $J_0$ is proportional to the area of the loop. As a result, the spectral density $\rho_C(E)$ should have power-law asymptotics. This property is essentially a consequence of the zig zag invariance, since the latter requires the Wilson loop of zero area to be the unit operator. This is not true for the boundary states in \rf{ampl}, which couple to all string modes even at $R=0$ \cite{Coh86,Mar89}. The modular transformation converts the Hagedorn divergency of the amplitude to the tachyon singularity in the open string channel \cite{Ole86}. Again, the tachyon instability is not expected to arise in gauge theories. \subsection*{Acknowledgments} I am grateful to Y.~Makeenko, P.~Olesen and G.~Semenoff for discussions and to J.~Abmj{\o}rn for hospitality at The Niels Bohr Institute. This work was supported by NATO Science Fellowship and, in part, by INTAS grant 96-0524, RFFI grant 97-02-17927 and grant 96-15-96455 for the promotion of scientific schools.
\section{Introduction} \label{intro} In a classic paper from 1977 \cite{Nam77}, a decade after the $SU(2)_L \times U(1)_Y$ model of electroweak interactions had been proposed \cite{GlaSalWei}, Nambu made the observation that, while the Glashow-Salam-Weinberg (GSW) model does not admit isolated, regular magnetic monopoles, there could be monopole-antimonopole pairs joined by short segments of a vortex carrying Z-magnetic field (a {\it Z-string}). The monopole and antimonopole would tend to annihilate but, he argued, longitudinal collapse could be stopped by rotation. He dubbed these configurations {\it dumbells}\footnote{or {\it monopolia}, after analogous configurations in superfluid helium \cite{Mak77}} and estimated their mass at a few TeV. A number of papers advocating other, related, soliton-type solutions \footnote{One example, outside the scope of the present review, are so-called vorticons, proposed by Huang and Tipton, which are closed loops of string with one quantum of Z boson trapped inside.} in the same energy range followed \cite{early}, but the lack of topological stability led to the idea finally being abandoned during the eighties. Several years later, and completely independently, it was observed that the coexistence of global and gauge symmetries can lead to stable non-topological strings called ``semilocal strings'' \cite{VacAch91} in the $\sin^2 \theta_w = 1$ limit of the GSW model that Nambu had considered. Shortly afterwards it was proved that Z-strings were stable near this limit \cite{Vac92}, and the whole subject made a comeback. This report is a review of the current status of research on electroweak strings. Apart from the possibility that electroweak strings may be the first solitons to be observed in the standard model, there are two interesting consequences of the study of electroweak and semilocal strings. One is the unexpected connection with baryon number and sphalerons. The other is a deeper understanding of the connection between the topology of the vacuum manifold (the set of ground states of a classical field theory) and the existence of stable non-dissipative configurations, in particular when global and local symmetries are involved simultaneously. In these pages we assume a level of familiarity with the general theory and basic properties of topological defects, in particular with the homotopy classification. There are some excellent reviews on this subject in the literature to which we refer the reader \cite{reviews,Col85a,SalVol87}. On the other hand, electroweak and semilocal strings are non-topological defects, and this forces us to take a slightly different point of view from most of the existing literature. Emphasis on stability properties is mandatory, since one cannot be sure from the start whether these defects will actually form. With very few exceptions, this requires an analysis on a case by case basis. Following the discussion in \cite{Col85}, one should begin with the definition of dissipative configurations. Consider a classical field theory with energy density $T_{00} \geq 0$ such that $T_{00} = 0$ everywhere for the ground states (or ``vacua'') of the theory. A solution of a classical field theory is said to be dissipative if \be \lim_{t \to \infty} {\rm max}_{\bf x} T_{00} ({\bf x},t) = 0 \end{equation} We will consider theories with spontaneous symmetry breaking from a Lie group $G$ (which we assume to be finite-dimensional and compact) to a subgroup $H$; the space $\cal V$ of ground states of the theory is usually called the vacuum manifold and, in the absence of accidental degeneracy, is given by ${\cal V}=G/H$. The classification of topological defects is based on the homotopy properties of the vacuum manifold. If the vacuum manifold contains non-contractible $n$-spheres then field configurations in $n+1$ spatial dimensions whose asymptotic values as $r \to \infty$ ``wrap around'' those spheres are necessarily non-dissipative, since continuity of the scalar field guarantees that, at all times, at least in one point in space the scalar potential (and thus the energy) will be non-zero. The region in space where energy is localized is referred to as a {\it topological defect}. Field configurations whose asymptotic values are in the same homotopy class are said to be in the same {\em topological sector} or to have the same {\em winding number}. In three spatial dimensions, it is customary to use the names {\it monopole}, {\it string}\footnote{The names {\it cosmic string} and {\it vortex} are also common. Usually, ``vortex'' refers to the configuration in two spatial dimensions, and ``string'' to the corresponding configuration in three spatial dimensions; the adjective ``cosmic'' helps to distinguish them from the so-called fundamental strings or superstrings. } and {\it domain wall} to refer to defects that are pointlike, one-dimensional or two-dimensional respectively. Thus, one can have topological domain walls only if $\pi_0({\cal V}) \neq 1$, topological strings only if $\pi_1({\cal V})\neq 1$ and topological monopoles only if $\pi_2({\cal V}) \neq 1$. Besides, defects in different topological sectors cannot be deformed into each other without introducing singularities or supplying an infinite amount of energy. This is the origin of the homotopy classification of topological defects. We should point out that the topological classification of textures based on $\pi_3({\cal V})$ has a very different character, and will not concern us here; in particular, configurations from different topological sectors can be continuously deformed into each other with a finite cost in energy. In general, textures unwind until they reach the vacuum sector and therefore they are dissipative. It is well known, although not always sufficiently stressed, that the precise relationship between the topology of the vacuum and the existence of stable defects is subtle. First of all, note that a trivial topology of the vacuum manifold does not imply the {\it non-existence} of stable defects. Secondly, we have said that a non-trivial homotopy of the vacuum manifold can result in non-dissipative solutions but, in general, these solutions need not be time independent nor stable to small perturbations. One exception is the field theory of a single scalar field in 1+1 dimensions, where a disconnected vacuum manifold (i.e. one with $\pi_0({\cal V}) \neq 1$) is sufficient to prove the existence of time independent, classically stable ``kink'' solutions \cite{GolJac75, Col85}. But this is not the norm. The $O(3)$ model, for instance, has topological global monopoles \cite{BarVil89} which are time independent, but they are unstable to angular collapse even in the lowest non-trivial winding sector\cite{Gol89}. It turns out that the situation is particularly subtle in theories where there are global and gauge symmetries involved simultaneously. The prototype example is the semilocal string, described in section \ref{semilocal}. In the semilocal string model, the classical dynamics is governed by a single parameter $\beta = m_s^2/m_v^2$ that measures the square of the ratio of the scalar mass, $m_s$, to the vector mass, $m_v$ (this is the same parameter that distinguishes type I and type II superconductors). It turns out that: - when $\beta>1$ the semilocal model provides a counterexample to the widespread belief that quantization of magnetic flux is tantamount to its localization, i.e., confinement. The vector boson is massive and we expect this to result in confinement of magnetic flux to regions of width given by the inverse vector mass. However, this is not the case! As pointed out by Hindmarsh \cite{Hin92} and Preskill \cite{Pre92}, this is a system where magnetic flux is topologically conserved and quantized, and there is a finite energy gap between the non-zero flux sectors and the vacuum, and yet there are {\it no stable vortices}. - when $\beta <1$ strings are stable\footnote{We want to stress that, contrary to what is often stated in the literature, the semilocal string with $\beta <1$ is {\it absolutely stable}, and not just {\it metastable}.} even though the vacuum manifold is simply connected, {$\pi_1 ({\cal V}) = 1$}. Semilocal vortices with $\beta < 1$ are a remarkable example of a non-topological defect which is stable both perturbatively and to semiclassical tunnelling into the vacuum \cite{PreVil92}. As a result, when { the global symmetries of a semilocal model} are gauged, dynamically stable non-topological solutions can still exist for certain ranges of parameters very close to stable semilocal limits. In the case of the standard electroweak model, for instance, strings are (classically) stable only when $\sin^2 \theta_w \approx 1$ and the mass of the Higgs is smaller than the mass of the Z boson. We begin with a description of the Glashow-Salam-Weinberg model, in order to set our notation and conventions, and a brief discussion of topological vortices (cosmic strings). It will be sufficient for our purposes to review cosmic strings in the Abelian Higgs model, with a special emphasis on those aspects that will be relevant to electroweak and semilocal strings. We should point out that these vortices were first considered in condensed matter by Abrikosov \cite{Abr57} in the non-relativistic case, in connection with type II superconductors. Nielsen and Olesen were the first to consider them in the context of relativistic field theory, so we will follow a standard convention in high energy physics and refer to them as Nielsen-Olesen strings \cite{NieOle73}. Sections \ref{semilocal} to \ref{zoo} are dedicated to semilocal and electroweak strings, and other embedded defects in the standard GSW model. Electroweak strings in extensions of the GSW model are discussed in section \ref{ewstringinextensions}. In section \ref{stability} the stability of straight, infinitely long electroweak strings is analysed in detail (in the absence of fermions). Sections \ref{superconductivity} to \ref{ewstringandsphaleron} investigate fermionic superconductivity on the string, the effect of fermions on the string stability, and the scattering of fermions off electroweak strings. The surprising connection between strings and baryon number, and their relation to sphalerons, is described in sections \ref{ewstringbaryonnumber} and \ref{ewstringandsphaleron}. Here we also discuss the possibility of string formation in particle accelerators (in the form of dumbells, as was suggested by Nambu in the seventies) and in the early universe. Finally, section \ref{he3section} describes a condensed matter analog of electroweak strings in superfluid helium which may be used to test our ideas on vortex formation, fermion scattering and baryogenesis. A few comments are in order: $\bullet$ Unless otherwise stated we take spacetime to be flat, 3+1 dimensional Minkowski space; the gravitational properties of embedded strings are expected to be the same as those of Nielsen-Olesen strings \cite{GibOrtRuiSam92} and will not be considered here. A limited discussion of possible cosmological implications can be found in sections \ref{networks} and \ref{cosmologicalapplications}. $\bullet$ We concentrate on regular defects in the standard model of electroweak interactions. Certain extensions of the Glashow-Salam-Weinberg model are {briefly} considered in section 6 but otherwise they are outside the scope of this review; the same is true of singular solutions. In particular, we do not discuss isolated monopoles in the GSW model \cite{GibOrtRuiSam92,ChoMai96}, which are necessarily singular. $\bullet$ No family mixing effects are discussed in this review and we also ignore $SU(3)_c$ colour interactions, even though their physical affects are expected to be very interesting, in particular in connection with baryon production by strings (see section \ref{ewstringbaryonnumber}). $\bullet$ Our conventions are the following: spacetime has signature $(+,-,-,-)$. Planck's constant and the speed of light are set to one, $\hbar = c = 1$. The notation $(x)$ is shorthand for all spacetime coordinates $(x^0, x^i), \ i = 1,2,3$; whenever the $x$-coordinate is meant, it will be stated explicitly. We also use the notation $(t, {\vec x})$. $\bullet$ Complex conjugation and hermitian conjugation are both indicated with the same symbol, ($^\dagger$), but it should be clear from the context which one is meant. For fermions, ${\overline \psi} = \psi^\dagger \gamma^0$, as usual. Transposition is indicated with the symbol $(^T)$. $\bullet$ One final word of caution: a gauge field is a Lie Algebra valued one-form $A = A_\mu dx^\mu = A_\mu^a T^a dx^\mu$, but it is also customary to write it as a vector. In cylindrical coordinates $(t,\rho,\varphi,z)$, $ A = A_t dt + A_\rho d\rho + A_\varphi d\varphi + A_z dz$ is often written ${\vec {A}} = {A}_t \hat{t} + {A}_\rho \hat{\rho} + ({A}_\varphi / \rho) \hat{\varphi} + {A}_z \hat{z} $, In spherical coordinates, $(t,r,\theta,\varphi)$, $ A = A_t dt + A_r dr + A_\theta d\theta + A_\varphi d\varphi $ is also written ${\vec {A}} = {A}_t \hat{t} + {A}_r \hat{r} + ({A}_\theta /r) \hat{\theta} + ({A}_\varphi / r\sin\theta) \hat{\varphi} $. We use both notations throughout. \subsection{The Glashow-Salam-Weinberg model.} \label{WS} In this section we set out our conventions, which mostly follow those of \cite{CheLi91}. The standard (GSW) model of electroweak interactions is described by the Lagrangian \be L = L_b \ + \sum_{\rm families} L_f \ + \ L_{fm} \label{GWSaction}\end{equation} The first term describes the bosonic sector, comprising a neutral scalar $\phi^0$, a charged scalar $\phi^+$, a massless photon $A_\mu$, and three massive vector bosons, two of them charged ($W^\pm_\mu$) and the neutral $Z_\mu$. The last two terms describe the dynamics of the fermionic sector, which consists of the three families of quarks and leptons \be \pmatrix{\nu_e \cr e \cr u\cr d\cr} \qquad \pmatrix{\nu_\mu \cr \mu \cr c\cr s\cr}\qquad \pmatrix{\nu_\tau \cr \tau \cr t\cr b\cr} \end{equation} \subsubsection{The bosonic sector} The bosonic sector describes an $SU(2)_L\times U(1)_Y$ invariant theory with a scalar field $\Phi$ in the fundamental representation of $SU(2)_L$. It is described by the Lagrangian: \begin{equation} L_b = L_W + L_Y + L_{\Phi} - V(\Phi ) \label{GSWb} \end{equation} with \begin{equation} \eqalign{ L_W &= - {1 \over 4} W_{\mu \nu}^{\ \ a} W^{\mu \nu a} \qquad, \qquad a = 1,2,3 \cr L_Y &= - {1 \over 4} Y_{ \mu \nu} Y^{\mu \nu} \cr} \label{2.3} \end{equation} where $\ W_{\mu \nu}^{\ \ a} = \partial_\mu W_\nu^a - \partial_\nu W_\nu^a + g \epsilon^{abc} W_\mu^b W_\nu^c \ $ and $\ Y_{\mu \nu} = \partial_\mu Y_\nu - \partial_\nu Y_\mu \ $ are the field strengths for the $SU(2)_L$ and $U(1)_Y$ gauge fields respectively. Summation over repeated $SU(2)_L$ indices is understood, and there is no need to distinguish between upper and lower ones. $\epsilon^{123} = 1$. Also, \begin{equation} L_\Phi = |D_\lambda \Phi |^2 \equiv \biggl |\biggl (\partial _\lambda - {{ig} \over 2} \tau ^a W_\lambda ^a - {{ig'} \over 2} Y_\lambda \biggr) \Phi \biggr| ^2 \label{2.4} \end{equation} \begin{equation} V(\Phi ) = \lambda (\Phi ^{\dag} \Phi - \eta ^2 /2 )^2 \ , \label{2.5} \end{equation} where $\tau^a$ are the Pauli matrices, \be \tau^1 = \pmatrix {0 \ \ 1 \cr 1 \ \ 0}\ \ , \qquad \tau^2 = \pmatrix {0 \ -i \cr i \ \ \ 0}\ \ , \qquad \tau^3 = \pmatrix {1 \ \ \ 0 \cr 0 \ -1} \ \ , \end{equation} from which one constructs the weak isospin generators $T^a = {1 \over 2}\tau^a $ satisfying $[ T^a, T^b ] = i\epsilon^{abc} T^c$. The classical field equations of motion for the bosonic sector of the standard model of the electroweak interactions are (ignoring fermions): \be D^\mu D_\mu \Phi + 2\lambda \left ( \Phi^{\dag} \Phi - {{\eta^2} \over 2} \right ) \Phi = 0 \label{phieqn} \end{equation} \be D_\nu W^{\mu \nu a} = j^{\mu a}_W = {i \over 2} g \left [ \Phi^{\dag} \tau^a D^\mu \Phi - (D^\mu \Phi )^{\dag} \tau^a \Phi \right ] \label{weqn} \end{equation} \be \partial_\nu Y^{\mu \nu} = j^{\mu}_Y = {i \over 2} g' \left [ \Phi^{\dag} D^\mu \Phi - (D^\mu \Phi )^{\dag} \Phi \right ] \ \ , \label{ye} \end{equation} where $D_\nu W^{\mu \nu a} = \partial_\nu W^{\mu \nu a} + g \epsilon^{abc} W_\nu ^b W^{\mu \nu c} \ $. When the Higgs field $\Phi$ acquires a vacuum expectation value (VEV), the symmetry breaks from $SU(2)_L\times U(1)_Y$ to $U(1)_{em}$. In particle physics it is standard practice to work in unitary gauge and take the VEV of the Higgs to be $\langle\Phi ^T \rangle = \eta(0,1)/\sqrt 2$. In that case the unbroken $U(1)$ subgroup, which describes electromagnetism, is generated by the charge operator \be Q \ \equiv \ T^3 + {Y \over 2} \ = \ \pmatrix{1 \ \ 0 \cr 0 \ \ 0} \end{equation} and the two components of the Higgs doublet are charge eigenstates \begin{equation} \Phi = \pmatrix{\phi^+\cr \phi^0\cr}\ \ . \end{equation} $Y$ is the hypercharge operator, which acts on the Higgs like the $2 \times 2$ identity matrix. Its eigenvalue on the various matter fields can be read off from the covariant derivatives $D_\mu = \partial_\mu - ig W_\mu^a T^a - i g' Y_\mu (Y/2) $ which are listed explicitly in equations (\ref{2.4}) and (\ref{DPsi})-(\ref{Ddr}). In unitary gauge, the $Z$ and $A$ fields are defined as \begin{equation} Z_\mu \equiv \cos\theta_wW_\mu^3 - \sin\theta_w Y_\mu \ , \ \ \ \ A_\mu \equiv \sin\theta_wW_\mu^3+ \cos\theta_w Y_\mu \ , \label{2.9} \end{equation} and $W_\mu^\pm \equiv (W_\mu^1 \mp i W_\mu^2) / \sqrt 2 $ are the W bosons. The weak mixing angle $\theta_w$ is given by $\tan\theta_w \equiv {{g'} / g} $; electric charge is $e = g_z \sin\theta_w \cos\theta_w$ with $g_z \equiv (g^2 + {g'} ^2 )^{1/2} $. However, unitary gauge is not the most convenient choice in the presence of topological defects, where it is often singular. Here we shall need a more general definition in terms of an arbitrary Higgs configuration $\Phi (x)$: \be Z_\mu \equiv \cos\theta_w ~n^a (x) W_\mu ^a - \sin\theta_w ~Y_\mu \ , \ \ \ \ A_\mu \equiv \sin\theta_w ~n^a (x) W_\mu ^a + \cos\theta_w ~Y_\mu \ , \label{zmuamu} \end{equation} where \be n^a (x) \equiv -{{\Phi^{\dag}(x) \tau^a \Phi (x)} \over {\Phi^{\dag}(x) \Phi}(x)} \ . \label{nadef} \end{equation} is a unit vector by virtue of the Fierz identity $\ \sum_a (\Phi^\dagger \tau^a \Phi)^2 = (\Phi^\dagger \Phi)^2\ $. In what follows we omit writing the $x$-dependence of $n^a$ explicitly. Note that $n^a$ is ill defined when $\Phi = 0$, so in particular at the defect cores. The generators associated with the photon and the Z-boson are, respectively, \be Q = n^aT^a + Y/2 \ \ , \qquad\qquad T_Z = \cos^2\theta_w n^aT^a - \sin^2 \theta_w {Y\over 2} = n^aT^a - \sin^2 \theta_w Q \ \ , \label{QandTzgenerators}\end{equation} while the generators associated with the (charged) W bosons are determined, up to a phase, by the conditions \be [Q, T^\pm] = \pm T^\pm \qquad\qquad [T^+,T^-] = n^aT^a = T_Z + \sin^2 \theta_w Q \ \ , \qquad\qquad (T^+)^\dagger = T^- \end{equation} (note that, if $n^a = (0,0,1)$ as is the case in unitary gauge, one would take $T^\pm = (T^1 \pm i T^2)/\sqrt 2$.) There are several different choices for defining the electromagnetic field strength but, following Nambu, we choose: \be A_{\mu \nu} = sin\theta _w ~n^a W_{\mu \nu} ^a + cos\theta _w ~Y_{\mu \nu} \label{amunu} \end{equation} where, $W_{\mu \nu}^a$ and $Y_{\mu \nu}$ are field strengths. The different choices for the definition of the field strength agree in the region where $D_\mu \Phi = 0$ where $D_\mu$ is the covariant derivative operator; in particular this is different from the well known 't Hooft definition which is standard for monopoles \cite{tHoPol74}. (For a recent discussion of the various choices see, e.g. \cite{HinJam94,Hin94,Tor98}). And the combination of $SU(2)$ and $U(1)$ field strengths orthogonal to $A_{\mu \nu}$ is defined to be the $Z$ field strength: \be Z_{\mu \nu} = cos\theta _w ~n^a W_{\mu \nu} ^a - sin\theta_w ~Y_{\mu \nu} \ . \label{zmunu} \end{equation} \subsubsection{The fermionic sector} \label{fermionicsector} The fermionic Lagrangian is given by a sum over families plus family mixing terms ($L_{fm}$). Family mixing effects are outside the scope of this review, and we will not consider them any further. Each family includes lepton and quark sectors \begin{equation} L_f = L_l + L_q \label{2.6} \end{equation} which for, say, the first family are \begin{equation} L_l = - i {\bar \Psi} \gamma^\mu D_\mu \Psi - i {\bar e}_R \gamma^\mu D_\mu e_R + h({\bar e}_R \Phi^{\dag} \Psi + {\bar \Psi} \Phi e_R ) \ \ , \qquad\qquad {\rm where} \ \ \Psi = \pmatrix{\nu_e \cr e\cr}_L \label{2.7} \end{equation} \be \eqalign{ L_q =& -i ({\bar u} , {\bar d})_L \gamma^\mu D_\mu \pmatrix{u\cr d\cr}_L -i {\bar u}_R \gamma^\mu D_\mu u_R -i {\bar d}_R \gamma^\mu D_\mu d_R \cr & -G_d \biggl [ ({\bar u}, {\bar d})_L \pmatrix{\phi^+\cr \phi^0\cr} d_R +{\bar d}_R (\phi^{-} , {\phi^*}) \pmatrix{u\cr d\cr}_L \biggr ] \cr & - G_u \biggl [ ({\bar u}, {\bar d})_L \pmatrix{-{\phi^*}\cr \phi^{-}\cr} u_R + {\bar u}_R (-\phi^0, \phi^+) \pmatrix{u\cr d\cr}_L \biggr ] \cr} \label{2.8} \end{equation} where $\phi^*$ and $\phi^-$ are the complex conjugates of $\phi^0$ and $\phi^+$ respectively. $h, \ G_d$ and $G_u$ are Yukawa couplings. The indices $L$ and $R$ refer to left- and right-handed components and, rather than list their charges under the various transformations, we give here all covariant derivatives explicitly: \begin{equation} D_\mu \Psi = D_\mu \pmatrix{\nu \cr e\cr}_L = \biggl ( \partial_\mu - {{ig} \over 2} \tau^a W_\mu ^a + {{ig'} \over 2} Y_\mu \biggr ) \pmatrix{\nu \cr e\cr}_L \label{DPsi} \end{equation} \begin{equation} D_\mu e_R = ( \partial_\mu + ig' Y_\mu ) e_R \label{Der} \end{equation} \begin{equation} D_\mu \pmatrix{u\cr d\cr}_L = \biggl ( \partial_\mu - {{ig} \over 2} \tau^a W_\mu ^a - {{ig'} \over 6} Y_\mu \biggr ) \pmatrix{u\cr d\cr}_L \label{Dul} \end{equation} \begin{equation} D_\mu u_R = ( \partial_\mu - {{i2g'} \over 3} Y_\mu ) u_R \label{Dur} \end{equation} \begin{equation} D_\mu d_R = ( \partial_\mu + {{ig'} \over 3} Y_\mu ) d_R \label{Ddr} \end{equation} \noindent $\bullet$ One final comment: Electroweak strings are non-topological and their stability turns out to depend on the values of the parameters in the model. In this paper we will consider the electric charge $e$, Yukawa couplings and the VEV of the Higgs, $\eta / \sqrt 2$, to be given by their measured values, but the results of the stability analysis will be given as a function of the parameters $\sin^2 \theta_w$ and $\beta = (m_H/ m_Z)^2$ (the ratio of the Higgs mass to the Z mass squared); we remind the reader that $\sin^2 \theta_w \approx 0.23$, $m_Z \equiv g_z\eta/2 = 91.2$ GeV, $m_W \equiv g\eta/2 = 80.41$ GeV and current bounds on the Higgs mass $m_H \equiv \sqrt{2\lambda} \eta$ are $m_H > 77.5 $ GeV. \section{Review of Nielsen-Olesen topological strings} \label{NO} We begin by reviewing Nielsen-Olesen (NO) vortices in the Abelian Higgs model, with emphasis on those aspects that are relevant to the study of electroweak strings. More detailed information can be found in existing reviews \cite{reviews}. \subsection{The Abelian Higgs model} The theory contains a complex scalar field $\Phi$ and a $U(1)$ gauge field which becomes massive through the Higgs mechanism. By analogy with the GSW model, we will call this field $Y_\mu$. The action is \be {\cal S} = \int d^4x \left[ |D_\mu\Phi|^2 - \lambda \Biggl(\Phi^\dagger\Phi - {\eta^2\over 2}\Biggr)^2 - {\mathchoice{{\textstyle{1\over 4}}}{1\over 4}{1\over 4 Y_{\mu\nu}Y^{\mu\nu} \right] \label{aHaction} \end{equation} where $D_\mu = \partial_\mu -iqY_\mu$ is the $U(1)$-covariant derivative, and $Y_{\mu \nu} = \partial_\mu Y_\nu -\partial_\nu Y_\mu $ is the $U(1)$ field strength. The theory is invariant under $U(1)$ gauge transformations: \be \Phi (x) \to e^{iq\chi (x)} \Phi (x) = \hat\Phi (x) \ , \qquad Y_\mu (x) \to Y_\mu (x) + \partial_\mu \chi (x) = \hat Y_\mu (x) \end{equation} which give $D_\mu \Phi (x) \to \hat D_\mu \hat \Phi(x) = e^{iq\chi (x)} D_\mu \Phi $. The equations of motion derived from this Lagrangian are: \be \eqalign{ &D_\mu D^\mu\Phi + 2 \lambda(|\Phi|^2 - {\eta^2\over 2})\Phi = 0 \cr &\partial^\mu Y_{\mu\nu} = - iq\Phi^\dagger{\Dbw}_{\nu}\Phi \cr} \label{eom} \end{equation} Before we proceed any further, we should point out that, up to an overall scale, the classical dynamics of the Abelian Higgs model is governed by a single parameter, $\beta = 2 \lambda / q^2$, the (square of the) ratio of the scalar mass to the vector mass \footnote{ $\beta$ is also the parameter that distinguishes superconductors or type I ($\beta <1$) from type II ($\beta >1$)}. The action (\ref{aHaction}) contains three parameters, $(\lambda, \eta, q)$, which combine into the scalar mass $\sqrt {2\lambda} \eta = m_s \equiv l_s^{-1} $, the vector mass $q \eta = m_v \equiv l_v^{-1} $, and an overall energy scale given by the vacuum expectation value of the Higgs, $\eta/ \sqrt 2$. The rescaling \be \Phi(x) = {{\eta} \over {\sqrt{2}} } \hat{\Phi}(x) \ \ , \qquad\qquad x = {\sqrt{2} \over q \eta} \hat{x} \ \ , \qquad\qquad Y_\mu = {\eta\over \sqrt{2}} \hat {Y}_\mu \label{natural} \end{equation} changes the action to \be {S} = {1\over q^2 } \int d^4 x\[ |D_\mu\Phi|^2 - \half \beta (\Phi^\dagger\Phi -1)^2 - {\mathchoice{{\textstyle{1\over 4}}}{1\over 4}{1\over 4 Y_{\mu\nu}Y^{\mu\nu} \] \ \ , \end{equation} where now ${D}_\mu = {\partial}_\mu - i {Y}_\mu$ and we have omitted hats throughout for simplicity. In physical terms this corresponds to taking $l_v$ as the unit of length (up to a factor of $\sqrt2$) and absorbing the $U(1)$ charge $q$ into the definition of the gauge field, thus \be \Phi \to {\Phi \over <|\Phi|>} \ \ , \qquad\qquad x \to {x \over \sqrt{2} l_v} \ \ , \qquad\qquad eY_\mu \to {Y_\mu \sqrt{2} l_v} \ \ , \qquad\qquad E \to {E \over <|\Phi|^2>} \ \ . \label{natural2} \end{equation} The energy associated with (\ref{aHaction}) is \be {\cal E} = \int d^3x \[|D_0\Phi|^2 + | D_i\Phi|^2 + \lambda \Biggl(\Phi^\dagger\Phi - {\eta^2\over 2}\Biggr)^2\ + \half {\vec E}^2 + \half {\vec B}^2\] \end{equation} where the electric and magnetic fields are given by $F_{0i} = E_i $ and $F_{ij} = \epsilon_{ijk} B^k$ respectively ($i,j,k = 1,2,3$). Modulo gauge transformations, the ground states are given by $Y_\mu = 0$, $\Phi = \eta e^{iC}/{\sqrt 2}$, where $C$ is constant. Thus, the vacuum manifold is the circle \be {\cal V} = \{ \Phi \in {\rm I\!\!\! C} \ | \ \Phi^\dagger \Phi - {\eta^2 \over 2} = 0 \} \cong S^1 \ . \label{vacman} \end{equation} A necessary condition for a configuration to have finite energy is that the asymptotic scalar field configuration as $r \to \infty$ must lie entirely in the vacuum manifold. Also, $D_\mu \Phi $ must tend to zero, and this condition means that scalar fields at neighbouring points must be related by an infinitesimal gauge transformation. Finally, the gauge field strengths must also vanish asymptotically. Note that, in the Abelian Higgs model, the last condition follows from the second, since $ 0 = [D_\mu, D_\nu]\Phi = -iqY_{\mu\nu} \Phi$ implies $Y_{\mu \nu} = 0$ But this need not be the case when the Abelian Higgs model is embedded in a larger model. Vanishing of the covariant derivative term implies that, at large $r$, the asymptotic configuration $\Phi(x)$ must lie on a gauge orbit; \be \Phi(x) = g(x) \Phi_0 \ , \qquad {\rm where} \ g(x) \in G \qquad {\rm and} \qquad\Phi_0 \in \cal V \ . \label{gaugeorbit} \end{equation} where $\Phi_0$ is a reference point in $\cal V$. Note that, since all symmetries are gauge symmetries, the set of points that can be reached from $\Phi_0$ through a gauge transformation (the gauge orbit of $\Phi_0$) spans the entire vacuum manifold. Thus, ${\cal V} = G/H = G_{\rm local} / H_{\rm local}$, where $G_{\rm local }$ indicates the group of gauge -- {\it i.e.} local -- symmetries. On the other hand the spaces $\cal V$ and $G_{\rm local}/H_{\rm local}$ need not coincide in models with both local and global symmetries, and this fact will be particularly relevant in the discussion of semilocal strings. \subsection{Nielsen-Olesen vortices} In what follows we use cylindrical coordinates $(t, \rho, \varphi, z)$. We are interested in a static, cylindrically symmetric configuration corresponding to an infinite, straight string along the $z$-axis. The ansatz of Nielsen and Olesen \cite{NieOle73} for a string with winding number $n$ is \be \Phi = {\eta \over \sqrt 2} f (\rho) e^{in\varphi}\ \ , \qquad\quad q Y_\varphi = n v (\rho) \ \ , \qquad\quad Y_\rho = Y_t = Y_z= 0 \label{NOansatz}\end{equation} (that is, $Y = v(\rho) ~d\varphi \ $ or $ \ \vec Y = {\hat \varphi} ~v(\rho) /\rho \ $), \ with boundary conditions \be f(0) = v(0) = 0\ , \qquad f(\rho) \to 1 \ , \qquad v(\rho) \to 1 \qquad {\rm as} \ \rho \to \infty . \label{NObc} \end{equation} Note that, since $Y_z = Y_t = 0$, and all other fields are independent of $t$ and $z$, the electric field is zero, and the only surviving component of the magnetic field $\vec B$ is in the $z$ direction. Substituting this ansatz into the equations of motion we obtain the equations that the functions $f$ and $v$ must satisfy: \be \eqalign{ f''(\rho) + {{f'(\rho)} \over \rho} - {n^2f (\rho) \over {\rho^2}} [1-v(\rho)]^2 + \lambda \eta^2 (1-f(\rho)^2) f(\rho) &= 0 \cr v''(\rho) - {{v'(\rho)} \over \rho} + {{q^2 \eta^2}} f^2(\rho)[1-v(\rho)] &= 0 \cr} \label{NOeqs} \end{equation} In what follows we will denote the solutions to the system (\ref{NOeqs},\ref{NObc}) by $f_{NO}$ and $v_{NO}$; they are not known analytically, but have been determined numerically; for $n=1$, $\beta = 0.5$, they have the profile in Fig. \ref{fvgraphs}. \begin{figure}[tbp] \caption{\label{fvgraphs} The functions $f_{NO}$, $v_{NO}$ for a string with winding number $n=1$ (top panel) and $n=50$ (bottom panel), for $\beta \equiv 2\lambda/q^2 = 0.5$. The radial coordinate has been rescaled as in eq. (\ref{natural}), $\hat{\rho} = q\eta \rho/\sqrt{2}$. } \vskip 1truecm \epsfxsize = \hsize \epsfbox{fvgraphs.eps} \end{figure} At small $\rho$, the functions $f$ and $v$ behave as $\rho^n$ and $\rho^2$ respectively; as $\rho \to \infty$, they approach their asymptotic values exponentially with a width given by the inverse scalar mass, $m_s$, and the inverse vector mass, $m_v$, respectively, if $\beta < 4$. For $\beta >4$ the fall-off of both the scalar and the vector is controlled by the vector mass \cite{Peri93}. One case in which it is possible to find analytic expressions for the functions $f_{NO}$ and $v_{NO}$ is in the limit $n\to \infty$ \cite{AchGreHarKui94}. Inside the core of a large $n$ vortex, the functions $f$ and $v$ are \be f(\rho) = \left ( {q \over 4n} m_s m_v \rho^2 \right )^{n\over 2} e^{-qm_sm_v \rho^2 /8} \ , \qquad\qquad v(\rho) = {1 \over 4n} m_sm_v \rho^2 \end{equation} to leading order in $1/n$, and the transition to their vacuum values is controlled by a first integral $\Psi( f,f',v,v') = {\rm const \ }$. Large $n$ vortices behave like a conglomerate of ``solid'' $n=1$ vortices. The area scales as $n$, so the radius goes like $\sqrt n L_0$, where $L_0 = 2 (\sqrt{m_sm_v})^{-1}$. The transition region between the core and asymptotic values of the fields is of the same width as for $n=1$ vortices Fig. \ref{fvgraphs} shows the functions $f_{NO}, \ v_{NO}$ for $n=50$, $\beta = 0.5$ (note that for $\beta>1$ these multiply winding solutions are unstable to separation into $n=1$ vortices which repel one another). \bigskip $\bullet$ Energy considerations: The energy per unit length of such configurations (static and z-independent) is therefore \be {E} = \int d^2x \[ |D_m\Phi|^2 + \half B^2 + \lambda (\Phi^\dagger\Phi - {\eta^2\over 2})^2 \] \end{equation} where $ m,n = 1,2$ and $B=\partial_mY_n - \partial_nY_m$ is the $z$-component of the magnetic field. In order to have solutions with finite energy per unit length we must demand that, as $\rho\to\infty$, $D_\mu \Phi$, $|\Phi|^2 - {\eta^2 / 2} $ and $Y_{mn} $ all go to zero faster than $1/\rho$. The vacuum manifold (\ref{vacman}) is a circle and strings form when the asymptotic field configuration of the scalar field winds around this circle. The important point here is that there is no way to extend a winding configuration inwards from $\rho = \infty$ to the entire $xy$ plane continuously while remaining in the vacuum manifold. Continuity of the scalar field implies that it must have a zero somewhere in the $xy$ plane. This happens even if the $xy$ plane is deformed, and at all times, and in three dimensions one finds a continuous line of zeroes which signal the position of the string (a sheet in spacetime). Note that the string can have no ends; it is either infinitely long or a closed loop. The zeroes of the scalar field are forced by the non-zero topological degree of the map \be \eqalign{ S^1 \quad &\to \quad {\cal V} \cr \varphi \ \quad &\to \quad \Phi(\rho = \infty, \varphi) \ \ , \cr} \end{equation} usually called the {\it winding number} of the vortex; the resulting vortices are called topological because they are labelled by non-trivial elements of the first homotopy group of the vacuum manifold (where non-trivial means ``other than the identity element''). Thus, $\pi_1( {\cal V}) = \pi_1(S^1) \neq 1$, is a necessary condition for the existence of topological vortices. Vortices whose asymptotic scalar field configurations are associated with the identity element of $\pi_1({\cal V})$ are called non-topological. In particular, if ${\cal V}$ is simply connected, i.e. $\pi_1 ({\cal V}) = 1$, one can only have non-topological vortices. A few comments are needed at this point: \bigskip$\bullet$ Quantization of magnetic flux: Recall that $B$ is the $z$-component of the magnetic field. The magnetic flux $F_Y$ through the $xy$-plane is therefore \be F_Y \equiv \int d^2x B = \int_{\rho=\infty} {\vec Y}_\infty \cdot {\vec dl} = \int_0^{2\pi} \partial_\varphi \chi d\varphi = {2\pi n\over q} \end{equation} and is quantized in units of $2\pi / q$. This is due to the fact that $\Phi(\rho=\infty, \varphi) = \eta e^{iq\chi (\varphi)}/\sqrt{2}$, $D_\varphi \Phi = {\eta / \sqrt 2} [ i q\partial_\varphi \chi - iq Y_\varphi ] = 0$ and $\Phi$ must be singlevalued, thus $q [\chi(2\pi) - \chi(0)] = 2\pi n$. The integer $n$ is, again, the winding number of the vortex. \bigskip$\bullet$ Magnetic pressure: In an Abelian theory, the condition $\vec{\nabla} \cdot \vec {B} = 0$ implies that parallel magnetic field lines repel. A two-dimensional scale transformation ${\vec x} \to \lambda {\vec x}$ where the magnetic field is reduced accordingly to keep the magnetic flux constant, $B_\Lambda = \Lambda^{-2} B( {{\vec x} / \Lambda})$, reduces the magnetic energy $\ \int d^2x~ B^2 /2 \ $ \ by $\Lambda^2$. What this means is that a tube of magnetic lines of area $S_0$ can lower its energy by a factor of $\Lambda^2$ by spreading over an area $\Lambda^2 S_0$. Note that later we will consider non-Abelian gauge symmetries, for which $\vec{\nabla} \cdot {\vec B}\neq 0$ and the energy can also be lowered in a different way. In this case, one can think of the gauge fields as carrying a magnetic moment which couples to the ``magnetic'' field and, in the presence of a sufficiently intense magnetic field, the energy can be lowered by the spontaneous creation of gauge bosons. In the context of the electroweak model, this process is known as W-condensation \cite{AmbOle90} and its relevance for electroweak strings is explained in section \ref{stability}. \bigskip$\bullet$ Meissner effect and symmetry restoration: In the Abelian Higgs model, as in a superconductor, it is energetically costly for magnetic fields to coexist with scalar fields in the broken symmetry phase. Superconductors exhibit the Meissner effect (the expulsion of external magnetic fields), but as the sample gets larger or the magnetic field more intense, symmetry restoration becomes energetically favourable. An example is the generation of Abrikosov lattices of vortices in type II superconductors, when the external magnetic field reaches a critical value. The same phenomenon occurs in the Abelian Higgs model. In a region where there is a concentration of magnetic flux, the coupling term $ q^2 A^2 \Phi^2$ in the energy will tend to force the value of the scalar field towards zero (its value in the symmetric phase). This will be important to understand the formation of semilocal (and possibly electroweak) strings, where there is no topological protection for the vortices, during a phase transition (see section \ref{networks}). The back reaction of the gauge fields on the scalars depends on the strength of the coupling constant $q$. When $q$ is large (in a manner that will be made precise in section \ref{networks}) semilocal strings tend to form regardless of the topology of the vacuum manifold. \subsection{Stability of Nielsen-Olesen vortices} Given a solution to the classical equations of motion, there are typically two approaches to the question of stability. One is to consider the stability with respect to infinitesimal perturbations of the solution. If one can establish that no perturbation can lower the energy, then the solution is called classically stable. Small perturbations that do not alter the energy are called zero modes, and signal the existence of a family of configurations with the same energy as the solution whose stability we are investigating (e.g. because of an underlying symmetry). If one can guess an instability mode, this approach is very efficient in showing that a solution is unstable (by finding the instability mode explicitly) but it is usually much more cumbersome to prove stability; mathematically the problem reduces to an eigenvalue problem and one often has to resort to numerical methods. A stability analysis of this type for Nielsen-Olesen vortices has only been carried out recently by Goodband and Hindmarsh \cite{GooHin95a}. An analysis of the stability of semilocal and electroweak strings can be found in later sections. A second approach, due to Bogomolnyi, consists in finding a lower bound for the energy in each topological sector and proving that the solution under consideration saturates this bound. This immediately implies that the solution is stable, although it does not preclude the existence of zero modes or even of other configurations with the same energy to which the solution could tunnel semiclassically. We will now turn to Bogomolnyi's method in the case of Nielsen-Olesen vortices. \bigskip$\bullet$ Bogomolnyi limit and bounds Consider the scalar gradients: \be \eqalign{ ({D_1\Phi})^\dagger D_1\Phi + ({D_2\Phi})^\dagger D_2\Phi &= [{(D_1+iD_2)\Phi}]^\dagger (D_1+iD_2) \Phi - i [({D_1\Phi})^\dagger D_2 \Phi - ({D_2\Phi})^\dagger D_1\Phi ] \cr &= |(D_1+iD_2)\Phi|^2 -i \[ \partial_1 (\Phi^\dagger D_2\Phi) - \partial_2 (\Phi^\dagger D_1 \Phi)\] + i \Phi^\dagger [D_1, D_2]\Phi \ \ . \cr} \label{Bog0} \end{equation} Note that the second term in the RHS of (\ref{Bog0}) is the curl of the current $J_i = -i \Phi^\dagger D_i \Phi $, and that $\oint {\vec J}\cdot{\vec dl}$ tends to zero as $\rho \to \infty$ for configurations with finite energy per unit length (because $D_i\Phi$ must vanish faster than $1/\rho$). Now use the identity $[D_1, D_2] \Phi = -iq F_{12} \Phi = -iqB\Phi$. to rewrite the energy per unit length as follows: \be \eqalign{E &= \int d^2 x \[ |(D_1 \pm iD_2 )\Phi|^2 + \half B^2 \pm qB\Phi^\dagger\Phi + \lambda \Biggl(\Phi^\dagger\Phi - {\eta^2 \over 2}\Biggr)^2 \] \cr &= \int d^2 x \[ |(D_1 \pm iD_2 )\Phi|^2 + \half \Biggl \{ B \pm q \Biggl(\Phi^\dagger\Phi - {\eta^2 \over 2}\Biggr) \Biggr \} ^2 + (\lambda - \half q^2)\Biggl(\Phi^\dagger\Phi - {\eta^2 \over 2}\Biggr)^2 \]\cr &\ \ \ \ \ \ \ \ \ \pm q{\eta^2\over 2} \int B d^2x \cr} \label{Bog1} \end{equation} The last integral is the total magnetic flux, and we saw earlier that it has to be an integral multiple of $2\pi/q$, so we can write, introducing $\beta = 2\lambda/q^2$, \be E = 2\pi (\pm n) {\eta^2 \over 2} + \int \[ |(D_1 \pm iD_2 )\Phi|^2 + \half \Biggl[B \pm q \Biggl(\Phi^\dagger\Phi - {\eta^2\over 2}\Biggr) \Biggr]^2 + \half q^2 (\beta -1)\Biggl(\Phi^\dagger\Phi - {\eta^2\over 2}\Biggr)^2 \] \label{Bog2} \end{equation} where the plus or minus signs are chosen so that the first term is positive, depending on the sign of the magnetic flux. Note that, if $\beta\geq 1$ the energy is bounded below by \be E \geq \langle \Phi^\dagger\Phi \rangle q F_Y \ \ , \label{Bogbound} \end{equation} where $F_Y$ is the magnetic flux. \footnote{ When $\beta= 1$, the masses of the scalar and the vector are equal, and the Abelian Higgs model can be made supersymmetric. In general, bounds of the form (Energy) $\geq$ (constant) $\times$ (flux) are called Bogomolnyi bounds, and their origin can be traced back to supersymmetry.} If $\beta = 1$, there are configurations that saturate this bound: those that satisfy the first order {\it Bogomolnyi equations} \be (D_1 \pm iD_2 )\Phi = 0 \ \ , \qquad\qquad B \pm q (\Phi^\dagger\Phi - {\eta^2\over 2}) = 0 \qquad . \label{Bog3}\end{equation} or, in terms of $f(\rho)$ and $v(\rho)$, \be f'(\rho) + (\pm n) {v(\rho)-1 \over \rho} f(\rho) \ = \ 0 \ \ , \qquad\qquad (\pm n) v'(\rho) + {q^2\eta^2 \over 2} \rho (f^2 (\rho) -1) \ = \ 0 \label{Bog4} \end{equation} However, when $\beta >1$ there does not exist a static solution with $E = \pi |n| \eta^2$ since requiring, {\it e.g.}, $ B + q (\Phi^\dagger\Phi - \eta^2/2) = 0 $ and $(\Phi^\dagger\Phi - \eta^2/2) = 0 $ simultaneously would imply $B=0$, which is inconsistent with the condition on the total magnetic flux, $\int B d^2x = 2\pi n /q \ $. This has an effect on the stability of higher winding vortices when $\beta >1$: if $n>1$ the solution breaks into $n$ vortices each with a unit of magnetic flux \cite{Bog76}, which repel one another. If $n=1$ there are stable static solutions, but with an energy higher than the Bogomolnyi bound. This is because the topology of the vacuum manifold forces a zero of the Higgs field, and then competition between magnetic and potential energy fixes the radius of the solution. The same argument shows that $n=1$ strings are stable for every value of $\beta$. One still has to worry about angular instabilities, but a careful analysis by \cite{GooHin95a} shows there are none. The dynamics of multivortex solutions is governed by the fact that when $\beta < 1$ vortices attract, but with $\beta >1$ they repel. This can be understood heuristically from the competition between magnetic pressure and the desire to minimise potential energy by having symmetry restoration in as small an area as possible. The width of the scalar vortex depends on the inverse mass of the Higgs, $l_s$, that of the magnetic flux tube depends on the inverse vector boson mass, $l_v$. If $\beta < 1$, have $m_v > m_s$ so $l_v < l_s$ (the radii of the scalar and vector tubes). The scalar tubes see each other first - they attract. Whereas if $\beta > 1$, the vector tubes see each other first - they repel. For $\beta = 1$ there is no net force between vortices, and there are static multivortex solutions for any $n$. In the Abelian Higgs case they were explicitly constructed by Taubes \cite {JafTau80} and their scattering at low kinetic energies has been investigated using the geodesic approximation of Manton \cite{Man82} by Ruback \cite{Rub88} and, more recently, Samols \cite{Sam92}. For $\beta <1$ Goodband and Hindmarsh \cite{GooHin95a} have found bound states of two $n=1$ vortices oscillating about their centre of mass. \section{Semilocal strings} \label{semilocal} The semilocal model is obtained when we replace the complex scalar field in the Abelian Higgs model by an $N$-component multiplet, while keeping only the overall phase gauged. In this section we will concentrate on $N=2$ because of its relationship to electroweak strings, but the generalisation to higher $N$ is straightforward, and is discussed below. \subsection{The model} Consider a direct generalization of the Abelian Higgs model where the complex scalar field is replaced by an $SU(2)$ doublet $\Phi^T \ = \ ( \phi_1, \phi_2 )$. The action is \be S = \int d^4 x \left [ |(\partial_\mu - iq Y_\mu ) \Phi |^2 \ - \ {1 \over 4} Y_{\mu \nu} Y^{\mu \nu} \ - \ {\lambda } \(\Phi^\dagger \Phi - {\eta^2\over 2} \)^2 \right ] \ , \label{semilocalaction} \end{equation} where $Y_{\mu} $ is the $U(1)$ gauge potential and $Y_{\mu \nu}= \partial_{\mu}Y_{\nu} - \partial_{\nu}Y_{\mu}$ its field strength. Note that this is just the scalar sector of the GSW model for $ g=0, \ g' = g_z = 2q$, \ {\it i.e.} for $\sin^2\theta_w = 1$, and $W_\mu^a = 0$. Let us take a close look at the symmetries. The action is invariant under $G = SU(2)_{global} \times U(1)_{local}$, with transformations \be \Phi \rightarrow e^{i q \gamma (x) }\Phi = \pmatrix {e^{i q \gamma(x)} & \ 0 \cr & \cr \ 0 &e^{i q \gamma (x)} \cr } \pmatrix {\phi_1 \cr \cr \phi_2 \cr} , \ \ \ \ \ \ \quad Y_{\mu} \rightarrow Y_{\mu} + \partial_{\mu} \gamma (x) \ , \end{equation} under $U(1)_{local}$, and \be \Phi \rightarrow e^{i\alpha_a \tau ^a} \Phi = \pmatrix{ {\rm cos }{\left(\alpha \over 2 \right)} +in_3{\rm sin}{\left(\alpha \over 2 \right)} &i(n_1 \! - \! i n_2){\rm sin}{\left( \alpha \over 2 \right)} \cr & \cr i(n_1 \! + \!i n_2){\rm sin}{\left(\alpha \over 2 \right)} &{\rm cos}{\left( \alpha \over 2 \right)} -in_3{\rm sin}{\left(\alpha \over 2 \right)} } \pmatrix {\phi_1 \cr \cr \phi_2 \cr} , \ \ Y_\mu \to Y_\mu \end{equation} under $SU(2)_{global}$, where $\alpha = \sqrt {\alpha_1^2 + \alpha_2^2 + \alpha_3^2} \in [0, 4\pi)$ is a positive constant and $n_a = \alpha_a / \alpha$ is a constant unit vector. Note that a shift of the function $\gamma(x)$ by $2\pi/q$ leaves the transformations unaffected. The model actually has symmetry $ G = [SU(2)_{global} \times U(1)_{local}] /Z_2$; the $Z_2$ identification comes because the transformation with $(\alpha, \gamma)$ is identified with that with $(\alpha + 2\pi, \gamma + \pi/q)$. Once $\Phi$ acquires a vacuum expectation value, the symmetry breaks down to $H = U(1)$ exactly as in the GSW model, except for the fact that the unbroken $U(1)$ subgroup is now {\it global} (for instance, if the VEV of the Higgs is $\langle\Phi^T \rangle = \eta (0,1)/ {\sqrt 2}$, the unbroken global $U(1)$ is the subgroup with $n_1 = n_2 = 0$, $n_3=1$, $q\gamma = \alpha/2$). Thus, the symmetry breaking is $ [SU(2)_{global} \times U(1)_{local}] /Z_2 \to U(1)_{global}$. Note also that, for any {\it fixed} $\Phi_0$ a global phase change can be achieved with either a global $U(1)_{\rm local} $ transformation or a $SU(2)_{\rm global} $ transformation. {The significance of this fact will become apparent in a moment} Like in the GSW model, the vacuum manifold is the three sphere \be {\cal V} \ = \ \{\Phi \in {\rm I\!\!\! C}^2 \ | \ \Phi^\dagger \Phi = {\eta^2 \over 2} \} \ \cong \ S^3 \ \ , \end{equation} which is simply connected, so there are no topological string solutions. On the other hand, if we only look at the {\it gauged} part of the symmetry, the breaking looks like $U(1) \to 1$, identical to that of the Abelian Higgs model, and this suggests that we should have local strings. After symmetry breaking, the particle content is two Goldstone bosons, one scalar of mass $m_s =\sqrt{2\lambda} \eta$ and a massive vector boson of mass $m_v = q\eta$. In this section it will be convenient to use rescaled units throughout; after the rescaling (\ref{natural}), and dropping hats, we find \be {q^2 S} = \int d^4 x \left [ |(\partial_\mu - i Y_\mu ) \Phi |^2 \ - \ {1 \over 4} Y_{\mu \nu} Y^{\mu \nu} \ - \ {\beta \over 2 } \(\Phi^\dagger \Phi - 1 \)^2 \right ] \ , \end{equation} and, as in the Abelian Higgs case, $\beta = {m_s^2 / m_v^2} = {2\lambda / q^2}$ is the only free parameter in the model. The equations of motion \be \eqalign{ D_\mu D^\mu\Phi + \beta(|\Phi|^2 - 1)\Phi &= 0 \cr \partial^\mu Y_{\mu\nu} &= - i\Phi^\dagger{\Dbw}_{\nu}\Phi \ \ .\cr} \label{eomnatural} \end{equation} are exactly the same as in the Abelian Higgs model but replacing the scalar field by the $SU(2)$ doublet, and complex conjugation by hermitian conjugation of $\Phi$. Therefore, any solution ${\hat \Phi}(x), \ {\hat Y}_\mu (x)$ of (\ref{eom}) (in rescaled units) extends trivially to a solution $\Phi_{sl} ({x}),\ (Y_\mu)_{sl} ({x})$ of the semilocal model if we take \be \Phi_{sl} ({x}) = {\hat \Phi} ({x}) \Phi_0 \qquad \qquad (Y_\mu)_{sl} ({x}) = {\hat Y}_\mu ({x}) \end{equation} with $\Phi_0$ a constant $SU(2)$ doublet of unit norm, $\Phi^\dagger_0 \Phi_0 = 1$. In particular, the Nielsen-Olesen string can be embedded in the semilocal model in this way. The configuration \be \Phi = f_{NO}(\rho)e^{in\varphi} \Phi _0 , \ \ \ \ \ \ Y = n v_{NO}(\rho) d{ \varphi} \label{nslsol} \end{equation} remains a solution of the semilocal model with winding number $n$ provided $f_{NO}$ and $ v_{NO}$ are the solutions to the Nielsen-Olesen equations (\ref{NOeqs}). In this context, the constant doublet $\Phi_0$ is sometimes called the `colour' of the string (do not confuse with $SU(3)$ colour!). One important difference with the Abelian Higgs model is that a scalar perturbation can remove the zero of $\Phi$ at the center of the string, thereby reducing the potential energy stored in the core. Consider the energy per unit length, in these units, of a static, cylindrically symmetric configuration along the $z$-axis: \be { E \over (\eta^2/2)} = \int d^2 x \left [ {1 \over 4} (\partial _m Y_n - \partial _n Y_m )^2 + |(\partial_m - iY_m ) \Phi |^2 + {\beta \over 2} (\Phi^\dagger \Phi - 1 ) ^2 \right ] \ \label{energynatural} \end{equation} Note, first of all, that any finite energy configuration must satisfy $$ (\partial _m - i Y_m )\phi_1 \rightarrow 0 \ \ ,\ \ \ \ \ \ \ (\partial _m - iY_m )\phi_2 \rightarrow 0 \ \ , \ \ \ \ \ \ \ \bar\phi_1 \phi_1 + \bar\phi_2 \phi_2 \rightarrow 1 \ \ \ \ \ \ \ \ \ \ {\rm as }\ \ \ \rho \rightarrow \infty$$ (As before, $m,n = 1,2$ and ($\rho , \varphi$) are polar coordinates on the plane orthogonal to the string). This leaves the phases of $\phi_1$ and $\phi_2$ undetermined at infinity and there can be solutions where both phases change by integer multiples of $2 \pi$ as we go around the string; however, there is only one $U(1)$ gauge field available to compensate the gradients of $\phi_1$ and $\phi_2$, and this introduces a correlation between the winding in both components: the condition of finite energy requires that the phases of $\phi_1$ and $\phi_2$ differ by, at most, a constant, as $\rho \rightarrow \infty$. Therefore, a finite energy string must tend asymptotically to a maximal circle on $S^3$ \be \Phi \rightarrow e^{in \varphi} \pmatrix {a e^{iC}\cr \cr \sqrt {1 - a^2} } \equiv e^{in \varphi} \Phi_0 \ \ \ \ \ \ \ \ \ \ Y \rightarrow n d {\varphi} \ \ \ \ \ \ \ \ \ \ \bigl( {\rm or} \ \ {\vec Y} \to {n \over \rho} {\hat \varphi} \bigr )\ \ , \end{equation} where $0 \leq a \leq 1$ and $C$ are real constants, and determine the `colour' of the string. A few comments are needed at this point. \bigskip$\bullet$ Note that the choice of $\Phi_0$ is arbitrary for an isolated string (any value of $\Phi_0$ can be rotated into any other without any cost in energy) but the relative `colour' between two or more strings is fixed. That is, the {\it relative} value of $\Phi_0$ is significant whereas the {\it absolute} value is not. \bigskip$\bullet$ The number $n$ is the winding number of the string and, although it is not a topological invariant in the usual sense (the vacuum manifold, $S^3$, is simply connected), it {\it is} topologically conserved. The reason is that, even though any maximal circle can be continuously contracted to a point on $S^3$, all the intermediate configurations have infinite energy. The space that labels finite energy configurations is not the vacuum manifold but, rather, the {\it gauge orbit} from any reference point $\Phi_0 \in{ \cal V}$, and this space $(G_{local} / H_{local})$, is not simply connected: $\pi_1 (G_{local} / H_{local}) = \pi_1 (U(1) / 1) = {\rm \angle \!\!\! Z}$. Thus, configurations with different winding numbers are separated by infinite energy barriers, but this information is not contained in $\pi_1({\cal V})$ \footnote{The fact that the gauge orbits sit inside ${\cal V} = G/H$ without giving rise to non-contractible loops can be traced back to the previous remark that every point in the gauge orbit of $\Phi_0$ can also be reached from $\Phi_0$ with a global transformation.}. \bigskip$\bullet$ On the other hand, because $\pi_1({\cal V}) = 1$, the existence of a topologically conserved winding number does not guarantee that winding configurations are non-dissipative either. In contrast with the Abelian Higgs model, a field configuration with non-trivial winding number at $\rho = \infty$ can be extended inwards for all $\rho$ without ever leaving the vacuum manifold. {Thus, the fact that $\pi_1 (G_{local} / H_{local)} \neq 1 $ only means that finite energy field configurations fall into inequivalent sectors, but it says nothing about the existence of stable solutions within these sectors. $\bullet$ Thus, we have a situation where \be \pi_1 ({\cal V}) = \pi_1 (G/H)= \pi_1(S^3) = 1 \qquad {\rm but} \qquad \pi_1 (G_{\rm local}/H_{\rm local})= \pi_1(S^1) = {\rm \angle \!\!\! Z} \ . \label{semilocalconditionforstring}\end{equation} and the effect of the global symmetry is to eliminate the topological reason for the existence of the strings. Notice that this subtlety does not usually arise because these two spaces are the same in theories where all symmetries are gauged (like GSW, Abelian Higgs, etc.). We will now show that, in the semilocal model, the stability of the string depends on the dynamics and is controlled by the value of the parameter $\beta = 2\lambda / q^2$. Heuristically we expect large $\beta$ to mimic the situation with only global symmetries (where the strings would be unstable) , whereas small $\beta$ resembles the situation with only gauge symmetries (where we expect stable strings). \subsection {Stability} Let us first prove that there are classically stable strings in this model. We can show this analytically for $\beta = 1$ \cite{VacAch91}. Recall the expression of the energy per unit length (\ref{energynatural}). The analysis in the previous section goes through when the complex field is replaced by the $SU(2)$ doublet, and we can rewrite \be {E \over (\eta^2/2)} = 2\pi |n| \ + \ \int d^2 x \Bigl[ |D_1 \Phi \pm i D_2 \Phi |^2 \ + \ {1\over 2 } (B \pm (\Phi^{\dag} \Phi - 1 ) )^2 \ + \ {1 \over 2} (\beta - 1) (\Phi^{\dag} \Phi - 1 )^2 \Bigr] \ , \end{equation} choosing the upper or lower signs depending on the sign of $n$. Since $n$ is fixed for finite energy configurations this shows that, at least for $\beta = 1$, a configuration satisfying the Bogomolnyi equations \be (D_1 \pm i D_2) \Phi = O \ \ \ \ \ \ \ \ \ \ \ \ B \pm (\Phi^{\dag} \Phi -1) = 0, \ \ \label{semilocalbogomolnyieq} \end{equation} is a local minimum of the energy and, therefore, automatically stable to infinitesimal perturbations. But these are the same equations as in the Abelian Higgs model, therefore the semilocal string (\ref{nslsol}) automatically saturates the Bogomolnyi bound (for any `colour' $\Phi_0$). Thus, it is classically stable for $\beta = 1$. This argument does not preclude zero modes or other configurations degenerate in energy. Hindmarsh \cite{Hin92} showed that, for $\beta = 1$ there are indeed such zero modes, described below in (\ref{semilocalzeromodes}). We have just proved that, for $\beta = 1$, semilocal strings are {\it stable}. This is surprising because the vacuum manifold is simply connected and a field configuration that winds at infinity may unwind without any cost in {\it potential} energy \footnote{ In the Nielsen-Olesen case a configuration with a non-trivial winding number {\it must} go through zero somewhere for the field to be continuous. But here, a configuration like $ \Phi^T ( \rho \! =\! {\infty}) = \eta( 0, e^{i {\varphi}}) /\sqrt 2$ can gradually change to $\Phi^T (\rho \! = \!0) = \eta (1,0) \sqrt 2$ as we move towards the centre of the ``string'' without ever leaving the vacuum manifold. This is usually called `unwinding' or `escaping in the third dimension' by analogy with condensed matter systems like nematic liquid crystals.}. The catch is that, because $\pi_1 (G_{local}/H_{local}) = \pi_1 (U(1)) = {\rm \angle \!\!\! Z}$ is non trivial, leaving the $U(1)$ gauge orbit is still expensive in terms of {\it gradient} energy. As we come in from infinity, the field has to choose between unwinding or forming a semilocal string, that is, between acquiring mostly gradient or mostly potential energy. The choice depends on the relative strength of these terms in the action, which is governed by the value of $\beta$, and we expect the field to unwind for large $\beta$, when the reduction in potential energy for going off the vacuum manifold is high compared to the cost in gradient energy for going off the $U(1)$ orbits. And vice versa. Indeed, we will now show that, for $\beta > 1$, the $n=1$ vortex is unstable to perturbations in the direction orthogonal to $\Phi_0$ \cite{Hin92} while, for $\beta <1$, it is stable. For $\beta = 1$, some of the perturbed configurations become degenerate in energy with the semilocal vortex and this gives a (complex) one-parameter family of solutions with the same energy and varying core radius \cite{Hin92}. \subsubsection {The stability of strings with $\beta >1$} Hindmarsh has shown \cite{Hin92} that for $\beta >1$ the semilocal string configuration with unit winding is unstable to perturbations orthogonal to $\Phi_0$, {which make the magnetic flux spread to infinity }. As pointed out by Preskill \cite{Pre92}, this is remarkable because the total amount of flux measured at infinity remains quantized, but the flux is {\it not} confined to a core of finite size (which we would have expected to be of the order of the inverse vector mass). The semilocal string solution with $n=1$ is, in rescaled units, \be \Phi_{sl} = f_{NO} (\rho) e^{i\varphi} \Phi_0 \ \ ,\qquad \qquad \ {Y_{sl}}= v_{NO}(\rho) { d\varphi} \ \ . \label{slsol} \end{equation} However, as pointed out in \cite{Hin92}, this is not the most general static one-vortex ansatz compatible with cylindrical symmetry. Consider the ansatz \be \Phi = f (\rho) e^{i\varphi} \Phi_0 + g(\rho) e^{i m\varphi}\Phi_\perp \ \ , \qquad\qquad {Y} = {v(\rho)} {d \varphi} \ \ , \label{perturbedsl} \end{equation} with $|\Phi_0| = |\Phi_\perp| = 1$ and ${\overline \Phi_0} \Phi_\perp = 0$. The orthogonality of $\Phi_0$ and $\Phi_\perp$ ensures that the effect of a rotation can be removed from $\Phi$ by a suitable $SU(2) \times U(1)$ transformation, therefore the configuration is cylindrically symmetric. For the configuration to have finite energy we require the boundary conditions $ f(0) = g'(0) = v(0) = 0$ and $f \to 1, \ g\to 0, \ v \to 1$ as $\rho \to \infty$ We know that if $g=0$ the energy is minimised by the semilocal string configuration $f=f_{NO}, \ v=v_{NO}$, because the problem is then identical to the Abelian Higgs case. The question is whether a non-zero $g$ can lower the energy even further, in which case the semilocal string would be unstable. The standard way to find out is to consider a small perturbation of (\ref{slsol}) of the form $g= \phi(\rho) e^{i\omega t}$ and look for solutions of the equations of motion where $g$ grows exponentially, that is, where $\omega^2 <0$. The problem reduces to finding the negative eigenvalue solutions to the Schr\"odinger-type equation \be \[ -{1\over \rho} {d\over d\rho} \left( \rho {d\over d\rho} \right) + {(v(\rho)-m)^2\over \rho^2} + \beta (f(\rho)^2 -1) \] \psi(\rho) = \omega^2 \psi (\rho)\end{equation} First of all, it turns out that it is sufficient to examine the $m=0$ case only. Note that, since $0\leq v(\rho) \leq 1$, for $m>1$ the second term is everywhere larger than for $m=1$, so if we one can show that all eigenvalues are positive for $m=1$ then so are the eigenvalues for $m>1$. But for $m=1$ the problem is identical to the analogous one for instabilities in $f$ in the Abelian Higgs model, and we know there are no instabilities in that case. Therefore it is sufficient to check the stability of the solution to perturbations with $m=0$ (negative values of $m$ also give higher eigenvalues than $m=0$.) If $m=0$, the above ansatz yields \be {E \over (\eta^2 / 2)} = 2\pi \int_0^\infty \rho \left[ (f')^2 + (g')^2 + {1 \over 2\rho^2} (v')^2 + {(1 - v)^2 \over \rho^2} f^2 + {v^2 \over \rho^2} g^2 + \half \beta (f^2 + g^2 - 1)^2 \right] d\rho \end{equation} for the (rescaled) energy functional (\ref{energynatural}). Notice that a non-zero $g$ at $\rho = 0$ (where $f \neq 1$) reduces the potential energy but increases the gradient energy for small values of $\rho$. If $\beta$ is large, this can be energetically favourable (conversely, for very small $\beta$, the cost in gradient energy due to a non-zero $g$ could outweigh any reduction in potential energy). Indeed, Hindmarsh showed that there are no minimum-energy vortices of finite core radius when $\beta >1$ by constructing a one-parameter family of configurations whose energy tends to the Bogomolnyi bound as the parameter $\rho_0$ is increased: \be f(\rho) = {\rho \over \rho_0} \[ 1 + {\rho^2 \over \rho_0^2} \]^{-1/2} \qquad g(\rho) = \[ 1 + {\rho^2 \over \rho_0^2} \]^{-1/2} \qquad v(\rho) = {\rho^2 \over \rho_0^2} \[ 1 + {\rho^2 \over \rho_0^2} \]^{-1} \end{equation} The energy per unit length of these configurations is ${E} = \pi\eta^2(1 + 1/3\rho_0^2)$ which, as $\rho_0 \to \infty$, tends to the Bogomolnyi bound. This shows that any stable solution must {\it saturate} the Bogomolnyi bound, but this is impossible because, when $\beta > 1$, saturation would require $B=0$ everywhere, which is incompatible with the total magnetic flux being $2\pi /q$ (see the comment after eq. (\ref{Bog4}). While this does not preclude the possibility of a metastable solution, numerical studies have found no evidence for it \cite{Hin92,AchKuiPerVac92}. All indications are that, for $\beta >1$, the semilocal string is unstable towards developing a condensate in its core which then spreads to infinity. \begin{figure}[tbp] \caption{\label{unstablestring} A two-dimensional simulation of the evolution of a perturbed isolated semilocal string with $\beta >1$, from \cite{AchKuiPerVac92}. The plot shows the (rescaled) energy density per unit length in the plane perpendicular to the string. $\beta = 1.1$ The initial conditions include a large destabilizing perturbation in the core, $\Phi^T(t=0) = (1, f_{NO}(\rho) e^{i\varphi})$, which is seen to destroy the string.} \smallskip \epsfxsize = \hsize \epsfbox{akpv4.eps} \vskip 1truecm \end{figure} \begin{figure}[tbp] \caption{\label{stablestring} The evolution of a string with $\beta <1$. The initial configuration is the same as in Fig. \ref{unstablestring} but now, after a few oscillations, the configuration relaxes into a semilocal string, $\Phi^T = (0, f_{NO}(\rho) e^{i\varphi})$. $\beta = 0.9$} \smallskip \epsfxsize = \hsize \epsfbox{akpv1.eps} \vskip 1 truecm \end{figure} Thus, the semilocal model with $\beta >1$ is a system where magnetic flux is quantized, the vector boson is massive and yet there is no confinement of magnetic flux \footnote{Preskill has emphasized that the ``mixing'' of global and local generators is a necessary condition for this behaviour, that is, there must be a generator of $H$ which is a non-trivial linear combination of generators of $G_{\rm global}$ and $G_{\rm local}$ \cite{Pre92}.}. \subsubsection {The stability of strings with $\beta <1$} Semilocal strings with $\beta < 1$ are stable to small perturbations. Numerical analysis of the eigenvalue equations \cite{Hin92,Hin93} shows no negative eigenvalues, and numerical simulations of the solutions themselves indicate that they are stable to $z$-independent perturbations \cite{AchKuiPerVac92,AchBorLid98}, including those with angular dependence. Note that the stability to $z$-dependent perturbations is automatic, as they necessarily have higher energy. These results are confirmed by studies of electroweak string stability \cite{GooHin95b,AchGreHarKui94} taken in the limit $\theta_w \to \pi/2$. \subsubsection {$\beta = 1$ zero modes and skyrmions} \label{semilocalzeromodes} Substituting the ansatz (\ref{perturbedsl}) into the (rescaled) Bogomolnyi equations for $n=1$ gives : \be \eqalign{ f'(\rho) + {v(\rho) -1\over \rho} f(\rho) &= 0 \cr g'(\rho) + {v(\rho) \over \rho} g(\rho) &= 0 \cr v'(\rho) + \rho(f^2(\rho) + g^2(\rho) - 1) &= 0\cr} \label{semilocalbogomolnyi} \end{equation} When $\beta = 1$ we showed earlier that the semilocal string $f=f_{NO}, \ g=0, v=v_{NO}$ saturates the Bogomolnyi bound, so it is necessarily stable (since it is a minimum of the energy). There may exist, however, other solutions satisfying the same boundary conditions and with the same energy. Hindmarsh showed that this is indeed the case by noticing that the eigenvalue equation has a zero-eigenvalue solution \cite{Hin92} \be \psi = \psi_0 \exp \[ -\int_0^\rho d{\hat\rho} {{v({\hat\rho})} \over {\hat\rho}} \] \ \ , \qquad \psi_0 = {\rm const} \ , \end{equation} which signals a degeneracy in the solutions to the Bogomolnyi equations. (Note that the `colour' at infinity, $\Phi_0$, is fixed, so this is not a zero mode associated with the global $SU(2)$ transformations; its dynamics have been studied in \cite{Lee92}.) It can be shown that the zero mode exists for any value of $g$, not just $g = 0$; the Bogomolnyi equations (\ref{semilocalbogomolnyi}) are not independent since, \be g(\rho) = q_0 {f(\rho) \over \rho} \end{equation} is a solution of the second equation for any (complex) constant $q_0$. Solving the other two equations leads to the most general solution with winding number one and centred at $\rho=0$. It is labelled by the complex parameter $q_0$, which fixes the size and orientation of the vortex: \be \left( \phi_1 \atop \phi_2 \right) ={1\over \sqrt{\rho^2+|q_0|^2}} \left( q_0 \atop \rho{\rm e}^{i\varphi} \right) \exp \left\{ {1\over 2}\, u(\rho;|q_0|) \right\} \> , \label{skyrmsol} \end{equation} where $u = \ln |\Phi|^2 $ is the solution to \be \nabla^2 u + 2(1 - {\rm e}^u ) = \nabla^2 \ln (\rho^2 + |q_0|^2) \, , \qquad\qquad u \to 0 \quad {\rm as} \quad \rho \to \infty \ \ . \end{equation} If $q_0 \ne 0$, the asymptotic behaviour of these solutions is very different from that of the Nielsen-Olesen vortex; the Higgs field is non-zero at $\rho=0$ and approaches its asymptotic values like $O(\rho^{-2})$. Moreover, the magnetic field tends to zero as $B \sim 2|q_0|^2 \rho^{-4}$, so the width of the flux tube is not as well-defined as in the $q_0 = 0$ case when $B$ falls off exponentially. These $q_0 \neq 0$ solutions have been dubbed `skyrmions'. In the limit $|q_0| \to 0$, one recovers the semilocal string solution (\ref{slsol}), with $u = \ln (f^2_{NO})$, the Higgs vanishing at $\rho=0$ and approaching the vacuum exponentially fast. On the other hand, when $|q_0| \gg 1$, $u\approx 0$ the scalar field is in vacuum everywhere and the solution approximates a ${\rm I\!\!\! C} P^1$ lump \cite{Hin92, LeeSam93}. Thus, in some sense, the `skyrmions' interpolate between vortices and ${\rm I\!\!\! C} P^1$ lumps. \subsubsection{Skyrmion dynamics} We have just seen that, for $\beta = 1$ the semilocal vortex configuration is degenerate in energy with a whole family of configurations where the magnetic flux is spread over an arbitrarily large area. It is interesting to consider the dynamics of these `skyrmions' when $\beta \neq 1$ \cite{Hin93,BenBuc93}: large skyrmions tend to contract if $\beta <1$ and to expand if $\beta>1$. The timescale for the collapse of a large skyrmion increases quadratically with its size \cite{Hin93}. Thus large skyrmions collapse very slowly. Benson and Bucher \cite{BenBuc93} derived the energy spectrum of delocalized `skyrmion' configurations in 2+1 dimensions as a function of their size. More precisely, they defined an `antisize' $\chi = E_{\rm magnetic}/E_{\rm total}$ as the ratio of the magnetic energy $\int d^2 x {B}^2/2$ to the total energy (\ref{energynatural}). Note that when the flux lines are concentrated, magnetic energy is high compared to the other contributions, and vice versa. Thus, $\chi \to 0$ corresponds to the limit in which the magnetic flux lines are spread over an infinitely large area, which explains the name `antisize'. For large skyrmions - those with $\chi \leq \beta/(1+\beta)$ - they concluded that the minimum energy configuration among all delocalized configurations with antisize $\chi$ satisfies \be E(\beta, \chi) \ = \ 2\pi {\eta^2 \over 2} ~ {\beta \over \beta - \chi (\beta - 1)}\end{equation} (if $\chi > \beta / (1 + \beta)$ the analysis does not apply). Therefore, energy decreases monotonically with decreasing $\chi$ for $\beta >1$ and increases monotonically for $\beta <1$, confirming that delocalized configurations tend to grow in size if $\beta >1$ and shrink if $\beta <1$. This behaviour is observed in numerical simulations \cite{AchBorLid97}. Benson and Bucher \cite{BenBuc93} have pointed out that in a cosmological setting the expansion of the Universe could drag the large skyrmions along with it and stop their collapse. The simulations in flat space are at least consistent with this, in that they show that delocalised configurations tend to live longer when artificial viscosity is increased, but a full numerical simulation of the evolution of semilocal string networks has not yet been performed and is possibly the only way to answer these questions reliably. Finally, we stress that the magnetic flux of a skyrmion does not change when it expands or contracts (the winding number is conserved) but this does not say anything about how localized the flux is. In contrast with the Abelian Higgs case, the size of a skyrmion can be made arbitrarily large with a finite amount of energy. \subsection{Semilocal string interactions} \subsubsection{Multivortex solutions, $\beta = 1$, same colour} Multi-vortex solutions in 2+1 dimensions corresponding to parallel semilocal strings with the same colour have been constructed by Gibbons, Ruiz-Ruiz, Ortiz and Samols \cite{GibOrtRuiSam92} for the critical case $\beta = 1$. Their analysis closely follows that of \cite{JafTau80} in the case of the Abelian Higgs model, and starts by showing that, as in that case, the full set of solutions to the (second order) equations of motion can be obtained by analysing the solutions to the (first order) Bogomolnyi equations. In the Abelian Higgs model, solutions with winding number $n$ are labelled by $n$ unordered points on the plane (those where the scalar field vanishes) which, for large separations, are identified with the positions of the vortices. In the semilocal model, the solutions have other degrees of freedom, besides position, describing their size and orientation. Assuming without loss of generality that the winding number $n$ is positive, and working in temporal gauge $Y_0 = 0$, any solution with winding number $n$ is specified (up to symmetry transformations) by two holomorphic polynomials \begin{eqnarray*} P_n(z) && =\prod_{r=1}^n (z-z_r) \\ &&\equiv z^n + p_{n-1} z^{n-1} + \ldots + p_1 z + p_0 \end{eqnarray*} and \be Q_n(z) \qquad \equiv q_{n-1} z^{n-1} + \ldots + q_1 z + q_0 \end{equation} where $z = x+iy$ is a complex coordinate on the $xy$ plane. The solution for the Higgs fields is, up to gauge transformations, \be \left( \phi_1 \atop \phi_2 \right) ={{\rm e}^{{1\over 2} u (z, \overline{z})}\over \sqrt{|P_n|^2+|Q_n|^2}} \left(Q_n \atop P_n \right) \end{equation} where the function $u(z, \overline{z}) = \ln (|\phi_1|^2 + |\phi_2|^2)$ must satisfy \be \nabla^2 u + 2( 1 - {\rm e}^u ) = \nabla^2 \ln (|P_n|^2 + |Q_n|^2) \, , \end{equation} and tend to 0 as $|z| \to \infty$. Although its form is not known explicitly, ref. \cite{GibOrtRuiSam92} proved the existence of a unique solution to this equation for every choice of $P_n$ and $Q_n$ (if $P_n$ and $Q_n$ have a common root then $\exp [u/2]$ has a zero there, so the expression for the Higgs field is everywhere well-defined). The gauge field can then be read off from the Bogomolny equations (\ref{semilocalbogomolnyieq}). This generalises (\ref{skyrmsol}) to arbitrary $n$. The coefficients of $P_n(z)$, $Q_n(z)$ parametrise the moduli space, ${\rm I\!\!\! C}^{2n}$. The Nielsen-Olesen vortex has $Q_n=0$. If $P_n \neq 0$, then in regions where $|Q_n| << |P_n|$ one finds \be |\phi_1| \sim 1 - \half \left| {Q_n \over P_n} \right|^2 \ \ , \qquad |\phi_2| \sim \left| {Q_n\over P_n} \right| \ \ , \qquad v \sim 1 - \left| {Q_n \over P_n} \right|^2 \end{equation} indicating that the scalar fields fall off as a power law, as opposed to the usual exponential fall off found in NO vortices. The same is true of the magnetic field. The low energy scattering of semilocal vortices and skyrmions with $\beta = 1$ was studied in \cite{LeeSam93} in the geodesic approximation of \cite{Man82}. The behaviour of these solitons was found to be analogous to that of ${\rm I\!\!\! C} P^1$ lumps but without the singularities, which are smoothed out in the core. \subsubsection{Interaction of parallel strings, $\beta <1$, different colours} Ref. \cite{AchKuiPerVac92} carried out a numerical study in two dimensions of the interaction between stable ($\beta<1$) strings with different ``colour'' with non-overlapping cores. It was found that the strings tend to radiate away their colour difference in the form of Goldstone bosons, and there is little or no interaction observed. The position of the strings remains the same during the whole evolution while the fields tend to minimize the initial relative $SU(2)$ phase (see figure \ref{interaction}). Thus, we expect interactions betwen infinitely long semilocal strings with different colours to be essentially the same as for Nielsen-Olesen strings. This expectation is confirmed by numerical simulations of two- and three-dimensional semilocal string networks \cite{AchBorLid97,AchBorLid98}, discussed in \ref{networks}. \begin{figure}[tbp] \epsfxsize = \hsize \epsfbox{akpv8.eps} \smallskip \caption{\label{interaction} A numerical simulation of the interaction between two parallel semilocal strings with different `colour', from Ref. \cite{AchKuiPerVac92}. The initial configuration has one string with $\Phi_1^T = (0, f(\rho_1) e^{i\varphi_1} )$ and the other with $\Phi_2^T = (if(\rho_2) e^{i\varphi_2}, 0)$, where $(\rho_i, \varphi_i)$ are polar coordinates centred at the cores of each string. The energy density of the string pair is plotted in the plane perpendicular to the strings. The colour difference is radiated away in the form of Goldstone bosons, and the strings cores remain at their initial positions. $\beta = 0.5$. } \vskip 1truecm \end{figure} \subsection {Dynamics of string ends} Note that, in contrast with Nielsen-Olesen strings, there is no topological reason that forces a semilocal string to continue indefinitely or form a closed loop. Semilocal strings can end in a ``cloud'' of energy, which behaves like a global monopole \cite{Hin92}. Indeed, consider the following asymptotic configuration for the Higgs field: \be \Phi = {\eta \over \sqrt 2} \(\begin{array}{c} \cos\half\theta \\ \sin\half\theta\, e^{i\varphi} \end{array}\) \end{equation} which is ill-defined at $\theta = \pi$ and at $r=0$. We can make the configuration regular by introducing profile functions such that the Higgs field vanishes at those points: \be \Phi = {\eta \over \sqrt 2} \(\begin{array}{c} h_1(r,\theta)\cos\half\theta \\ h_2(r,\theta)\sin\half\theta\, e^{i\varphi} \end{array}\) \end{equation} where $h_1$ and $h_2$ vanish at $r=0$ and $h_2(r,\pi)=0$. This configuration describes a string in the $z<0$ axis ending in a monopole at $z=0$. At large distances, $r>>1$, the Higgs field is everywhere in vacuum (except at $\theta \approx \pi$) and we find $\Phi^\dagger {\vec \tau} \Phi \sim \vec x$, just like for a Hedgehog in $O(3)$ models. On the other hand, the configuration for the gauge fields resembles that of a semi-infinite solenoid; the string supplies U(1) flux which spreads out from $z=0$. This is the $\theta_w \to \pi/2$ limit of a configuration first discussed by Nambu \cite{Nam77} in the context of the GSW model -see section \ref{zoo} - but here the energy of the monopole is linearly divergent because there are not enough gauge fields to cancel the angular gradients of the scalar field. Angular gradients provide an important clue to understand the dynamics of string ends. If $\beta < 1$, numerical simulations show that string segments grow to join nearby segments or to form loops (see figures \ref{loopformation} and \ref{semi4}) \cite{AchBorLid98}. This confirms analytical estimates in refs. \cite{GibOrtRuiSam92,Hin93}. In other cases the string segment collapses under its own tension, with the monopole and antimonopole at the ends annihilating each other. \begin{figure}[t!] \centering \leavevmode\epsfysize=7.5cm \epsfbox{sframe70.ps}\\ \vspace*{5pt} \leavevmode\epsfysize=7.5cm \epsfbox{sframe80.ps}\\ \vspace*{5pt} \caption[looopformation]{\label{loopformation} Loop formation from semilocal string segments. The figure shows two snapshots, at $t = 70$ and $t = 80$, of a $64^3$ numerical simulation of a network of semilocal strings with $\beta = 0.05$ from Ref.\cite{AchBorLid98}, where the ends of an open segment of string join up to form a closed loop (see section \ref{networks} for a discussion of the simulations). Subsequently the loops seem to behave like those of topological cosmic string, contracting and disappearing.} \vskip 1 truecm \end{figure} \begin{figure}[t!] \centering \leavevmode\epsfysize=7.5cm \epsfbox{zoom60_bw.eps}\\ \vspace*{5pt} \leavevmode\epsfysize=7.5cm \epsfbox{zoom70_bw.eps}\\ \vspace*{5pt} \caption[semi4]{\label{semi4} {\it The growth of string segments to form longer strings}. The figure shows two snapshots, at time $t=60$ and $t=70$ of a large $256^3$ numerical simulation of a network of semilocal strings with $\beta = 0.05$ from Ref. \cite{AchBorLid98}. Note several joinings of string segments, e.g.~two separate joinings on the long central string, and the disappearance of some loops. The different apparent thickness of strings is entirely an effect of perspective. The simulation was performed on the Cray T3E at the National Energy Research Scientific Computing Center (NERSC). See section \ref{networks} for a discussion.} \vskip 1 truecm \end{figure} \subsection{Numerical simulations of semilocal string networks} \label{networks} As the early Universe expanded and cooled to become what we know today it is very likely that it went through a number of phase transitions where topological (and possibly non-topological) defects are expected to have formed according to the Kibble mechanism \cite{Kib76,Zur85,reviews}. Although the cosmological evidence for the existence of such defects remains unclear \cite{cosm}, there is plenty of experimental evidence from condensed matter systems that networks of defects do form in symmetry breaking phase transitions \cite{NATO}, the most recent confirmation coming from the Lancaster-Grenoble-Helsinki experiments in vortex formation in superfluid Helium \cite{nature}. An important question is whether semilocal (and electroweak) strings are stable enough to form in a phase transition. We defer discussion of the electroweak case to section \ref{cosmologicalapplications}. Here we want to review recent numerical simulations of the formation and evolution of a network of $\beta<1$ semilocal strings \cite{AchBorLid97,AchBorLid98,AchBorLid98b} which show that such strings should indeed form in appreciable numbers in a phase transition. The results suggest that, even if no vortices are formed immediately after $\Phi$ has acquired a non-zero vacuum expectation value, the interaction between the gauge fields and the scalar fields is such that vortex formation does eventually occur simply because it is energetically favourable for the random distribution of magnetic fields present after the phase transition to become concentrated in regions where the Higgs field has a value close to that of the symmetric phase. Even though they do not account for the expansion of the Universe, these simulations represent a first step towards understanding semilocal string formation in cosmological phase transitions and they have already provided very interesting insights into the dynamical evolution of such a network. \subsubsection{Description of the simulations} >From a technical point of view, the numerical simulation of a network of semilocal strings has additional complications over that of $U(1)$ topological strings. Because there are not enough gauge degrees of freedom to cancel all of the scalar field gradients, the existence of string cores depends crucially on the way the fields (scalar and gauge) interact. Another problem, generic to all non-topological strings, is that the winding number is not well defined for configurations where the scalar is away from a maximal circle in the vacuum manifold, and this makes the identification of strings much more difficult than in the case of topological strings. The strategy proposed in \cite{AchBorLid97} to circumvent these problems is to follow the evolution of the gauge field strength in numerical simulations, since the field strength provides a gauge invariant indicator for the presence of vortices. The initial conditions are obtained by an extension of the Vachaspati-Vilenkin algorithm \cite{VacVil84} appropriate to non-topological defects, plus a short period of dynamical evolution including a dissipation term (numerical viscosity) to aid the relaxation of configurations in the `basin of attraction' of the semilocal string. As with any new algorithm, it is essential to check that it reproduces previously known results accurately, and this has been done in \cite{AchBorLid97}. Note that setting $\phi_2=0$ in the semilocal model obtains the Abelian Higgs model, thus comparison with topological strings is straightforward, and it is used repeatedly as a test case, both to check the simulation techniques and to minimise systematic errors when quoting formation rates. In particular, the proposed technique is tested in a two-dimensional toy model (representing parallel strings) in three different ways: a) restriction to the Abelian Higgs model gives good agreement with analytic and numerical estimates for cosmic strings in \cite{VacVil84}; b) the results are robust under varying initial conditions and numerical viscosities (see Figure \ref{semi2}), and c) they are in good agreement with previous analytic and numerical estimates for semilocal string formation in \cite{AchKuiPerVac92,Hin93}. The results are summarized in Fig. \ref{semi6}. We refer the reader to refs. \cite{AchBorLid97,AchBorLid98,AchBorLid98b} for details; however, a few comments are needed to understand those figures. $\bullet$ The study takes place in flat spacetime. Temporal gauge and rescaled units (\ref{natural}) are chosen. Gauss' law, which here is a constraint derived from the gauge choice $Y_0=0$, is used to test the stability of the code. $\bullet$ Space is discretized into a lattice with periodic boundary conditions. The equations of motion (\ref{eomnatural}) are solved numerically using a standard staggered leapfrog method; however, to reduce its relaxation time an {\it ad hoc} dissipation term was added to each equation ($\eta \dot{\Phi}$ and $\eta \dot{Y}_i$ respectively). A range of strengths of dissipation was tested, and it did not significantly affect the number densities obtained. The simulations displayed in this section all have have $\eta = 0.5$. $\bullet$ The number density of defects is estimated by an extension of the Vachaspati-Vilenkin algorithm \cite{VacVil84} by first generating a random initial configuration for the scalar fields drawn from the vacuum manifold, which is not discretised, and then finding the gauge field configuration that minimizes the energy associated with (covariant) gradients\footnote{In fact, it turns out that the energy-minimization condition is redundant, since the early stages of dynamical evolution carry out this role anyway.}. If space is a grid of dimension $N^3$, the correlation length is chosen to be some number $p$ of grid points ($p=16$ in \cite {AchBorLid97,AchBorLid98}; the size of the lattice is either $N=64$ or $N=256$.) To obtain a reasonably smooth configuration for the scalar fields, one throws down random vacuum values on a $(N/p)^3$ subgrid; the scalar field is then interpolated onto the full grid by bisection. Strings are always identified with the location of magnetic flux tubes. For cosmic strings, the two-dimensional toy model accurately reproduces the formation rates of \cite{VacVil84}. For semilocal strings, on the other hand, the initial configurations generated in this way have a complicated flux structure with extrema of different values (top panel of Fig. \ref{semi1}), and it is far from clear which of these, if any, might evolve to form semilocal vortices; in order to resolve this ambiguity, the initial configurations are evolved forward in time. As anticipated, in the unstable regime $\beta > 1$ the flux quickly dissipates leaving no strings. By contrast, in the stable regime $\beta < 1$ stringlike features emerge when configurations in the ``basin of attraction''of the semilocal string relax unambiguously into vortices (bottom panel of Fig. \ref{semi1}). Since the initial conditions are somewhat artificial, the results were checked against various other choices of initial conditions, in particular different initial conditions for the gauge field and also thermal initial conditions for the scalar field (see Fig. \ref{semi2}). All the initial conditions in \cite{AchBorLid97,AchBorLid98} had zero initial velocities for the fields. Initial conditions with non-zero field momenta have not yet been investigated. \begin{figure*} \centering \leavevmode\epsfysize=10.9cm \epsfbox{semi_fig2a_bw.eps}\\ \leavevmode\epsfysize=10.9cm \epsfbox{semi_fig2b_bw.eps}\\ \caption[semi1]{\label{semi1} The flux tube structure in a two-dimensional semilocal string simulation with $\beta = 0.05$, from Ref \cite{AchBorLid97}. The upper panel ($t=0$) shows the initial condition after the process described in the text. The lower panel shows the configuration resolved into five flux tubes by a short period of dynamical evolution ($t=100$). These flux tubes are semilocal vortices.} \vskip 1 truecm \end{figure*} \subsubsection {Results and discussion} These simulations give very important information on the dynamics and evolution of a network of semilocal strings. In particular, they confirm our discussion in the previous subsection of the behaviour of the ends of string segments, and of strings with different colours. String segments are seen to grow in order to join nearby ones or form closed loops, and very short segments are also observed to collapse and disappear. The colour degrees of freedom do not seem to introduce any new forces between strings. Because the strings tend to grow or form closed loops, time evolution makes the network resemble more and more a network of topological strings (NO vortices) but with lower number densities\footnote{However, one important point is that no intersection events were observed in the semilocal string simulations, so the rate of reconnection has not been determined.} Note that the correlation length in the simulations is constrained to be larger than the size of the vortex cores, to avoid overlaps. This results in a minimal value of the parameter $\beta$ of around 0.05 (if $\beta$ is lowered further, the scalar string cores become too wide to fit into a correlation volume, in contradiction with the vacuum values assumed in a Vachaspati-Vilenkin algorithm). Figure \ref{semi6} shows the results for seven different values of $\beta$ by taking several initial configurations on a $64^3$ grid smoothed over every $16$ grid-points. As expected, for $\beta < 1$ the formation rate depends on $\beta$, tending to zero as $\beta$ tends to 1. The ratio of semilocal string density to cosmic string density in an Abelian Higgs model for the same value of $\beta$ is {less than} but of order one. {For} the lowest value of $\beta$ simulated ($\beta = 0.05$), the semilocal string density is about one third {of} that of cosmic strings. \begin{figure} \centering \leavevmode\epsfysize=6cm \epsfbox{semi_fig3.eps}\\ \caption[semi2]{\label{semi2} A test of the sensitivity of the results to the choice of initial conditions in a two-dimensional simulation with the algorithm proposed in section \ref{networks}. The plot shows the number of semilocal strings formed per initial two-dimensional correlation volume. Each point is an average over ten simulations. Squares indicate that the vacuum initial conditions described in the text were used, while open circles indicate that non-vacuum (thermal) initial conditions were used. Both sets of initial conditions are seen to give comparable results. Statistical results are derived from a large suite of simulations (700 in all) carried out on a $64^3$ grid (from Ref. \cite{AchBorLid97}) } \vskip 1 truecm \end{figure} \begin{figure}[t] \centering \leavevmode\epsfysize=6cm \epsfbox{semi3_fig3.eps}\\ \caption[semi6]{\label{semi6} The ratio of lengths of semilocal and cosmic strings as a function of the stability parameter $\beta$, from \cite{AchBorLid98}.} \vskip 1 truecm \end{figure} One final word of caution about the possible cosmological implications of these simulations. We mentioned above that numerical viscosity was introduced to aid the relaxation of configurations close to the semilocal string. In an expanding Universe the expansion rate would provide some viscosity, though $\eta$ would typically not be constant. This may have an important effect on the production of strings. Indeed, note the different numbers of upward and downward pointing flux tubes in Fig. \ref{semi1}, despite the zero net flux boundary condition. The missing flux resides in the smaller `nodules', made long-lived by the numerical viscosity; these are none other than the `skyrmions' described in section \ref{semilocal}. As was explained there, the natural tendency of skyrmions when $\beta<1$ is to collapse into strings, but the timescale for collapse increases quadratically with their size and Benson and Bucher \cite{BenBuc93} have argued that the effect of the expansion could stop the collapse of large skyrmions almost completely. On the other hand one expects skyrmions to be formed with all possible sizes, so the effect of the expansion on the number density of strings remains an open question. Another important issue that has not yet been addressed is whether these semilocal networks show scaling behaviour, and whether reconnections are as rare as the above simulations suggest. Both would have important implications for cosmology. However, the answer to these and other questions may have to wait until full numerical simulations are available. \subsection{Generalisations and final comments} i) Charged semilocal vortices The semilocal string solution described earlier in this section is {\it strongly} static and z-independent, by which we mean that $D_t(\Phi) = D_z(\Phi) = 0$. It is possible to relax these conditions while still keeping the Lagrangian and the energy independent of $z$ The idea is that, as we move along the $z$-direction, the fields move along the orbit of the global symmetries; in other words, Goldstone bosons are excited. Abraham has shown that it is possible to construct semilocal vortices with finite energy per unit length carrying a global charge \cite{Abr93} in the Bogomolnyi limit $\beta = 1$ \footnote{By contrast, charged solutions with $D_0(\Phi) \neq 0$ in the Abelian Higgs model have infinite energy per unit length \cite{JulZee75}.}. They satisfy a Bogomolnyi-type bound and are therefore stable. Perivolaropoulos \cite{Per94} has constructed spinning vortices (however these have infinite energy per unit length). ii) {Semilocal models with $SU(N)_{\rm global} \times U(1)_{\rm local}$ symmetry} The generalization of semilocal strings to so-called Extended Abelian Higgs models with an N-component multiplet of scalars whose overall phase is gauged is straightforward \cite{VacAch91,Hin92}, and has been analysed in detail in \cite{Hin93,GibOrtRuiSam92}. The strings are stable (unstable) for $\beta < 1$ ($\beta >1$) and for $\beta = 1$ they are degenerate in energy with skyrmionic configurations labelled by an $N-1$ complex vector. For winding $n$, and widely separated vortices, the $Nn$ complex parameters that characterize the configurations can be thought of as the $n$ positions in ${\rm I\!\!\, R}^2 \sim {\rm I\!\!\! C}$ and the $(N-1)n$ `orientations'. iii) Semilocal monopoles and generalized semilocality We have seen that semilocal strings have very special properties arising from the fact that $\pi_1(G/H) = 0$ but $\pi_1(G_{\rm local}/H_{\rm local}) \neq 0$. An immediate question is whether it is possible to construct other non-topological defects such that \be \pi_k(G/H) = 0 \qquad\qquad {\rm but} \qquad\qquad \pi_k(G_{\rm local}/H_{\rm local}) \neq 0 \ \ . \label{semilocalcondition} \end{equation} This possibility would be particularly interesting in the case of monopoles, $k=2$, since they might retain some of the features of global monopoles, in particular a higher annihilation rate in the early Universe. Surprisingly, the answer seems to be negative. Within a very natural set of assumptions, it was shown in \cite{VacAch91} that the condition (\ref{semilocalcondition}) can only be satisfied if the gauge group $G_{\rm local}$ is Abelian, and therefore one cannot have semilocal monopoles (nor any other defects satisfying conditions (\ref{semilocalcondition}) with $k>1$). However, Preskill has remarked that it is possible to define a wider concept of semilocality \cite{Pre92} by considering the larger approximate symmetry $G_{\rm approx}$ which is obtained in the limit where gauge couplings are set to zero. The symmetry $G_{\rm approx} $ is partially broken to the exact symmetry $ G \sim G_{\rm local} \times G_{global}$ (modulo discrete transformations) when the gauge couplings are turned on It is then possible to have generalized semilocal monopoles associated with non-contractible spheres in $G_{\rm local} /H_{\rm local}$ which are contractible in the approximate vacuum manifold $G_{\rm approx}/H_{\rm approx}$ even though they are still non-contractible in the exact vacuum manifold $G/H$. Another obvious possibility is to have topological monopoles with ``colour'', by which we mean extra global degrees of freedom, if the symmetry $G \sim G_{\rm global} \times G_{\rm local}$ is such that the gauge orbits are non-contractible two-spheres, $\pi_2 (G_{\rm local}/H_{\rm local}) \neq 1$. Given that there are no semilocal monopoles \cite{VacAch91}, these monopoles must have $\pi_2 (G/H) \neq 1$, so they are topologically stable, and they have additional global degrees of freedom. iv) {Semilocal defects and Hopf fibrations} In the semilocal model, the action of the gauge group fibres the vacuum manifold $S^3$ as a non-trivial bundle over $S^2 \sim {\rm I\!\!\! C} P^1$, the Hopf bundle. The fact that this bundle is non-trivial is at the root of conditions (\ref{semilocalconditionforstring}), and is ultimately the reason why the topological criterion for the existence of strings fails. In view of this, Hindmarsh \cite{Hin93} has proposed an alternative definition of a semilocal defect: it is a defect in a theory whose vacuum manifold is a non-trivial bundle with fibre $G_{\rm local}/H_{\rm local}$. Extended Abelian Higgs models \cite{Hin93} are similarly related to the fibrations of the odd-dimensional spheres $S^{2N-1}$ with fibre $S^1$ and base space ${\rm I\!\!\! C} P^{N-1}$. A natural question to ask is if the remaining Hopf fibrations of spheres can also be realised in a field theoretic model. This question was answered affirmatively in \cite{HinHolKepVac93} for the $S^7 \ {S^3 \atop \rightarrow}\ S^4$ fibration in a quaternionic model. Other non-trivial bundles were also implemented in this paper, but to date the field theory realisation of the $S^{15} \ {S^7 \atop \rightarrow} \ S^8$ Hopf bundle remains an open problem. v) Monopoles and textures in the semilocal model: Since the gauge field is Abelian, ${\rm div} \vec B =0$, and isolated magnetic monopoles are necessarily singular in semilocal models. The only way to make the singularity disappear is by embedding the theory in a larger non-Abelian theory which provides a regular core, or by putting the singularity behind an event horizon \cite{GibOrtRuiSam92}. One important question that has not yet been addressed is if the scalar gradients in these spherical monopoles make them unstable to angular collapse into a flux tube. A related system where this happens is in $O(3)$ global monopoles where the spherically symmetric configuration is unstable. In the semilocal case, it is possible that the pressure from the magnetic field might prevent the instability towards angular collapse. Finally, note that, because $\pi_3(S^3) = {\rm \angle \!\!\! Z}$, there is also the possibility of textures in the semilocal model (\ref{semilocalaction}). In contrast with purely scalar $O(4)$ models, their collapse seems to be stopped by the pressure from the magnetic field \cite{Hin93}. Of course they can still unwind by tunnelling. vi) We should point out that systems related to the semilocal model have been studied in condensed matter. In \cite{BurKop87}, the system was an unconventional superconductor where the role of the global SU(2) group was played by the spin rotation group. In \cite{Vol84} the hypothetical case of an ``electrically charged'' A-phase of $^3 \rm He$, {\it i.e.} a superconductor with the properties of $^3 \rm He$-A, was considered (see section \ref{lightening} for a brief discussion of the A and B phases of $^3 \rm He$). In this case the global group was SO(3), the group of orbital rotations. Both papers discussed continuous vortices in such superconductors, which correspond to the ``skyrmions'' discussed here. \section{Electroweak strings} \label{ewsection} In this section we introduce electroweak strings. There are two kinds: one, more precisely known as the Z-string, carries Z-magnetic flux, and is the type that was discussed by Nambu and that becomes stable as it approaches the semilocal limit. It is associated with the subgroup generated by $$ T_Z = n^aT^a - \sin^2\theta_w Q \ . $$ There are other strings in the GSW model that carry $SU(2)$ magnetic flux, called $W$-strings. There is a one-parameter family of W strings which are all gauge equivalent to one another, and they are all unstable. They are generated by a linear combination of the SU(2) generators $T^+$ and $T^-$. These will be discussed in more detail in the next section. \subsection{The Z string} Modulo gauge transformations, the configuration describing a straight, infinitely long Z-string along the $z$-axis is \cite{Vac92}: \be \eqalign{ \Phi &= {\eta \over {\sqrt{2}}} f_{NO}(\rho) e^{i\varphi} \pmatrix{0\cr 1\cr} \ , \cr \ \ \ Z &= - {2 \over g_z} {{v_{NO}(\rho)}} d\varphi \quad{\rm or, \ in \ vector \ notation, \ } {\vec Z} = \( -{2 \over g_z} \){{v_{NO}(\rho)} \over {\rho}} {\hat \varphi} \cr A_\mu &= W_\mu^\pm = 0 \cr} \label{Zstring} \end{equation} where $f$ and $v$ are the Nielsen-Olesen profiles that solve the equations (\ref{NOeqs}). It is straightforward to show that this is a solution of the bosonic equations of motion (alternatively, one can show that it is an extremum of the energy \cite{Vac92}). Equations (\ref{Zstring}) describe a string with unit winding. The solutions with higher winding number can be constructed in an analogous way, but note that the winding number is not a topological invariant. The unstable string can decay by unwinding until it reaches the vacuum sector. The solution (\ref{Zstring}) reduces to the semilocal string in the limit $\sin^2\theta_w = 1$, and therefore it is classically stable for $\beta < 1$ and unstable for $\beta > 1$ (see section \ref{stability}), where $\beta$ is now the ratio between the Higgs mass, $\sqrt 2 \lambda \eta$ and the Z-boson mass $g_z \eta /2$, thus \be \beta = { 8 \lambda \over g_z^2 }\end{equation} The Z-string configuration is axially symmetric, as it is invariant under the action of the generalised angular momentum operator \be M_z = L_z + S_z + I_z\end{equation} where $L_z$, $S_z$ and $I_z$ are the orbital, spin and isospin parts, respectively, defined in section \ref{chernsimons}. The Z-string carries a Z-magnetic flux \be F_Z = {4\pi \over g_z }\end{equation} thus particles whose Z charge is not an integer multiple of $g_z / 2$ will have Aharonov-Bohm interactions with the string (see section \ref{scattering}). The Z-string can terminate on magnetic monopoles (such configurations are discussed in section \ref{zoo}). When a string terminates, the discrete Aharonov-Bohm interaction can be smoothly deformed to the trivial interaction. The smoothness is provided by the presence of the magnetic flux of the monopole. \bigskip$\bullet$ Note that, in the background given by (\ref{Zstring}), the covariant derivative becomes \be {\bf d}_\mu \equiv D_\mu|_{\rm Z-string} = \partial_\mu + i{g_z \over 2} \[ -2(T^3 - Q\sin^2 \theta_w) \] Z_\mu \label{smalldmu} \end{equation} in particular, left and right fermion fields couple to $Z_\mu$ with different strengths, since the effective Z-charge \be {\bf q} = -2(T^3 - Q\sin^2 \theta_w) \label{zcharge} \end{equation} has different values, $q_R = q_L \pm 1$. (Note that $\bf q$ is proportional to the string generator $T_z$, defined in equation (\ref{QandTzgenerators}). The proportionality factor has been introduced for later convenience). This will be important in section \ref{scattering}. Note also that, for the Higgs field, \be {\bf q} = {\rm diag} (-\cos 2\theta_w, 1) \end{equation} \bigskip $\bullet$ Ambj{\o}rn and Olesen \cite{AmbOle89} and, more recently, Bimonte and Lozano \cite{BimLoz94a} have derived Bogomolnyi-type bounds for periodic configurations in the GSW model. They consider static configurations such that physical observables are periodic in the $xy$-plane and cylindrically symmetric in each cell. If $A$ is the area of the basic cell, they find that the energy (per unit length) satisfies \be \cases{ {E} \geq ({1 / 2e^2}) m_W^2(2 g'F_Y - m_W^2 A) &if $m_H \geq m_Z$ \cr &\cr {E} \geq ({1 /2e^2}) ({m_W m_H / m_Z})^2 (2 g'F_Y - m_W^2 A) &if $m_H < m_Z$ \cr } \label{BiLobound} \end{equation} where $F_Y$ is the magnetic flux of the hypercharge field through the cell. Note that the top line of (\ref{BiLobound}) reduces to the familiar $E \geq \langle \Phi^\dagger\Phi \rangle q F_Y$ for the Abelian Higgs and semilocal case in the $g \to 0$ limit (with $q = g'/2$). In the non-Abelian case the bound involves an area term and therefore does not admit a topological interpretation. In the Bogomolnyi limit, $m_H = m_Z$, the bound is saturated for configurations satisfying the first order Bogomolnyi equations \be \eqalign{ D_1+iD_2 \Phi &= 0 \cr Y_{12} + {g'\over 2} \left( \Phi^\dagger\Phi - {\eta^2 \over 2\sin^2\theta_w} \right) &= 0 \cr W_{12}^a + {g \over 2}\Phi^\dagger \tau^a \Phi &= 0 \cr} \end{equation} A solution to these equations describing a lattice of Z-strings was constructed in \cite{BimLoz94a}. Other periodic configurations with symmetry restoration had been previously found in the presence of an external magnetic field in \cite{AmbOle89}. \section{The zoo of electroweak defects} \label{zoo} The electroweak Z-string is one member in the zoo of electroweak defects. Other members include the electroweak monopole, dyon and the W-string. The latter fall in the class of ``embedded defects'' and this viewpoint provides a simple way to characterize them. The electroweak sphaleron is also related to the electroweak defects. \subsection{Electroweak monopoles} \label{ewmonopoles} To understand the existence of magnetic monopoles in the GSW model, recall the following sequence of facts: \begin{itemize} \item The Z-string does not have a topological origin and hence it is possible for it to terminate. \item As the hypercharge component of the Z-field in the string is divergenceless it cannot terminate. Therefore it must continue from within the string to beyond the terminus. \item However, beyond the terminus, the Higgs is in its vacuum and the hypercharge magnetic field is massive. Then, if the massive hypercharge flux was to continue beyond the string, it would cost an infinite amount of energy and this is not possible. \item The only means by which the hypercharge field can continue beyond the terminus is in combination with the SU(2) fields such that it forms the massless electromagnetic magnetic field. \end{itemize} So the terminus of the Z-string is the location of a source of electromagnetic magnetic field, that is, a magnetic monopole \cite{Nam77}. We now make this argument more quantitative. Assume that we have a semi-infinite Z-string along the $-z$ axis with terminus at the origin (see Fig. \ref{fluxbalance}). Let us denote the A- and Z- magnetic fluxes through a spatial surface by $F_A$ and $F_Z$. These are given in terms of the W- and Y- fluxes by taking surface integrals of the field strengths (see eqs. (\ref{amunu}), (\ref{zmunu})). Therefore \be F_Z = \cos\theta_w F_n - \sin\theta_w F_Y \ , \ \ \ \ F_A = \sin\theta_w F_n + \cos\theta_w F_Y \ , \label{fzfa} \end{equation} where we have denoted the SU(2) flux (parallel to $n^a$ in group space) by $F_n$ and the hypercharge flux by $F_Y$. \begin{figure}[tbp] \caption{\label{fluxbalance} The outgoing hypercharge flux of the monopole passing through the surface ${\Sigma -S}$ should equal the incoming hypercharge flux through the Z-string. } \vskip 1truecm \epsfxsize = \hsize \epsfbox{fluxbalance.eps} \end{figure} Now consider a large sphere $\Sigma$ centered on the string terminus. The field configuration is such that there is only A-flux through $\Sigma$ except near the South pole ($S$) of $\Sigma$, where there is only a $Z$ magnetic flux. Hence, \be F_Z |_{\Sigma -S} = 0 \ , \ \ F_A |_S = 0 \ . \label{fzfaonsurfaces} \end{equation} Together with (\ref{fzfa}) this gives, \be F_n |_{\Sigma -S} = \tan\theta_w F_Y |_{\Sigma -S} \ , \ \ \ F_n |_S = - \cot \theta_w F_Y |_S \ . \label{fnfyfnfy} \end{equation} The hypercharge flux must be conserved as it is divergenceless. So \be F_Y |_{\Sigma -S} = - F_Y |_S \equiv F_Y \ , \label{fy} \end{equation} and, inserting this and (\ref{fnfyfnfy}) in (\ref{fzfa}) yields \be F_A |_{\Sigma -S} = {{F_Y} \over {\cos\theta_w}} \ , \ \ \ F_Z |_S = {{F_Y} \over {\sin\theta_w}} \ . \label{fafyfzfy} \end{equation} Now the flux in the $Z-$string along the $-z$ axis is quantized in units of $4\pi /g_z$ (recall $g_z = e/\cos\theta_w\sin\theta_w$ gives the coupling of the Z boson to the Higgs field). Therefore, for the unit winding string, \be F_Z |_S = {{4\pi} \over {g_z}} \ . \label{fzs} \end{equation} Then (\ref{fafyfzfy}) yields, \be F_Y = {{4\pi} \over {g_z}} \sin\theta_w \ , \ \ \ F_A |_{\Sigma -S}= {{4\pi} \over {g_z}} \tan\theta_w = {{4\pi} \over {e}} \sin^2 \theta_w \ \label{fyfa} \end{equation} Hence the terminus of the string has net A-flux emanating from it and hence it is a magnetic monopole. The electromagnetic flux of the electroweak monopole appears to violate the Dirac quantization condition. However this is not true since one must also take the Z-string into account when deriving the quantization condition relevant to the electroweak monopole This becomes clearer when we work out the magnetic flux for the $SU(2)$ fields. Using (\ref{fnfyfnfy}) with (\ref{fyfa}), the net non-Abelian flux is: \be F_n = F_n |_S + F_n |_{\Sigma -S} = {{4\pi} \over {g}} \label{fn} \end{equation} just as we would expect for a 't Hooft-Polyakov monopole \cite{tHoPol74}. That is, the Dirac quantization condition works perfectly well for the SU(2) field and the monopole charge is quantized in units of $4\pi /g$. Another way of looking at (\ref{fn}) is to say that the electroweak monopole is a genuine SU(2) monopole in which there is a net emanating $U(1)_n \subset SU(2)$ flux. The structure of the theory, however, only permits a linear combination of this flux and hypercharge flux to be long range and so there is a string attached to the monopole. But this string is made of Z field which is orthogonal to the electromagnetic field and so the string does not surreptitiously return the monopole electromagnetic flux. Also, the magnetic charge on the monopole is conserved and electroweak monopoles can only disappear by annihilating with antimonopoles. It is useful to have an explicit expression describing the asymptotic field of the electroweak monopole and string. Nambu's monopole-string configuration, denoted by $({\bar \Phi}, {\bar W}_\mu ^a, {\bar Y}_\mu )$, is \be {\bar \Phi} = {{\eta} \over {\sqrt{2}}} \pmatrix{ \cos(\theta /2) \cr \sin(\theta /2) e^{i\varphi }} \ \label{barphi} \end{equation} where, $\theta$ and $\varphi$ are spherical coordinates centred on the monopole, and the gauge field configuration is, \be g {\bar W}_\mu ^a = - \epsilon^{abc} n^b \partial_\mu n^c + i \cos^2 \theta_w n^a ({\bar \Phi}^{\dag} ~ \partial_\mu {\bar \Phi} - \partial_\mu {\bar \Phi}^{\dag} ~{\bar \Phi} ) \label{gwbarmua} \end{equation} \be g' {\bar Y}_\mu = - i \sin^2 \theta_w ({\bar \Phi}^{\dag} ~ \partial_\mu {\bar \Phi} - \partial_\mu {\bar \Phi}^{\dag} ~{\bar \Phi} ) \ \label{gpbarymu} \end{equation} where, $n^a$ is given in eq. (\ref{nadef}). Note that there is no electroweak configuration that represents a magnetic monopole surrounded by vacuum. \subsection{Electroweak dyons} \label{ewdyons} Given that the electroweak monopole exists, it is natural to ask if dyonic configurations exist as well. We now write down dyonic configurations that solve the asymptotic field equations \cite{Vac95}. The existence of such configurations is implicit in Nambu's original paper in the guise of what he called ``external'' potentials \cite{Nam77}. Essentially, the dyon solution is an electroweak monopole together with a particular external potential. The ansatz that describes an electroweak dyon connected by a semi-infinite $Z$ string is: \be \Phi = {\bar \Phi} \label{phibarphi} \end{equation} \be W ^a = {\bar W} ^a - dt {{ n^a {\dot \zeta}} \over \cos\theta_w} \label{wadyon} \end{equation} \be Y = {\bar Y} - dt {{\dot \zeta} \over \sin\theta_w} \label{ydyon} \end{equation} where, $\zeta= \zeta (t, \vec x )$, overdots denote partial time derivatives and barred fields have been defined in the previous subsection. We now need to insert this ansatz into the field equations and to find the equation satisfied by $\zeta$. Some algebra leads to \be \partial ^i \partial_i {\dot \zeta} = 0 \ , \ \ \ \partial _t \partial ^i {\dot \zeta} =0 \label{dotzeta} \end{equation} which can be solved by separating variables, \be \zeta = \xi (t) f(\vec x ) \ . \label{zetaxif} \end{equation} This leads to \be {\ddot \xi} = 0 \ , \ \ \ \nabla ^2 f = 0 \ . \label{ddotxi} \end{equation} The particular solution that we will be interested in is the solution that gives a dyon. Hence, we take: \be \xi = \xi_0 t \ , \ \ \ f(r) = - {{q \sin\theta_w \cos\theta_w} \over {4\pi \xi_0}} {1 \over r} \ , \label{xisoln} \end{equation} where, $\xi_0$ and $q$ are constants. Now, using (\ref{xisoln}), together with (\ref{wadyon}), (\ref{ydyon}) and (\ref{zetaxif}), we get the dyon electric field: \be {\vec E}_A = {q \over {4\pi}} {{\vec r} \over {r^3}} \ . \label{ecoulomb} \end{equation} For a long segment of string, the monopole and the antimonopole at the ends are well separated and we can repeat the above analysis for both of them independently. Therefore, the electric charge on the antimonopole at one end of a Z-string segment is uncorrelated with the charge on the monopole at the other end of the string. This means that we can have dyons of arbitrary electric charge at either end of the string. The situation will change with the inclusion of fermions since these can carry currents along the string and transport charge from monopole to antimonopole. This completes our construction of the dyon-string system in the GSW model. As of now, the charge $q$ on the dyon is arbitrary. Quantum mechanics implies that the electric charge must be quantized. If we include a $\theta$ term in the electroweak action (but no fermions): \be S_\theta = {{g^2 \theta} \over {32 \pi^2}} \int d^4 x W_{\mu \nu}^a {\tilde W}^{\mu \nu ^a} \ \label{thetaaction} \end{equation} where \be {\tilde W}^{\mu \nu ^a} = {1\over 2} \epsilon^{\mu \nu \lambda \sigma} W_{\mu \nu}^a \ , \end{equation} then the charge quantization condition becomes \be q = \left ( n + {\theta \over {2\pi}} \right ) e \ . \label{dyonquant} \end{equation} This agrees with the standard result for dyons \cite{Wit79}. In the GSW model with fermions, it is known that the $\theta$ term can be eliminated by a rotation of the fermionic fields. This argument can be turned around to argue that the CP violation in the mass matrix of the fermions will lead to an effective $\theta$ term and so the electroweak monopoles will indeed have a fractional charge with $\theta$ being related to the CP violation in the mass matrix. The precise value of the fractional electric charge on electroweak monopoles has not yet been calculated and remains an open problem. It should be mentioned that, even though the electric charge on an electroweak dyon can be fractional as in (\ref{dyonquant}), the total electric charge on the dyon-string system is always integral because the CP violating fractional charge on the monopole is equal and opposite to that on the antimonopole. \subsection{Embedded defects and W-strings:} \label{embeddeddefects} A very simple way of understanding the existence of electroweak string solutions is in terms of embedded defects. While this method does not shed any light on the stability of the electroweak string, it does provide a scheme for finding other solutions. The idea is that the electroweak symmetry group contains several $U(1)$ subgroups which break completely when the electroweak symmetry breaks. Corresponding to each such breaking, one might have a string solution. A more complete analysis tells us when such a solution can exist \cite{VacBar92,BarVacBuc94,DavLep95,LepDav95}. Consider the general symmetry breaking \be G \rightarrow H \label{gtoh} \end{equation} Suppose $G_{emb}$ is a subgroup of $G$ which, in this process, breaks down to $G_{emb} \cap H$. Then we ask the question: when are topological defects in the symmetry breaking \be G_{emb} \rightarrow G_{emb} \cap H \label{gembtogembinth} \end{equation} also solutions in the full theory? An answer to this question requires separating the gauge fields into those that transform within the $G_{emb}$ subgroup and those that do not. Similarly, the Higgs field components are separated into those that lie in the embedded vector space of scalar fields and those that do not. Then, it is possible to write down general conditions under which solutions can be embedded \cite{BarVacBuc94,DavLep95}. Here we shall not describe these conditions but remark that the Z-string is due to the embedded symmetry breaking \be U(1)_Z \rightarrow 1 \label{u1zto1} \end{equation} where the $U(1)_Z$ is generated by $T_Z$, defined in eq. (\ref{QandTzgenerators}). Now, there are other $U(1)$'s that can be embedded in the GSW model which lie entirely in the $SU(2)$ factor. For example, we can choose $U(1)_1$ which is generated by $T^1$ (one of the off-diagonal generators of $SU(2)$). Since we have \be U(1)_1 \rightarrow 1 \label{u11to1} \end{equation} when the electroweak symmetry breaks, there is the possibility of another string solution in the GSW model. Indeed, it is easily checked that this string can be embedded in the GSW model and the solution is called a W-string. By considering a one parameter family of $U(1)$ subgroups generated by \be T_\zeta = \cos(\zeta ) ~T^1 + \sin(\zeta ) ~T^2 \label{tgenerator} \end{equation} we can generate a one parameter family of W-strings \be \Phi = {\eta \over {\sqrt{2}}} f_{NO} (\rho ) \pmatrix{ \cos\varphi \cr ie^{-i\zeta} \sin\varphi} \label{phiforwstring} \end{equation} \be W^1 = -{2 \over g} \cos\zeta ~{v_{NO}(\rho )} d\varphi ~ , \ \ W^2 = -{ 2\over g} \sin\zeta ~{v_{NO} (\rho )} d\varphi ~ , \label{wforwstring} \end{equation} and all other fields vanish. Although the string solutions are gauge equivalent for different values of $\zeta$, the parameter does take on physical meaning when considering multi-string configurations in which the value of $\zeta$ is different for different strings \cite{BarVacBuc94}. Note that the generator (\ref{tgenerator}) can be obtained from $T^1$ by the action of the unbroken (electromagnetic) group, \be T_\zeta = e^{i\zeta Q} T^1 e^{-i\zeta Q} \label{tzetadef} \end{equation} With this in mind, Lepora {\it et al} \cite{LepDav95, LepKib99a} have classified embedded vortices. The idea is that, for a general symmetry breaking $G \to H$,the Lie algebra of $G$, $\cal G$, decomposes naturally into a direct sum of the space $\cal H$ of generators of the unbroken subgroup $H$ (the ones associated with massless gauge bosons) and the space $\cal M$ of generators associated with massive gauge bosons: ${\cal G} = {\cal H} + {\cal M}$. The action of $H$ on the subspace $\cal M$ further decomposes $\cal M$ into irreducible subspaces. The classification of embedded vortices is based on this decomposition, as we now explain. Recall (eq. \ref{gaugeorbit}) that finite energy vortices are associated with gauge orbits on the vacuum manifold\footnote{The gauge orbits are geodesics of a squashed metric on the vacuum manifold which is different from the isotropic metric relevant to the scalar sector \cite{LepKib99b}.}. Choosing a base point $\Phi_0$ in the vacuum manifold, each embedded vortex can be associated to a Lie algebra generator which is tangent to the gauge orbit describing the asymptotic scalar field configuration of the vortex. The unbroken subgroup $H$ at $\Phi_0$ ``rotates'' the various gauge orbits among themselves as in eq. (\ref{tzetadef}). Thus, the action of $H$ splits the space of gauge orbits into irreducible subspaces. Except for critical values of the coupling constants (which could lead to so-called {\it combination vortices}), it can be shown \cite{BarVacBuc94,LepDav95} that embedded vortices have to lie entirely in one of these irreducible subspaces. If the subspaces have dimension greater than one, then there may be a {\it family} of gauge equivalent vortices. In the GSW model, for instance, the Lie Algebra decomposes into ${\cal H} + {\cal M}_1 + {\cal M}_2 \ $ where $\cal H$ is spanned by the charge $Q$, ${\cal M}_1$ is a one-dimensional subspace spanned by $T_Z$ (corresponding to the Z-string) and ${\cal M}_2 $ is a two-dimensional subspace comprising all W-string generators $T_\zeta$. Both the W- and the Z-string are embedded string solutions in the GSW model. What makes the Z-string more interesting is its unexpected stability properties. It can be shown \cite{LepDav95} that only those vortices lying in one-dimensional subspaces can have a stable semilocal limit. Thus, embedded vortices belonging to a family are always unstable. Another important difference is that the Z-string is known to terminate on magnetic monopoles but this is not true of the W-string. The W-string can terminate without any emanating electromagnetic fields since it is entirely within the $SU(2)$ sector of the GSW model. It is straightforward to embed domain walls in the GSW model. There are no embedded monopoles in the GSW model since there is no SU(2) subgroup that is broken to U(1). \section{Electroweak strings in extensions of the GSW model} \label{ewstringinextensions} Electroweak strings have been discussed in various extensions of the GSW model. We describe some of this work below. We do not, however, discuss extensions in which topological strings are produced at the electroweak scale \cite{DvaSen94,BimLoz94b}. \subsection{Two Higgs model} \label{twohiggs} As discussed in Sec. \ref{embeddeddefects}, the Z-string is an embedded string in the GSW model. The general conditions that enable the embedding are valid even with a more complicated Higgs structure. Here we will consider the two Higgs doublet model which is inspired by supersymmetric extensions of the GSW model. In a two Higgs doublet model, the Higgs structure of the GSW model is doubled so that we have scalars $\Phi_1$ and $\Phi_2$ and the scalar potential is \cite{Kas93} \begin{eqnarray} V(\Phi_1,\Phi_2) & = & \lambda_1(\Phi_1^{\dagger}\Phi_1 - \frac{\nu_1^2}{2})^2 + \lambda_2(\Phi_2^{\dagger}\Phi_2 - \frac{\nu_2^2}{2})^2 + \lambda_3[(\Phi_1^{\dagger}\Phi_1 - \frac{\nu_1^2}{2}) + (\Phi_2^{\dagger}\Phi_2 - \frac{\nu_2^2}{2})]^2 \nonumber \\ & + & \lambda_4[(\Phi_1^{\dagger}\Phi_1)(\Phi_2^{\dagger}\Phi_2) - (\Phi_1^{\dagger}\Phi_2)(\Phi_2^{\dagger}\Phi_1)] + \lambda_5[{\rm Re}(\Phi_1^{\dagger}\Phi_2) - \frac{\nu_1\nu_2}{2}\cos\xi]^2 \nonumber \\ & + & \lambda_6[{\rm Im}(\Phi_1^{\dagger}\Phi_2) - \frac{\nu_1\nu_2}{2}\sin\xi]^2 \ \nonumber \\ & + & \lambda_7[{\rm Re}(\Phi_1^{\dagger}\Phi_2) - \frac{\nu_1\nu_2}{2}\cos\xi]~[{\rm Im}(\Phi_1^{\dagger}\Phi_2) - \frac{\nu_1\nu_2}{2}\sin\xi] \ \ . \label{twohiggspotential} \end{eqnarray} Here $\nu_1$ and $\nu_2$ are the respective VEVs of the two doublets, $\lambda_i$ are coupling constants and the parameter $\xi$ is a phase In polar coordinates, the solution for the two Higgs Z-string is: \be \Phi _1 = \nu_1 f_1 (\rho ) e^{i\varphi} \pmatrix{0\cr 1\cr} \label{phi1for2higgs} \end{equation} \be \Phi_2 = \nu_2 f_2 (\rho ) e^{i\varphi } \pmatrix{0\cr 1\cr} \label{phi2for2higgs} \end{equation} \be {\vec Z} = -{2 \over {g_z}} {{v(\rho )}\over {\rho}} {\hat \varphi} \label{afor2higgs} \end{equation} with the profile functions satisfying differential equations similar to the Abelian-Higgs case. These have been studied in Ref. \cite{EarJam93} where the stability has also been analyzed (also see \cite{Tom97}). \subsection{Adjoint Higgs model} \label{adjointhiggs} The GSW model with an additional $SU(2)$ field in the adjoint representation, $\vec \chi$, is what we shall refer to as the ``adjoint Higgs model''. The impact of the adjoint field on electroweak defects was considered in Ref. \cite{KepVac96}. The bosonic sector of the adjoint Higgs model is: \begin{equation} L = T_{ew} + |(\partial_\mu +ig \epsilon^a W_\mu ^a ) {\vec \chi}|^2 - V(\Phi , {\vec \chi}) + L_f\ \label{adjointmodellagrangian} \end{equation} where, $T_{ew}$ is the gradient part of the bosonic sector of the electroweak Lagrangian, $L_f$ is the fermionic part of the Lagrangian, $\epsilon^a_{ij} = \epsilon_{aij}$ ($a,i,j = 1,2,3$) and, \begin{equation} V(\Phi , {\vec \chi}) = - \mu_2^2 \Phi^{\dag}\Phi - \mu_3^2 {\vec \chi}^2 + \lambda_2 (\Phi^{\dag}\Phi)^2 +\lambda_3 {\vec \chi}^4 + a {\vec \chi}^2 \Phi^{\dag}\Phi + b {\vec \chi} \cdot \Phi^{\dag} {\vec \tau} \Phi \; . \label{adjointmodelpotential} \end{equation} If we impose an additional $Z_2$ symmetry on the Lagrangian under $ \Phi \rightarrow +\Phi \ , \ \ \ {\vec \chi} \rightarrow -{\vec \chi} \ . $ the symmetry is $([SU(2)_L\times U(1)_Y]/Z_2) \times Z_2$ and we must set $b=0$. In what follows, we shall only consider this case and henceforth ignore the last (cubic) term in the potential. In this case, an additional simplification is that the leptons and quarks do not couple to $\vec \chi$ and so $L_f$ is identical to the fermionic Lagrangian of the GSW model. (If the $Z_2$ symmetry is absent, the cubic term in the potential is allowed but is constrained to be small by experiment.) In a cosmological context, as the universe cools down from high temperatures, if the parameters lie in a certain range \cite{KepVac96} there will first be a phase transition in which the adjoint field gets a VEV. The VEV of the adjoint will break the $SU(2)$ factor of the high temperature symmetry group to $U(1)$. If the VEV of ${\vec \chi}$ is along the $(0,0,1)$ direction, the generator of this $U(1)$ will be $T^3$ and we will denote the unbroken subgroup as $U(1)_3$. So the symmetry breaking pattern at this stage is \be ([SU(2) \times U(1)_Y]/Z_2) \times Z_2 \rightarrow ([U(1)_3 \times U(1)_Y]/Z_2) \times Z_2 \label{adjointstage1} \end{equation} and topological magnetic monopoles will be produced with pure $U(1)_3$ flux At a lower temperature, the doublet field will also get a VEV with the effect, \be ([U(1)_3 \times U(1)_Y]/Z_2) \times Z_2 \rightarrow U(1)_{em} \ . \label{adjointstage2} \end{equation} where, as usual, the electromagnetic charge operator is \be Q = T^3 + {Y\over 2} \label{qt3y} \end{equation} The electromagnetic component (A) from the monopoles is massless but the orthogonal part (Z) of the flux is massive and gets confined to a string. This is the Z-string. In addition, the breaking of the $Z_2$ factor gives domain walls. In the second stage of symmetry breaking, the Z-string is topological and hence is stable. The presence of magnetic monopoles from the earlier symmetry breaking means that the Z-strings can break by terminating on monopoles. But, as the monopoles form at a higher energy scale, their mass is much larger than the energy scale at which strings form and which sets the scale for the tension in the string. So the string can only break by instanton processes. At a yet lower temperature, the VEV of the adjoint turns off. This makes no difference to the symmetry structure of the model (apart from restoring the $Z_2$ symmetry and eliminating the domain walls) and hence no significant difference to the monopoles connected by strings. However, it does affect the stability of the strings since the monopoles are no longer topological. \section{Stability of electroweak strings} \label{stability} \subsection{Heuristic stability analysis} \label{heuristic} As described in \cite{Vac92}, the Z-string goes over into the semilocal string in the limit $\theta_w \rightarrow \pi /2$ and hence the stability of the Z-string should match on continuously to that of the semilocal string. Therefore we expect that Z-strings should be stable if $\theta_w$ is close to $\pi/2$ and $m_H \le m_Z$. The stability analysis to certain subsets of perturbations can be carried out much more easily than to the completely general perturbations. The subset includes perturbations in the Higgs field and W-fields separately. Such analyses may be found in \cite{Vac92,Vac93,BarVacBuc94} and \cite{Perk93}. \noindent (i) {\it Higgs field perturbations:} Perturbations in the Higgs field alone have maximum destabilizing effect for $\theta_w = \pi /4$ \cite{BarVacBuc94} and, in this case, it is easy to see that the Z-string is unstable. Consider the one parameter family of field configurations \be \Phi (\vec x ; \xi ) = \cos\xi ~\Phi_0 (\cos\xi ~{\vec x}) + \sin\xi ~\Phi_\perp \label{phiconfigpiover4} \end{equation} \be Z_j (\vec x ; \xi ) = \cos\xi ~Z_{(0)j} (\cos\xi ~\vec x ) \label{zconfigpiover4} \end{equation} where, the string solution is denoted by the $0$ subscript, $\xi \in [0,\pi /2]$ and \be \Phi_\perp = {\eta \over {\sqrt{2}}}\pmatrix{1\cr 0\cr} \ . \label{phiperp} \end{equation} For $\xi=0$ the field configuration is the unperturbed Z-string while for $\xi =\pi/2$ it is the vacuum. The energy per unit length of this field configuration can be evaluated and is found to be: \be E(\xi ) = \cos^2\xi ~ E(\xi =0) \ . \label{energywithxi} \end{equation} Hence the energy per unit length of the string is a monotonically decreasing function of $\xi$ and so the string is unstable to decay into the vacuum. \noindent (ii) {\it Incontractible two spheres:} James \cite{Jam94}, and, Klinkhamer and Olesen \cite{KliOle94} have constructed the Z- and W-string solutions by considering incontractible two spheres in the space of electroweak field configurations in two spatial dimensions. The idea was introduced by Taubes \cite{Tau82} and was used by Manton to construct the sphaleron \cite{Man83,KliMan84}. The procedure (known as the ``minimax'' procedure) is to construct a set of field configurations that are labelled by some parameters $\mu_i$. If this set is incontractible in the space of field configurations, then there exist (subject to certain assumptions \cite{Man83}) values of the parameters for which the field configuration extremizes the energy functional. For example, Klinkhamer and Olesen \cite{KliOle94} give the following construction for the Z-string in terms of a two parameter ($\mu , \nu$) family of field configurations \begin{eqnarray} \pi /2 \le [ \mu \nu ] \le \pi \, : && W=0 \ , \ \ \ Y=0 \nonumber \\ &&\Phi = ( 1-\{ 1-h(\rho )\} \sin [ \mu \nu ] ) {\eta \over {\sqrt{2}}} \pmatrix{0\cr 1\cr} \label{fkpowyphi1} \end{eqnarray} \begin{eqnarray} 0 \le [ \mu \nu ] \le \pi /2 \, : && W= -f(\rho ) G^a T^a \ , \ \ \ Y= f(\rho ) \sin^2\theta_w F^3 \nonumber \\ &&\Phi = h(\rho ) {\eta \over {\sqrt{2}}} \Omega U \pmatrix{0\cr 1\cr} \label{fkpowyphi2} \end{eqnarray} where, $W$ and $Y$ are Lie algebra valued 1-forms ({\it e.g.} $W = W_\mu ^a T^a dx^\mu$), $[\mu \nu] \equiv {\rm max}(|\mu |, |\nu|)$, \be F^a T^a = 2i U^{-1} dU \ , \label{fataua} \end{equation} \be G^a T^a = \Omega U [ F^1 T^1 + F^2 T^2 + \cos^2\theta_w F^3 T^3 ] U^{-1} \Omega^{-1} \ , \label{gataua} \end{equation} \be U(\mu , \nu , \varphi ) = -i \sin\mu \tau_1 -i \cos\mu \sin\nu \tau_2 -i \cos\mu \cos\nu \sin\varphi \tau_3 + \cos\mu \cos\nu \cos\varphi {\bf 1} \ , \label{umatrix} \end{equation} \be \Omega = U(\mu , \nu , \varphi = 0 )^{-1} \ , \label{omegadefn} \end{equation} and the functions $f(\rho )$ and $h(\rho )$ satisfy the boundary conditions \be f(0) = 0 = h(0) \ , \ \ \ f(\infty )=1=h(\infty ) \ . \label{fkpofhbcs} \end{equation} This set of field configurations labelled by the parameters $\mu , \nu \in [-\pi , \pi ]$ defines an incontractible two sphere in the space of field configurations. This is seen by considering the fields as if they were defined on the three-sphere on which the coordinates are $\varphi$, $\mu$ and $\nu$ and then showing that the field configurations define a topologically non-trivial mapping from this $S^3$ to the vacuum manifold which is also an $S^3$. Then the minimax procedure says that there is an extremum of the energy at some value of the parameters. By inserting the field configurations into the energy functional, it can be checked that the extremum occurs at $\mu =0=\nu$, when the configuration coincides with that of the Z-string. Furthermore, for $\theta_w \le \pi /4$, the extremum is a maximum and hence the Z-string is unstable. A very similar analysis has been done \cite{Jam94,KliOle94} for the W-string confirming the result \cite{BarVacBuc94} that it is always unstable. \noindent (iii) {\it W-condensation:} There is also a well-known \cite{AmbOle90} instability to perturbations in the W-fields alone called ``W-condensation''. Application of this instability to the Z-string may be found in \cite{Perk93,Vac93,Vac94,AchGreHarKui94}. A heuristic argument goes as follows. The energy of a mass $m$, charge $e$ and spin $s$ particle in a uniform magnetic field $\vec B$ along the z-axis is given by: \be E^2 = p_z^2 +m^2 +(2n+1) e B -2e {\vec B}\cdot {\vec s} \label{energyspininb} \end{equation} where $n=0,1,2,...$ labels the Landau levels and $p_z$ is the momentum along the z-axis. Now, if $s=1$, the right-hand side can be negative for $p_z =0$, $n=0$ provided \be B > {{m^2}\over e} ~ . \label{bcritical} \end{equation} This signals an instability towards the spontaneous creation of spin one particles in sufficiently strong magnetic fields \cite{AmbOle90}. In our case, the magnetic field is a Z-magnetic field and this couples to the spin one W-particles. If the string thickness is larger than the Compton wavelength of the W-particles, the Z-magnetic field may be considered uniform. Also, the relevant charge in this case is the Z-charge of the W-bosons and is $g_Z\cos^2\theta_w$. The constraint that the string be thick so that the Z-magnetic field appears uniform and that the charge not be too small means that $\theta_w$ should be small. Hence the instability towards W-condensation applies for small $\theta_w$. This analysis can be performed more quantitatively \cite{Perk93} with the result that there is a relatively hard bound $\sin^2\theta_w > 0.8$ for the string to be stable to W-condensation. \subsection{Detailed stability analysis} \label{detailedstability} To analyse the stability of electroweak strings, we perturb the string solution, extract the quadratic dependence of the energy on the perturbations and then determine if the energy can be lowered by the perturbations by solving a Schr\"odinger equation. The analysis is quite tedious \cite{JamPerVac92,AchGreHarKui94,GooHin95b,MacTor94,MacTor95} and here we will only outline the main steps. We use the vector notation in this section for simplicity. The general perturbations of the Z-string are \be (\phi_\perp,\phi_\parallel, \delta {\vec Z}, {\vec W}^{\bar a}, {\vec A}) \label{pertlist} \end{equation} where, ${\bar a}=1,2$, $\phi_\perp$ and $\phi_\parallel$ are scalar field fluctuations defined by \be \Phi = \pmatrix{\phi_\perp\cr \phi_{NO} + \phi_\parallel\cr} \ , \label{perturbedphi} \end{equation} $\delta {\vec Z}$ is defined by \be {\vec Z} = {\vec Z}_{NO} + \delta {\vec Z} \ . \label{perturbedz} \end{equation} (The subscript $NO$ means that the field is the unperturbed Nielsen-Olesen solution for the string as described in Sec. \ref{NO}.) The fields ${\vec W}^{\bar a}, {\vec A}$ are perturbations since the unperturbed values of these fields vanish in the Z-string. The perturbations can depend on the $z-$coordinate and the $z-$components of the vector fields can also be non-zero. However, since the vortex solution has translational invariance along the $z-$direction, it is easy to see that it is sufficient to consider $z$ independent perturbations and to ignore the $z-$components of the gauge fields. This follows from the expression for the energy resulting from the Lagrangian in eq. (\ref{GSWb}) where the relevant $z-$dependent terms in the integrand are: \be {1\over 2} G_{i3} ^a G_{i3} ^a + {1\over 4} F_{Bi3} F_{Bi3} + (D_3 \Phi ) ^{\dag} (D_3 \Phi ) \label{zdepenergy} \end{equation} and explicitly provide a positive contribution to the energy. Hence we drop all reference to the $z-$coordinate with the understanding that the energy is actually the energy {\it per unit length} of the string. Now we calculate the energy of the perturbed configuration, discarding terms of cubic and higher order in the infinitesimal perturbations. We find, \begin{equation} E = ( E_{NO} + \delta E_{NO} ) + E_\perp + E_c + E_W \label{perturbede1} \end{equation} where, $E_{NO}$ is the energy of the Nielsen-Olesen string and $\delta E_{NO}$ is the energy variation due to the perturbations $\phi_\parallel$ and $\delta {\vec Z}$. The term $E_\perp$ is due to the perturbation $\phi_\perp$ in the upper component of the Higgs field, $E_c$ is the cross-term between perturbations in the Higgs and gauge fields, while $E_W$ is the contribution from perturbing the gauge fields alone: \begin{equation} E_{\perp} = \int d^2 x \left [ |{\bar d}_j \phi_\perp |^2 + \lambda \eta^2 ( f^2 - 1 ) | \phi_\perp |^2 \right ] \ , \label{e1} \end{equation} \begin{equation} E_c = i {{g_z}\over 2} \cos\theta_w \int d^2 x \left [ \Phi^{\dag} T^{\bar a} {\bf d}_j \Phi - ( {\bf d}_j \Phi ) ^{\dag} T^{\bar a} \Phi \right ] W_j ^{\bar a} \ , \label{ec} \end{equation} with ${\bf d}_j$ defined in \ref{smalldmu}, \begin{eqnarray} E_W = & \int d^2 x \biggl [ \gamma {\vec W} ^1 \times {\vec W} ^2 \cdot \vec \nabla \times \vec Z + {1\over 2} |\vec \nabla \times {\vec W} ^1 + \gamma {\vec W} ^2 \times \vec Z | ^2 \nonumber \\ &+ {1\over 2} |\vec \nabla \times {\vec W} ^2 + \gamma \vec Z \times {\vec W} ^1 | ^2 + {1\over 4} g^2 f^2 ( {\vec W } ^{\bar a} ) ^2 + {1\over 2} ( \vec \nabla \times \vec A )^2 \biggr ] \ , \label{ew} \end{eqnarray} where $\gamma \equiv g \cos\theta_w$, \be {\bar d} _j \equiv \partial _j - i{g_z \over 2} \cos(2\theta_w ) Z_j \label{barddefn} \end{equation} and, the $f$ and $\vec Z$ fields in the above equations are the unperturbed fields of the string. The two instabilities discussed in the previous subsection can also be seen in eq. (\ref{perturbede1}). First consider perturbations in the Higgs field alone. Then only $E_\perp$ is relevant. For $\theta_w =\pi/4$, ${\bar d} _j = \partial_j$, and $E_\perp$ is the energy of a particle described by the wavefunction $\phi_\perp$ in a purely negative potential in two dimensions since $f^2 \le 1$ everywhere. It is known that a purely negative potential in two dimensions always has a bound state. \footnote{For some potentials though, the wavefunction of the bound state may have singular (though integrable) behaviour at the origin and such bound states would be inadmissible for us since we require that the perturbations be small. This turns out not to be the case for the potential in eq. (\ref{e1}).} Hence the energy can be lowered by at least one perturbation mode and so the string is unstable when $\theta_w = \pi/4$. The instability towards W-condensation can be seen in $E_W$. The term with $\gamma$ can be negative and its strength is largest for small $\theta_w$. Hence W-condensation is most relevant for small $\theta_w$. Returning to the full stability analysis, we first note that the perturbations of the fields that make up the string do not couple to the other available perturbations {\it i.e.} the perturbations in the fields $f$ and $v$ only occur inside the variation $\delta E_{NO}$. Now, since we know that the Nielsen-Olesen string with unit winding number is stable to perturbations for any values of the parameters then necessarily, $\delta E _{NO} \ge 0$ and the perturbations $\phi_\parallel$ and $\delta {\vec Z}$ cannot destabilize the vortex. Then, we are justified in ignoring these perturbations and setting $\delta E _{NO} = 0$. Also we note that the ${\vec A}$ boson only appears in the last term of eq. (\ref{ew}) and is manifestly positive. So we can set ${\vec A}$ to zero. The remaining perturbations can be expanded in modes: \be \phi _\perp = \chi (\rho ) e^{im\varphi} \label{phi1modes} \end{equation} for the $m^{th}$ mode where $m$ is any integer. For the gauge fields we have, \be {\vec W}^1 = \left [ \left \{ {\bar f}_{1} (\rho ) \cos(n\varphi ) + f_{1} (\rho ) \sin(n\varphi ) \right \} {\hat e} _\rho + {1 \over \rho} \left \{ - {\bar h}_{1} (\rho )\sin(n\varphi ) + h_{1} (\rho )\cos(n \varphi ) \right \} {\hat e} _\varphi \right ] \label{w1modes} \end{equation} \be {\vec W}^2 = \left [ \left \{ - {\bar f}_{2} (\rho ) \sin(n\varphi ) + f_{2} (\rho )\cos(n\varphi ) \right \} {\hat e} _\rho + {1 \over \rho } \left \{ {\bar h}_{2} (\rho )\cos(n\varphi ) + h_{2} (\rho ) \sin(n \varphi ) \right \} {\hat e} _\varphi \right ] \label{w2modes} \end{equation} for the $n^{th}$ mode where $n$ is a non-negative integer. The most unstable mode is the one with $m=0$ and $n=1$. This is because these have the lowest gradient energy and are the only perturbations that can be non-vanishing at $\rho =0$. Further analysis shows that the string is most unstable to the $h_1 + h_2$ mode. Hence, we can ignore $f_i$, $h_1-h_2$ and the barred variables. A considerable amount of algebra then yields: \be \delta E[ \chi , \xi_+ ] = 2\pi \int d\rho \ \rho ~ ( \chi , \xi _+ ) {\bf O} \pmatrix{\chi\cr \xi_+\cr} \label{deltaeofchiandxi} \end{equation} where, {\bf O} is a $2\times 2$ matrix differential operator and \be \xi_+ = {{h_1+h_2} \over 2} \ . \label{xiplusdefn} \end{equation} Before proceeding further, note that a gauge transformation on the fields does not change the energy. However, we have not fixed the gauge in the preceding analysis and hence it is possible that some of the remaining perturbations, $(\chi, \xi_+)$, might correspond to gauge degrees of freedom and may not affect the energy. So we now identify the combination of perturbations $\chi$ and $\xi_+$ that are pure gauge transformations of the string configuration. The $SU(2)$ gauge transformation, $\exp (ig\psi )$, of an electroweak field configuration leads to first order changes in the fields of the form \be \delta\Phi=ig\psi\Phi_0 ~ ,{\ {\delta W}_i=-iD^{(0)}_{i}\psi} \ , \label{gaugeonfields} \end{equation} where $W_i = W_i^a T^a$, $\psi =\psi^a T^a$, and the $0$ index denotes the unperturbed field and covariant derivative. In our analysis above, we have fixed the form of the unperturbed string and so we should restrict ourselves to only those gauge transformations that leave the Z-string configuration unchanged. (For example, $\delta \Phi$ should only contain an upper component and no lower component.) This constrains $\psi$ to take the form \be \psi=s(\rho )\pmatrix{0&ie^{-i\varphi }\cr -ie^{i\varphi }&0\cr} \label{psiform} \end{equation} where $s(\rho )$ is any smooth function. This means that perturbations given by \begin{equation} \pmatrix{\chi (\rho) \cr\xi_+ (\rho)}\ = \ s(\rho )\pmatrix{-g\eta f(\rho)/\sqrt{2}\cr 2(1-2 \cos^2\theta_w v(\rho ))\cr} \label{puregaugemode} \end{equation} are pure gauge perturbations that do not affect the string configuration. Therefore, such perturbations cannot contribute to the energy variation and must be annihilated by ${\bf O}$. Then, in the two-dimensional space of $(\chi , \xi_+ )$ perturbations, we can choose a basis in which one direction is pure gauge and is given by (\ref{puregaugemode}) and the other orthogonal direction is the direction of physical perturbations. The physical mode can now be written as, \be \zeta (\rho )\ =\ (1-2 \cos^2\theta_w v(\rho ))\chi (\rho )\ + \ {g\eta f(\rho )\over {2\sqrt{2}}}\xi_+ (\rho ) ~ . \label{zetadefn} \end{equation} So now the energy functional reduces to one depending only on $\zeta (\rho )$: \be \delta E[ \zeta ] = 2\pi \int d\rho ~ \rho ~ \zeta {\overline O} \zeta \label{deltaefinal} \end{equation} where ${\overline O}$ is the differential operator \be {\overline O} = - {1 \over \rho} {d \over {d\rho }} \left ( {\rho \over {P_+}} {d \over {d\rho }} \right ) + U(\rho ) \label{overlineo} \end{equation} and \be U(\rho ) = {{{f'}^2} \over {P_+ f^2}} + {{4 S_+} \over {g^2 \eta^2 \rho^2 f^2}} + {1 \over \rho } {d \over {d\rho }} \biggl ( {{\rho f'} \over {P_+ f}} \biggr ) \ , \label{schrodingerudefn} \end{equation} where \be P_{+} = (1 - 2 \cos^2\theta_w v )^2 + {g^2 \eta^2 \rho^2 f^2\over 4} \label{pplusdefn} \end{equation} \be S_+ (\rho ) = {g^2 \eta^2 f^2\over 4} - {{4 \cos^4\theta_w {v'}^2} \over {P_+ (\rho )}}+ \rho {{d \ } \over {d\rho }} \left [ 2 \cos^2\theta_w {{v'} \over \rho } {{(1 - 2 \cos^2\theta_w v )} \over {P_+ (\rho )}} \right ] \ . \label{splusdefn} \end{equation} The question of Z-string stability reduces to asking if there are negative eigenvalues $\omega$ of the Schr\"odinger equation, \begin{equation} {\overline O} \zeta =\omega\zeta \ . \label{eigenvalueproblem} \end{equation} The eigenfunction $\zeta (\rho )$ must also satisfy the boundary conditions $\zeta (\rho =0) = 1$ and $\zeta ' (0) =0$ where prime denotes differentiation with respect to $\rho$. In this way the stability analysis reduces to a single Schr\"odinger equation which can be solved numerically. The results of the stability analysis are shown in Fig. (\ref{stabilityregion}) as a plot in parameter space $(m_H/m_Z, \sin^2 \theta_w)$, demarcating regions where the Z-string is unstable (that is, where negative $\omega$ exist) and stable (negative $\omega$ do not exist). It is evident that the experimentally unconstrained values: $\sin^2 \theta_w=0.23$ and $m_H / m_Z > 0.9$ lie entirely inside the unstable sector. Hence the Z-string in the GSW model is unstable. \begin{figure}[tbp] \caption{\label{stabilityregion} The Z-string is stable in the triangular shaded region of parameter space. At $\sin^2\theta_w =0.5$, the string has a scaling instability. The experimentally allowed parameters are also shown. } \vskip 1truecm \epsfxsize = \hsize \epsfbox{stabilityplot.eps} \end{figure} The stability analysis of the Z-string described above leaves open the possibility that the string might be stable in some special circumstances such as, the presence of extra scalar fields, or a magnetic field background, or fermions. We now describe some circumstances in which the Z-string stability has been analyzed. \subsection{Z-string stability continued} \label{stabilitycontd} The stability of Z-strings has been studied in various other circumstances: (i) {\it Thermal effects:} In \cite{HolHsuVacWat92} the authors examined thermal effects on Z-string stability using the high temperature effective potential and found slight modifications to the stability. The conclusion is that Z-strings in the GSW model are unstable at high temperatures as well. In the same paper, left-right symmetric models were studied and it was found that these could contain stable strings that are similar to the Z-string. (ii) {\it Extra scalar fields:} It is natural to wonder if the presence of extra scalar fields in the model can help provide stability. In \cite{EarJam93} the stability was examined in the physically motivated two Higgs doublet model with little advantage. In \cite{VacWat93} it was shown that an extra (globally) charged scalar field could enhance stability. The extra complex scalar field, $\psi$, is coupled to the electroweak Higgs by a term $|\psi|^2 \Phi^\dag \Phi$ and hence the charges have lower energy on the string where $\Phi^\dag \Phi \sim 0$ than outside the string where $\Phi$ has a non-zero VEV. the background of a string and hence a sufficient amount of charge can stabilize the string. This is exactly as in the case of non-topological solitons or Q-balls \cite{Ros68,FriLeeSir76,Col85}. However, scalar global charges attract and this can cause an instability of the charge distribution along the string \cite{CopKolLee88,VacWat93}. For realistic parameters, stable Z-strings do not seem likely even in the presence of extra scalar fields. (iii) {\it Adjoint scalar field:} A possible variant of the above scheme is that an $SU(2)$ adjoint can be included in the GSW model as described in Sec. (\ref{adjointhiggs}). Now, since the Z-string is topological within the second symmetry breaking stage in eq. (\ref{adjointstage2}), it is stable. However, to be consistent with current experimental data the VEV of the $SU(2)$ adjoint must vanish at a lower energy scale. At this stage the Z-string becomes unstable. Hence, in this scheme, there could be an epoch in the early universe where Z-strings would be stable. (iii) {\it External magnetic field:} An interesting possibility was studied by Garriga and Montes \cite{GarMon95} when they considered the stability of the Z-string placed in an external electromagnetic magnetic field of field strength $B$ parallel to the string. First, note that $B$ should be less than $B_c = m_W^2/e$, otherwise the vacuum outside the string is unstable to W-condensation \cite{AmbOle90}. Then they found that the Z-string could be stable if $B > \sqrt{\beta} B_c$, where $\beta = m_H^2/m_Z^2$ should be less than 1 for stability of the ambient vacuum. The region of stability for a few values of the magnetic field (given by $K=g_z B/2m_Z^2$) is sketched in Fig. \ref{stabilityregionb}. For a certain range of $K \sim 0.85$, stable Z-strings in the GSW model are still just possible. \begin{figure}[tbp] \caption{\label{stabilityregionb} The triangular regions depict the parameter range for which the electroweak vacuum and the Z-string are both stable in the presence of a uniform external magnetic field whose strength is proportional to $K$. For a range of magnetic field ($K \sim 0.85$), stable strings are possible even with the experimentally constrained parameter values. } \vskip 1truecm \epsfxsize = \hsize \epsfbox{stabilitywithb.eps} \end{figure} A way to understand the enhanced stability of the Z-string in a magnetic field is to realize that the W-condensation instability is due to the interaction of $W^3_ \mu = \sin \theta_w A_\mu + \cos \theta_w Z_\mu$ and $W^\pm_\mu$. The Z-string itself has a Z magnetic flux. Then the external electromagnetic flux can serve to lower the net $W^3$ flux. This reduces the efficiency of W-condensation and makes the string more stable. Another viewpoint can be arrived at if we picture the Z-string instability to be one in which the string breaks due to the production of a monopole- antimonopole pair on the string. If the external magnetic field is oriented in a direction that prevents the nucleated magnetic monopoles from accelerating away from each other, it will suppress the monopole pair production process, leading to a stabilization of the string for sufficiently strong magnetic fields. (iv) {\it Fermions:} The effect of fermions on the stability of the Z-string has been considered in Refs. \cite{EarPer94,Nac95,KonNac95,LiuVac96}. Naculich \cite{Nac95} found that fermions actually make the Z-string unstable. In \cite{LiuVac96} it was argued that this effect of fermions is quite general and also applies to situations where the strings form at a low energy scale due to topological reasons but can terminate on very massive monopoles formed at a very high energy scale. This most likely indicates that the Z-string solution itself should be different from the Nielsen-Olesen solution when fermions are included. We shall describe these results in greater detail in Sec. \ref{superconductivity} after discussing fermion zero modes on strings. Z-strings have also been considered in the presence of a cold bath of fermions \cite{BimLoz95}. The effect of the fermions is to induce an effective Chern-Simons term in the action which then leads to a long range magnetic field around the string. \subsection{Semiclassical stability} \label{semiclassicalstability} Preskill and Vilenkin \cite{PreVil92} have calculated the decay rate of electroweak strings in the region of parameter space where they are classically stable. The instability is due to quantum tunneling and is calculated by finding the semiclassical rate of nucleation of monopole-antimonopole pairs on electroweak strings. The bounce action is found to be \be S \sim {{4\pi^2}\over {g^2}} {{a_\infty} \over {a_s}} \label{bounceaction} \end{equation} where, the strings are classically stable if the ratio of parameters ${{a_\infty} / {a_s}}$ is larger than 1. ($a_\infty /a_s$ is the ratio of energy in the magnetic flux when it is spread over an infinite area to that if it is confined within the string.) The semiclassical decay probability of the string per unit length per unit time is proportional to $\exp[-S]$. The decay rate gets suppressed as we approach the semilocal string ($g \rightarrow 0$) thus the semilocal string is also stable semiclassically. \section{Superconductivity of electroweak strings} \label{superconductivity} \subsection{Fermion zero modes on the Z-string} \label{zeromodes} Here we shall consider the fermionic sector of the GSW model in the fixed background of the unit winding Z-string for which the solution is given in eq. (\ref{Zstring}). The Dirac equations for a single family of leptons and quarks are obtained from the Lagrangian in Sec. \ref{fermionicsector}. These have been solved in the background of a straight Z string in \cite{EarPer94,GarVac95,MorOakQui95}. The analysis is similar to that for $U(1)$ strings \cite{JacRos81} since the Z-string is an embedded $U(1)$ string in the GSW model (see Sec. \ref{embeddeddefects}). A discussion of the fermion zero modes in connection with index theorems can be found in \cite{KliRup97,Kli98} In polar coordinates with the Z-string along the $z$-axis, a convenient representation for the $\gamma$ matrices is: \begin{equation} \gamma^\rho = \pmatrix{0&e^{-i\varphi }&0&0\cr -e^{i\varphi }&0&0&0\cr 0&0&0& -e^{-i\varphi }\cr 0&0&e^{i\varphi}&0\cr}\ , \ \ \ \gamma^\varphi = \pmatrix{0&-ie^{-i\varphi}&0&0\cr -ie^{i\varphi}&0&0&0\cr 0&0&0&ie^{-i\varphi}\cr 0&0&ie^{i\varphi}&0\cr}\ , \label{gamma1} \end{equation} \begin{equation} \gamma^t = \pmatrix{\tau^3&0\cr 0&-\tau^3\cr}\ , \ \ \ \gamma^z = \pmatrix{0&{\bf 1}\cr -{\bf 1}&0\cr}\ , \ \ \ \gamma^5 = \pmatrix{0&{\bf 1}\cr {\bf 1}&0\cr}\ . \label{gamma2} \end{equation} (Note that the derivative $\gamma^\mu \partial_\mu$ is given by $\gamma^t\partial_t + \gamma^\rho \partial_\rho +\gamma^\phi \partial_\phi /\rho +\gamma^z\partial_z$.) Then the electron has a zero mode solution \begin{equation} e_L = \pmatrix{1\cr 0\cr -1\cr 0\cr} \psi_1 (\rho ) \ , \ \ \ e_R = \pmatrix{0\cr 1\cr 0\cr 1\cr} i \psi_4(\rho ) \label{eler} \end{equation} where, \begin{equation} \psi_1 ' + {{qv} \over \rho} \psi_1 = -h{\eta \over \sqrt{2}} f \psi_4 \label{psi1eom} \end{equation} \begin{equation} \psi_4 ' - {{(q-1)v} \over \rho } \psi_4 = -h {\eta \over \sqrt{2}} f \psi_1 \ . \label{psi4eom} \end{equation} In these equations $q$ is the eigenvalue of the operator ${\bf q}$ defined in eq. (\ref{zcharge}) and denotes the Z-charge of the various left-handed fermions. (For the electron, $q = \cos(2\theta_w)$.) The boundary conditions are that $\psi_1$ and $\psi_4$ should vanish asymptotically. This means that there is only one arbitrary constant of integration in the solution to eqns. (\ref{psi1eom}) and (\ref{psi4eom}). This may be taken to be a normalization of $\psi_1$ and $\psi_4$. For the $d$ quark, the solution is the same as in eqns. (\ref{eler}), (\ref{psi1eom}) and (\ref{psi4eom}) except that $q=1 - (2/3)\sin^2\theta_w$. For the $u$ quark the solution is: \begin{equation} u_L = \pmatrix{0\cr 1\cr 0\cr -1\cr} \psi_2 (\rho ) \ , \ \ \ u_R = \pmatrix{1\cr 0\cr 1\cr 0\cr} i \psi_3(\rho ) \label{ulur} \end{equation} where, \begin{equation} \psi_2 ' - {{qv} \over \rho } \psi_2 = -G_u {\eta \over \sqrt{2}} f \psi_3 \label{psi2eom} \end{equation} \begin{equation} \psi_3 ' + {{(q+1)v} \over \rho } \psi_3 = -G_u {\eta \over \sqrt{2}} f \psi_2 \label{psi3eom} \end{equation} where, $q = -1 + (4/3) \sin^2\theta_w$. Note that (\ref{psi1eom}), (\ref{psi4eom}) are related to (\ref{psi2eom}), (\ref{psi3eom}) by $q \rightarrow -q$. The right-hand sides of the neutrino Dirac equations (corresponding to eqns. (\ref{psi1eom}) and (\ref{psi4eom})) vanish since the neutrino is massless. The solutions can be found explicitly in terms of the string profile equations in the case when the Higgs boson mass ($m_H = \sqrt{2\lambda }\eta$) equals the Z boson mass ($m_Z = g_z\eta /2$) \cite{GarVac95}. Recall that the string equations in the $m_H=m_Z$ case are \cite{Bog76}: \begin{equation} f' = {f \over \rho} (1-v) \label{bogo1} \end{equation} \begin{equation} v' = {{m_Z^2} \over 2} \rho (1-f^2 ) \label{bogo2} \end{equation} yielding the useful relation: \begin{equation} \int d\rho {v \over \rho} = {\rm ln} \biggl ( {m_Z \rho \over f} \biggr ) \label{vr} \end{equation} where we have included a factor of $m_Z$ to make the argument of the logarithm dimensionless. Now the zero mode profile functions for the massless fermions are: \begin{equation} \psi _1 = c_1 m_Z^{3/2} \biggl ( {{m_Z \rho} \over f} \biggr )^{-q} \ , \ \ \ \psi _4 = c_4 m_Z^{3/2} \biggl ( {{m_Z \rho} \over f} \biggr )^{q-1} \label{nupsi1psi4} \end{equation} where, $c_1$ and $c_4$ are independent constants that can be chosen to normalize the left- and right-handed fermion states and the spinors are given in (\ref{eler}). The boundary condition that the left-handed fermion wavefunction should vanish at infinity is only satisfied if $q > 0$. Hence (\ref{nupsi1psi4}) can only give a valid solution for $q > 0$ for the left-handed fermion. If we also require normalizability, we need $q > 1$. (Note that there is no singularity at $\rho =0$ because $f \propto \rho $ when $\rho \sim 0$.) If we have a left-handed fermion with $q \le -1$, the correct equations to use are the equations corresponding to the up quark equations given in (\ref{psi2eom}) and (\ref{psi3eom}) and these are solved by letting $q \rightarrow -q$ in (\ref{nupsi1psi4}). In this case, the spinors are given in (\ref{ulur}). For the electroweak neutrino, the right-handed component is absent and $q = -1$. This means that the neutrino has the same spinor structure as the left-handed up quark and the solution is that in (\ref{nupsi1psi4}) with $q$ replaced by $+1$. Therefore the wave function falls off as $1/\rho$ and the state is strictly not normalizable - the normalization integral diverges logarithmically. However, depending on the physical situation, one could be justified in imposing a cut-off. For example, when considering closed loops of string, the cutoff is given by the radius of the loop. Next we give the explicit solutions to the Dirac equations in (\ref{psi1eom}) and (\ref{psi4eom}) in the case when the fermion mass ($m_f = h\eta /\sqrt{2}$) is equal to the scalar mass which is also equal to the vector mass. This so-called ``super-Bogomolnyi'' limit is not realized in the GSW model but may be of interest in other situations (for example, in supersymmetric models). Then, if the charge on the left-handed fermion vanishes ($q=0$), the solution can be verified to be: \begin{equation} \psi _1 (\rho ) = N m_Z ^{3/2} (1-f(\rho )^2) \label{3.5a} \end{equation} \begin{equation} \psi _4 (\rho ) = 2 N m_Z^{1/2} {f(\rho ) \over \rho} (1-v(\rho )) \label{3.5b} \end{equation} where $N$ is a dimensionless normalization factor. For the same set of parameters, the solution for the up quark equations can be written by using the transformation $q \rightarrow -q$ in the above solutions. Further, this solution can also be derived using supersymmetry arguments \cite{VecFer77,DavDavTro97}. The left-handed fermion wave-functions found above can be multiplied by a phase factor ${\rm exp}[i(E_p t - pz)$ and the resulting wave-function will still solve the Dirac equations provided \begin{equation} E_p = \epsilon_i p \label{ep} \end{equation} where, $i$ labels the fermions, and, \begin{equation} \epsilon_\nu = +1 = \epsilon_u \ , \ \ \ \epsilon_e = -1 = \epsilon_d \ . \label{epsi1} \end{equation} In other words, $\nu_L$ and $u$ travel parallel to the string flux while $e$ and $d$ travel anti-parallel to the string flux. \begin{figure}[tbp] \caption{\label{zeromodedirections} The direction of propagation of quark and lepton zero modes on the Z-string. } \vskip 1 truecm \epsfxsize = \hsize \epsfbox{zeromodesonstring.eps} \vskip 0.5 truecm \end{figure} We should mention that the picture of quarks travelling along the Z-string may be inaccurate since QCD effects have been totally ignored. At the present time it is not known if the strong forces of QCD will confine the quarks on the string into mesons and baryons (for example, pions and protons). Further, the electromagnetic interactions of the particles on the string might lead to bound states of electrons and protons on the string. This would imply a picture where hydrogen (and other) atoms are the fundamental entities that live on the string. \subsection{Stability of Z-string with fermion zero modes} \label{stabilitywithfermions} In Fig. \ref{pertonfermions} we show the effect that perturbations of order $\epsilon$ in the Z-string fields have on the fermion (u and d quarks) zero modes. The zero momentum modes acquire an $O(\epsilon )$ mass while the non-zero momentum modes get an $O(\epsilon^2 )$ mass. For the perturbation analysis to make sense, we require that the u and d quark zero momentum modes are either both filled or both empty. In that case, the $O(\epsilon )$ terms in the variation in the energy will cancel and we will be left with something that is $O(\epsilon^2 )$. In fact, \be \Delta E = - {{\epsilon^2}\over 2} \vert m_1 \vert ^2 L \sum_{k=1}^N {1 \over k} \end{equation} where $m_1$ is a matrix element having to do with the interactions of the u and d quarks, $L \rightarrow \infty$ is the length of the string on which periodic boundary conditions have been imposed, and $N \rightarrow \infty$ is a cut-off on the energy levels which are labeled by $k$. The crucial piece of this formula is the minus sign which shows that the energy of the string is lowered due to perturbations \cite{Nac95}. In Ref. \cite{LiuVac96} it was argued that an identical calculation could be done for any classically stable string that could terminate on (supermassive) magnetic monopoles. However, in the low energy theory, the strings are effectively topological and hence, it seems unlikely that fermions can lead to an instability. This suggests that the bosonic string configuration gets modified by the fermions and the stability analysis around the Nielsen-Olesen solution may be inappropriate. So far, the stability analysis with fermions presented here only considered the zero modes and ignored the infinitely many massive fermion modes. Very recently, Groves and Perkins \cite{GroPer99} have analysed the full spectrum of massless and massive fermionic modes in the background of the electroweak string. They then calculate the effect of the Dirac sea on the stability of electroweak strings by calculating the renormalised energy shift of the Dirac sea when a Z-string is perturbed by introducing a non-zero upper component to the Higgs doublet. This energy shift is negative and so destabilises the string, but it is small, leading them to conclude that if positive energy fermionic states are populated, it is conceivable that the total fermionic contribution could be to stabilise the string. This work is still in progress. In the meantime, the stability of Z--strings remains an open question. \begin{figure}[tbp] \caption{\label{pertonfermions} The effect of perturbations of the Z-string on fermion zero modes. } \vskip 1 truecm \epsfxsize = \hsize \epsfbox{pertonfermions.eps} \end{figure} \subsection{Scattering of fermions off electroweak strings} \label{scattering} The elastic scattering of fermions off semilocal and electroweak strings has been considered in \cite{Gan93,DavMarGan94,Lo95}. The main feature of the cross section is that the scattering violates helicity \cite{Gan93}. It is straightforward to show that the helicity operator $\Sigma \cdot \Pi$, where $\Sigma^i = \epsilon^{ijk} \gamma^i\gamma^j$ is the spin operator and $\Pi^i$ are the canonical momenta, does not commute with the hamiltonian. If $\Phi^T = (\phi^+, \phi^0)$, the commutator is proportional to $(D\phi^0)$ terms. Consider for a moment the usual representation of Dirac matrices, \be \gamma^0 = \pmatrix{0 \ \ \ \ 1 \cr 1 \ \ \ \ 0} \qquad\qquad \gamma^i = \pmatrix{0 \ \ -\tau^i \cr \tau^i \ \ \ \ 0} \qquad\qquad \gamma_5 = \pmatrix{1 \ \ \ \ \ 0 \cr 0 \ \ -1} \end{equation} Then, for an incoming electron, one finds \be [{\rm H},{ \Sigma}\cdot{\Pi}]=i h \left( \begin{array}{cc} 0 & \tau^{j}(D^{j}\phi^{0})^{\dagger}\\ \tau^{j}D^{j}\phi^{0} & 0\\ \end{array} \right), \end{equation} where $h$ is the Yukawa coupling and $(D_j\phi^0)$ is given in eq. (\ref{smalldmu}). Therefore helicity-violating processes can take place in the core of the string. A preliminary calculation by Ganoulis in ref. \cite{Gan93} showed that, for an incoming plane wave, the dominant mode of scattering gives identical cross sections for positive and negative helicity scattered states. More precisely, for an incoming electron plane wave of momentum $k$, energy $\omega$ and positive helicity it was found that, to leading order, \be { d\sigma \over dk}\biggr |_\pm \sim {1 \over k} \left( \omega - k \over 2\omega \right)^2 \sin^2 (\pi q_R) \end{equation} where $\omega^2 = k^2 + m_e^2$, $q_R$ is the Z-charge of the right fermion field, given in eq. (\ref{zcharge}) (recall that right and left fermion fields have different Z-charges, $q_R = q_L \pm 1$). A more detailed calculation was done by Davis, Martin and Ganoulis \cite{DavMarGan94}, and later extended by Lo \cite{Lo95}, using a `top hat' model \be f(r) = \cases{ 0 &$r<R$ \cr {\eta / \sqrt{2}} &$r>R$ \cr} \qquad\qquad\qquad v(r) = \cases{ 0 &\ \ $r<R$ \cr {2 / g_z} &\ \ $r>R$ \cr} \ \ , \end{equation} which is expected to be a reasonable approximation since the scattering cross section in the case of cosmic strings has been shown to be insensitive to the core model \cite{PerPerDavBraMat91}. Note that there is a discontinuous jump in the fermion mass and string flux; however the wave functions are matched so that they are continuous at $r=R$. Note that the left and right fields decouple in the core of the string, so helicity violating processes are concentrated at $r=R$. The authors of \cite{DavMarGan94,Lo95} confirmed that, in the massive case, there are helicity-conserving and helicity-flip scattering cross sections of equal magnitude. The latter goes to zero in the massless limit (in that case, the left and right fields decouple, and no helicity violation is possible), suggesting that helicity violation may be stronger at low energies. For ``fractional string flux'' ({\it i.e.} for fractional $q$) the cross section is of a modified Aharonov-Bohm form, and independent of string radius. For integer $q$ it is of Everett form \cite{Eve81} (the strong interaction cross section is suppressed by a logarithmic term). Another interesting feature has to do with the amplification of the fermionic wave function in the core of the string. Lo \cite{Lo95} has remarked that there is a regime in which the scattering cross section for electroweak strings is much less sensitive to the fermion charge (that is, to $\sin^2 \theta_w$) than for cosmic strings. In contrast with, {\it e.g.}, baryon number violating processes, which show maximal enhancement only for discrete values of the fractional flux, the helicity violating cross section for electroweak strings in the regime $k \sim m, \ kR << 1$ shows a plateau for $0< \sin^2 \theta_w < 1/2$ where amplification is maximal and the cross section becomes of order $m_f^{-1}$. This can be traced back to the asymmetry between left and right fields; while the wave function amplification is a universal feature, different components of the fermionic wave function acquire different amplification factors in such a way that the total enhancement of the cross section is approximately independent of the fermionic charge, $q$ (or, equivalently, of $\sin^2 \theta_w$). Elastic scattering is independent of the string radius for both electroweak and semilocal strings (for integral flux there is only a mild dependence on the radius coming from the logarithmic suppresion factor in the Everett cross section). Since the cross section is like that of $U(1)$ strings, we would expect electroweak and semilocal strings to interact with the surrounding plasma in a way that is analogous to topological strings. \section{Electroweak strings and baryon number} \label{ewstringbaryonnumber} As first shown by Adler \cite{Adl69}, and, Bell and Jackiw \cite{BelJac69}, currents that are conserved in a classical field theory may not be conserved on quantization of the theory. In the GSW model, one such current is the baryon number current and the anomalous current conservation equation is: \begin{equation} \partial_\mu j^\mu _{B} = {{N_F} \over {32\pi^2}} [ -g^2 W^a _{\mu \nu} {\tilde W}^{a \mu \nu} + {g'}{}^2 Y_{\mu \nu} {\tilde Y}^{\mu \nu} ]. \label{anomalyeq} \end{equation} where $j^\mu _B$ is the expectation value of the baryon number current operator $\sum_s b_s :{\bar \psi} \gamma^\mu \psi:$ where the sum is over all the species of fermions labeled by $s$, $\psi$ is the fermion spinor and $b_s$ is the baryon number for species $s$ and the operator product is normal ordered. Also, $N_F$ denotes the number of families, and tilde the dual of the field strengths. The anomaly equation can be integrated over all space leading to \begin{equation} \Delta Q_{B} = {{N_F} \over {32\pi^2}} \int dt d^3 x [ -g^2 W^a _{\mu \nu} {\tilde W}^{a \mu \nu} + {g'}{}^2 Y_{\mu \nu} {\tilde Y}^{\mu \nu} ] = \Delta Q_{CS}. \label{intanomeq} \end{equation} with, \be Q_{CS} = {{N_F} \over {32\pi^2}} \int d^3 x \epsilon_{ijk} \biggl [ g^2 \biggl ( W^{a ij} W^{a k} - {g \over 3} \epsilon_{abc} W^{ai} W^{bj} W^{ck} \biggr ) - {g'}{}^2 Y^{ij} Y^{k} \biggr ] \ . \label{qcs} \end{equation} Here, $\Delta ( \cdot )$ denotes the difference of the quantities evaluated at two different times, $Q_B$ is the baryonic charge and the surface currents and integrals at infinity are assumed to vanish. $Q_{CS} $ is called the Chern-Simons, or topological, charge and can be evaluated if we know the gauge fields. The left-hand side of eq. (\ref{intanomeq}) evaluates the baryon number by counting the fermions directly. We describe the evaluation of both the right- and left-hand side for fermions on certain configurations of Z-strings in the following subsections. Finally, in Sec. \ref{cosmologicalapplications} we briefly comment on possible applications to cosmology. \subsection{Chern-Simons or topological charge} \label{chernsimons} We will be interested in the Chern-Simons charge contained in configurations of Z-strings. Then, we set all the gauge fields but for the Z-field to zero in the expression for the Chern-Simons charge, yielding \begin{equation} Q_{CS} = N_F {{\alpha ^2 } \over {32\pi^2}} \cos(2\theta_w) \int d^3 x {\vec Z}\cdot {\vec B}_Z \label{qcsbrief} \end{equation} where, ${\vec B}_Z$ denotes the magnetic field in the Z gauge field: $B^i_Z = \epsilon^{ijk} \partial_jZ_k $. The terms on the right-hand side have a simple interpretation in terms of a concept called ``helicity'' in fluid dynamics \cite{BerFie84}. Essentially, if a fluid flows with velocity $\vec v$ and vorticity $\vec \omega = \vec \nabla \times \vec v$, then the helicity is defined as: \be h = \int d^3 x {\vec v} \cdot {\vec \omega} \label{fluidhelicity} \end{equation} Since the helicity measures the velocity flow along the direction of vorticity, it measures the corkscrew motion (or twisting) of the fluid flow. A direct analog is defined for magnetic fields: \be h_B = \int d^3 x {\vec A} \cdot {\vec B} \label{magnetichelicity} \end{equation} which is of the same form as the terms appearing in (\ref{qcsbrief}). Hence the Chern-Simons charge measures the twisting of the magnetic lines of force. The helicity associated with the Z field alone is given by: \be H_Z = \int d^3 x {\vec Z}\cdot {\vec B}_Z \ . \label{zhelicity} \end{equation} If we think in terms of flux tubes of $Z$ magnetic field, $H_Z$ measures the sum of the link and twist number of these tubes: \be H_Z = L_Z + T_Z \ . \label{linktwist} \end{equation} For a pair of unit winding Z flux tubes that are linked once as shown in Fig. \ref{linkedloops} the helicity is: \be H_Z = 2 F_Z ^2 \label{helicityoflink} \end{equation} where, $F_Z$ is the magnetic flux in each of the two tubes Note that the helicity is positive for the strings shown in Fig. \ref{linkedloops}. If we reversed the direction of the flux in one of the loops, the magnitude of $H_Z$ would be the same but the sign would change. For the Z-string, we also know that \be F_Z = {{4\pi} \over g_z} \label{zfluxforlink} \end{equation} and so eq. (\ref{qcsbrief}) yields \cite{VacFie94}: \be Q_{CS} = N_F \cos (2\theta_w ) \ . \label{qcsresult} \end{equation} \begin{figure}[tbp] \caption{\label{linkedloops} A pair of linked loops.} \vskip 1 truecm \epsfxsize = \hsize \epsfbox{linkedloops.eps} \end{figure} \subsection{Baryonic charge in fermions} \label{baryoninfermion} The baryon number associated with linked loops of Z-string has been evaluated in Ref. \cite{GarVac95} by studying the fermionic zero modes on such loops. This corresponds to evaluating the left-hand side of eq. (\ref{intanomeq}) directly in terms of the fermions that carry baryon number. The calculation involves adding the baryonic charges of the infinite Dirac sea of fermions living on the string together with zeta function regularization. To understand why the linking of loops leads to non-trivial effects, note that the quarks and leptons have a non-trivial Aharanov-Bohm interaction with the Z-string. So the Dirac sea of fermions on a loop in Fig. \ref{linkedloops} is affected by the Z-flux in the second loop. This shifts the level of the Dirac sea in the ground state leading to non-trivial baryonic and other charges. Instead of considering the linked loops as shown in Fig. \ref{linkedloops} it is simpler to consider a large circular loop of radius $a \rightarrow \infty$ in the $xy$-plane threaded by $n$ straight infinite strings along the $z$-axis (Fig. \ref{threading}). Then the fermionic wave-functions take the form: \begin{equation} \psi_L = e^{- i(E_p t - p \sigma )} \psi_L^{(0)}(r) \ , \ \ \ \psi_R = e^{- i(E_p t - (p-n/a) \sigma )} \psi_R^{(0)}(r) \ \label{propmodes} \end{equation} where the functions with superscript $(0)$ are the zero mode profile functions described in Sec. \ref{zeromodes} and $\sigma$ is a coordinate along the length of the circular loop. From these wavefunctions, the dispersion relation for a zero mode fermion on the circular loop is \begin{equation} \omega_k = \epsilon_i ( k - qZ ) \ . \label{omega} \end{equation} where, $q$ is the Z-charge of the fermion, $\epsilon_i$ is defined in eq. (\ref{epsi1}), $\omega$ is related to the energy $E$ by $\omega \equiv aE$, and $k$ to the momentum $p$ by $k \equiv a p \in {\rm \angle \!\!\! Z}$. $Z$ is the component of the gauge field along the circular loop multiplied by $a$ and is given by, \begin{equation} Z \equiv {{2 n} \over {g_z}} \ . \label{zna} \end{equation} The crucial property of the dispersion relation is that, if there is an Aharanov-Bohm interaction between the Z-string and the fermion, $\omega_k$ cannot be zero for any value of $k$ since $k$ is an integer but $qZ$ is not. The Z- , A- and baryon number (B) charges of the leptons and quarks are shown in Table \ref{tab:effluents}. Note that we use $2q_Z/g_z$ to denote the Z-charge and this is identical to the eigenvalue of the operator ${\bf q}$ defined in eq. \ref{zcharge} and also to $q$ used in the previous section. \begin{figure}[tbp] \caption{\label{threading} A circular Z-string loop of radius $a$ threaded by $n$ Z-strings.} \vskip 1 truecm \epsfxsize = \hsize \epsfbox{threading.eps} \end{figure} \begin{table*}[hbt] \setlength{\tabcolsep}{1.5pc} \newlength{\digitwidth} \settowidth{\digitwidth}{\rm 0} \catcode`?=\active \def?{\kern\digitwidth} \caption{Summary of $Z-$, electric and baryonic charges for the leptons and quarks. The charges $q_Z$ are for the left-handed fermions and $s^2 \equiv \sin^2 \theta_w$.} \smallskip \label{tab:effluents} \begin{tabular*}{\textwidth}{@{}l@{\extracolsep{\fill}}cccc} \hline & \multicolumn{1}{c}{$\nu_L$} & \multicolumn{1}{c}{$e$} & \multicolumn{1}{c}{$d$} & \multicolumn{1}{c}{$u$} \\ \hline \smallskip $2q_Z/g_z$ &-1 &$ 1-2s^2 $ &$1-{{2s^2}/ 3}$ &$-1+{{4s^2}/ 3}$ \\ $q_A/e$ &0 &-1 & $-1 /3$ & $2 / 3$ \\ $q_B$ &0 &0 &$1/ 3$ & $1/ 3$ \\ \hline \multicolumn{5}{@{}p{120mm}}{} \end{tabular*} \end{table*} The energy of the fermions is found by summing over the negative frequencies - that is, the Dirac sea - and so the energy $E$ due to a single fermion species is: \begin{equation} E = {1 \over a} \sum \omega_k = \epsilon_i {1 \over a} \sum_{k=k_F}^{-\epsilon_i \infty} ( k - q Z ) \label{energy} \end{equation} where, $k_F$ denotes the Fermi level - the value of $k$ for the highest filled state. Therefore we need to sum a series of the type: \begin{equation} S = \sum_{k=k_F}^{\infty} ( k - q Z ) = \sum_{k=0}^{\infty} ( k+k_F - q Z ) \ . \label{series} \end{equation} The sum is found using zeta function regularization: \begin{equation} S = \zeta (-1, k_F -q Z) = - {1 \over {12}} - {1 \over 2} (k_F-qZ)(k_F-q Z-1) \label{sum} \end{equation} With this result, the energy contribution from the $i^{th}$ species of fermions takes the form: \begin{equation} E_i = -{1 \over {24a}} + {1 \over {2a}} \biggl [ k_F^{(i)} - q_i Z + {\epsilon_i \over 2} \biggr ] ^2 \equiv -{1 \over {24a}} + {1 \over {2a}} K_i ^2 \ . \label{ei} \end{equation} Adding the contributions due to different members of a single fermion family, we get \begin{equation} 2 a E = K_\nu ^2 +K_e^2 + 3K_d^2 + 3 K_u^2 \ . \label{gene} \end{equation} Next we can calculate the angular momentum of the fermions in the circular loop background. The system has rotational symmetry about the $z-$axis and this enables us to define the generalized angular momentum operator as the operator that annihilates the background field configuration \cite{JamPerVac92}: \begin{equation} M_z = L_z + S_z + n I_z \label{gam} \end{equation} where, \begin{equation} L_z = -i {\bf 1} {{\partial \ } \over {\partial \varphi}} \ , \label{lz} \end{equation} $S_z$ is the spin operator, and, the isospin operator is given in terms of the $U(1)$ (hypercharge) and $SU(2)$ charges - $q_1$ and $q_2$ respectively - of the field in question: \begin{equation} I_z = {1 \over 2} \biggl [ \biggl ( {{2q_2} \over g} \biggr ) T^3 - \biggl ( {{2q_1} \over {g'}} \biggr ) {\bf 1} \biggr ] \ . \label{iz} \end{equation} The isopin operator acts via a commutator bracket on the gauge fields and by ordinary matrix multiplication on the Higgs field and fermion doublets. We are interested in the angular momentum of the chiral fermions on the circular loop which lies entirely in the $xy$-plane. The fermions in the zero modes therefore have $S_z =0$. (The spin of the fermions is oriented along their momenta which lies in the $xy$-plane.) The action of $L_z$ is found by acting on the fermion wave-functions such as in eq. (\ref{propmodes}) (remembering to let $n \rightarrow -n$ for the neutrino and up quark). The action of $I_z$ is found by using the charges of the fermions given in the GSW model defined in Sec. \ref{fermionicsector}. We then find: \begin{equation} M_z \pmatrix{\nu_L\cr e_L\cr} = \pmatrix{(k^{(\nu )} +n)\nu_L \cr k^{(e)} e_L\cr} \ , \ \ \ M_z \pmatrix{u_L\cr d_L\cr} = \pmatrix{(k^{(u)} +{n/ 3})u_L \cr (k^{(d)} -{{2n}/ 3}) d_L\cr} \ , \label{mzpsil} \end{equation} \begin{equation} M_z e_R = k^{(e)} e_R \ , \ \ M_z u_R = (k^{(u)} +{n/ 3}) u_R \ , \ \ M_z d_R = (k^{(d)} - {{2n}/ 3}) d_R \ \label{mzul} \end{equation} where the $k^{(i)}$ are defined above eqn. (\ref{omega}). Now summing over states, as in the case of the energy, we find the total generalized angular momentum of the fermions on the circular loop: \begin{equation} {\cal M} = {1\over 2} \biggl [ k_F ^{(\nu )} +n +{1\over 2} \biggr ] ^2 - {1\over 2} \biggl [ k_F ^{(e)}-{1\over 2} \biggr ]^2 - {3\over 2} \biggl [ k_F ^{(d)}-{{2n}\over 3} -{1\over 2}\biggr ] ^2 + {3\over 2} \biggl [ k_F ^{(u)}+{{n}\over 3} +{1\over 2}\biggr ] ^2 \ . \label{totgam} \end{equation} Note that though the gauge fields do not enter explicitly in the generalized angular momentum, they do play a role in determining the angular momentum of the ground state through the values of the Fermi levels. The calculation of the electromagnetic and baryonic charges and currents on the linked loops is similar but has a subtlety. To find the total charge, a sum over the charges in all filled states must be done. This leads to a series of the kind: \begin{equation} S_q = \sum_{k=k_F}^{ \infty} 1 \ . \label{sq1} \end{equation} To regularize the divergence of the series, it is written as \begin{equation} S_q =\lim_{\lambda\to 0} \sum_{k=k_F}^{ \infty} (k-q Z)^{\lambda} \ . \label{sq2} \end{equation} The subtlety is that the gauge invariant combination $k-qZ$ is used as a summand rather than $k$ or some other gauge non-invariant expression \cite{Man85}. Once again zeta function regularization is used to get: \begin{equation} S_q = \sum_{k=0}^{\infty} (k+k_F -q Z)^0 = \zeta(0, k_F-q Z ) = - \biggl [ k_F -qZ - {1 \over 2} \biggr ] \ . \label{sqdone} \end{equation} With this result, the contribution to the charge due to fermion $i$ is: \begin{equation} Q_i = \epsilon_i {\bar q}_i \biggl [ k_F^{(i)} -q_i Z + {\epsilon_i \over 2} \biggr ] = \epsilon_i {\bar q}_i K_i \label{qi} \end{equation} where, ${\bar q}_i$ is the charge carried by the $i^{th}$ fermion of the kind that we wish to calculate. (Note that ${\bar q}_i$ can represent any charge - electric, baryonic {\it etc.} - and is, in general, different from the Z-charge $q_i$.) The currents along the string are given by ${\bar \psi} \gamma^z \psi$ where $\gamma^z$ is given in eq. (\ref{gamma2}). This gives \begin{equation} J_i = \epsilon_i {{Q_i} \over {2\pi a}} \ . \label{ji} \end{equation} By adding the contributions due to each variety of fermion, expressions for the energy, angular momentum, charges and currents for one loop threaded by $n$ have been found in \cite{GarVac95}. These results are reproduced in Table \ref{tab:results}. It is reassuring to note that in the ground state, the baryon number of the single loop is given by $nN_F \cos2\theta_W$ in agreement with the calculation of the Chern-Simons number. \begin{table*}[hbt] \setlength{\tabcolsep}{.99pc} \catcode`?=\active \def?{\kern\digitwidth} \caption{Expressions for the energy, generalized angular momentum, charges and currents in terms of $x = 2n\sin^2\theta_w /3$. We have omitted the multiplicative factor $N_F$ in all the expressions for convenience. } \smallskip \label{tab:results} \begin{tabular*}{\textwidth}{@{}l@{\extracolsep{\fill}}cccc} \hline \smallskip & \multicolumn{1}{c}{$x \in (0,1/3)$} & \multicolumn{1}{c}{$(1/3,1/2)$} & \multicolumn{1}{c}{$(1/2,2/3)$} & \multicolumn{1}{c}{$(2/3,1)$} \\ \hline \smallskip $aE$ &$12x^2-6x+1$ &$12x^2-9x+2$ &$12x^2-15x+5$ &$12x^2-18x+7$ \\ $M$ &0 &$n-1$& $2-n$ & 0 \\ $Q_A /e$ &0 &-1 & +1 & 0 \\ $B$ &$-3x+1$ &$-3x+1$ &$-3x+2$ &$-3x+2$ \\ $2\pi a J_A /e$ &$-8x+2$ &$-8x+3$ &$-8x+5$ &$-8x+6$ \\ $2\pi a J_B$ &$-x$ &$-x$ &$1-x$ &$1-x$ \\ \hline \multicolumn{5}{@{}p{120mm}}{} \end{tabular*} \end{table*} \begin{figure}[tbp] \caption{\label{evsx} The energy of the ground state of linked loops versus $x = 2n\sin^2\theta_w /3$.} \vskip 1 truecm \epsfxsize = \hsize \epsfbox{evsx.ps} \end{figure} The energy of the fermionic ground state shows a complicated dependence on $x$ as is demonstrated in Fig. \ref{evsx}. Note that $E(x)$ does not have a monotonic dependence on $x$ and the energy of strings that are linked $n$ times bears no simple relation to those linked $m$ times. In particular, the energy does not continue to decrease as we consider strings that have higher linkage. The lowest energy possible, however, is when $x=1/4$ and for $n=1$, this corresponds to $\sin^2 \theta_w = 3/8$, which is also the value set by Grand Unified models. It is not clear if this is simply a coincidence or if there is some deeper underlying reason \cite{Kep98}. \subsection{Dumbells} \label{dumbells} In his 1977 paper, Nambu discussed the possible occurrence of electroweak monopoles and strings in particle accelerators. There are two issues in this discussion: the first is the production crossection of solitonic states in particle collisions, and the second is the signatures of such states if they are indeed produced in an accelerator. The answer to the first question is not known though it is widely believed that the process is suppressed not only by the large amount of energy required but also due to the coherence of the solitonic state. The second question was addressed by Nambu \cite{Nam77} and he estimated the energy and lifetime of electroweak strings that may be possible to detect in accelerators. To find the energy of a Z-string segment, Nambu treated the monopoles at the ends as hollow spheres of radius $R$ inside which all fields vanish. A straightforward variational calculation in units of $\eta \approx 246$ GeV then gives the monopole mass \be M = {4 \pi \over 3 e} \sin ^{5/2}\theta_w \sqrt{m_H \over m_W} \end{equation} and radius \be R = \sqrt{ \sin\theta_w \over m_H m_W} \end{equation} The string segment is approximated by a cylindrical tube with uniform $Z$ magnetic flux with all other fields vanishing. This gives \be \rho = {2 \over \sqrt{m_H m_Z}} \ \ , \qquad\qquad \tau = \pi \left( {m_H \over m_Z} \right ) \end{equation} for the core radius and string tension, respectively. Now, if the monopoles are a distance $l$ apart, the total energy of the system is \be E = 2M -{Q^2 \over 4\pi l} + \tau l \end{equation} which is clearly minimised by $l=0$ {\it i.e.} the string can minimize its energy by collapsing. The tendency to collapse can be countered by a centrifugal barrier if the string segment (``dumbell'') is rotating fast enough about a perpendicular axis. The energy and angular momentum of a relativistic dumbell has been estimated by Nambu to be: \be E \sim \half \pi l \tau \qquad L \sim {1\over 8} \pi l^2 \tau \end{equation} where, \be {l \tau \over 2 M} = {v^2 \over 1-v^2} \end{equation} with $v \sim 1$ being the velocity of the poles. The expressions for $E$ and $L$ imply the existence of asymptotic Regge trajectories, \be L \sim \alpha_0 ' E^2 \end{equation} with slope \be \alpha_0 ' = {1 \over {2\pi \tau}} \sim \left( {m_Z \over m_H} \right) {\rm TeV}^{-2}\ . \end{equation} which, if found, would be a signature of dumbells. The orbiting poles at the ends of the rotating dumbell will radiate electromagnetically and this energy loss provides an upper bound to the lifetime of the configuration. An estimate of the radiated power from the analysis of synchrotron radiation in classical electrodynamics ({\it eg.} see \cite{Jac75}) gives \be P \sim {{8\pi} \over 3} \times 137 \left ({\tau \over M}\right) ^2 \sin^4 \theta_w \end{equation} Therefore the decay width $\Gamma = P/E$ is given by \be \Gamma \simeq {E \over L} \end{equation} and for large angular momentum, can lead to significant lifetimes (compared to $E^{-1}$). To obtain numerical estimates, note that the above estimates are valid only if the dumbell length is much greater than the width of the Z-string. This imposes a lower bound on the angular momentum: \be L >> {\pi \over 2} \times 137 \sin^2 \theta_w \cos^2\theta_w \sim 36 \end{equation} Using the relation between the energy and the angular momentum, such an object has $E >> 6 (m_H/m_Z)^{1/2}$ TeV. The estimates above assume that the lifetime of the dumbell is dictated by the energy emission in photons. In reality, there are other decay channels as well, though it is likely that these will be comparitively suppressed since the photon is the only massless boson present in the system. The dumbell can also decay by fragmenting due to field-theoretic instabilities of the kind discussed in Sec. \ref{stability}. These may be suppressed due to the finite size of the dumbell, and as Nambu points out, due to the angular momentum of the dumbell \footnote{In the stability analysis for a finite piece of string of length $L$, the eigenvalues of the stability equation are shifted by a contribution of order $\pi^2/ L^2$ with respect to the infinitely long case, thus for sufficiently short segments the radial decay mode could become stable. Longitudinal collapse might then be stabilized by rotation, as explained above}. A careful analysis of these factors has not yet been performed and is a vital open problem that may become experimentally relevant with the next generation of accelerators. \subsection{Possible cosmological applications} \label{cosmologicalapplications} The role of electroweak strings in cosmology depends on their abundance during and after the electroweak phase transition. If this abundance is negligible, electroweak strings may at best only be relevant in future accelerator experiments (see Sec. \ref{dumbells}). If, however, there is a cosmological epoch during which segments and loops of electroweak strings were present, they could impact on two observational consequences: the first is the presence of a primordial magnetic field, and the second is the generation of a cosmological baryon number. What is perhaps most remarkable is that the two consequences might be related - the baryonic density of the universe would be related to the helicity of the primordial magnetic field \cite{Vac94,Rob89}. (i) {\it Primordial magnetic fields:} A gas of electroweak segments is necessarily accompanied by a gas of electroweak monopoles. The eventual collapse and disappearance of electroweak strings removes all the electroweak monopoles but the long range magnetic field emanating from the monopoles is expected to remain trapped in the cosmological plasma since that is a very good electrical conductor. This will then lead to a residual primordial magnetic field in the present universe. A quantitative estimate of the resulting primordial magnetic field cannot be made with confidence but a dimensional estimate is possible. An estimate for the average flux through an area $L^2 = N^2/T^2$, where $N$ is a dimensionless number that relates the length scale of interest, $L$, to the cosmological thermal correlation length $T^{-1}$, was obtained in \cite{Vac91,Vac94}, and then translated into the average magnetic field through that area. The result is: \be B\vert_{area} \sim T^2 /N \ . \label{magneticfieldresult} \end{equation} (Magneto-hydrodynamical considerations provide a lower bound $\sim 10^{12}$ cms on $L$ at the present epoch.) It is important to remember that the above is an areal ({\it i.e.} flux) average, defined by \cite{EnqOle93} \be B\vert_{area} \equiv \biggl \langle \biggl ( {1\over A} \int d{\vec S} \cdot {\vec B} \biggr ) ^2 \biggr \rangle ^{1/2} \label{averagebdefn} \end{equation} where the surface integral is over an area $A$ and $\langle \cdot \rangle$ denotes ensemble averaging. (ii) {\it Baryon number:} A gas of electroweak string segments and loops would, in general, contain some helicity density of the Z-field. When the electroweak strings eventually annihilate, it is possible that the helicity gets converted into baryon number \cite{VacFie94,Vac94}. However, in Ref. \cite{FarGolGutRajSin95,FarGolLueRaj96} it is argued that fractional quantum numbers of a soliton are unrelated to the number of particles produced when the soliton decays. Instead, only the change in the winding of the Higgs field in a process that starts out in the vacuum and ends up in the vacuum can be related to the particle number. This would imply that we would have to consider the formation of electroweak strings together with their decay before we can find the resulting baryon number. Such a calculation has not yet been attempted. An interesting question is to consider what happens to the helicity in the Z-field after the strings disappear. One possibility is that the helicity gets transferred to a frozen-in residual magnetic field after the strings have decayed. To see this, consider a linked pair of loops as in Fig. \ref{linkedloops}. The strings can break by nucleating monopole-antimonopole pairs, and then the string segments can shrink, finally leading to monopole annihilation. If this process happens in the early universe, the loops will be surrounded by the ambient plasma which will freeze-in the magnetic field lines. Hence, after the strings have disappeared, we will be left with a linked pair of magnetic field lines. In other words, the original helicity in the Z-field has been transferred to helicity in the A-field. This argument relies on the freezing-in of the magnetic field emanating from the monopoles and in the real setting the physics can be much more complicated. However, a connection between the baryon abundance of the universe and the properties of a primordial magnetic field seems tantalizing. {\it Stable strings at the electroweak scale:} If in more exotic models, strings at the electroweak scale were stable and had the superconducting properties discussed above, they could be responsible for baryogenesis \cite{Bar95} and the presence of primary antiprotons in cosmic rays \cite{StaVac96}. The production of antiprotons follows on realizing that any strings tangled in the galactic plasma would be moving across the galactic magnetic field. In the rest frame of the string, the changing magnetic field causes an electric field along the string according to Faraday's law. The electric field along the string raises the levels of the u- and d-quark Dirac seas (see Fig. \ref{zeromode}), as well as the electron Dirac sea (not shown in the figure). This means that the electric field produces quarks and leptons on the string. The electric charges of the particles are in the ratio $e:u:d::-1:+2/3:-1/3$ and the rate of production of these particles due to the applied electric field is proportional to the charges. Furthermore, the quarks come in three colors and so for every electron that is produced, $3\times 2/3 =2$ u-quarks and $3\times 1/3=1$ d-quark are also produced. As a result, the net electric charge produced is $1\times (-1)+2\times (2/3)+1\times (-1/3) = 0$. However, net baryon number $2\times (1/3)+1\times (1/3)=1$ is produced because the quarks carry baryonic charge $1/3$ while the baryonic charges of the leptons vanish. Depending on the orientation of the string, either baryons or antibaryons will be produced. Some of these would then be emitted from the string and would arrive on earth as cosmic rays. {\it Formation of strings in the electroweak phase transition.} Early attempts to understand the formation rates of electroweak strings were made in \cite{Vac94} based on the statistical mechanics of strings. The estimates indicate that a density of strings will be formed immediately after the phase transition. However, the application of string statistical mechanics to electroweak strings may not be justified and so other avenues of investigation are needed. An alternative approach to study electroweak string formation was taken by Nagasawa and Yokoyama \cite{NagYok96}. They assumed a thermal distribution of scalar field values and gradients, and estimated the probability of obtaining a string-like scalar field configuration. The conclusion was that electroweak vortex formation in a thermal system is totally negligible. One possible caveat is that the technique used in \cite{NagYok96} ignores the effect of gauge fields, which we know are significant in the formation of related objects such as semilocal strings. In \cite{SafCop97}, Saffin and Copeland have evolved the classical equations of motion to study the formation of electroweak strings, and they found the presence of the gauge fields led to larger string densities than one would have inferred from the scalar fields alone, at least when $\sin^2 \theta_w = 0$. However, this study does not directly address the question of string formation in a phase transition because no measure has been placed on the choice of initial conditions and their choice may be too restrictive. Most recently, a promising development has taken place \cite{Che98} - calculations in lattice gauge theory have been done to study the electroweak phase transition and there is evidence that electroweak strings will form. Further studies along these lines will provide important and quantitative insight into the formation of electroweak strings. Using the results on the formation of semilocal strings, we can gain some intuition about the formation of electroweak strings in the region of parameter space close to the semilocal limit (the region of stability in Fig. \ref{stabilityregion}). We have seen that semilocal strings with $\beta <1$ have a non-zero formation rate, increasing as $\beta \to 0$. Initially short segments of string are seen to grow and join nearby ones because this reduces the gradient energy at the ends of the strings. The ends of electroweak strings are proper magnetic monopoles, and therefore the scalar gradients are cancelled much more efficiently by the gauge fields, but as $\sin \theta_w \to 0$ the cores of the monopoles get larger and larger, and they could begin to overlap with nearby monopoles, so it is possible that short segments of electroweak string will also grow into longer ones. \begin{figure}[tbp] \caption{\label{zeromode} The dispersion relations for the u and d quark zero modes are shown. The filled states are denoted by solid circles while dashes denote unfilled states. For convenience, periodic boundary conditions are assumed along the string and so the momentum takes on discrete values. } \vskip 1 truecm \epsfxsize = \hsize \epsfbox{udzeromodes.eps} \vskip 0.5 truecm \end{figure} \section{Electroweak strings and the sphaleron} \label{ewstringandsphaleron} The sphaleron is a classical solution in the GSW model that carries baryon number $N_F/2$, where $N_F$ is the number of fermion families \cite{Man83,KliMan84}. For $\theta_w =0$, the asymptotic form of the sphaleron Higgs field is: \be \Phi_{sph} = \pmatrix{\cos\theta\cr \sin\theta ~ e^{i\varphi}\cr} \ . \label{phisphaleron} \end{equation} while the gauge fields continue to be given by eq. (\ref{gwbarmua})-(\ref{gpbarymu}) in which ${\bar \Phi}$ should be replaced by $\Phi_{sph}$. (Note that the hypercharge gauge field vanishes for $\theta_w =0$.) Inside the sphaleron, the Higgs field vanishes at one point. The sphaleron also has a magnetic dipole moment that has been evaluated for small values of $\theta_w$. The reason that the sphaleron is important for particle physics is that its energy defines the minimum energy required for the classical violation of baryon number in the GSW model. As has already been described in Sec. \ref{ewstringbaryonnumber}, non-trivial baryon number can be associated with linked and twisted segments of electroweak string. Further, for specific values of the link and twist, the baryon number of a configuration of Z-strings can also be $N_F/2$. This raises the question: are sphalerons related to Z-string segments? An early paper to draw a connection between the various solutions in the GSW model is Ref. \cite{FujOtsToy89}. In \cite{VacFie94,HinJam94,Vac94,Hin94}, however, a direct correspondence between the field configuration of the Z-string and the sphaleron was made. \subsection{Content of the sphaleron} \label{sphaleroncontent} In \cite{HinJam94} Hindmarsh and James evaluated the magnetic charge density and current density within the sphaleron. A subtlety in this calculation is that there is no unique definition of the electromagnetic field when the Higgs field is not everywhere in the vacuum. The choice adopted in \cite{HinJam94} (and also the choice in this review) is \be F_{ij}^{em} = \sin\theta_w W_{ij}^{a} n^a + \cos\theta_w Y_{ij} \ . \label{fijem} \end{equation} The evaluation of the magnetic charge density (which is proportional to the divergence of the magnetic field strength) clearly shows that the sphaleron contains a region with positive magnetic charge density and a region with negative magnetic charge density. Furthermore, the total charge in, say, the positive charge region agrees with the magnetic charge of a monopole. In addition, there is a flux of Z magnetic field connecting the two hemispheres. This would seem to confirm that the sphaleron consists of a Z-string segment. However, this is not the full picture. In addition to the string segment, Hindmarsh and James find that the electric current is non-zero in the equatorial region and is in the azimuthal (${\hat e}_\varphi$) direction. \subsection{From Z-strings to the sphaleron} \label{stringstosphaleron} The scalar field configuration for a finite segment of Z-string was given in Sec. \ref{dumbells}: \begin{equation} \Phi_{m\bar m} = \pmatrix{\cos(\Theta /2) \cr \sin(\Theta /2) ~ e^{i\varphi} \cr } \label{phimbarm} \end{equation} where, \be \cos\Theta \equiv \cos\theta_m - \cos\theta_{\bar m} +1 \label{angledefn} \end{equation} and the angles $\theta_m$ and $\theta_{\bar m}$ are measured from the monopole and antimonopole respectively, as shown in Fig. \ref{coords}. \begin{figure}[tbp] \caption{\label{coords} Definition of the coordinate angles $\theta_m$ and $\theta_{\bar m}$. The azimuthal angle, $\varphi$, is not shown. } \vskip 1 truecm \epsfxsize = \hsize \epsfbox{dumbellcoords.eps} \end{figure} It is straightforward to check that (\ref{phimbarm}) yields the monopole field configuration close to the monopole ($\theta_{\bar m} \rightarrow 0$) and the antimonopole configuration close to the antimonopole ($\theta_{m} \rightarrow \pi$). It also yields a string singularity along the straight line joining the monopole and antimonopole ($\theta_m = \pi , ~ \theta_{\bar m} = 0$). However, there are other Higgs field configurations that also describe monopoles and antimonopoles: \begin{equation} \Phi_m = e^{i\gamma} \pmatrix{ \cos(\theta_m /2) \cr \sin(\theta_m /2) ~e^{i\varphi }\cr } \ , \ \ \ \ \Phi_{\bar m} = e^{i\gamma} \pmatrix{ \sin(\theta_{\bar m} /2) \cr \cos(\theta_{\bar m} /2) ~ e^{i\varphi}} \ . \label{phimphibarm} \end{equation} Next consider the Higgs field configuration: \begin{equation} \Phi_{m \bar m} (\gamma ) = \pmatrix{ \sin(\theta_m /2) \sin(\theta_{\bar m} /2) e^{i\gamma} + \cos(\theta_m /2) \cos(\theta_{\bar m} /2) \cr \sin(\theta_m /2) \cos(\theta_{\bar m} /2) e^{i\varphi } - \cos(\theta_m /2) \sin(\theta_{\bar m} /2) e^{i(\varphi - \gamma )} \cr } \label{twistedzstring} \end{equation} together with the gauge fields given by eq. (\ref{gwbarmua})-(\ref{gpbarymu}) with ${\bar \Phi}$ replaced by $\Phi_{m\bar m} (\gamma )$. When we take the limit $\theta_{\bar m} \rightarrow 0$ we find the monopole configuration (with $\gamma =0$) and when we take $\theta_m \rightarrow \pi$ the configuration is that of an antimonopole (with arbitrary $\gamma$) provided we perform the spatial rotation $\varphi \rightarrow \varphi + \gamma$. Note that the asymptotic gauge fields agree since these are determined by the Higgs field. The monopole and antimonopole in (\ref{twistedzstring}) also have the usual string singularity joining them. This means that the configuration in eq. (\ref{twistedzstring}) describes a monopole and antimonopole pair that are joined by a Z-string segment that is twisted by an angle $\gamma$. The Chern-Simons number of one such segment can be calculated \cite{VacFie94} and is \be Q_{CS} = N_F \cos2\theta_w ~ {{\gamma} \over {2\pi}} \ . \label{qcsonesegment} \end{equation} If $\gamma = \pi /\cos(2\theta_w )$ then the Chern-Simons number of the twisted segment of string is $N_F /2$ and is precisely that of the sphaleron. Given that the segment with twist $\pi /\cos(2\theta_w )$ has Chern-Simons number equal to that of the sphaleron, it is natural to ask if some deformation of it will yield the sphaleron. This deformation is not hard to guess for the $\theta_w = 0$ case. In this case, if we let the segment size shrink to zero, we have $\theta_m = \theta_{\bar m} = \theta$ and the Higgs field configuration of eq. (\ref{twistedzstring}) gives: \be \Phi_{m \bar m} (\gamma = \pi ) = \pmatrix{\cos\theta\cr \sin\theta ~ e^{i\varphi}\cr} \ . \label{phimbarmatpi} \end{equation} This is exactly the scalar field configuration of the sphaleron for $\theta_w =0$ (eq. (\ref{phisphaleron})). Note that the asymptotic gauge fields continue to be given by eq. (\ref{gwbarmua})-(\ref{gpbarymu}) and satisfy the requirement that the covariant derivatives of the Higgs field vanish. Encouraged by this successful connection in the $\theta_w =0$ case, it was conjectured in \cite{VacFie94,Vac94} that the sphaleron can also be obtained by collapsing a twisted segment of Z-string with Chern-Simons number $N_F /2$ {\it for any $\theta_w$}. If true, this would mean that the asymptotic Higgs field configuration, $\Phi_S$, for the sphaleron for arbitrary $\theta_w$ is given by \be \Phi_S = \pmatrix{ \sin^2(\theta /2) ~ e^{i\gamma_S} + \cos^2(\theta /2) \cr \sin(\theta /2) \cos(\theta /2) ~ e^{i\varphi} (1 - e^{-i\gamma_S})\cr } \label{phisphaleronthetaw} \end{equation} where $\gamma_S = \pi /\cos(2\theta_w )$. The twisting of the magnetic field lines in the sphaleron configuration has been further clarified in \cite{Hin94}. The direction of magnetic field lines is shown for a dumbell in Fig. \ref{marksintra2} and for a ``stretched'' sphaleron in Fig. \ref{marksintra3}. (The asymptotic fields for the stretched sphaleron are identical to those for the sphaleron and the twisted Z-string.) In the stretched sphaleron case, the magnetic field line twists around the vertical string segment by an angle $\pi$ (for $\theta_w \rightarrow 0$) as one goes from monopole to antimonopole. This twist provides non-trivial Chern-Simons number to the configuration \cite{VacFie94}. \begin{figure}[tbp] \caption{\label{marksintra2} The thick solid line is the location of the Z-string for a dumbell configuration and the dashed curves lie in the equatorial plane and are drawn to guide the eye. The dotted lines depict lines of magnetic flux. The arrows show the orientation of the vector ${\hat n} \propto - \Phi^{\dag} {\vec \tau} \Phi$. } \vskip 1 truein \epsfxsize = \hsize \epsfbox{marksintra_f2.eps} \end{figure} \begin{figure}[tbp] \caption{\label{marksintra3} The field configuration for a stretched sphaleron as in Fig. \ref{marksintra2}. Only one magnetic field line is shown. } \vskip 1 truein \epsfxsize = \hsize \epsfbox{marksintra_f3.eps} \end{figure} On physical grounds it seems reasonable that there should be a critical value of twist at which one can get a static solution for a Z-string segment. This is because the segment likes to shrink under its own tension but the twist prevents the shrinkage and is equivalent to a repulsive force between the monopole and antimonopole. (This idea owes its origin to Taubes \cite{Tau82} who discovered a solution containing a monopole and an antimonopole in an O(3) model in which the Coulomb attraction is balanced by the relative misorientation of the magnetic poles.) Then, if the string is sufficiently twisted, the attractive force due to the tension and the repulsive force due to the twist will balance and a static solution can exist. So far we have been assuming that the only dynamics of the segment is towards collapsing or expanding of the string segment. However, since we are dealing with twisted segments, we should also include the rotational dynamics associated with twisting and untwisting. So, while any twist greater than a certain critical twist might successfully prevent the segment from collapsing, only a special value of the twist can give a static solution to the rotational dynamics. Furthermore, we expect that this solution will be unstable towards rotations that twist and untwist the string segment. This would be the unstable mode of the sphaleron. Similar connections between the W--string and the sphaleron have also been constructed in \cite{AxeJohNieTor96}. \section{The $^3$He analogy} \label{he3section} The symmetry structure of $^3$He closely resembles the electroweak symmetry group and hence we expect the analog of electroweak strings to exist in $^3$He \cite{VolWol90,Vol92,VolVac96}. Indeed, this analog is called the $n=2$ vortex. We now explain this correspondence in greater detail. \subsection{Lightning review of $^3$He} \label{lightening} $^3$He nuclei have spin $1/2$ and two such nuclei form a Cooper pair which is the order parameter for the system. Unlike $^4$He, the pairing is a spin triplet ($S=1$) as well as an orbital angular momentum triplet ($L=1$). As a result there are $3\times 3$ components of the wavefunction of the Cooper pair - that is the order parameter has 9 complex components. Hence, the order parameter is written as a 3 by 3 complex valued matrix: $A_{\alpha i}$ with $\alpha$ (spin index) and $i$ (spatial index) ranging from 1 to 3. At temperatures higher than a few milli Kelvin the system is invariant under spatial rotations ($SO(3)_L$) as well as rotations of the spin degree of freedom of the Cooper pair ($SO(3)_S$). Another symmetry is under overall phase rotations of the wavefunction ($U(1)_N$) and the corresponding conserved charge is particle number ($N$). Hence the symmetry group is: \be G = SO(3)_L \times SO(3)_S \times U(1)_N \ . \label{he3symmetry} \end{equation} There are several possible phases of $^3$He corresponding to different expectation values of the order parameter. In the A-phase, the orbital angular momenta of the Cooper pairs are all aligned and so are the spin directions. This corresponds to \be A_{\alpha i} = \Delta_0 {\hat d}_\alpha \psi_i \label{he3aorderparameter} \end{equation} where $\Delta_0 \sim 10^{-7}$ eV is the temperature dependent gap amplitude, the real unit vector ${\hat d}_\alpha$ is the spin part of the order parameter, and \be \psi_i = {{{\hat m}_i + i{\hat n}_i} \over {\sqrt{2}}} \label{psiimini} \end{equation} with ${\hat m}$ and ${\hat n}$ being orthogonal unit vectors, is the orbital part of the order parameter. This expectation value of the order parameter leads to the symmetry breaking: \be G \rightarrow U(1)_{S_3}\times U(1)_{L_3-N/2} \times Z_2\ . \label{he3asymbreaking} \end{equation} The reason why a $U(1)$ subgroup of $SO(3)_L \times U(1)_N$ survives the symmetry breaking can be derived from the expectation value in eq. (\ref{he3aorderparameter}). A spatial rotation of the order parameter is equivalent to a phase rotation of $\psi_i$ and this phase can be absorbed by a corresponding $U(1)_N$ rotation of the order parameter. Hence, just as in the electroweak case, a diagonal $U(1)$ subgroup remains unbroken. The $U(1)_{S_3}$ survives since rotations about the ${\hat d}$ axis leave the order parameter invariant. The non-trivial element of the residual discrete $Z_2$ symmetry corresponds to a sign inversion of both $\psi_i$ and ${\hat d}$. A depiction of the A- and B- phases is shown in Fig. \ref{he3phases} (after \cite{Liu82}). \begin{figure}[tbp] \caption{\label{he3phases} Depiction of the A- and B- phases of $^3$He. In the A-phase, the spin orientations of all the Cooper pairs are parallel and so are the orbital orientations. In the B-phase, the relative orientation of the spin and orbital orientations are fixed in all the Cooper pairs but neither the spin nor the orbital orientations of the various Cooper pairs are aligned. } \vskip 1truecm \epsfxsize = \hsize \epsfbox{he3phases.eps} \end{figure} In the B-phase, neither the orbital angular momenta nor the spin directions of the different Cooper pairs are aligned. But the angle between the direction of the angular momenta and the spin direction is fixed throughout the sample. Hence in the B-phase, independent rotations of the orbital angular momenta and of spin are no longer symmetries. However, a simultaneous rotation of both orbital angular momenta and spin remains an unbroken symmetry. In other words, a diagonal subgroup of $SO(3)_S \times SO(3)_L$ remains unbroken. Therefore, in the B-phase the order parameter is written as: \be A_{\alpha i} = 3^{-1/2} e^{i\phi} R_{\alpha i} ({\hat n}, \theta ) \label{he3borderparameter} \end{equation} where, $\phi$ is a phase and the $3\times 3$ matrix $R_{\alpha i}$ describes relative rotations of the spin and orbital degrees of freedom about an axis ${\hat n}$ and by angle $\theta$. The symmetry breaking pattern is: \be G \rightarrow SO(3)_{L_3+S_3} \ . \label{he3bsymbreaking} \end{equation} This symmetry breaking resembles the chiral symmetry breaking transition studied in QCD (with two flavors of quarks) and may be useful for experimentally investigating phenomenon such as the formation of ``disoriented chiral condensates'' \cite{Bjo97}. The B-phase does not resemble the electroweak model and hence we will not discuss it any further. We shall also not discuss the various other phases of $^3$He (for example, the ${\rm A}_1$ phase) which are known to occur. (For a useful chart of the phases, see Sec. 6.2 of Ref. \cite{VolWol90}.) In addition to the continuous symmetries, there are a number of discrete symmetries that arise in the phases of $^3$He. These are important for the classification of topological defects in $^3$He. A description may be found in \cite{SalVol87}. \subsection{Z-string analog in $^3$He} \label{zstringinhe3} Clearly the A-phase closely resembles the electroweak symmetry breaking because of the mixing of the generator of the non-Abelian group ($SO(3)_L$) and the Abelian group ($U(1)_N$). The orbital part of the order parameter is responsible for this pattern of symmetry breaking and hence $\psi_i$ plays the role of the electroweak Higgs field $\Phi$. The connection, however, is indirect since $\psi_i$ is a complex 3 vector while $\Phi$ is a complex doublet. The idea is that the $^3$He-A real vector \be {\hat l}_{HeA} = i{ {{\vec \psi} \times {{\vec \psi}^{\dag}}} \over {{\vec \psi}^{\dag} {\vec \psi}} } = {\hat m} \times {\hat n} \label{lhathea} \end{equation} is analogous to the electroweak real vector \be {\hat l}_{ew} = - {{\Phi^{\dag} {\vec \tau} \Phi} \over {\Phi^{\dag} {\vec \tau} \Phi} } \ . \label{hlatew} \end{equation} The electroweak Z-string is a non-topological solution for which the Higgs field configuration is: \be \Phi = {\eta \over {\sqrt{2}}} f(r) e^{i\varphi} \pmatrix{0\cr 1\cr} \ . \end{equation} For this configuration ${\hat l}_{ew} = {\hat z}$. The vacuum manifold $M_A$ of $^3$He-A has \be \pi_1 ( M_A ) = Z_4 \label{pi1he3a} \end{equation} and hence there are topological $Z_4$ vortices in $^3$He-A. The vortices occur in classes labeled by $n=\pm 1/2,1$. The vortices with $n$ equal to an even integer are topologically equivalent to the vacuum. The non-trivial topological vortices (labeled by $n=\pm 1/2,1$) cannot be the equivalent of the non-topological Z-string. However, the topologically trivial $n=-2$ vortex is also seen in $^3$He-A. The order parameter for this vortex is: \be A_{\alpha j}(\rho ,\varphi)= \Delta_0 \hat z_{\alpha }[ \ e^{in\varphi} f_1(\rho )(\hat x_j+i \hat y_j) +\ e^{i(n+2)\varphi} f_2(\rho )(\hat x_j-i \hat y_j)]~~. \label{opforneq2} \end{equation} where $f_1(\rho )$ and $f_2(\rho )$ are two profile functions with $f_1 (\infty )=1$, $f_2(\infty ) =0$, $f_1 (0) =0$ and $f_2 (0)$ depending on $n$. In correspondence with the electroweak Z-string, the $n=2$ vortex has ${\hat l}_{HeA} = {\hat z}$. However, the order parameter need not vanish at the center of the vortex for certain members of the $n=2$ class of vortices. For example, with $n=-2$, we may have $f_2(0)\ne 0$. The $n=2$ vortex is not topological and can be continuously deformed into the vacuum manifold. The configuration at the terminus of the $n=2$ vortex is called the hedgehog or monopole ${\hat l}_{HeA} = {\hat r}$ (the radial unit vector). This texture is the direct analog of the electroweak magnetic monopole (${\hat l}_{ew} = {\hat r}$) at the terminus of a Z-string. The $n=2$ discontinuous vortex is unstable but even so has been observed in $^3$He. In the laboratory, the rotation of the sample stabilizes the $n=2$ vortex. This seems to be closely analogous to the result of Garriga and Montes \cite{GarMon95} who find that electroweak strings can be stabilized by external magnetic fields (Sec. \ref{stabilitycontd}). Before proceeding further, it is prudent to remind ourselves of some important differences between the (bosonic sector of the) GSW model and $^3$He. The symmetries in $^3$He are all global whereas the symmetries in the GSW model are all local. So the $n=2$ discontinuous vortex is like a global analog of the Z-string. Another important difference is in the discrete symmetries in the two systems. The symmetry structure of the GSW model is really $[SU(2)\times U(1)]/Z_2$ since the $Z_2$ elements ${\bf 1}$ and $- {\bf 1}$ which form the center of $SU(2)$ also occur in $U(1)$. On the contrary, the symmetry group of $^3$He-A has a multiplicative $Z_2$ factor which gives rise to the non-trivial topology of the vacuum manifold. It is important to note that we cannot expect $^3$He to provide an exact replica of the GSW model. However, the similar structures of the two systems means that certain issues can be experimentally addressed in the $^3$He context while they are far beyond the reach of current particle physics experiments. An issue of this kind is the baryon number anomaly in the GSW model and the anomalous generation of momentum in $^3$He. As described in Sec. \ref{superconductivity}, there are fermionic zero modes on the Z-string and an electric field applied along the Z-string leads to the anomalous production of baryon number. What is the corresponding analog in $^3$He? At first sight, $^3$He does not have the non-Abelian gauge fields that the electroweak string has and so it seems that the analogy is doomed. But this is not true. The point is that the physics of fermionic zero modes has to do with the dynamics of fermions on the {\em fixed} background of the Z-string. Likewise, in $^3$He we can be interested in the dynamics of quasiparticles in the fixed background of the $n=2$ vortex. As far as the interaction of quasiparticles with the order parameter background is concerned, one can think of the $^3$He-A vortex as being due to a to a (fictitious) gauge field ${{Z^\prime}^\mu}$. Then the interaction of quasiparticles with the order parameter is of the form $j_\mu {{Z^\prime}^\mu}$ which is exactly analogous to the interaction of quarks and leptons with the Z-boson. Just as in the electroweak case, the $^3$He quasiparticles have zero modes on the vortex. In close analogy with the scenario where the motion of a superconducting string through an external magnetic field leads to currents along the string (Sec. \ref{cosmologicalapplications}), the velocity of the $^3$He vortex through the superfluid leads to an anomalous flow of quasiparticles but this time in the direction perpendicular to the vortex. This flow causes an extra force on the vortex as it moves through the superfluid that can be monitored experimentally. Such a force was measured in the Manchester experiment \cite{Bevetal97,Vol98} and is in excellent agreement with theoretical predictions. Hence the Manchester experiment verifies the anomalous production of quasiparticle momentum on moving vortices and the corresponding production of baryon number on electroweak strings moving through a magnetic field. \section{Concluding remarks and open problems} \label{conclude} Quantum field theory has been very successful in describing particle physics. Yet the successes have mostly been relegated to perturbative phenomena. A more spectacular level of success will be achieved when our field theoretic description of particle physics is confirmed at the non-perturbative level. The first non-perturbative objects that are likely to be encountered in this quest are topological defects and their close cousins that we have described in this review. The search for topological defects can be conducted in accelerator experiments or in the cosmological realm via astronomical surveys. These searches are complementary - only supermassive topological defects can be evident in astronomical surveys, while only the lightest defects can potentially be produced in accelerators. Foreseeable accelerator experiments give us access only to topological defects at the electroweak symmetry breaking scale. So it is very important to understand the defects present in the standard electroweak model and all its viable extensions. One may hope that the structure of defects will yield important clues about the underlying symmetry of the standard model. With this hope, we have described wide classes of defects present in field theories. These defects are not all topological and this is relevant to the standard electroweak model which also lacks the non-trivial topology needed to contain topological defects. The absence of topology in the model means that the defect solutions cannot be enumerated in topological terms and neither can their stability be guaranteed. We have described, however, how the existence of defect solutions may still be derived by examining the topological defects occurring in subspaces of the model. The electroweak defects can be thought of as being topological defects that are embedded in the electroweak model. The issue of stability of the defect solution is yet more involved and has not yet been fully resolved in the presence of fermions. That the electroweak Z-string is stable for large $\theta_w$ (within the bosonic sector) was inspired by the discovery of semilocal strings and their stability properties. The explicit stability analysis of the electroweak string marks out the region of parameter space in which the Z-string is stable. Then it is clear that the Z-string is unstable for the parameters of the standard model. In certain viable extensions of the standard model and under some external conditions (such as an external magnetic field), the standard electroweak Z-string can still be stable. Even if the Z-string is unstable, it is possible that the lifetime of segments of string is long enough so that they can be observed in accelerators. This possibility was discussed in the first paper on the subject by Nambu \cite{Nam77}. The discovery of Z-string segments would truly be historic since it would confirm the existence of magnetic monopoles in particle physics. However, the rate of formation of Z-string segments and their lifetime has not yet been studied in detail. Some of the difficulties in this problem lie quite deep since they involve the connection of perturbative particle physics to the non-perturbative solitonic features. Additionally, the influence of fermions on electroweak strings needs further investigation. Electroweak strings may play a cosmological role in the genesis of matter over antimatter as is evident since configurations of electroweak string have properties that are similar to the electroweak sphaleron. The challenge here is to determine the number density of electroweak strings formed during the electroweak phase transition and their decay rate. Note that the formation of topological strings has been under constant examination over the last two decades and only now, with some experimental input, are we beginning to understand their formation. The cosmological formation of electroweak defects has not been addressed with as much vigour. Recently though, there have been spurts of activity in this area, with lattice calculations beginning to shed interesting insight \cite{Che98}. It is very likely that further lattice results will be able to give quantitative information about the formation of electroweak strings at the electroweak phase transition. While particle physics experiments to detect electroweak strings are quite distant, experiments in condensed matter systems to study topological defects are becoming more feasible and can be used to test theoretical ideas that are relevant to both particle physics and condensed matter physics. Already there are experiments that test theories of the formation of topological vortices. We can also expect that condensed matter experiments might some day test the formation of defects that are not topological. The experiments on He$^3$ are most relevant in this regard since it contains close analogs of electroweak strings. Furthermore, ideas relating to the behaviour of fermions in the background of electroweak strings can also be tested in the realm of He$^3$. This makes for exciting physics in the years to come which will stimulate the growth of particle physics, cosmology and both, theoretical and experimental, condensed matter physics. \section{Acknowledgements} We are grateful to G. Volovik for very useful comments on section \ref{he3section}, to M. Groves and W. Perkins for an early draft of their paper, and to M. Hindmarsh for the sphaleron figures in section \ref{ewstringandsphaleron}. AA thanks K. Kuijken and L. Perivolaropoulos for help with some of the figures in sections \ref{NO} and \ref{semilocal}, and J. Urrestilla for pointing out several typos in an earlier draft. This work was supported by a NATO Collaborative Research Grant CRG 951301, and our travel was also partially supported by NSF grants PHY-9309364 (AA), as well as by UPV grant UPV 063.310-EB187/98 and CICYT grant AEN-93-0435. TV was partially supported by the Department of Energy, USA. TV thanks the University of the Basque Country, and AA thanks Case Western Reserve University, Pierre Van Baal and Leiden University for their hospitality.
\section{Introduction} The determination of the stellar locus in the HR diagram is a subject of the prime importance in astrophysics, as well as it has wide applications. For instance, the determination of the distance scale relies much on the uniqueness of the stellar locus. Such work has often resorts to the knowledge of theoretical isochrones, since the observations alone do not span sufficiently large parameter space. On the other hand, theoretical isochrones, expressed in the colour-magnitude space, may suffer from errors of three origins. The errors may arise from (i) evolution track calculations which depend on opacity, nuclear reaction rate, equation of states and the treatment of convection, (ii) conversion of temperature to colour index (colour-temperature relation), and (iii) conversion of luminosity to magnitude in a specific passband (bolometric corrections). The recent findings of the difference in the distance to open clusters by Hipparcos parallax as compared to the traditional zero age main sequence (ZAMS) fitting (van Leeuwen \& Hansen Luiz 1997; van Leeuwen 1999; Mermilliod et al. 1997; Pinsonneault et al. 1998) have tempted us to study the problem of reliability concerning theoretical isochrones. An analysis of Nordstr\"om, Andersen \& Andersen (1997) indicates that the discrepancy of the isochrones among different authors can be as much as 0.4$-$0.6 mag, although in most practical applications the isochrones are used so that such a large error does not directly affect the results. Their figures also indicate that a dominant part of the errors may arise from the colour-temperature relation, especially when $B-V$ colour is used, as we have also confirmed from our own analysis. The error arising from the bolometric correction is rather small, and the evolution track on the luminosity temperature plane is reasonably converged among authors, in so far as we are concerned with the region far from the turn-off point where convective overshooting may start making a difference. Another particularly important application of the stellar track is stellar population synthesis of galaxy colours, which play an important role in cosmology (e.g., Tinsley \& Gunn 1976; Bruzual \& Charlot 1993; Kodama \& Arimoto 1997). If colour of giant stars would contain errors as much as $\Delta (B-V)>0.05$, the interpretation of elliptical galaxies could significantly be disturbed. In this paper we focus on the problem of the colour-temperature ($T_{\rm eff}$) relation. We attempt to construct the relation, which we think the least model dependent, and study what errors are contained in the existing relations. There are a lot of work for the colour-temperature relation. The early authority is the one given by Johnson (1966), and the work to 1980 is summarised by B\"ohm-Vitense (1980). We also quote several representative examples, which include Code et al. (1976), Blackwell \& Shallis (1977), Bessell (1979), Ridgway et al. (1980), Saxner \& Hammarba\"ck (1985), Arribas \& Mart\'inez-Roger (1988), Tsuji et al (1995) and Di Benedetto \& Rabbia (1987). Most recent work includes Flower (1996), Alonso et al. (1996b), Blackwell \& Lynas-Gray (1998; hereafter BL98), and Lejeune, Cuisinier \& Buser (1998). The prime difficulty lies in estimating precise temperature. The most direct method employs the measurement of angular diameter of stars using interferometry or lunar occultations. The number of stars which are given accurate angular diameters are increasing (Davis 1998), especially with the advancement of the Michelson interferometry technique, but not yet sufficiently many to explore large parameter space. One of the methods which are supposed to be accurate and often used in modern literature is the infrared flux method (IRFM), in which $T_{\rm eff}$ is estimated from the measurement of $R=F_{\rm bol}/F_\lambda=\sigma T_{\rm eff}^4/\psi(\lambda)$ for infrared $\lambda$, where $\psi(\lambda)$ is calculated from model atmospheres (Blackwell \& Shallis 1977). This method has been developed to reduce the model dependence using the fact that $F_\lambda$ in the near infrared regions is smoothly proportional to $T_{\rm eff}$ and model dependence is fairly small. Alonso, Arribas \& Mart\'inez-Roger (1996a,b) and BL98 have given the latest and most extensive work employing this method. M\'egessier (1994) and Alonso et al. (1996a) have examined the error associated with this method. The authors of both papers claim that the derived temperature differs as much as 100K depending on the model atmosphere employed in the work. For example use of ATLAS9 (Kurucz 1993) gives temperature 100K higher than ATLAS8. Similar difference is also reported between MARCS (Gustafsson et al. 1975 and their updates) and ATLAS9. An important advancement is brought by Di Benedetto (1998; hereafter B98) who found an empirically very tight relationship between $T_{\rm eff}$ and $V-K$ colour, which is calibrated with the direct angular diameter measurement. He has shown that this relationship depends very little on luminosity class and metal abundance. This makes possible to estimate temperature for F-K dwarfs, for which direct measurements of diameters are still lacking. This method would offer the least model-dependent method to deal with a fairly large sample without a direct angular size measurement for each star. We have examined the accuracy of the B98 temperature, and are convicted that it is perhaps the best method available to us for the time being. B98, however, has not discussed much with $B-V$ colour, which is very sensitive to line blanketing and also to surface gravity, giving a large scatter around the surface brightness colour relation. This is also true with a recent comparative study of Bessell, Castelli \& Plez (1998), who extensively compared model atmosphere calculations with empirically estimated $T_{\rm eff}$ for various colour bands, but except for $B-V$. On the other hand, most of the applications of the stellar locus still rely predominantly on $V$ magnitude and $B-V$ colour. For this reason we attempt to construct a $B-V$ colour temperature relation adopting the $T_{\rm eff}$ data from B98 together with an extensive photometric database compiled by Hauck \& Mermilliod (1998), combined with the metallicity and log $g$ data compiled by de Strobel et al. 1997 (hereafter SSFRF) for F0-K7 stars. We also examine the available $B-V$ colour temperature relations, especially those published recently, against what we have constructed in order to find external errors of these works. Further examination is also made for the colour-temperature relation used by theoretical work of stellar evolution, which is often used as a basis to discuss the cosmic distance scale and age, as well as taken as a fiducial for stellar population synthesis for colour of galaxies. As for the required accuracy for the distance work, if our goal is to obtain a 3\% accuracy in the distance estimation, one would need to achieve the accuracy of $B-V$ to be 0.01 mag, which in turn is translated to temperature error of 20$-$45 K. In our paper we attempt to document the error budget arising from many components of the input data. This would clarify the limiting factor to the accuracy of the $B-V$ colour temperature relation and tell us what improvement should be done to go further. We shall see that the accuracy that we can achieve is worse than this goal by about a factor of 2$-$3 for each star. When many stars in clusters are averaged, however, the accuracy is of the order of 0.015 mag in $B-V$. One of our additional aims is to extract the dependence of the $B-V$ colour temperature relation on metallicity, without resorting to theoretical grids, and in particular to examine the accuracy of synthetic results from model atmospheres. As a byproduct this makes possible to estimate colour of the sun from solar analogues (e.g., Taylor 1998; de Strobel 1996 for reviews). In section 2, we derive the $B-V$ colour temperature relation, after examining the quality and reliability of input data. We also present a table of the estimated error budget. In section 3, an extensive comparison is made with the existing $B-V$ colour temperature relations based on either empirical or theoretical ground. We discuss in section 4 $B-V$ colour of the sun. Metallicity dependence is discussed in section 5. Our conclusion is given in section 6. \section{Construction of the $B-V$ colour temperature relation} \subsection{Data} To derive a relation between $T_{\rm eff}$ and $B-V$ colour, we start with 537 ISO standard stars for which B98 has given accurate estimates of $T_{\rm eff}$. Of those 537 starts, 270 stars are given estimates of [Fe/H] and $\log g$ by SSFRF. Among them about 40\% (110 stars) are dwarfs or subdwarfs and 60\% (160 stars) are giant stars. For a majority of entries SSFRF gives more than one data for [Fe/H] and $\log g$ for a given star; in such cases we take median values of [Fe/H] and $\log g$. The distributions of the median values of [Fe/H] and $\log g$ are shown in Figs. 1 (a) and (b). The data are distributed widely from [Fe/H] =$-$2.0 to +0.5; only 10\% of stars have [Fe/H] $<-1.0$, but we still have a reasonable number of stars to constrain our analysis in this region. The ten percentile for the metal rich side is [Fe/H] $= 0.12$. All stars of our sample are given $B-V$ colours (Hauck \& Mermilliod 1998) and parallaxes with Hipparcos (ESA 1997). We give in Table 1 the derivative of $T_{\rm eff}$ against $B-V$ to obtain an idea about the propagation of errors from $T_{\rm eff}$ to $B-V$. The table gives $\Delta T_{\rm eff}$ that causes a change of $\Delta (B-V)$=0.01. For instance, $\Delta T_{\rm eff}=45$K is an allowance for F0 stars (7000K); it is 30 K for G2 stars, and 16K for K5 stars at 4000 K. \subsection{Examination of B98's temperature estimates} \label{sec_teff} None of the temperature estimates are completely free from the atmosphere model. Spectroscopic determinations of temperature directly rely on the details of the atmosphere model, and IRFM uses the prediction of the atmosphere model in the near infrared region. Even the direct measurement of angular diameters should be supplemented with the atmosphere model, albeit with a minimal extent, in order to estimate the bolometric flux in the invisible regions. The method proposed by B98 basically belongs to this last category. He found a tight correlation between surface brightness and $V-K$ colour (dispersion being 0.03 mag) to estimate angular size $\phi$, and used another tight relationship between the bolometric flux $F_{\rm bol}$ and $V-K$ colour found by BL98 to estimate effective temperature of stars for which the direct measurements are not available for $\phi$ and $F_{\rm bol}$. Since it is of crucial importance to examine the accuracy of effective temperature estimated with this method, we plot in Fig. 2 the difference of temperature given by B98 from $V-K$ colours and that obtained directly using \begin{equation} T_{\rm eff}^4 = \frac{4}{\sigma} \phi^{-2} F_{\rm bol}, \label{teff_eq} \end{equation} which is equivalent to the defining equation for effective temperature. Here $\sigma$ is the Stefan-Boltzmann constant. The table given by B98 (Table 3) contains all angular size measurements which could be used for our test (21 stars and the sun). We take the bolometric flux from direct evaluations of several sources, as presented in Table 2. The bolometric flux is an integration of flux from $U$ to $K$ bands with shorter and longer wavelength ranges evaluated with the aid of atmosphere models. The direct integration accounts for 93\% of flux at $T$=4500K, and 83\% at 7000K (Alonso et al 1995). Therefore, the dependence on model atmosphere is expected to be quite small. It is expected that the error is no more than 2\% for the bolometric flux, which means a 0.5\% error in $T_{\rm eff}$. We also note that there is an error of this order (1.5\%) in the absolute calibration of flux of $\alpha$ Lyr at 5556\AA~(Hayes 1985). We explicitly document the calibrations employed by the respective authors in Table 2 (ref/norm), but do not dare to adjust to the same scale, since the difference is smaller than errors of other origins. Among 22 stars of B98 direct bolometric flux estimates are available for 16 stars. We take each estimate by different authors as an independent data, so that 32 data points are contained in Fig. 2. The errors attached to our $T_{\rm eff}$ are obtained by a quadrature of the errors for $F_{\rm bol}$ and $\phi$ given by the authors. We also added the data points from a similar test of B98 himself (Table 5 of B98) and from the $T_{\rm eff}$ estimate of BL98 when available. For stars later than F5 type ($V-K\ge 1.0$), errors in $\phi$ and those in $F_{\rm bol}$ are comparable, and their quadratures ($\delta T_{\rm eff}$) are also comparable to the difference between the B98 estimate and our ``true'' (reference) temperature ($\Delta T_{\rm eff}({\rm B98})=T_{\rm eff}({\rm B98})-T_{\rm eff}({\rm ref)}$). We see some systematic trend that B98 temperature is slightly lower than the reference temperature by 30$-$40K. We note an excellent agreement between BL98 and B98. For stars earlier than F0 type the angular size measurement yields too large errors to carry out an accurate test. For all stars formal errors exceed the difference between of the two temperature estimates. Except for $\varepsilon$ Sgr and one point for $\alpha$ Lyr, however, B98 temperature agrees very well with our reference within $\approx \pm$ 50K, although error bars are larger. For $\alpha$ Lyr (BS 7001) the two independent flux determinations disagree by 4.8\%, which is not explained merely by the different absolute calibrations they take. On the other hand, BL98 tend to give temperature 100-200K higher than ours in this region. From this test we conclude that B98's estimate of temperature is correct allowing for errors of 30$-$40K at least for F0-K8 stars, ($V-K=0.7-3.5$). We confine ourselves to the range $T_{\rm eff}<7000$ (F0 or later), for which the error of $T_{\rm eff}$ is small and the result is reliable to a high accuracy. \subsection{Errors of $B - V$} \label{sec_bv} We take the mean values of $B-V$ compiled by Hauck \& Mermilliod (1998). 81\% of the stars in our sample have a dispersion of less than 0.01 among multiple observations, and 93\% of the stars have less than 0.014 mag dispersion. As a conservative estimate we infer the average photometric error to be 0.01 mag, which corresponds to 15-45 K in our $T_{\rm eff}$ range. Most of stars in our sample are located nearby: 61\% of our stars are within 50 pc, 21\% lie between 50 pc and 100 pc, and 11\% between 100 pc and 150 pc. Therefore, the necessary extinction correction is a minimum amount. Nevertheless, we apply the extinction correction $E(B-V)/d=0.235$ mag/kpc, or $A_V/d=0.8$ mag/kpc ($d$ being the distance) with $R=3.4$, taken from Blackwell et al. (1990). We examine the validity of this extinction correction when we obtain a fit of the form $T_{\rm eff}=T_{\rm eff}((B-V)_0, [{\rm Fe/H]}, \log g)$, and confirm that extinction correction is indeed indispensable to obtain a good fit; selective extinction of $E(B-V)/d=0.269$/kpc gives a minimum to the residuals of the fit (see below). Since the adoption of this value hardly modifies the results of the fit, we take 0.235 mag/kpc for our final results and we take the difference of the two values evaluated at 100 pc, i.e., 0.0034 mag, as a representative error from the extinction correction. \subsection{Errors of [Fe/H] and $\log g$ estimates} \label{sec_feh} The scatter of the [Fe/H] values documented in a catalogue of SSFRF is $<0.15$ dex for 80\% of our sample, and $<0.2$ dex for 87\% of the sample. Our fit given below shows a derivative $\partial T_{\rm eff}/\partial$[Fe/H]$\simeq$ 320 K/dex. Therefore, the error of the [Fe/H] measurement causes 50 K in the determination of $T_{\rm eff}$. The uncertainty of $\log g$ does not cause much errors in $T_{\rm eff}$. The scatter of SSFRF data is about 0.2 in log $g$ units; the derivative of our fit $\partial T_{\rm eff}/\partial \log g=-30$K means the error being about 6 K. We also expect this order of scatter in $\log g$ from a star to a star at a given stellar mass (Andersen 1991). \subsection{Results of the Fit} We present in Fig. 3 $T_{\rm eff}$ as a function of unreddened $B-V$ colour for all 283 stars. 13 out of 16 calibrating stars in Table 2 have [Fe/H] and $\log g$ data in SSFRF, and they are also included in the plot. In this figure we classified stars according to metallicity: solid circles are stars with [Fe/H]$\le-$0.75, open squares for $-0.75<$[Fe/H]$\le-0.25$, cross symbols for $-0.25<$[Fe/H]$\le$+0.25, and solid triangles are for [Fe/H]$>$+0.25. The sample covers the range $0.3\le B-V \le 1.5$, which is the range of our analysis. In Fig. 4 are selected only dwarfs (and sub-dwarfs). In this sample stars with $B-V>0.9$ are scanty, and we must limit our study to the range between $B-V=0.3$ and 0.9. In carrying out our fitting, we exclude 17 stars, 10 of them as having $B-V<0.3$, one of them is too distant (700 pc away) and remaining 6 being located more than 4$\sigma$ away from the locus of the fit. Our fitting is made in the following steps. First we fit the samples (full and dwarf samples) simply with $T_{\rm eff}=T_{\rm eff}((B-V)_0)$ ignoring the [Fe/H] and $\log g$ dependence; we find the rms residual of the fit to be 123K for the full sample and 151 K for the dwarfs. We then carry out a fit with the form \begin{equation} \log T_{eff} = c_0 + c_1 (B - V)_0 + c_2 (B - V)_0^2 + c_3 (B - V)_0^3 + f_1 {\rm [Fe/H]} + f_2 {\rm [Fe/H]}^2 + g_1 \log g. \label{teff_fit} \end{equation} with an equal weight given to all data points. This largely reduces the rms of residuals. We find 66 K (1.1\%) for the full sample and 71 K (1.2\%) for dwarfs. This rms is somewhat smaller than that of BL98 (80$-$90 K), and smaller by a factor of two than is given by Alonso et al. (1996b) (130 K). In fitting the dwarf sample, we fixed $\log g$ to be constant ($\log g=4.3$) (see below), and $g_1$ to the value derived from the fit to the full sample, since the variation of $\log g$ for main sequence stars in this colour range is too small to constrain the fit. The parameters of this fit are given in Table 3. Although we have a strong correlation among coefficients $c_0-c_3$, and between $f_1$ and $f_2$, cross correlations are quite small ($<0.1$ when diagonals are normalized to unity) between $c_i$ and $f_j$. Cross correlations between $c_i$ and $g_1$, and between $f_i$ and $g_1$ are also small ($<0.25$ and $<0.15$, respectively). The fit is well constrained expect for a high temperature range $\log T >3.82$: the adoption of variables $T$ or $1/T$ in the right hand of equation (2) does not modify the shape of the curve and the quality of the fit, except at the weakly constrained very end of the high temperature edge where we see a change equivalent to $B-V\leq 0.01$. The parameters for the dwarf sample are consistent with those for the full sample, though the former set has larger errors. The consistency of the two fits means that the two samples are well controlled by simply different $\log g$. Hence, we adopt the fit with the full sample as our best result and use this for further analysis in what follows including that for dwarfs. We further proceed with our fitting tests. We fit our samples with adding a cross term [Fe/H] $\times$ (B-V)$_0$. This does not reduces rms residuals at all, but merely increase the error estimate of the parameters (in particular this doubles the error of $f_1$), indicating that the cross term is not very well determined. Actually the cross term thus obtained is rather small, and it changes $\partial T_{\rm eff}/\partial$ [Fe/H] only by 14\% between $(B-V)=0.4$ and 0.8. So we can drop this term. The next is a fit ignoring extinction corrections. This raises the residual temperature from 66 K to 73 K, indicating the necessity of extinction corrections. The significance is demonstrated in Fig. 5, where we plot the residual rms as a function of the selective extinction per unit distance. We see that the minimum is attained with $E(B-V)=0.269$/kpc, and our adopted value 0.235 increases a residual only less than 0.2 K \footnote[1]{Within 50 pc of the solar neighbourhood, the extinction seems to be somewhat smaller than this value. For the stars within 50 pc, we obtain $0.21\pm0.06$ mag/kpc from a sample of 151 stars. This means that extinction increases to $\sim0.35$ mag/kpc at $\approx$ 100pc. We thank Bohdan Paczy\'nski for attracting our attention to this problem}. We obtain an error of 0.030/kpc for this parameter when this is allowed to vary as a free parameter. This gives not only an excellent confirmation of the selective extinction per distance used in the literature (e.g. Blackwell et al. 1990), but also shows that our fit would differentiate such a small changes in the data, indicating an overall consistency of the fitting procedures. The overall quality of our best fit is shown in Fig. 6, where the effective temperature data are corrected for metallicity and surface gravity according to $T_{eff} -( f_1 {\rm [Fe/H]} + f_2 {\rm [Fe/H]}^2 + g_1 \log g )$, and those points plotted are supposed to give data for [Fe/H]=0 and $\log g$=4.3 as a function of $(B-V)_0$. Fig. 7 shows residuals in more detail, where we see that giants and dwarfs are indeed on the same family simply with different $\log g$. This would justify to use the same family of curves for the entire range of our colour space. The data points with squares are stars used for examination of temperature above (we plot only medians when a number of bolometric flux estimates are available). \subsection{Error budget} We have already discussed the source of errors. Our estimate of the size of errors is summarized in Table 4 for F5, G2 and K5 stars. The sources we discussed above all contribute to the dispersion of the final fit. The quadrature of internal error budget amounts to 67$-$77 K, which is consistent with the actual dispersion of the fit 60$-$80 K (global value is 66 K). This means that the error propagation is well controlled in our data processing procedures, and intrinsic scatter of the $(B-V)$ colour temperature relation is substantially smaller than $\approx$40 K. We note that the error of [Fe/H] and that of $T_{\rm eff}$ are comparable and are the dominant source of errors. Photometry error might compete with this for early type stars, where the curve gets steeper. Errors from other entries are smaller. This dispersion of temperature corresponds to $\delta (B-V)\simeq 0.02$ and increases to 0.03 for low temperature stars. Of course, the locus of the relation is better determined. We anticipate a systematic error up to about 30-40 K in the B98 temperature estimate, and the normalisation error of the bolometric flux, which is used as an external calibrator in our work, of the order of 1.5\% (0.37\% in temperature), and there may be a systematic trend of metallicity scale on the order of 0.05 dex depending on authors. This makes overall systematic error to be about 45 K (these systematic errors are significantly smaller than the random errors, and are supposed to be already included in the error budget shown in Table 4). This seems to be the best we can achieve with the present data. \section{Comparison of the $B-V/T_{\rm eff}$ relations in the literature} \subsection{Dwarfs} We discuss the $B-V$ colour temperature relations ($B-V/T_{\rm eff}$ relations) available in the literature, taking the one we obtained above as a reference. Fig. 8 is a compilation of the colour temperature relations for the main sequence stars with solar metallicity. To draw the locus of our colour temperature relation we assume $\log g=129.34 - 64.66\log T + 8.347(\log T)^2$, which is obtained by fitting data of SSFRF for dwarfs used in our analysis. This relation differs from what would be obtained by fitting the data from binary stars (Popper 1980) by an amount of $\Delta \log g\simeq 0.2$ for G stars, but the scatter indicated by the B98 sample (with SSFRF data for $\log g$) and that by the data of Andersen (1991) are somewhat larger than the offset. In any case, the difference caused by the difference of $\log g$ of this amount is small and it changes the resulting temperature by no more than $\approx$ 10 K, or $B-V$ by 0.005. Now, in order to examine the detail, we display in Fig. 9 the difference of various relations against the one obtained by Lejeune, Cuisinier \& Buser (1998; hereafter LCB): $\Delta(B-V)=(B-V)_0 - (B-V)_{\rm 0,LCB}$. The adoption of their relation as a fiducial zero point is motivated by the fact that their relation covers the widest $B-V$ range, while the range of the relation we obtained is not as wide as theirs. In this figure we have plotted 9 relations, which we are going to discuss in detail: Flower (1996), Alonso et al. (1996b), BL98, Code et al. (1976), Demarque et al. (1985; Yale isochrone), and Bertelli et al. (1994), and the relation derived using ATLAS9 (Kurucz 1993) atmosphere, together with LCB and ours. The Yale isochrone and Bertelli et al. are theoretical estimates based on model atmospheres, and we have computed colours for ATLAS9 using the response functions of Azusienis and Straizys (1969) for the $B$ and $V$ pass bands. Table 4 summarizes briefly the methods adopted by the respective authors. We also plot the position of the sun taking ($T_{\rm eff}=5777\pm6$ K and $B-V=0.64\pm0.01$, which we discuss in the next section. It is clear at a glance that LCB and Flower (1996) are largely deviated from the others including our newly obtained $B-V/T_{\rm eff}$ relation. Flower uses temperature information collected from various sources: some from direct measurements of angular diameters, some from IRFMs, and others from spectroscopic analysis. For $\log T_{\rm eff}<3.75$, where a large departure starts, Flower's data mostly rely on temperature from an IRFM analysis of Bell \& Gustafsson (1989). To study a possible problem with Flower's temperature, we show in Fig. 10 the stars he used. In the range of our interest $B-V>0.65$, we find new temperature determinations by Soubiran, Katz \& Cayrel (1998) for 7 stars (indicated by arrows). The new temperature for these stars are significantly lower as we go to redder stars, and they fall on the curve of $B-V/T_{\rm eff}$, we have obtained in this paper. This lends an additional support for our $B-V/T_{\rm eff}$, and at the same time indicates that temperature obtained with the Bell \& Gustafsson atmosphere suffers from errors for stars later than the G5 type ($T_{\rm eff}<5500$ K). Flower's relation reflects this overestimate of temperature for late type stars. LCB adopts Flower's $B-V/T_{\rm eff}$ for $T_{\rm eff}>4250$ K (log $T_{\rm eff}>3.63$), so that it is identical with Flower's for this temperature range. Accordingly, LCB's $B-V/T_{\rm eff}$ relation inherits the same problem as Flower's. BL98 agree with ours for a rather wide range $3.68< \log T_{\rm eff}<3.78$ to within 0.02 mag. Beyond $3.78< \log T_{\rm eff}$ (6000 K) BL98 shows a sudden break and turns away from our curve, giving significantly bluer colour; At log $T_{\rm eff}>3.80$ it is bluer by 0.04 mag than ours. B98 examined his surface brightness against IRFM of BL98 for the main-sequence stars and found that BL98 give angular diameter larger by 4\% at log($T_{\rm eff}$)=3.95, while there is no offset between the two at log($T_{\rm eff}$)=3.72. This 4\% offset is consistent with BL98's $B-V$ bluer by 0.03-0.04 mag than ours. Alonso et al (1996b)'s $B-V/T_{\rm eff}$ relation, which is based on IRFM with ATLAS9 atmosphere supplemented with a calibration against angular diameter measurements, is closely parallel to ours. The agreement between the two is $<0.01$ mag in $B-V$ for 3.75$<$log($T_{\rm eff}$)$<$3.85. The difference gradually increases as temperature decreases, and it becomes 0.03 mag at the end point of Alonso et al., log($T_{\rm eff}$)=3.70. We have retained in Fig. 9 old Code et al. (1976)'s $B-V/T_{\rm eff}$ relation, which is based on a direct angular size measurement employing intensity interferometry. Their curve smoothly matches with ours with the difference is no more than 0.01 mag in the overlapping range 3.76$<$log($T_{\rm eff}$)$<$3.85. Our final assessment concerns the theoretical colour temperature relations used by the Yale isochrone (Demarque et al. 1996) and by Bertelli et al. (1994), and the one derived from ATLAS9. The departure from our empirical $B-V/T_{\rm eff}$ relation is significant, and the difference can be 0.04 mag. The discrepancy is even larger with the $B-V/T_{\rm eff}$ relation derived from the Kurucz (1991, unpublished) atmosphere (Bertelli et al. 1994): at log($T_{\rm eff}$)=3.75, it gives 0.05 mag redder than our relation. It is interesting to note that these theoretical $B-V/T_{\rm eff}$ relations give values in agreement with ours at log($T_{\rm eff}$)=3.65. This implies that the theoretical relations, {\it if they are used to connect K5 stars with G5 stars}, raise systematic errors of 0.04 to 0.05 mag for relative colours of stars between these two types. When translated into $M_V$, this systematic offset amounts to 0.24-0.30 mag for a given $B-V$. We see the same trend with the calculation using ATLAS9 (Kurucz 1993) atmosphere, although it shows larger deviations at both lower and higher temperatures from ours than the curve of Bertelli et al. who used 1991 version of Kurucz' atmosphere. \subsection{Giants} A similar analysis is carried out for giants. We have plotted $\Delta(B-V)=(B-V)_0 - (B-V)_{\rm 0,Flower}$ in Fig. 11, taking this time Flower's (1996) relation that covers the widest range as the zero point. The figure includes classical work by Johnson (1966) and Ridgway et al. (1980), which have been taken as the standard for long, and recent work by LCB and BL98; a synthetic calculation using theoretical atmosphere of Kurucz (1993) is also included. As for Ridgway et al. (1980)'s data points, we assign $B-V$ colours from Hauck \& Mermilliod (1998) for the stars they used. Since the scatter is to large to obtain a sensible fit, however, we instead plot the points of individual stars. We have also indicated the metallicity correction, when [Fe/H] data are available, by arrows using the metallicity gradient given in eq. (2). It is seen in Fig. 11 that Johnson (1966) and LCB are 0.03$-$0.04 mag bluer than our $B-V/T_{\rm eff}$ relation for a $T=4000-4500$ K range, and Flower (1996) are 0.03$-$0.04 redder in the same range. Unfortunately, our formula does not reliably apply to the temperature lower than 4000 K. BL98 give a $B-V/T_{\rm eff}$ relation with slope somewhat steeper than ours, and the disagreement increases to $>$0.3 mag for $T<4300$ K ($\log T<3.63$). Ridgway et al.'s data are too noisy to make an accurate comparison, but it is likely that their colour temperature relation, when transformed to the $B-V$ colour band, giving too blue colours, say by 0.10$-$0.15 mag for K giants (see also Flower 1996). A good agreement ($\Delta(B-V)<0.02$ mag) is seen between our curve and Kurucz (1993) for a range 4200 K (the lowest temperature) and 5100 K. Kurucz (1993) gives redder colours only for giants earlier than G type. \section{Colour of the Sun} $B-V$ colour of the sun has been playing an important role as a normalization point for the stellar evolution models, yet observationally an accurate measurement of solar colours is notoriously difficult. Photometric observations yield 0.63 (Stebbins \& Kron 1957) to 0.69 (T\"ug \& Schmidt-Kaler 1982). The method often adopted by observers is to use observations of other stars, and interpolate and translate them to the sun, which typically leads to 0.633$\pm$0.009 or 0.665$\pm$0.003 (Taylor 1998), or look for ``solar twins'' (de Strobel 1996, for a summary), rather than to work directly with the sun. For this solar analogue method to work properly, it is essential to control the accuracy of temperature and also metallicity of these stars; this is not easy a task, as we have seen in the preceding section. In Fig. 9 the zero point at $T_{\rm eff}= 5777$ K is adjusted to LCB's value $B-V=0.633$. Our compilation shows that all modern determinations of the $B-V/T_{\rm eff}$ relation give colours equal to or about 0.1 mag bluer than this value. In particular, our newly obtained curve gives $B-V=0.627$. The bluest value is given by Alonso et al. (1996b)'s curve, which yields 0.621. On the other hand, synthetic colors from atmosphere models are significantly redder, 0.65 with the colour-temperature relation of the Yale isochrone, and 0.67 with Kurucz' atmosphere (Bertelli et al 1994; Bessell, Castelli and Plez 1998). Many solar analogue analyses in the past gave rather redder colour, such as 0.66 (e.g., Hardrop 1978; Wamstecker 1981). de Strobel (1996) has argued that Hardrop's sample is significantly metal rich, leading to redder colour. This is also true with, e.g., Wamstecker's sample. The stars in his sample have either luminosity lower than the sun or metallicity higher than the sun by +0.15 dex. After careful selection de Strobel concluded that $B-V=0.642\pm 0.004$. Here we re-examine the case with de Strobel (1996)'s analysis in view of our assessment for the temperature estimate. She has given 26 stars on the list of effective-temperature-selected solar analogues. Among those stars 8 stars are given temperature by B98, and BL98's temperature estimate is available for additional two stars (at the solar temperature, BL98 and B98 agree very well). For 6 stars among these 10 stars, de Strobel has given temperature based on spectroscopic studies much higher than BL98 and B98; the difference amounts to 100-160 K. The average of the offset in the two $T_{\rm eff}$ estimates amounts to 63 K with de Strobel's temperature higher. The adjustment of temperature can easily modify de Strobel's estimate of $(B-V)_\odot$ into a bluer value by an amount of 0.02 or so. We have tried to find solar colour by fitting the 8 stars with B98 data to the form $T_{\rm eff}=c_0+c_1(B-V)+f_1[{\rm Fe/H}]$ using the B98 temperature data. In spite of a small sample and a small range, the fit is well constrained, yielding $(B-V)_\odot=0.61$, with metallicity gradient $\partial T_{\rm eff}/ \partial {\rm [Fe/H]}=$ 220 K/dex and temperature gradient of the right order of magnitude, although this temperature-selected sample is clearly too narrow in the temperature range to find the correct gradient (see Fig. 12a). When we replace the B98 temperature with de Strobel's ([Fe/H] data are not replaced), however, we obtain the temperature basically constant at 5820 K and metallicity gradient with the sign opposite to what we have obtained with the B98 temperature (Fig. 12b). This indicates that the spectroscopic temperature does not agree with the bolometry-surface brightness estimate we used in this paper. This uncertainty in temperature estimations tells us a difficulty of an accurate estimate of solar colour from solar analogues: one must know star's temperature to an absolute accuracy of $\pm 50$ K and [Fe/H]$<$ 0.1. The former is the accuracy one can barely achievable with the bolometry-surface brightness estimate as we have seen in this paper. From our $B-V/T_{\rm eff}$ relation we conclude $(B-V)_\odot=0.627\pm0.014\pm0.012$ with the two errors standing for uncertainty of the estimate of the locus of the relation and possible intrinsic dispersion for the sun around the relation. All modern $B-V/T_{\rm eff}$ relations (other than synthetic) documented in Fig. 9 give $(B-V)_\odot$ within this error. We are not able to reconcile our value with a red colour 0.66-0.67, often referred to in the literature. If the sun would really be this red, it is significantly off from normal G2 stars. Our final remark concerns colour of the sun from spectroscopic synthesis that synthesis calculations using the measured solar spectrum tend to give bluer colours 0.61-0.65 (see Fukugita, Ichikawa \& Sekiguchi 1999 for details), and this agrees with the value we have inferred above. \section{Metallicity dependence} \label{sec_feh_dep} We present the metallicity dependence of our $B-V/T_{\rm eff}$ relation in Fig. 13. Three solid curves represent the relations for [Fe/H]=$-$1.5, $-$0.5, 0 and 0.3 for main sequence. We limit our plot to $0.3<B-V<1.0$, for which our metallicity dependence is well constrained by data. We overlay the curves of Alonso et al. (1996b) and the one we computed with Kurucz (1993) atmosphere, where Kurucz' temperature is scaled down by 200 K to make figure ease a comparison in the same figure. With our curve $\partial T_{\rm eff}/\partial {\rm [Fe/H]}=325\pm20$ K/dex, owing to the absence of a [Fe/H]*$(B-V)$ term as discussed in section 2.5. This means for G2 stars $\partial (B-V)/\partial [{\rm Fe/H}]= 0.9$ mag/dex for the portion contributed from the atmosphere. If we adopt the fit where the cross term is taken as a free parameter the partial derivative increases towards bluer side by 50 degree between $B-V=0.8$ and 0.4. Apparently a good gross agreement is seen between Alonso et al. and ours. A more careful examination, however, shows that metallicity gradient of the former increases from 330 K/dex at $B-V=0.8$ to 480 K/dex at 0.4. We have not seen this large change with our fit, even if we allow for the cross term: our analysis shows that it is at most one third this value. The family of curves calculated with Kurucz (1993)'s atmosphere is generally shifted by about 200 K to a lower temperature. With shifting by this amount, the curve derived from Kurucz's atmosphere agrees well with ours for the range $0.5<B-V<0.7$. On the other hand, the metallicity gradient shows a very good agreement with ours: $\partial (B-V)/\partial {\rm [Fe/H]}$ stays between 350$-$370 K/dex in the range $B-V=0.4-0.8$ with the Kurucz atmosphere. The metal dependence is well accounted for with the Kurucz atmosphere for F-K stars. \section{Conclusion} In this paper we have used the modern, least model-dependent determination of effective temperature of stars and the best available data for $B-V$ colours, metallicity and surface gravity in order to derive a $B-V$ colour temperature relation with metallicity and gravity as auxiliary parameters. We have achieved the smallest residual temperature over those available in the literature. The fit we obtained is well constrained and a number of tests assure the quality of data and show a consistency of the data processing procedures. Our relation covers the range from F0 to K5 stars ($T_{\rm eff}=4000-7000$ K) and [Fe/H] from $-$1.5 to +0.3 both for dwarfs and giants. The dispersion of the fit, 66 K, perhaps represents the limit we can achieve with the present quality of data. The most important limiting factors are temperature determinations and metallicity measurements. This means that we can attain accuracy of 0.02 mag in $B-V$ colours for a given temperature for F0-K0 stars, and slightly worse for later type stars. This is still a significant error, but it is not as large as the disagreement recognized among various isochrone works. We have examined various $B-V/T_{\rm eff}$ relations available in the literature. Our relation smoothly joins the Code et al. (1976)'s relation given for high temperature stars, and also shows a close match with Alonso et al. (1996b)'s $B-V/T_{\rm eff}$ relation which is based on the IRFM with additional calibrations. BL98 give a relation that agrees reasonably well with ours, but a significant discrepancy is observed for stars earlier than F8 stars. The relations of Flower (1996) and LCB are largely off from ours, as much as 0.1 mag for low temperature stars. Colour becomes significantly redder for G5 or later type stars, if these $B-V/T_{\rm eff}$ relations are used. Our comparison demonstrates that some calibration of temperatures against those obtained from angular diameter measurements consists an essential element for a high accuracy. We also clarified systematic errors with the colour-temperature relation obtained by synthetic computation using model atmospheres: they deviate significantly from our empirical relation. In particular, the offset changes by 0.04-0.05 in $B-V$ across K0 to K5 stars, which directly induces an error in the slope of the colour-magnitude diagram by this amount. This means that, for instance, if the distance to one open cluster is calibrated with K5 stars and if another is with K0 or some earlier stars, we would be led to an error of 10-15\% in distance. This error also makes intermediate age galaxies, which contains G stars as a major source of the bluer component, appreciably redder in stellar population synthesis model of galaxies. Another implication important for cosmology is that the population synthesis model of Bruzual \& Charlot (1993) (see also Charlot, Worthey \& Bressan 1996) would give too blue colour for early type galaxies in their late stage of evolution. Typically 5 Gyr after the initial burst, G and K giants start dominating the light from elliptical galaxies, and Bruzual \& Charlot take Ridgway et al (1980)'s temperature scale to assign colours to tracks. We have shown that Ridgway et al.'s scale gives typically 0.1 mag bluer at a given temperature for K giants. Therefore, we should make Bruzual \& Charlot's $B-V$ colour prediction redder by this amount. With this revision, the burst model would give $B-V=0.95$ already at 5 Gyr from the burst, rather than 9 Gyr in their original model. This offset also explains the discrepancy, at least in part, between the predictions of Bruzual \& Charlot and of Bertelli et al. (1994), the latter, using the Kurucz atmosphere, giving 0.05 magnitude redder than the former. We have also studied the problem of colour of the sun. Our $B-V/T_{\rm eff}$ relation gives $(B-V)_\odot=0.63\pm0.02$ which agrees with ``long wavelength group'', but disagree with ``short wavelength group'' of solar colour (Taylor 1998). We have emphasized the importance to accurately estimate temperature and metallicity when colour of the sun is inferred from solar analogues. The quality of temperature and metallicity determinations of the presently available sample is probably insufficient to determine colour of the sun within an error of 0.02 mag. \acknowledgements We are grateful to Drs. R. L. Kurucz, J.-C. Mermilliod and S. Yi for kindly providing us with their data in machine readable form. One of us (MF) thanks the Raymond and Beverly Sackler Fellowship and the Alfred P. Sloan Foundation for the support for the work in Princeton. \newpage \include{Sekiguchi.tab1} \newpage \include{Sekiguchi.tab2} \newpage \include{Sekiguchi.tab3} \newpage \include{Sekiguchi.tab4} \newpage \include{Sekiguchi.tab5} \newpage
\section{Introduction} The relation between the jets and the accretion processes in the central 'engine' is a crucial ingredient in our understanding of the physics of active galactic nuclei (AGN). In some theoretical models of the formation of the jet, the power is generated through accretion and then extracted from the disc/black hole rotational energy and converted into the kinetic power of the jet (Blandford \& Znajek; Blandford \& Payne 1982). Recently, the concept of jet-disc symbiosis was introduced and the inhomogeneous jet model plus mass and energy conservation in the jet-disc system was applied to study the relation between disc and jet luminosities (Falcke \& Biermann 1995; Falcke, Malkan \& Biermann 1995; Falcke \& Biermann 1998). An effective approach to study the link between these two phenomena is to explore the relationship between luminosity in line emission and kinetic power of jets in different scales (Rawlings \& Saunders 1991; Celotti, Padovani \& Ghisellini 1997; Cao \& Jiang 1998). Rawlings \& Saunders (1991) derived the total jet kinetic power $Q_{jet}$ and found a correlation between $Q_{jet}$ and the narrow line luminosity $L_{NLR}$. There are some indications that the narrow-line emission could be partly caused by the power supplied by the jet (Lacy \& Rawlings 1994; Capetti et al. 1996). Celotti, Padovani \& Ghisellini (1997) considered the luminosity in broad emission lines and explored its relation with kinetic power for a sample of radio-loud objects. They also found evidence for a link between jets and discs. The kinetic power of jets can be estimated by using radio data on very long-baseline interferometry (VLBI) scales and the standard synchrotron self-Compton theory (Celotti \& Fabian 1993), or in a similar way based on K\"onigl's inhomogeneous jet model (Cao \& Jiang 1998). The kinetic power of the jet can be estimated in this way only for a small fraction of quasars due to the lack of the necessary observational data for many sources. Serjeant et al. (1998) found a significant correlation between radio and optical emission for a sample of steep-spectrum quasars, which presents direct evidence for a close link between accretion on to black holes and the fuelling of relativistic jets. Their sample is limited to steep-spectrum quasars to reduce the effects of relativistic beaming for optical continuum. The luminosity in the broad-line region can be taken as an indicator of accretion power of the source (Celotti, Padovani \& Ghisellini 1997), and the radio luminosity is believed to be a straightforward indicator of jet power (Serjeant et al. 1998). In this paper we present a correlation between radio and broad-line emission for a sample of radio-loud quasars. The sample and the estimate of total broad-line flux are described in Sect. 2. Section 3 contains the results. The last section is devoted to discussion. The cosmological parameters $H_{0}=50$ km s$^{-1}$ Mpc$^{-1}$ and $q_{0}=$0.5 have been adopted in this work. \section{The sample} The 1 Jy catalogue is an all-sky survey covering 9.81 sr with a flux density limit of $S_{\rm 5 GHz}\ge 1$ Jy lying outside the galactic plane ($|b|\ge 10^{\circ}$) (K\"uhr et al. 1981b). Optical counterparts have been found for 97 \% of the radio sources in 1 Jy catalogue (Stickel, Meisenheimer \& K\"uhr 1994). The S4 survey covers the region between $35^{\circ} \le \delta \le 70^{\circ}$ with $S_{\rm 5 GHz}\ge 0.5$ Jy (Pauliny-Toth et al. 1978), while the S5 survey contains the sources with $S_{\rm 5 GHz}\ge 0.25$ Jy in the region $\delta \ge 70^{\circ}$ (K\"uhr et al. 1981a). For the S4 catalogue, about 90 \% of the radio sources have known optical counterparts (Stickel \& K\"uhr 1994), while about 75 \% of the sources have optical counterparts in S5 catalogue (Stickel \& K\"uhr 1996a). In this work, we only consider the quasars and BL Lac objects in the 1 Jy, S4 and S5 radio source catalogues (all sources identified as galaxies have not been considered here). Combining these three catalogues, we find 378 sources with available redshifts including 358 quasars and 20 BL Lac objects (only those identified as BL/QSO). Complete information on the line spectra is available for very few sources in our sample, since different lines are observed for the sources at different redshifts. We have to estimate the total broad-line flux from the available observational data. There is not a solidly established procedure to derive the total broad-line flux and we therefore adopt the method proposed by Celotti, Padovani \& Ghisellini (1997). The following lines: Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$, and H$\alpha$, which contribute the major parts in the total broad-line emissions, are used in our estimate. We use the line ratios reported by Francis et al. (1991) and add the contribution from line H$\alpha$ to derive the total broad-line flux (see Celotti, Padovani \& Ghisellini 1997 for details). We then search the literature to collect data on broad-line fluxes. We only consider values of line fluxes (or luminosities) given directly or the equivalent width and the continuum flux at the corresponding line frequency which are reported together in the literatures (slightly different from Celotti, Padovani \& Ghisellini (1997)). When more than one value of the same line flux was found in the literature, we take the most recent reference. This finally leads to a sample of 198 sources, consisting of 184 quasars and 14 BL Lac objects. Our sample covers more than 50 \% of the 378 sources in the starting samples. The radio data in the sample are taken from some related references (Stickel \& K\"uhr 1996a, 1996b; Stickel \& K\"uhr 1994; Stickel, Meisenheimer \& K\"uhr 1994). The radio data and broad-line fluxes are listed in Table 1. \begin{figure} \centerline{\psfig{figure=mz003f1.ps,width=10.0cm,height=10.0cm}} \caption{The radio luminosity $\nu L_{\nu{5G}}$, redshift $z$ plane for the sample. The dashed line denotes the low flux density limit. The open circles represent the quasars with $\alpha_{11-6}> -0.7$, and the squares represent the quasars with $\alpha_{11-6}\le -0.7$, while the full circles represent BL Lac objects. } \end{figure} \begin{figure} \centerline{\psfig{figure=mz003f2.ps,width=10.0cm,height=10.0cm}} \caption{The total broad-line luminosity $L_{\rm line}$, redshift $z$ plane for the sample. (Symbols as in Fig. 1. } \end{figure} \begin{figure} \centerline{\psfig{figure=mz003f3.ps,width=10.0cm,height=10.0cm}} \caption{The radio and broad-line flux relation (symbols as in Fig. 1.).} \end{figure} \begin{figure} \centerline{\psfig{figure=mz003f4.ps,width=10.0cm,height=10.0cm}} \caption{The radio and broad-line luminosity relation (symbols as in Fig. 1.).} \end{figure} \newpage \begin{table} \begin{minipage}{150mm} \caption{Radio and BLR data of the sample.} \begin{tabular}{ccccllrr}\hline Source & Class. & z & log f$_{BLR}$ & Lines & Refs. & f$_{5{\rm G}}$ & $\alpha_{11-6}$ \\ (1) & (2) & (3) & (4) & (5) &(6) & (7) & (8)\\ \hline 0003$-$066& Q &0.347 & -13.46& H$\alpha$ & S89& 1.48 & 0.02\\ 0014+813& Q &3.384 & -12.37& Ly$\alpha$, C\,{\sc iv} &O94 & 0.55 & -0.16\\ 0016+731& Q &1.781 & -13.36& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}& L96& 1.65 & 0.16\\ 0022$-$297& Q &0.406 & -12.62& Mg\,{\sc ii}, H$\beta$, H$\alpha$ & S93b& 1.03 & -0.74\\ 0024+348& Q &0.333 & -13.76& H$\beta$, H$\alpha$ & SK93b & 0.73 & -0.30\\ 0035+413& Q &1.353 & -13.40& Mg\,{\sc ii} & SK93b & 0.64 & 0.75\\ 0056$-$001& Q &0.717 & -12.50& Mg\,{\sc ii}& B89& 1.40 & -0.39\\ 0106+013& Q &2.107 & -12.36& Ly$\alpha$, C\,{\sc iv}& B89& 2.28 & -0.08\\ 0112$-$017& Q &1.381 & -12.80& C\,{\sc iv}, Mg\,{\sc ii}&B89 & 1.20 & -0.01\\ 0119+041& Q &0.637 & -12.61& Mg\,{\sc ii}, H$\gamma$, H$\beta$&JB91,RS80 & 1.67 & 0.04\\ 0119+115& Q &0.570 & -13.47& Mg\,{\sc ii} &S89 & 1.01 & 0.33\\ 0133+207& Q &0.425 & -11.79& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$ &W95,JB91 & 1.08 &-1.11\\ 0133+476& Q &0.859 & -13.09& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &L96 & 3.22 & 0.62\\ 0134+329& Q &0.367 & -11.71& Ly$\alpha$, H$\gamma$, H$\beta$, H$\alpha$ &JB91,K85 & 5.28 & -0.86\\ 0135$-$247& Q &0.831 & -12.16& H$\gamma$, H$\beta$ & JB91& 1.70 & 0.35\\ 0153+744& Q &2.338 & -12.48& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}& L96& 1.52 & -0.31\\ 0159$-$117& Q &0.669 & -11.88& Mg\,{\sc ii}, H$\beta$ & O84& 1.39 & -0.59\\ 0210+860& Q &0.186 & -13.39& H$\gamma$, H$\beta$, H$\alpha$ & L96& 1.72 &-1.13\\ 0212+735& Q &2.367 & -13.69& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}&L96 & 2.20 & -0.12\\ 0229+131& Q &2.065 & -12.49& Ly$\alpha$, C\,{\sc iv} &O94 & 1.00 & -0.33\\ 0234+285& Q &1.210 & -12.66& C\,{\sc iv} & W84& 1.44 & -0.24\\ 0235+164& BL& 0.940 & -13.79& Mg\,{\sc ii} & S93a & 2.85 & 1.03\\ 0237$-$233& Q &2.224 & -11.78& H$\beta$, H$\alpha$ & B94& 3.40 & -0.68\\ 0248+430& Q &1.316 & -12.74& Mg\,{\sc ii} & S93b& 1.20 & 0.37\\ 0256+075& Q &0.893 & -14.08& Mg\,{\sc ii} & S89& 1.04 & 0.45\\ 0336$-$019& Q &0.852 & -12.55 & Mg\,{\sc ii}, H$\gamma$, H$\beta$ &B89,JB91 & 2.86 & 0.30\\ 0400+258& Q &2.109 & -13.31& C\,{\sc iv} &W84 & 1.79 & 0.11\\ 0403$-$132& Q &0.571 & -11.86& C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$& JB91,M96,O84& 2.88 & -0.05\\ 0405$-$123& Q &0.574 & -11.22& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\beta$, H$\alpha$ &O94,O84,M96,N79 & 1.96 & -0.30\\ 0414$-$189& Q &1.536 & -13.62& C\,{\sc iv}, Mg\,{\sc ii} &H78 & 1.35 & 0.22\\ 0420$-$014& Q &0.915 & -12.70& Mg\,{\sc ii} &B89 & 1.46 & 0.01\\ 0437+785& Q &0.454 & -12.35& H$\gamma$, H$\beta$ & S93b & 0.26 & -0.72\\ 0440$-$003& Q &0.844 & -13.00& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &B89,JB91 & 2.61 & -0.29\\ 0440+732& Q &1.290 & -13.08& Mg\,{\sc ii} & S93b & 0.30 & -0.75\\ 0444+634& Q &0.781 & -13.54& Mg\,{\sc ii}, H$\beta$ & SK93b & 0.50 & -0.04\\ 0451$-$282& Q &2.559 & -12.44& Ly$\alpha$, C\,{\sc iv} & F83& 2.26 & -0.04\\ 0454$-$810& Q &0.444 & -13.24& Mg\,{\sc ii}, H$\beta$, H$\alpha$ & S89& 1.40 & 0.29\\ 0454$-$463& Q &0.858 & -12.18& Mg\,{\sc ii} & F83 & 2.32 & -0.03 \\ 0454$-$234& Q &1.003 & -13.29& Mg\,{\sc ii} & S89& 2.06 & 0.16\\ 0457+024& Q &2.384 & -12.83& Ly$\alpha$, C\,{\sc iv} & B89& 1.15 & -0.64\\ 0458$-$020& Q &2.286 & -13.28& Ly$\alpha$, C\,{\sc iv} & B89& 1.74 & -0.08\\ 0511$-$220& Q &1.296 & -12.94& C\,{\sc iv}, Mg\,{\sc ii} &S89 & 1.31 & 0.13\\ 0514$-$459& Q &0.194 & -12.45& H$\beta$, H$\alpha$ & S93b& 1.06 & -0.32\\ 0518+165& Q &0.759 & -12.94& H$\gamma$, H$\beta$ & JB91& 4.11 & -0.59\\ 0528$-$250 & Q& 2.765 & -15.03 & Ly$\alpha$ & WM96 & 1.16 & -0.20 \\ 0537$-$441 &BL & 0.896 & -12.55& Mg\,{\sc ii} & S93a & 4.00 & 0.06\\ 0537$-$286 & Q& 3.119 & -13.54 & Ly$\alpha$, C\,{\sc iv} & O94 & 1.02 & 0.52\\ 0538+498& Q &0.545 & -12.53& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &L96 & 8.30 & -0.72\\ 0539$-$057& Q &0.839 & -13.42& Mg\,{\sc ii} & SK93b& 1.55 & 1.41\\ 0602$-$319& Q &0.452 & -12.39& H$\beta$ &R84 & 1.25 & -0.67\\ 0605$-$085& Q &0.872 & -12.95& Mg\,{\sc ii} & S93b& 3.49 & 0.09\\ 0607$-$157& Q &0.324 & -12.98& H$\beta$, H$\alpha$ &H78 & 1.82 & -0.01\\ 0636+680& Q &3.180 & -12.23& Ly$\alpha$, C\,{\sc iv} & O94 & 0.54 & 0.87\\ 0637$-$752& Q &0.654 & -12.01& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$ & W98,H78,T93& 5.85 & 0.19\\ 0642+449& Q &3.396 & -12.91& Ly$\alpha$, C\,{\sc iv} & O94 & 0.78 & -0.79\\ 0646+600& Q &0.455 & -13.57& H$\gamma$, H$\beta$ & SK93b & 0.79 & 0.23\\ 0707+689& Q &1.139 & -13.71& Mg\,{\sc ii} & SK93b & 0.75 & -0.67\\ 0711+356& Q &1.620 & -12.43& C\,{\sc iv}, Mg\,{\sc ii} &L96 & 1.52 & -0.29\\ 0723+679& Q &0.846 & -12.74& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &L96 & 1.32 & -0.34\\ 0736+017& Q &0.191 & -11.80& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$, H$\alpha$ & JB91,O94& 1.99 & -0.13\\ 0738+313& Q &0.631 & -11.45& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &W85,JB91 & 2.49 & 0.27\\ \end{tabular} \end{minipage} \end{table} \begin{table} \begin{minipage}{150mm} \contcaption{Radio and BLR data of the sample.} \begin{tabular}{ccccllrr}\hline 0740+380& Q &1.063 & -12.07& Ly$\alpha$, C\,{\sc iv} & W95 & 0.27 &-1.24\\ 0740+828& Q &1.041 & -12.96& Mg\,{\sc ii} & S93b & 0.93 & -0.53\\ 0743$-$673& Q &1.511 & -12.30& C\,{\sc iv}, Mg\,{\sc ii}, H$\alpha$ & E89,d94 & 1.79 & -0.69\\ 0743$-$006& Q &0.994 & -12.83& Mg\,{\sc ii} & S89& 1.99 & 0.57\\ 0743+744& Q &1.629 & -13.18& Mg\,{\sc ii} & S93b & 0.33 & -0.05\\ 0804+499& Q &1.433 & -12.71& C\,{\sc iv}, Mg\,{\sc ii} &L96 & 2.05 & 0.47\\ 0809+483& Q &0.871 & -12.34& Mg\,{\sc ii}, H$\gamma$, H$\beta$ & L96& 4.42 & -0.89\\ 0814+425& BL &0.258 & -14.50& Mg\,{\sc ii} & L96& 1.69 & -0.10\\ 0820+225& BL &0.951 & -14.58& Mg\,{\sc ii} & S93a& 1.60 & -0.20\\ 0823+033& BL &0.506 & -13.60& Mg\,{\sc ii} & S93a& 1.32 & 0.77\\ 0825$-$202& Q &0.822 & -12.40& Mg\,{\sc ii} & S93b& 1.18 & -0.93\\ 0834$-$201& Q &2.752 & -12.89& Ly$\alpha$, C\,{\sc iv} & F83& 3.72 & 0.03\\ 0836+710& Q &2.172 & -12.12& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii} & L96& 2.59 & -0.32\\ 0838+133& Q &0.684 & -12.18& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$ &W95,JB91,M96 & 1.39 & -0.39\\ 0842$-$754& Q &0.524 & -11.84& H$\beta$ & T93& 1.42 & -0.67\\ 0850+581& Q &1.322 & -12.36& Mg\,{\sc ii} &L96 & 1.39 & 0.75\\ 0851+202& BL &0.306 & -12.88& Mg\,{\sc ii}, H$\beta$, H$\alpha$& S89,S93a& 2.62 & 0.11\\ 0858$-$279& Q &2.152 & -12.83 & Ly$\alpha$, C\,{\sc iv} & d94& 1.42 & -0.55\\ 0859$-$140& Q &1.339 & -12.29& H$\alpha$ & E89& 2.30 & -0.42\\ 0859+470& Q &1.462 & -12.86& C\,{\sc iv}, Mg\,{\sc ii} &L96 & 1.78 & -0.15\\ 0906+015& Q &1.018 & -12.62& Mg\,{\sc ii} & B89& 1.04 & 0.04\\ 0906+430& Q &0.668 & -13.95& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &L96 & 1.80 & -0.38\\ 0917+624& Q &1.446 & -13.06& Mg\,{\sc ii} & SK93b& 1.00 & 0.08\\ 0923+392& Q &0.698 & -11.55& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$ & W95,L96& 8.73 & 1.03\\ 0945+408& Q &1.252 & -12.37& C\,{\sc iv}, Mg\,{\sc ii} &L96 & 1.39 & 0.11\\ 0953+254& Q &0.712 & -12.38& H$\gamma$, H$\beta$ & JB91& 1.82 & 0.83\\ 0954+556& Q &0.901 & -12.63& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$ & W95,L96& 2.28 & -0.19\\ 0954+658& BL &0.367 & -14.04& H$\alpha$ & L96& 1.46 & 0.35\\ 1007+417& Q &0.6123 & -11.48& C\,{\sc iv}, H$\gamma$, H$\beta$ & JB91,M96& 0.71 & -0.63\\ 1007+716& Q &1.192 & -12.30& Mg\,{\sc ii} & S93b & 0.59 & -0.95\\ 1017$-$426& Q &1.280 & -12.80& Mg\,{\sc ii} & S93b& 1.27 & -0.98\\ 1034$-$293& Q &0.312 & -13.25& H$\beta$, H$\alpha$ &S89 & 1.51 & 0.35\\ 1040+123& Q &1.029 & -12.64& Mg\,{\sc ii} & N79& 1.47 & -0.61\\ 1045$-$188& Q &0.595 & -13.77& H$\beta$ & S93b& 1.14 & 0.32\\ 1055+018& Q &0.892 & -13.06& Mg\,{\sc ii} & B89& 3.47 & 0.32\\ 1100+772& Q &0.3115 & -11.67& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$, H$\alpha$ & JB91,W98,M96& 0.77 & -0.88\\ 1111+408& Q &0.734 & -11.82& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$ & JB91,W95 & 0.79 & -0.98\\ 1127$-$145 & Q& 1.187 & -12.13& Ly$\alpha$ & W95 & 6.57 & 0.03\\ 1136$-$135& Q &0.554 & -11.89& Mg\,{\sc ii}, H$\beta$ &O84,T93 & 2.11 & -0.42\\ 1137+660& Q &0.6563 & -11.42& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$ & W95,M96,JB91& 1.06 & -0.82\\ 1144$-$379& BL &1.048 & -13.39& Mg\,{\sc ii} &S89 & 1.61 & 0.66\\ 1148$-$001& Q &1.982 & -12.04& Ly$\alpha$, C\,{\sc iv} &B89 & 1.90 & -0.44\\ 1150+497& Q &0.334 & -12.18& H$\beta$ & SM87& 1.12 & -0.48\\ 1151$-$348& Q &0.258 & -12.68& H$\beta$ &T93 & 2.83 & -0.65\\ 1226+023& Q &0.158 & -10.27& Ly$\alpha$, C\,{\sc iv}, H$\beta$, H$\alpha$ &O94,JB91,M96 &42.85 & 0.15\\ 1229$-$021& Q &1.045 & -12.08& Mg\,{\sc ii} & B89& 1.07 & -0.29\\ 1237$-$101& Q &0.753 & -12.44& Mg\,{\sc ii} &S93b & 1.31 & -0.25\\ 1245$-$197 &Q&1.275 &-14.51 &Mg\,{\sc ii} &d94 & 2.50 &-0.70 \\ 1250+568& Q &0.321 & -11.96& C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$, H$\alpha$ & JB91,W95& 1.06 & -0.51\\ 1253$-$055& Q &0.536 & -12.42& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\beta$, H$\alpha$ & W95,M96,N79&14.95 & 0.30\\ 1258+404& Q &1.6656 & -12.47& C\,{\sc iv}, Mg\,{\sc ii} & C91 & 0.36 & -0.88\\ 1302$-$102& Q &0.286 & -11.51& Ly$\alpha$, C\,{\sc iv}, H$\beta$ &O94,M96 & 1.17 & 0.17\\ 1308+326& BL &0.997 & -12.60& Ly$\alpha$, Mg\,{\sc ii} & S93a,O94& 1.53 & 0.20\\ 1328+307& Q &0.846 & -13.22& H$\gamma$, H$\beta$ &GW94 & 7.40 & -0.53\\ 1334$-$127& Q &0.539 & -12.88& Mg\,{\sc ii} & S93b& 2.24 & 0.17\\ 1340+606& Q &0.961 & -12.34& Ly$\alpha$, C\,{\sc iv} & W95 & 0.40 &-1.14\\ 1354+195& Q &0.720 & -11.44& Ly$\alpha$, C\,{\sc iv} & W95&1.56 & -0.07\\ 1355$-$416& Q &0.313 & -11.24& Ly$\alpha$, H$\beta$ &T93,O94 & 1.44 & -0.88\\ 1416$-$067& Q &1.439 & -12.26& H$\alpha$ & E89& 1.53 & -0.96\\ 1424$-$418& Q &1.522 & -13.20& Mg\,{\sc ii} & S89& 3.13 & 0.28\\ 1442+101& Q &3.5305 & -13.13& C\,{\sc iv} & C91& 1.22 & -0.61\\ 1451$-$375& Q &0.314 & -11.34& Ly$\alpha$ & K85& 1.89 & 0.37\\ 1452+502& Q &2.849 & -13.43& Ly$\alpha$, C\,{\sc iv} & SK93a & 0.54 & -0.78\\ 1458+718& Q &0.905 & -12.14& Mg\,{\sc ii}, H$\gamma$, H$\beta$ & L96& 3.39 & -0.71\\ 1504$-$166& Q &0.876 & -12.64& Mg\,{\sc ii} & H78& 1.98 & -0.16\\ \end{tabular} \end{minipage} \end{table} \begin{table} \begin{minipage}{150mm} \contcaption{Radio and BLR data of the sample.} \begin{tabular}{ccccllrr}\hline 1510$-$089& Q &0.361 & -12.00& Ly$\alpha$, Mg\,{\sc ii}, H$\gamma$, H$\beta$, H$\alpha$ & N79,T93,O94,BK84 & 3.08 & 0.31\\ 1512+370& Q &0.371 & -12.21& Ly$\alpha$, C\,{\sc iv}, H$\beta$ & M96,W98& 0.41 & -0.45\\ 1518+046& Q &1.296 & -14.17& Mg\,{\sc ii} & d94 & 1.06 & -1.30\\ 1531+722& Q &0.899 & -12.78& Mg\,{\sc ii} & S93b & 0.46 & -0.04\\ 1532+016& Q &1.435 & -13.26& C\,{\sc iv}, Mg\,{\sc ii} &B89 & 1.14 & 0.06\\ 1538+149& BL &0.605 & -14.07& Mg\,{\sc ii} & S93a& 1.96 & 0.34\\ 1546+027& Q &0.412 & -12.10& Mg\,{\sc ii} & B89& 1.45 & 0.46\\ 1555+001& Q &1.770 & -13.37& Ly$\alpha$, C\,{\sc iv} & B81 & 2.24 & 0.34\\ 1611+343& Q &1.401 & -12.17& Ly$\alpha$, C\,{\sc iv} & W95& 2.67 & 0.10\\ 1622$-$253 &Q& 0.786 & -13.80 & Mg\,{\sc ii}, H$\beta$ &d94 &2.08 &-0.14 \\ 1624+416& Q &2.550 & -14.75& Ly$\alpha$, C\,{\sc iv} & L96& 1.32 & -0.26\\ 1633+382& Q &1.814 & -12.52& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii} &L96 & 4.02 & 0.73\\ 1634+628& Q &0.988 & -13.34& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &L96 & 1.52 & -0.92\\ 1637+574& Q &0.750 & -11.84& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$ & W95,L96& 1.42 & 0.35\\ 1638+398& Q &1.666 & -13.68& C\,{\sc iv}, Mg\,{\sc ii} &S89 & 1.15 & 0.17\\ 1641+399& Q &0.594 & -11.69& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$ & L96,W95&10.81 & 0.54\\ 1642+690& Q &0.751 & -13.56& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &L96 & 1.39 & -0.10\\ 1704+608& Q &0.371 & -11.79& Ly$\alpha$, C\,{\sc iv}, H$\beta$, H$\alpha$ &N79,W98,M96 & 1.23 & -0.81\\ 1721+343& Q &0.206 & -11.52& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$, H$\alpha$ & W98,R84,S81 & 0.93 & -0.43\\ 1725+044& Q &0.293 & -12.35& H$\beta$ & R84& 1.24 & 0.76\\ 1732+389& Q &0.976 & -13.95& Mg\,{\sc ii} & S89& 1.13 & 0.85\\ 1739+522& Q &1.379 & -12.90& C\,{\sc iv}, Mg\,{\sc ii} & L96& 1.98 & 0.68\\ 1741$-$038& Q &1.057 & -13.27& Mg\,{\sc ii} & S89& 3.68 & 0.75\\ 1745+624& Q &3.886 & -13.23& Ly$\alpha$, C\,{\sc iv} & SK93a & 0.58 & -0.30\\ 1803+784& BL &0.684 & -12.75& Mg\,{\sc ii}, H$\beta$ & L96& 2.62 & 0.25\\ 1823+568& BL &0.664 & -13.96& Mg\,{\sc ii}, H$\beta$ & L96& 1.66 & 0.22\\ 1828+487& Q &0.691 & -12.07& Mg\,{\sc ii}, H$\gamma$, H$\beta$ &L96 & 6.19 & -0.75\\ 1830+285& Q &0.594 & -11.76& Mg\,{\sc ii} & RS80& 1.06 & -0.30\\ 1849+670& Q &0.657 & -12.83& Mg\,{\sc ii}, H$\beta$ & SK93a & 0.39 & -0.76\\ 1928+738& Q &0.302 & -11.29& C\,{\sc iv}, H$\gamma$, H$\beta$, H$\alpha$ &M96,L96 & 3.34 & -0.01\\ 1936+714& Q &1.864 & -13.11& C\,{\sc iv} & S93b & 0.34 & -0.35\\ 1945+725& Q &0.303 & -12.99& H$\gamma$, H$\beta$, H$\alpha$ & SK93a & 0.28 & -0.81\\ 1954$-$388& Q &0.626 & -13.02& H$\beta$ & T93& 2.06 & 0.43\\ 1954+513& Q &1.230 & -12.55& Mg\,{\sc ii}, H$\gamma$ &L96 & 1.43 & -0.01\\ 1958$-$179& Q &0.650 & -13.09& Mg\,{\sc ii} & O84& 1.20 & 0.13\\ 2000$-$330& Q &3.777 & -12.44& Ly$\alpha$, C\,{\sc iv} & O94& 1.15 & 0.78\\ 2015+657& Q &2.845 & -13.44& Ly$\alpha$, C\,{\sc iv} & SK93b & 0.51 & -0.70\\ 2029+121& BL &1.215 & -14.04& Mg\,{\sc ii} & SK93a& 1.33 & 0.74\\ 2044$-$027& Q &0.942 & -12.57& Mg\,{\sc ii} & SS80& 1.02 & -0.52\\ 2111+801& Q &0.524 & -13.37& Mg\,{\sc ii}, H$\gamma$, H$\beta$ & SK93a & 0.26 & -0.37\\ 2113+293& Q &1.514 & -13.36& Mg\,{\sc ii} & S93b & 1.47 & 0.62\\ 2126$-$158& Q &3.266 & -12.25& Ly$\alpha$, C\,{\sc iv} & O94 & 1.28 & 0.14\\ 2128$-$123& Q &0.501 & -11.57& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii},H$\beta$ &T93,O84,O94 & 2.07 & 0.11\\ 2134+004& Q &1.936 & -12.13& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii} &B89,O94 &12.30 & 0.83\\ 2135$-$147& Q &0.200 & -11.34& Ly$\alpha$, C\,{\sc iv}, H$\gamma$, H$\beta$, H$\alpha$ & JB91,O94,M96& 1.41 & -0.70\\ 2136+141& Q &2.427 & -12.65& Ly$\alpha$, C\,{\sc iv} &B81 & 1.11 & -0.10\\ 2145+067& Q &0.990 & -11.93& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii} & N79,O94& 4.00 & 0.25\\ 2155$-$152& Q &0.672 & -13.59& Mg\,{\sc ii}, H$\beta$ & S89& 1.77 & 0.15\\ 2201+315& Q &0.298 & -11.00& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$, H$\alpha$ &JB91,W95 & 2.32 & 0.24\\ 2203$-$188& Q &0.619 & -13.05& Mg\,{\sc ii} & S89& 4.38 & -0.28\\ 2209+080& Q &0.484 & -12.53& Mg\,{\sc ii}, H$\beta$ & RS80& 1.08 & -0.39\\ 2216$-$038& Q &0.901 & -11.82& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii} &B89,W95 & 1.50 & 0.48\\ 2218+395& Q &0.655 & -12.98& Mg\,{\sc ii}, H$\beta$ & SK93b & 0.69 & -0.60\\ 2223$-$052& Q &1.404 & -12.46& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii} &P89,W95 & 4.51 & -0.11\\ 2223+210& Q &1.949 & -11.92& Ly$\alpha$, C\,{\sc iv} & RS80,W84& 1.24 & -0.65\\ 2230+114& Q &1.037 & -11.87& Ly$\alpha$, C\,{\sc iv} & W95& 3.61 & -0.50\\ 2234+282& Q &0.795 & -12.95& Mg\,{\sc ii}, H$\gamma$, H$\beta$ & RS80,JB91& 1.06 & 0.08\\ 2240$-$260& BL &0.774 & -13.92& Mg\,{\sc ii} & S93a& 1.03 & -0.08\\ 2243$-$123& Q &0.630 & -11.94& Mg\,{\sc ii}, H$\beta$ & O84,T93& 2.45 & -0.18\\ 2247+140& Q &0.237 & -12.39& H$\gamma$, H$\beta$, H$\alpha$ & GW94& 1.02 & -0.63\\ 2251+158& Q &0.859 & -11.88& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$ & N79,W95,JB91&17.42 & 0.64\\ 2255+416& Q &1.149 & -13.81& Mg\,{\sc ii} & SK93a& 1.00 & -0.55\\ 2311$-$452& Q &2.884 & -13.79& Ly$\alpha$, C\,{\sc iv} & S89& 1.47 & -0.38\\ 2311+469& Q &0.741 & -12.46& Mg\,{\sc ii}, H$\gamma$, H$\beta$ & SK93b & 0.73 & -0.69\\ \end{tabular} \end{minipage} \end{table} \begin{table} \begin{minipage}{150mm} \contcaption{Data of the sample.} \begin{tabular}{ccccllrr}\hline 2318+049& Q &0.623 & -12.51& Mg\,{\sc ii} & RS80& 1.13 & -0.11\\ 2319+272& Q &1.253 & -13.07& Mg\,{\sc ii} & S93b& 1.07 & 0.20\\ 2326$-$477& Q &1.306 & -12.14& H$\alpha$ & E89& 2.33 & -0.06\\ 2328+107& Q &1.489 & -12.92& C\,{\sc iv} & W84 & 1.01 & -0.05\\ 2342+821& Q &0.735 & -13.99& Mg\,{\sc ii}, H$\gamma$, H$\beta$ & L96 & 1.33 & -0.89\\ 2344+092& Q &0.673 & -11.80& Ly$\alpha$, C\,{\sc iv}, Mg\,{\sc ii}, H$\gamma$, H$\beta$ & B89,O94,JB91 & 1.43 & -0.09\\ 2345$-$167& Q &0.576 & -12.75& H$\gamma$, H$\beta$ & JB91 & 3.66 & 0.51\\ 2351+456& Q &1.992 & -13.58& C\,{\sc iv}, Mg\,{\sc ii} & L96 & 1.42 & -0.05\\ \hline \end{tabular} \end{minipage} \medskip Notes for the table 1. Q: quasars; BL: BL Lac objects.\\ Column (1): IAU source name. Column (2): classification of the source. Column (3): redshift. Column (4): estimated total broad-line flux (erg s$^{-1}$ cm$^{-2}$). Column (5): lines from which the total $f_{line}$ has been estimated. Column (6): references for the line fluxes. Column (7): radio flux density at 5 GHz. Column (8): two-point spectral index between 6$-$11 cm.\\ \vskip 1mm References: B81: Baldwin et al. (1981). B89: Baldwin et al. (1989). B94: Baker et al. (1994). BK84: Bergeron \& Kunth (1984). C91: Corbin (1991). CF94: Corbin \& Francis (1994). d94: di Serego Alighieri et al. (1994). E89: Espey et al. (1989). F83: Fricke et al. (1983). GW94: Gelderman \& Whittle (1994). H78: Hunstead et al. (1978). JB91: Jackson \& Browne (1991). K85: Kinney et al. (1985). L96: Lawrence et al. (1996). M96: Marziani et al. (1996). N79: Neugebauer et al. (1979). O94: Osmer et al. (1994). O84: Oke et al. (1984). P89: Perez et al. (1989). R84: Rudy (1984). RS80: Richstone \& Schmidt 1980. S81: Steiner (1981). S89: Stickel et al. (1989). S93a: Stickel et al. (1993a). S93b: Stickel et al. (1993b). SJ85: Stiko \& Junkkarinen (1985). SK93a: Stickel \& K\"uhr (1993a). SK93b: Stickel \& K\"uhr (1993b). SM87: Stockton \& MacKenty (1987). SS80: Smith \& Spinrad (1980). T93: Tadhunter et al. (1993). V95: Vermeulen et al. (1995). W84: Wampler et al. (1984). W85: Wills et al. (1985). W95: Wills et al. (1995). W98: Wang et al. (1998). WM96: Warren \& Moller (1996). \end{table} \section{Results} We present the radio luminosity at 5 GHz and the total broad-line luminosity as functions of redshift $z$ for our sample in Figs. 1 and 2 respectively. The radio luminosity has been K-corrected, and the low flux limit is indicated by the dashed line in Fig. 1 assuming the spectral index $\alpha=$ 0. We note that our sample provide wide dispersion in both the radio and broad-line luminosities at any redshift $z>$0.3. Figure 3 shows the correlation between radio flux $\nu f_{\nu 5G}$ at 5 GHz (K-corrected) and the total broad-line flux $f_{\rm line}$. We test the radio flux and the total broad-line flux against each other and find that the distribution of the radio flux is significantly different from that of the total broad-line flux (99.98 per cent confidence by Kolmogorov-Smirnov test). We find the radio and the total broad-line fluxes are significantly correlated for quasars in the sample ($\gg$99.99 per cent confidence using Spearman's correlation coefficient $\rho$). The result is same for the whole sample (we exclude the source 1226+023 in the analyses). The BL Lac objects seem to follow the similar statistical behaviour of the quasars, but with relatively lower broad-line fluxes (see Fig. 3). Fig. 4 shows strong correlation between the radio and broad-line luminosities. We also performed analyses on a subsample in the restricted redshift range: $0.5<~z<1.5$. A correlation between the radio and broad-line fluxes is found at 99.93 per cent, while the correlation between luminosities is at 99.7 per cent. There are 32 steep-spectrum quasars with radio spectral index $\alpha_{11-6}\le -0.7$ in our sample(marked by squares in figures). The sample without these steep-spectrum quasars is similar to those for the whole sample. No significant correlation is found for this steep-spectrum quasar subsample between fluxes, while a weak correlation is found between luminosities. However, we have not found obvious different behaviours between the steep-spectrum quasars and the remains of the sample, though detailed comparisons are not possible due to the small number of steep-spectrum quasars. \section{Discussion} The radio emission for steep-spectrum quasars is believed to be unbeamed emission from the lobes. The dominant influence on the radio luminosity is $Q_{jet}$, and not the large-scale radio source environment (Serjeant et al. 1998), though $Q_{jet}$ also depends on the gaseous environment of the radio source (Ellingson et al. 1991; Rawlings \& Saunders 1991). Therefore the radio emission can be a measure of the bulk power in the jets $Q_{jet}$ (Serjeant et al. 1998). The optical continuum is a good indicator of the disc surrounding a black hole for the steep-spectrum quasars, since relativistic beaming does not affect the optical continuum in these sources. Thus, the radio-optical correlation gives evidence of a disc-jet link (Serjeant et al. 1998). The optical radiation may be enhanced by relativistically beamed synchrotron radiation for some flat-spectrum quasars, which prevent us from using the optical continuum as an indicator of the disc. The broad-line region is photoionized by a nuclear source (probably radiation from the disc), so the broad-line emission instead of the optical continuum can be used as an indictor of the accretion power for both steep and flat-spectrum quasars (Celotti, Padovani \& Ghisellini 1997). The $Q_{jet}$ is proportional to the bulk Lorentz factor of the jet, and can be estimated if the size of the jet and density of the electrons in the jet are available(Celotti \& Fabian 1993). The radio emission for flat-spectrum quasars is beamed and is thought to be mainly from the core of the jet. The Doppler factor of the jet is the dominant influence on the observed radio luminosity. The radio luminosity is also determined by the size of the jet, the electron density, and the magnetic energy density in the jet. The Doppler factor depends on both the values of Lorentz factor and viewing angle of the jet. The viewing angle of the jet for flat-spectrum quasar is believed to be smaller that that of steep-spectrum quasar. The radio luminosity of flat-spectrum quasars can be naturally linked to the jet power, though some uncertainties are induced mainly by the unknown viewing angle. Part of the scatter in the radio and broad-line emission correlations can be attributed to the uncertainty of viewing angle for flat-spectrum quasars. The radio luminosity seems to be a good indicator for the jet in both steep or flat-spectrum quasars, but in different ways. The correlation found in this work between radio and broad-line emission suggests a close link between formation of the jets and accretion on to the central black hole. Another possible interpretation of the correlation is that the broad-line region is photoionized by the relativistic beamed radiation from the jets. However, the similarity of the broad-line component in radio-quiet and radio-loud sources and no systematically larger equivalent widths for lines in steep-spectrum quasars than that in flat-spectrum quasars are found (Celotti, Padovani \& Ghisellini 1997; Ghisellini 1998), which seems to rule out this possibility. The present radio and broad-line correlation gives a new evidence for a close link between the jets and discs, though how the disc and jet is coupled is still not clear. \section*{Acknowledgments} The support from NSFC and Pandeng Project is gratefully acknowledge. We thank Peter Scheuer for his helpful suggestions and the corrections on the manuscript. The anonymous referee is thanked for the useful suggestions that improved the presentation of this paper, and for his prompt reply. This research has made use of the NASA/IPAC Extragalactic Database (NED), which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautic and Space Administration.
\section*{Introduction} \def\thesection.\arabic{subsection}{\arabic{subsection}} In the last decade, there has been much interest in establishing asymptotics for the number of points of bounded height on algebraic varieties defined over number fields. Yu.~Manin and V.~Batyrev~\cite{batyrev-m90} have formulated conjectures describing such asymptotics in geometrical terms. These conjectures have been further refined by E.~Peyre in \cite{peyre95}. More precisely, let $X$ be a smooth projective algebraic variety defined over a number field $F$ and $H:X(F)\ra{\mathbf R}_{>0}$ an exponential height function on the set of rational points of $X$ defined by some metrized ample line bundle~$\mathscr L$. One wants to relate the asymptotic behaviour of the counting function \[ N(U,\mathscr L,B) = \# \{ x\in U(F)\,;\, H(x)\leq B\} \] to geometric invariants of $X$, such as the cone of effective line bundles and the (anti)-canonical line bundle of $X$. Here, $U$ is a sufficiently small Zariski dense open subset; its presence is made necessary by possible ``accumulating subvarieties'', which contain more rational points than their complement in $X$. If $X$ is a Fano variety and $\mathscr L=K_X^{-1}$, one expects that \[ N(U,K_X^{-1},B) \sim \frac{\Theta(X)}{(r-1)!} B (\log B)^{r-1} \] where $r=\rk \Pic(X)$ and $\Theta(X)$ is the product of three numbers: a Tamagawa constant which measures the volume of the closure of rational points in the adelic points $\overline{X(F)}\subset X({\mathbf A}_F)$ with respect to the metrization, a rational number defined in terms of the cone of effective divisors and the order of the non-trivial part of the Brauer group of $X$. Such a description cannot hold universally (see the example by V.~Batyrev and Yu.~Tschinkel~\cite{batyrev-t96}), but there are two classes of algebraic varieties where it does hold: those for which the \emph{circle method} in analytic number theory applies, and those possessing many symmetries, such as an \emph{action} (with a dense orbit) of a linear algebraic group. The circle method is concerned with complete intersections of small degree and small codimension in projective space. They have moduli, but only few projective embeddings; the Picard group is ${\mathbf Z}$. As a reference, let us mention the papers by B.~Birch~\cite{birch62} and W.~Schmidt~\cite{schmidt85}. The other approach leads, via harmonic analysis on the adelic points of the corresponding group, to a proof of conjectured asymptotic formulas for toric varieties (see~\cite{batyrev-t98b}) or for generalized flag varieties (using Langlands' work on Eisenstein's series, see~\cite{franke-m-t89}). These have Picard groups of higher ranks, but no deformations due to the \emph{rigidity} of reductive groups. In this paper we treat certain equivariant compactifications of vector groups. In a previous paper~\cite{chambert-loir-t99c}, we had established asymptotic formulas for blow-ups of ${\mathbf P}^2$ in any number of points \emph{on a line}. Here we work out the case of blow-ups of a projective space ${\mathbf P}^n$ of dimension at least $3$ in a smooth codimension 2 subvariety contained in a hyperplane. It should be clear to the reader that these varieties admit deformations (they are parametrized by an open subset of an appropriate Hilbert scheme). More precisely, let $f\in{\mathbf Z}[x_1,\dots,x_n]$ be a homogeneous polynomial of degree~$d$ and $X\rightarrow {\mathbf P}^n=\Proj({\mathbf Z}[x_0,\dots,x_n])$ be the blow-up of the ideal generated by $(x_0,f)$. Suppose that the hypersurface defined by $f$ in ${\mathbf P}^{n-1}_{\mathbf C}$ is smooth and let $U\simeq{\mathbf A}^n$ be the inverse image in $X$ of ${\mathbf A}^n\subset{\mathbf P}^n$. Then, $X_{\mathbf C}$ is a smooth projective variety, with Picard group ${\mathbf Z}^2$ and trivial Brauer group. Moreover, $X_{\mathbf C}$ is an equivariant compactification of ${\mathbf G}_a^n$. There is a natural metrization on $K_X^{-1}$ (recalled below) which allows to define the height function and the \emph{height zeta function} \[ Z(U,K_X^{-1},s) = \sum_{x\in U({\mathbf Q})} H_{K_X^{-1}}(x)^{-s}, \] The series converges absolutely for $\Re(s)\gg 0$. Our main theorem is: \begin{thm} There exists a function $h$ which is holomorphic in the domain $\Re(s)> 1-\frac 1n$ such that \[ Z(U,K_X^{-1},s) = \frac{h(s)}{(s-1)^2}\quad\text{and}\quad h(1)=\Theta(X)\neq 0. \] \end{thm} A standard Tauberian theorem implies that $X$ satisfies Peyre's refinement of Manin's conjecture: \begin{cor} We have the following asymptotic formula: \[ N(U,K_X^{-1},B) \sim {\Theta(X)} B \log(B)\] as $B$ tends to infinity. \end{cor} In fact, we will prove asymptotics for every $\mathscr L$ on $X$ such that its class is contained in the interior of the effective cone $\Lambda_{\text{\upshape eff}}(X)$. Moreover, we will prove estimates for the growth of $Z(s)$ in vertical strips in the neighbourhood of $\Re(s)=1$. It is well known that this implies a more precise asymptotic expansion for the counting function $N(U,\mathscr L,B)$, see Theorem~\ref{thm:peyre} and its corollary at the end of the paper. \newpage \section{Geometry, heights} \def\thesection.\arabic{subsection}{\thesection.\arabic{subsection}} Let $f\in{\mathbf Z}[x_1,\dots,x_n]$ be a homogeneous polynomial of degree~$d$ with coprime coefficients and $\pi:X\ra{\mathbf P}^n$ the blow-up of the ideal $(x_0,f)$ in ${\mathbf P}^n=\Proj({\mathbf Z}[x_0,\dots,x_n])$. We denote by $Z_f$ the hypersurface defined by $f$ in ${\mathbf P}^{n-1}$. Throughout the paper, we assume that $Z_{f,{\mathbf C}}$ is smooth, irreducible and that it doesn't contain any hyperplane. In other words, $n\geq 3$ and $d\geq 2$. The universal property of blowing up implies that the scheme $X$ is an equivariant compactification of the additive group ${\mathbf G}_a^n=\Spec({\mathbf Z}[x_1,\dots,x_n])$. Denote by $D_1$ the exceptional divisor in $X$ and by $D_0$ the strict transform of the divisor $x_0=0$ in ${\mathbf P}^n$. Let $U\simeq{\mathbf G}_a^n$ be the inverse image of ${\mathbf G}_a^n$ under $\pi$. We identify rational points in $U$ with their image in the affine space ${\mathbf G}_a^n\subset{\mathbf P}^n$. If $\mathbf s\in{\mathbf C}^2$, denote $D(\mathbf s)=s_0 [D_0]+s_1[D_1]\in\Pic(X)\otimes_{\mathbf Z}{{\mathbf C}}$. The following proposition summarizes the geometric facts needed in the sequel. \begin{prop} The classes of the divisors $D_0$ and $D_1$ form a basis of $ \Pic (X)$. For $\mathbf s=(s_0,s_1)\in{\mathbf Z}^2$, the divisor class $D(\mathbf s)$ is effective iff $s_0\geq 0$ and $s_1\geq 0$. The variety $X_{\mathbf Q}$ is smooth; its anticanonical line bundle has class $D(n+1,n)$. \end{prop} \begin{proof} See~\cite{chambert-loir-t99c}, Prop.~1.3 and Prop.~1.6 or~\cite{hartshorne77}, chap.~II, \S\,8. \end{proof} We now define height functions on $X$. We denote by $\Val({\mathbf Q})=\{2,3,\dots,\infty\}$ the set of places of ${\mathbf Q}$. If $p$ is a prime number and $\mathbf x\in{\mathbf G}_a^n({\mathbf Q}_p)$, let $\norm{\mathbf x}_p=\max(\abs{x_1}_p,\dots,\abs{x_n}_p)$ and define the functions $H_{D_1,p}$ and $H_{D_0,p}$ by \begin{align} H_{D_1,p}(\mathbf x)^{-1} & = \max\big( \frac{1}{\max(1,\norm{\mathbf x}_p)}, \frac{\abs{f(\mathbf x)}_p}{\max(1,\norm{\mathbf x}_p)^d} \big) \\ H_{D_0,p}(\mathbf x)^{-1} & = \frac{H_{D_1,p}(\mathbf x)}{\max(1,\norm{\mathbf x})}. \end{align} At the archimedian place of ${\mathbf Q}$, define the local height functions by replacing maximums by the square root of the sum of squares. For any place $v$ of ${\mathbf Q}$ and any $\mathbf s=(s_0,s_1)\in{\mathbf C}^2$, we set \begin{equation} H_v(\mathbf s;\mathbf x)=H_{D_0,v}(\mathbf x)^{s_0} H_{D_1,v}(\mathbf x)^{s_1}. \end{equation} Finally, we define a global height pairing \begin{equation} H:\Pic(X)_{\mathbf C} \times {\mathbf G}_a^n({\mathbf A}_{\mathbf Q}) \ra{\mathbf C}^*, \quad H(\mathbf s;\mathbf x) = \prod_{v\in\Val({\mathbf Q})} H_v(\mathbf s;\mathbf x_v). \end{equation} \begin{prop} If $\mathscr L\in\Pic(X)$, the function $\mathbf x\mapsto H(\mathscr L;\mathbf x)$ on ${\mathbf G}_a^n({\mathbf Q})$ is an exponential height in the sense of Weil. \end{prop} \begin{proof} See~\cite{chambert-loir-t99c}, (1.12), (1.13) and (2.2). \end{proof} The height zeta function is then defined by the series \begin{equation} Z(\mathbf s) = \sum_{\mathbf x\in {\mathbf G}_a^n({\mathbf Q})} H(\mathbf s;\mathbf x)^{-1}. \end{equation} It converges \emph{a priori} for all $\mathbf s\in{\mathbf C}^2$ such that $D(\mathbf s)$ is sufficiently ample, i.e.~if $\Re(s_0-s_1)$ and $\Re(s_0)$ are big enough. Let $\psi=\prod_v\psi_v:{\mathbf G}_a({\mathbf A}_{\mathbf Q})\ra{\mathbf C}^*$ be the standard additive character of ${\mathbf A}_{\mathbf Q}$. If $\mathbf a\in{\mathbf Q}^n$, we define \[ \psi_{\mathbf a}(\mathbf x)=\psi(\langle\mathbf a,\mathbf x\rangle).\] We use the standard self-dual Haar measure $\,{\mathrm d\mathbf x}$ on ${\mathbf G}_a^n({\mathbf A}_{\mathbf Q})$. For any $\mathbf a\in{\mathbf Q}^n$, define the Fourier transform \[ \hat H(\mathbf s;\psi_{\mathbf a}) = \int_{{\mathbf G}_a^n({\mathbf A}_{\mathbf Q})} H(\mathbf s;\mathbf x)^{-1}\, \,{\mathrm d\mathbf x}. \] It is the product of the local Fourier transforms $\hat H_v(\mathbf s;\psi_{\mathbf a})$. For $\mathbf s\in{\mathbf C}^2$ such that both sides converge absolutely, we have the following identity: \begin{equation} Z(\mathbf s) = \sum_{\mathbf a\in{\mathbf Z}^n} \hat H(\mathbf s;\psi_{\mathbf a}). \end{equation} This is a consequence of the usual Poisson formula, see~\cite{chambert-loir-t99c}, end of~\S\,2. In the following sections we determine the domain of absolute convergence of the right hand side and prove that $Z(\mathbf s)$ admits a meromorphic continuation beyond this domain. \section{The local Fourier transform at the trivial character} We denote by $S$ the minimal set of primes such that $Z_f\subset{\mathbf P}^{n-1}_{\mathbf Z}$ is smooth over $\Spec {\mathbf Z}[S^{-1}]$. Let $p$ be a prime number. \subsection{Decomposition of the domain} We define subsets of ${\mathbf Q}_p^n$ as follows: \begin{itemize} \item $U(0)={\mathbf Z}_p^n$; \item if $0\leq \beta < \alpha$, $U_1(\alpha,\beta)$ is the set of $\mathbf x\in{\mathbf Q}_p^n$ such that $\norm{\mathbf x}=p^\alpha$ and $\abs{f(\mathbf x)}=p^{d\alpha-\beta}$; \item if $\alpha\geq 1$, $U_1(\alpha)$ is the set of $\mathbf x\in{\mathbf Q}_p^n$ such that $\norm{\mathbf x}=p^\alpha$ and $\abs{f(\mathbf x)}\leq p^{(d-1)\alpha}$; \item if $\alpha\geq 1$, $U(\alpha)$ is the set of $\mathbf x\in{\mathbf Q}_p^n$ such that $\norm{\mathbf x}=p^\alpha$ and $\abs{f(\mathbf x)}=p^{d\alpha}$. \end{itemize} The local height function is constant on each of these subsets. Namely, if $\mathbf x\in U(0)$, $H_{D_0,p}=H_{D_1,p}=1$. If $\mathbf x\in U_1(\alpha,\beta)$, $H_{D_0,p}= p^{\alpha-\beta}$ and $H_{D_1,p}=p^\beta$. On $U(\alpha)$, $H_{D_0,p}=p^\alpha$ and $H_{D_1,p}=1$. Finally, if $\mathbf x\in U(\alpha)$, then $H_{D_0,p}=1$ and $H_{D_1,p}=p^\alpha$. \subsection{Volumes} Denote by \[ \tau_p(f) = \big( 1-\frac1p \big) \frac{\# Z_f({\mathbf F}_p)}{p^{n-2}}. \] The Weil conjectures proved by Deligne imply that $\tau_p(f)=1+O(1/p)$. In a much more elementary way, it follows from Lemma~\ref{lem:weil} below that $\tau_p(f)$ is bounded as $p$ varies. \begin{lem} For $p\not\in S$, we have \addtocounter{equation}{-1} \begin{subequations} \begin{align} \vol(U(0)) &= 1 \\ \vol (U_1(\alpha,\beta)) &= \frac{p-1}p \tau_p(f) p^{n\alpha -\beta} \\ \vol (U_1(\alpha)) &= \tau_p(f) p^{(n-1)\alpha} \\ \vol (U(\alpha)) & = \big(1-p^{-n} - p^{-1}\tau_p(f)\big)p^{n\alpha}. \end{align} \end{subequations} \end{lem} \begin{proof} For $\beta\geq 1$, let $\Omega(\beta)$ be the set of $\mathbf x\in{\mathbf Z}_p^n$ such that $\norm{\mathbf x}=1$ and $\abs{f(\mathbf x)}\leq p^{-\beta}$. By definition, \[ \vol(\Omega(\beta)) = p^{-n\beta} p^{\beta-1}(p-1) \# Z_f({\mathbf Z}/p^\beta{\mathbf Z}).\] As $Z_f$ is smooth of pure dimension $n-2$ over ${\mathbf Z}_p$, Hensel's lemma implies that \[ \# Z_f({\mathbf Z}/p^\beta{\mathbf Z})=p^{(\beta-1)(n-2)}\# Z_f({\mathbf F}_p). \] Consequently, \[ \vol(\Omega(\beta)) = (p-1) p^{-\beta-1} \frac{\# Z_f({\mathbf F}_p)}{p^{n-2}} = \tau_p(f)p^{-\beta} . \] As $ U_1(\alpha)=p^{-\alpha} \Omega(\alpha)$, we have \[ \vol(U_1(\alpha)) = \tau_p(f) p^{(n-1)\alpha}. \] Now, \[ U_1(\alpha,\beta)=p^{-\alpha} U_1(0,\beta)=p^{-\alpha}\big( \Omega(\beta)-\Omega(\beta+1) \big), \] therefore \[ \vol (U_1(\alpha,\beta)) = \frac{p-1}p \tau_p(f) p^{n\alpha-\beta} . \] Finally, $U(\alpha)=p^{-\alpha} ({\mathbf Z}_p^n\setminus (p{\mathbf Z}_p^n\cup \Omega(1)) )$, hence \[ \vol(U(\alpha))=\big(1-p^{-n} - p^{-1}\tau_p(f)\big)p^{n\alpha} . \] \end{proof} \begin{prop} Assume that $p\not\in S$. Then, \[ \hat H_p(\mathbf s;\psi_0) = \hat H_{{\mathbf P}^n,p}(s_0) + \tau_p(f) \frac{p^{s_0-n}-p^{s_1-n}}{(p^{s_0-n}-1)(p^{s_1-n+1}-1)} \] where \[ \hat H_{{\mathbf P}^n,p}(s_0) = \frac{1-p^{-s_0}}{1-p^{n-s_0}} \] denotes the Fourier transform (with respect to the trivial character $\psi_0$) of the local height function of ${\mathbf P}^n$ for the tautological line bundle at $s_0$. \end{prop} \begin{proof} By definition, \begin{align*} \hat H_p(\mathbf s;\psi_0) & = \int_{{\mathbf Q}_p^n} H(\mathbf s;\mathbf x)^{-1}\, \,{\mathrm d\mathbf x} \\ &= \int_{U(0)} + \sum_{1\leq\beta <\alpha} \int_{U_1(\alpha,\beta)} + \sum_{1\leq\alpha} \int_{U_1(\alpha)} + \sum_{1\leq\alpha}\int_{U(\alpha)}. \end{align*} We compute these sums separately. The integral over $U(0)$ is equal to~$1$. Then \begin{align*} \sum_{1\leq\beta <\alpha} \int_{U_1(\alpha,\beta)} &= \frac{p-1}p \tau_p(f) \sum_{1\leq\beta<\alpha} p^{-\alpha s_0} p^{-\beta(s_1-s_0)} p^{\alpha n-\beta} \\ &= \frac{p-1}p \tau_p(f) \sum_{\beta=1}^\infty p^{-\beta(s_1-s_0+1)} \sum_{\alpha=\beta+1}^\infty p^{-\alpha(s_0-n)} \\ &= \frac{p-1}p \tau_p(f) \sum_{\beta=1}^\infty p^{-\beta(s_1-s_0+1)} p^{-\beta(s_0-n)} \frac{1}{p^{s_0-n}-1} \\ &= \frac{p-1}p \tau_p(f) \frac{1}{p^{s_0-n}-1} \sum_{\beta=1}^\infty p^{-\beta(s_1-n+1)} \\ &= \frac{p-1}p \tau_p(f) \frac{1}{p^{s_0-n}-1} \frac{1}{p^{s_1-n+1}-1}. \end{align*} Concerning the integrals over $U_1(\alpha)$, we have \[ \sum_{1\leq\alpha} \int_{U_1(\alpha)} = \tau_p(f) \sum_{\alpha=1}^\infty p^{-\alpha s_1} p^{(n-1)\alpha} = \tau_p(f) \frac{1}{p^{s_1-n+1}-1}. \] Finally, \begin{align*} \sum_{1\leq\alpha} \int_{U(\alpha)} & = (1-p^{-n}-p^{-1}\tau_p(f)) \sum_{\alpha=1}^\infty p^{-s_0\alpha} p^{n\alpha} \\ & = \big( 1-p^{-n}-p^{-1}\tau_p(f) \big) \frac{1}{p^{s_0-n}-1}. \end{align*} Adding all these terms gives \begin{align*} \hat H_p(\mathbf s;\psi_0) &= 1 + (1-p^{-1})\tau_p(f) \frac{1}{p^{s_0-n}-1} \frac{1}{p^{s_1-n+1}-1} \\ & \qquad{} +\tau_p(f) \frac{1}{p^{s_1-n+1}-1} + \big( 1-p^{-n}-p^{-1}\tau_p(f) \big) \frac{1}{p^{s_0-n}-1} \\ &= 1 + (1-p^{-n}) \frac{1}{p^{s_0-n}-1} \\ & \qquad{} + \tau_p(f) \left( (1-p^{-1}) \frac{1}{p^{s_0-n}-1} \frac{1}{p^{s_1-n+1}-1} + \frac{1}{p^{s_1-n+1}-1} \right. \\ & \qquad\qquad \left. {} - p^{-1} \frac{1}{p^{s_0-n}-1} \right) \\ & = \hat H_{{\mathbf P}^n,p}(s_0) + p^{-1} \tau_p(f) \frac{p-1 + p^{s_0-n+1}-p - p^{s_1-n+1}} {(p^{s_0-n}-1)(p^{s_1-n+1}-1)} \\ & = \hat H_{{\mathbf P}^n,p}(s_0) + p^{-1} \tau_p(f) \frac{p^{s_0-n+1}-p^{s_1-n+1}}{(p^{s_0-n}-1)(p^{s_1-n+1}-1)} \\ & = \hat H_{{\mathbf P}^n,p}(s_0) + \tau_p(f) p^{n-1} \frac{p^{-s_1}-p^{-s_0}}{(1-p^{n-s_0})(1-p^{n-1-s_1})} \end{align*} \end{proof} \section{The local Fourier transform at a non-trivial character} In this subsection we evaluate the local Fourier transform at $p$ for a non-trivial character $\psi_{\mathbf a}$. Let $S(\mathbf a)$ be the union of $S$ and of the set of primes $p$ such that $\mathbf a\in p{\mathbf Z}^n$; We assume that $p\not\in S(\mathbf a)$. Recall that $Z_f\subset{\mathbf P}^{n-1}_{\mathbf Z}$ denotes the subscheme defined by $f$ and define $Z_{f,\mathbf a}=Z_f\cap H_{\mathbf a}$, where $H_{\mathbf a}$ is the hyperplane of ${\mathbf P}^{n-1}$ defined by $\mathbf a$. Finally, let $Z_{f,\mathbf a}^t$ (resp.\ $Z_{f,\mathbf a}^{nt}$) be the locus of points in $Z_{f,\mathbf a}$ where the intersection $Z_f\cap H_{\mathbf a}$ is \emph{transverse} (resp.\ is \emph{not transverse}). By assumption, $Z_f$ and $H_{\mathbf a}$ are smooth over ${\mathbf Z}_p$. Let $I(\alpha,\beta)$ be the integral of $\psi_{\mathbf a}$ over the set of $\mathbf x\in{\mathbf Q}_p^n$ such that $\norm{\mathbf x}=p^\alpha$ and $\abs{f(\mathbf x)}\leq p^{d\alpha -\beta}$. Then, according to our partition of ${\mathbf Q}_p^n$, we have \begin{align*} \hat H_p(\mathbf s;\psi_{\mathbf a} ) & = 1 + \sum_{\alpha=1}^\infty \sum_{\beta=0}^{\alpha-1} p^{-\alpha s_0} p^{-\beta(s_1-s_0)} \int _{\substack{\norm{\mathbf x}=p^\alpha \\ \abs{f(\mathbf x)}=p^{d\alpha-\beta}}} \psi_{\mathbf a} \\ &\hskip .5\textwidth + \sum_{\alpha=1}^\infty p^{-\alpha s_1} \int_{\substack{\norm{\mathbf x}=p^{\alpha} \\ \abs{f (\mathbf x)}\leq p^{-\alpha(d-1)}}} \psi_{\mathbf a} \\ &= 1+ \sum_{\alpha=1}^\infty \sum_{\beta=0}^{\alpha-1} p^{-\alpha s_0} p^{-\beta(s_1-s_0)} \big( I(\alpha,\beta)- I(\alpha,\beta+1)\big) \\ &\hskip .5\textwidth + \sum_{\alpha=1}^\infty p^{-\alpha s_1} I(\alpha,\alpha). \\ &= 1+ \sum_{\alpha=1}^\infty p^{-\alpha s_0} I(\alpha,0) - (p^{s_1-s_0}-1) \sum_{\alpha=1}^\infty \sum_{\beta=1}^\alpha p^{-\alpha s_0} p^{-\beta(s_1-s_0)} I(\alpha,\beta) \end{align*} \begin{lem} If $t\in{\mathbf Q}_p$, the mean value over ${\mathbf Z}_p^*$ of $\psi(t\cdot)$ is equal to \[ \frac{\int_{{\mathbf Z}_p^*} \psi(tu)\, du} {\int_{{\mathbf Z}_p^*} du } = \begin{cases} 1 & \text{if $t\in{\mathbf Z}_p$;} \\ -1/(p-1) & \text{if $v_p(t)=-1$;} \\ 0 & \text{if $v_p(t)\leq -2$.} \end{cases} \] \end{lem} \begin{proof} Indeed, we have \begin{align*} \int_{{\mathbf Z}_p^*} \psi(tu)\, du &= \int_{{\mathbf Z}_p} \psi(tu)\, du - \int_{p{\mathbf Z}_p} \psi(tu)\, du \\ &= \int_{{\mathbf Z}_p} \psi(tu)\, du - \frac1p \int_{{\mathbf Z}_p}\psi(ptu)\, du . \end{align*} The integral of a non-trivial character over a compact group is $0$, hence this integral equals $0$ if $t\not\in p^{-1}{\mathbf Z}_p$, equals $-\frac 1p$ if $t\in p^{-1}{\mathbf Z}_p\setminus {\mathbf Z}_p$ and equals $1-\frac1p$ if $t\in{\mathbf Z}_p$. This proves the lemma. \end{proof} Using the change of variables $\mathbf x=p^{-\alpha} \mathbf y$, this implies the following formula: \begin{multline} I(\alpha,\beta) = p^{n\alpha} \left( \frac p{p-1} \vol\big(\norm{x}=1;\, p^{\beta}|f(\mathbf x);\, p^\alpha|\langle \mathbf a,\mathbf x\rangle \big)\right. \\ \left. - \frac1{p-1} \vol\big(\norm{x}=1;\, p^{\beta}|f(\mathbf x);\, p^{\alpha-1}|\langle \mathbf a,\mathbf x\rangle \big) \right). \end{multline} \begin{lem} If $1\leq\beta\leq\alpha$, one has \[ \vol\big(\norm{\mathbf x}=1;\, p^{\beta}| f(\mathbf x);\, p^\alpha| \langle\mathbf a,\mathbf x\rangle \big) = p^{-\alpha} p^{(2-n)\beta }\big(1-\frac1p\big) \# Z_{f,\mathbf a}({\mathbf Z}/p^\beta{\mathbf Z}). \] \end{lem} In particular, \begin{equation} I(\alpha,\beta) = 0 \quad\text{if $1\leq\beta <\alpha$} . \end{equation} Moreover, if $\alpha\geq 2$, \begin{align*} \vol\big( \norm{\mathbf x}=1;\, p^{\alpha}| f(\mathbf x);\, p^{\alpha-1}| \langle\mathbf a,\mathbf x\rangle\big) &= \frac1p \vol\big( \norm{\mathbf x}=1;\, p^{\alpha-1}| f(\mathbf x);\, p^{\alpha-1}| \langle\mathbf a,\mathbf x\rangle\big) \\ &= \frac1p p^{(1-\alpha)(n-1)} \big(1-\frac1p) \#Z_{f,\mathbf a}({\mathbf Z}/p^{\alpha-1}) . \end{align*} If $\alpha=1$, one has \[ \vol\big(\norm{\mathbf x}=1;\, p | f(\mathbf x)\big) = \big(1-\frac1p\big) p^{1-n} \# Z_f({\mathbf Z}/p{\mathbf Z}). \] We had computed in~\cite{chambert-loir-t99c}, proof of Lemma 3.5, the integral \[ \int_{\norm{\mathbf x}=p^\alpha}\psi_{\mathbf a} = \begin{cases} -1 & \text{if $\alpha=1$;} \\ 0 & \text{if $\alpha\geq 2$.} \end{cases} \] so that \begin{multline} \label{formula:hathp} \hat H(\mathbf s;\psi_{\mathbf a}) = 1- p^{-s_0} + \frac{p^{s_1-s_0}-1}{p-1} p^{-s_1} \# Z_f({\mathbf F}_p) \\ - \frac{p^{s_1-s_0}-1}{p-1} (1-p^{n-s_1-2}) \sum_{\alpha=1}^\infty p^{-\alpha(s_1-1)} \# Z_{f,\mathbf a}({\mathbf Z}/p^\alpha{\mathbf Z}). \end{multline} \begin{lem} For all $\alpha\geq 1$, \[ \# Z_{f,\mathbf a} ({\mathbf Z}/p^\alpha{\mathbf Z}) \leq p^{(n-3)(\alpha-1)} \# Z_{f,\mathbf a}^t ({\mathbf Z}/p{\mathbf Z}) + p^{(n-2)(\alpha-1)} \# Z_{f,\mathbf a}^{nt} ({\mathbf Z}/p{\mathbf Z}). \] \end{lem} \begin{proof} The inequality is trivially true for $\alpha=1$. We prove it for any $\alpha$ by induction: to lift a point in $Z_{f,\mathbf a}({\mathbf Z}/p^\alpha{\mathbf Z})$ to a point in $Z_{f,\mathbf a}({\mathbf Z}/p^{\alpha+1}{\mathbf Z})$, one needs to solve two equations in $\mathbf u\in{\mathbf F}_p^n$: \[ \langle \nabla f(\mathbf x),\mathbf u\rangle \equiv p^{-\alpha} f(\mathbf x) , \quad \langle {\mathbf a},\mathbf u \rangle \equiv p^{-\alpha} \langle {\mathbf a},\mathbf x\rangle \pmod{p} .\] A point in $Z_{f,\mathbf a}({\mathbf Z}/p^\alpha{\mathbf Z})$ which reduces to a point in $Z_{f,\mathbf a}^t$ modulo $p$ has $p^{n-3}$ lifts in $Z_{f,\mathbf a}({\mathbf Z}/p^{\alpha+1}{\mathbf Z})$. On the other hand, a point reducing to a point in $Z_{f,\mathbf a}^{nt}$ has $p^{n-2}$ or $0$ lifts according to the two linear equations being compatible or not. This implies the lemma. \end{proof} \begin{prop} \label{prop:37} If not empty, the set $Z_{f,\mathbf a}^{nt}$ is a closed subscheme of bounded degree of $Z_{f,\mathbf a}$ and of dimension~$0$. There exist a constant $C$, independent of $\mathbf a$ and $p$ such that \[ \# Z_{f,\mathbf a}^t ({\mathbf Z}/p{\mathbf Z}) \leq C p^{n-3}, \quad \# Z_{f,\mathbf a}^{nt}({\mathbf Z}/p{\mathbf Z}) \leq C . \] \end{prop} As a corollary, one gets: \begin{cor} There exist a constant $C$ such that for all $\alpha$ and $p\not\in S(\mathbf a)$, \[ \# Z_{f,\mathbf a}({\mathbf Z}/p^\alpha{\mathbf Z}) \leq C p^{(n-3)\alpha} + C p^{(n-2)(\alpha-1)}. \] \end{cor} \begin{proof}[Proof of Prop.~\ref{prop:37}] The set $Z_{f,\mathbf a}$ is defined by the two equations $f(\mathbf x)=\langle \mathbf a,\mathbf x\rangle=0$. Fix the coordinates $x_1,\dots,x_n$ so that $\mathbf a$ is the first vector. Up to a constant, one may write \[ f(\mathbf x)=x_1^d + g_1(x_2,\dots,x_n)x_1^{d-1} + \cdots + g_{d_1} x_1 + g_d (x_2,\dots,x_n) \] for some homogeneous polynomials $g_i$ of degree $i$. Then, denoting $\mathbf x= (x_1,\mathbf x')$, $Z_{f,\mathbf a}$ is defined by the equations \[ x_1=g_d(\mathbf x')= \partial_2 g_d(\mathbf x') = \dots =\partial_n g_d(\mathbf x') = 0. \] On $Z_{f,\mathbf a}$, $\partial_1 f(0,\mathbf x')=g_{d-1}(\mathbf x')$ and on $Z_{f,\mathbf a}^{nt}\subset Z_{f,\mathbf a}$, $\partial_i f(0,\mathbf x')=\partial_i g_d(\mathbf x')$. As $Z_f$ is smooth, $g_{d-1}(\mathbf x')$ doesn't vanish on $Z_{f,\mathbf a}^{nt}$ which must therefore be either empty or of dimension $0$. Its degree cannot exceed $d (d-1)^{n-1}$. The bound on the number of ${\mathbf F}_p$-rational points are a consequence of the following (certainly well-known) easy lemma. \end{proof} \begin{lem} \label{lem:weil} Let $k={\mathbf F}_q$ be a finite field, $X$ a closed subscheme of ${\mathbf P}^n_k$ of dimension $d$. Then \[ \# X({\mathbf F}_q) \leq {\mathbf P}^d({\mathbf F}_q) \deg X. \] \end{lem} \begin{proof} We prove this by induction on $d$. If $d=0$, the result is clear. Then, one can assume that $X$ is reduced, irreducible and not contained in any hyperplane. For any hyperplane $H\subset{\mathbf P}^n$ which is rational over $k$, $X\cap H$ is a closed subscheme of $H$ of dimension $d-1$ and of degree $\leq\deg X$. By induction, we have \[ \# (X\cap H)({\mathbf F}_q) \leq \# {\mathbf P}^{d-1} ({\mathbf F}_q) \deg X. \] Finally, any point of $X({\mathbf F}_q)$ is contained in exactly $\# {\mathbf P}^{n-1}({\mathbf F}_q)$ rational hyperplanes in ${\mathbf P}^n$, so that \[ \# X({\mathbf F}_q) \#{\mathbf P}^{n-1}({\mathbf F}_q) \leq \# {\mathbf P}^{d-1}({\mathbf F}_q) \# {\mathbf P}^n({\mathbf F}_q) \deg X .\] As $n\geq d$, this implies \[ \# X({\mathbf F}_q)\leq \frac{q^{n+1}-1}{q^n-1} \frac{q^d-1}{q-1} \deg X \leq {\mathbf P}^{d}({\mathbf F}_q) \deg X. \] \end{proof} \section{The height zeta function} From now on, we fix some $\eps>0$ and consider only $\mathbf s$ in the subset $\Omega$ of ${\mathbf C}^2$ defined by the inequalities $\Re(s_0)>n+\eps$ and $\Re(s_1)>n-1+\eps$. \begin{prop} There exist a holomorphic function $g$ on $\Omega$ which has polynomial growth in vertical strips such that \[ \hat H(\mathbf s,\psi_0) = g(\mathbf s) \frac{1}{(s_0-n-1)(s_1-n)}.\] \end{prop} \begin{proof} Indeed, we see from 2.3 that for $p\not\in S$, \[ \hat H_p(\mathbf s,\psi_0) = 1+ p^{n-s_0} + p^{n-s_1-1} + O(p^{-1-\eps}), \] the $O$ being uniform in $p$. Consequently, \[ \prod_{p\not\in S} \hat H_p(\mathbf s,\psi_0)(1-p^{n-s_0})(1-p^{n-1-s_1}) \] converges to a holomorphic bounded function on $\Omega$. As the finite number of remaining factors converge uniformly in $\Omega$, the existence of $g$ is proven. The growth of $g$ in vertical strips follows from Rademacher's estimates for the Riemann zeta function. \end{proof} \begin{lem} There exist a constant $C>0$ such that for all $\mathbf a\in{\mathbf Z}^n\setminus\{0\}$, all $p\not\in S(\mathbf a)$ and all $(s_0,s_1)\in\Omega$, one has \[ \abs{ \hat H_p(\mathbf s,\psi_{\mathbf a}) -1} \leq C p^{-1-\eps}. \] \end{lem} \begin{proof} Recall the formula~\ref{formula:hathp}: \begin{multline*} \hat H_p(\mathbf s,\psi_{\mathbf a}) - 1 = - p^{-s_0} + \frac{p^{-s_0}-p^{-s_1}}{p-1} p^{n-2}(1-\frac1p)^{-1}\tau_p(f) \\ - \frac{p^{s_1-s_0}-1}{p-1} (1-p^{n-s_1-2}) \sum_{\alpha=1}^\infty p^{-\alpha(s_1-1)}\# Z_{f,\mathbf a}({\mathbf Z}/p^\alpha{\mathbf Z}) \end{multline*} the right hand side of which we have to estimate all terms. The first one is $p^{-s_0}=O(p^{-1-\eps})$. Then, as $\tau_p(f)$ is bounded, the second one is \[ O(p^{n-3-\Re(s_0)}) + O(p^{n-3-\Re(s_1)}) = O(p^{-2}). \] For the last term $T_3$, we use Lemma 3.8 so that, denoting $\sigma_1=\Re(s_1)$, \begin{multline*} \sum_{\alpha=1}^\infty p^{-\alpha(s_1-1)}\# Z_{f,\mathbf a}({\mathbf Z}/p^\alpha{\mathbf Z}) \\ \leq C \sum_{\alpha=1}^\infty p^{-\alpha(\sigma_1-1)} p^{(n-3)\alpha} + C \sum_{\alpha=1}^\infty p^{-\alpha(\sigma_1-1)} p^{(n-2)(\alpha-1)} \\ \leq C \frac{1}{p^{\sigma_1-n+2}-1} + C p^{2-n} \frac{1}{p^{\sigma_1-n+1}-1}. \end{multline*} Moreover, \[ \abs{1-p^{n-s_1-2}} \leq 2 \] so that \begin{align*} \abs{T_3}& \ll \frac{1}{p-1}\, \frac{p^{\sigma_1-\sigma_0}+ 1}{p^{\sigma_1-n+2}-1} +2C \frac{p^{2-n}}{p-1} \frac{p^{\sigma_1-\sigma_0}+1}{p^{\sigma_1-n+1}-1} \\ &\ll \frac{1}{p} \big( p^{n-2-\sigma_0} + p^{n-2-\sigma_1}\big) + p^{1-n} \big( p^{n-1-\sigma_0}+p^{n-1-\sigma_1}\big) \\ &\ll p^{-2}. \end{align*} The lemma is proved. \end{proof} \begin{prop} For each $\mathbf a\in{\mathbf Z}^n\setminus\{0\}$, $\hat H(\mathbf s,\psi_{\mathbf a})$ is a holomorphic function on $\Omega$. Moreover, there exist constants $C>0$ and $\nu$ (which are independent of $\mathbf s$ and $\mathbf a$) such that \[ \abs{\hat H(\mathbf s,\psi_{\mathbf a})} \leq C (1+\norm{\Im(s)})^\nu (1+\norm{\mathbf a})^{-n-1}.\] \end{prop} \begin{proof} Write \[ \hat H(\mathbf s,\psi_{\mathbf a}) = \prod_{p\not\in S(\mathbf a)} \hat H_p \times \prod_{p\in S(\mathbf a)} \hat H_p \times \hat H_\infty. \] The convergence of the first infinite product to a bounded holomorphic function follows from the preceding lemma. As in Lemma~3.7 of \cite{chambert-loir-t99c}, there exists a constant $\kappa>0$ such that \[ \abs{\prod_{p\in S(\mathbf a)} \hat H_p (\mathbf s,\psi_{\mathbf a})} \ll (1+\norm{\mathbf a})^{\kappa}. \] Using the rapidly decreasing behaviour of $\hat H_\infty$ as a function of $\mathbf a$ \[ \abs{\hat H_\infty(\mathbf s,\psi_{\mathbf a})} \ll (1+\norm{\mathbf a})^{-n-\kappa-1} \] established in Prop.~2.13 of \emph{loc. cit.}, the proposition is proved. \end{proof} \begin{thm} \label{thm:peyre} The height zeta function converges in the domain $\Re(s_0)>n+1$, $\Re(s_1)>n$. Moreover, there exists a holomorphic function $g$ in the domain $\Re(s_0)>n$, $\Re(s_1)>n-1$ such that \[ Z(\mathbf s) = g(\mathbf s) \frac{1}{(s_0-n-1)(s_1-n)}. \] The function $g$ has polynomial growth in vertical strips and $g(n+1,n)\neq 0$. \end{thm} Specializing to $\mathbf s=s(n+1,n)$ and using a standard Tauberian theorem, one obtains the following corollary. \begin{cor} \label{cor:peyre} There exist a polynomial $P_X$ of degree $1$ and a real number $\alpha>0$ such that the number of points of $U({\mathbf Q})\subset X({\mathbf Q})$ of anticanonical height $\leq B$ satisfies \[ N(U,K_X^{-1},B)=B P(\log B) + O(B^{1-\alpha}). \] Moreover, if $\tau(K_X)$ denotes the Tamagawa number, the leading coefficient of $P_X$ is equal to \[ \frac{\tau(K_X)}{(n+1)n}, \] as predicted by Peyre's refinement of Manin's conjecture. \end{cor} \def\noop#1{\ignorespaces} \bibliographystyle{smfplain}
\section{Introduction} It is remarkable that the simple fireball model for cosmological gamma-ray bursts can explain the many major features of the gamma-ray bursts and their afterglows (Rees \& M\'ezs\'aros 1992, 1994; Paczynski \& Rhodes 1993; Wijers, Rees, \& M\'ezs\'aros 1997; Vietri 1997a,b; Waxman 1997a,b; Reichart 1997; Katz \& Piran 1997; Sari 1997). It is noted, however, that the fireball model may produce multiple spectral components due to the existence of both the forward and reverse shocks at least at some time period (e.g., M\'ezs\'aros, Rees, \& Papathanassiou 1994). Moreover, reprocessing of the primary spectrum by, for example, inverse Compton scattering (e.g., Pilla \& Loeb 1998), introduces additional features to the spectra, even if it is initially featureless. Finally, resonant line features or recombination lines due to heavy elements, for example, in dense blobs as discussed in M\'ezs\'aros \& Rees (1998), may add additional features to the continuum. In this {\it Letter} it is pointed out that spectral smoothing due to differentially varying Doppler shift for different patches of the fireball at varying angles to the line of sight provides an intrinsic, unavoidable smoothing mechanism. While this effect is fairly well known, a more quantitative analysis focusing on the case of GRBs is useful. This {\it Letter} presents such a semi-quantitative analysis. \section{Differential Doppler Shift Across a Fireball Front} For the present illustration I assume that the GRB fireball is spherical and homogeneous, for which a single parameter, $\theta$, the angle to the burster-observer vector, is sufficient to characterize the direction of a traveling patch of shock heated, radiation emitting material. It is convenient to use the time measured in the rest-frame of the burster, $t$, as the independent variable to express other quantities. First, one has to find the relation between time $t$ for a fireball patch at $\theta$ and the time measured by the observer on the Earth (called ``{\it O}" hereafter), $t_{obs}$, i.e., the arrival time. Since the {\it apparent} perpendicular traveling speed of a patch with $\theta$ seen by {\it O} is \begin{equation} \beta_{\perp} (t)={\beta (t) \sin\theta \over 1-\beta (t)\cos\theta} \end{equation} \noindent in units of the speed of light, where $\beta (t)$ is the spherical expansion speed of the fireball in the rest-frame of the burster. By definition, $\beta_\perp(t)$ is \begin{equation} \beta_{\perp}(t)={dr \over dt_{obs}}\sin\theta. \end{equation} \noindent Combining equations (1, 2) gives \begin{equation} {dr \over dt_{obs}} = {\beta(t)\over 1-\beta(t)\cos\theta}. \end{equation} \noindent Since one also has the following relation \begin{equation} {dr \over dt} = \beta (t), \end{equation} \noindent one finds the equation relating $t_{obs}$ to $t$: \begin{equation} {dt_{obs} \over dt} = 1-\beta(t)\cos\theta. \end{equation} Next, to have a tractable treatment, it is assumed that the Lorentz factor, $\Gamma(t)\equiv 1/\sqrt{1-\beta(t)^2}$, has the following simplified evolution: it is constant (equal to $\Gamma_i$) at $t\le t_{dec}$ and decays at $t > t_{dec}$ as \begin{equation} \Gamma = \Gamma_i ({t\over t_{dec}})^{-\alpha}, \end{equation} \noindent where $\Gamma_i$ is the initial Lorentz factor and $t_{dec}$ (measured in the rest-frame of the burster) characterizes the transition time after which deceleration of the fireball expansion becomes significant (hence the fireball kinetic energy can be converted into radiation) and can be expressed approximately as $t_{dec}=({3E\over 4\pi \Gamma_i^2 c^5 m_p n})^{1/3}$ (Blandford \& McKee 1976), where $E$ is the initial fireball energy, $n$ is the density of the circumburster medium (which is assumed to be uniform, for simplicity) and other notations are conventional. Note that $\alpha=3$, if the fireball cools radiatively efficiently, and $\alpha=3/2$, if the fireball cools only adiabatically. To make the point in a simple way it is assumed that the total luminosity per unit frequency (in the comoving frame) of the shock front at a fixed time $t$ is a delta function in frequency (i.e., a monochromatic spectrum): \begin{equation} L_\nu ({\nu^\prime},t) = C(t)\delta [\nu^\prime -\nu_0(t)] \end{equation} \noindent in the frame comoving with the blastwave. The characteristic frequency $\nu_0(t)$ at time $t$ in the comoving frame is parameterized as \begin{equation} \nu_0(t) = A ({\Gamma\over\Gamma_i})^\psi, \end{equation} \noindent where $A$ is a constant. Note that, in the case of synchrotron radiation and assuming that both the electron thermal energy density and the magnetic energy density are fixed fractions of the post-shock nucleon thermal energy density, one has $\psi=3$. $C(t)$ is expressed as \begin{equation} C(t) = B ({t\over t_{dec}})^\xi, \end{equation} \noindent where $B$ is another constant and the $\xi$ parameterizes the temporal profile of the amplitude of the radiation. At a given time {\it O} receives radiation from {\it different parts across the fireball surface}, emitted at varying times in the burster frame; i.e., radiation from regions of varying $\theta$ at varying $t$ [see equation (5) for the relation between $t$ and $t_{obs}$] is seen by {\it O} at the same time (Sari 1998). The received frequency of the radiation at $t_{obs}$ from region with $\theta$ emitted at $t$ is \begin{equation} \nu(t_{obs}) = \nu_0(t) D(\theta,t), \end{equation} \noindent where $D(\theta,t)$ is the Doppler factor for regions with $\theta$ at time $t$: \begin{equation} D(\theta,t) = {1\over\Gamma(t) [1-\beta(t)\cos\theta]}. \end{equation} \noindent The flux density observed by {\it O} at frequency $\nu$ at time $t_{obs}$ is \begin{eqnarray} S(\nu,t_{obs}) = {1\over 8\pi d^2} \int_{-1}^1 L_\nu(\nu^\prime,t) D^3(\theta,t) d\mu \end{eqnarray} \noindent where $\nu'$($=\nu/D$) is the frequency in the blastwave frame, $d$ is the distance of the GRB from $O$ and $\mu\equiv\cos\theta$. Combining equations (10,11) gives \begin{equation} {\rm d} \mu = \left( \frac{1 - \beta \mu}{\beta} \right) \: \frac{{\rm d} \nu}{\nu} \; . \end{equation} \noindent Inserting equations (7,13) into equation (12) and integrating over $\nu$ give \begin{eqnarray} S(\nu,t_{obs}) = {C(t)\over 8\pi d^2 \nu_0(t)} D^3(\theta,t) {1-\beta(t)\cos\theta\over\beta(t)}. \end{eqnarray} \noindent Note that $S(\nu,t_{obs})$ is in a parametric form; given $t_{obs}$, one can determine the burster frame time $t$ for a given patch at $\theta$ using equation (5). Then, one determines $S(\nu,t_{obs})$ using equation (14) combined with equations (6,8,9,11), given $t$ and $\theta$. Meantime, $\nu(t_{obs})$ is related to $t$ and $\theta$ by equation (10). Thus, one can find $S(\nu,t_{obs})$ as a function of $\nu(t_{obs})$ at $t_{obs}$. Let us consider a few simple but relevant cases to illustrate the effect. \noindent Case 1: $\alpha=3/2$, $\psi=0$ and $\xi=0$. This case may have some bearing on such radiation features as atomic line features whose intrinsic frequencies are independent of $t$ (i.e., $\psi=0$). \noindent Case 2: $\alpha=3$, $\psi=3$ and $\xi=1$. This case may be related to the epoch where radiative cooling is efficient and intensity at the frequency in question is still rising, which could be relevant for radiation at frequencies lower than the peak of the synchrotron radiation spectrum (due to a truncated power-law electron distribution) at early times of a GRB event. \noindent Case 3: $\alpha=3$, $\psi=3$ and $\xi=-1$. This may be related to the epoch where radiative cooling is efficient and intensity at the frequency in question has started to decrease, which could be relevant for radiation at frequencies higher than the peak frequency of the spectrum at early times of a GRB event. \noindent Case 4: $\alpha=3/2$, $\psi=3$ and $\xi=1$, This case is similar to Case (2) with the primary difference that radiative cooling is unimportant here. This may be relevant for GRB afterglows such as the radio afterglows when electron cooling time is likely to be significantly longer than the dynamic time of the expanding fireball at frequencies lower than the peak frequency of the spectrum. \noindent Case 5: $\alpha=3/2$, $\psi=3$ and $\xi=-1$, This case is similar to Case 4 but for frequencies higher than the peak frequency of the spectrum. While it is convenient to express various quantities using $t$ as the independent time variable, one needs to express the final observables using $t_{obs}$, which is related to $t$ (for $\theta=0$; see equation 5) as \begin{equation} t_{obs} = {t_{dec}\over 2\Gamma_i^2}\left[1+{1\over 2\alpha+1} \left({t\over t_{dec}}\right)^{2\alpha +1}\right] \end{equation} \noindent for the simplified solution of $\Gamma(t)$ given by equation (6). Figure (1a) shows the flux density as a function of frequency at $t_{obs}=t_{dec}/\Gamma_i^2$ for the five cases. The frequency is normalized such that unity corresponds to the radiation from regions with $\theta=0$ and the flux density is normalized to be unity at unity frequency. The sharp turns to the left for cases (ii,iii,iv,v) correspond to the sharp turn of the evolution of the Lorentz factor at $t=t_{dec}$. Figure (1b) shows the flux density as a function of frequency at $t_{obs}=5000t_{dec}/\Gamma_i^2$ for the five cases. Also shown in both panels are two straight lines in the upper right corner indicating the spectral slope of $-0.75$ and $1.25$, respectively, which bracket the range of the spectral slope for various cases shown. Note that sharp turns as seen in (1a) are not visible simply because they appear at much lower intensity level than the displayed range in the figure. Note that $t_{dec}=16\hskip -0.1cm\left({E\over 10^{52}{\rm erg}}\right)^{1/3}\hskip -0.2cm \left({\Gamma_i\over 300}\right)^{-2/3}\hskip -0.2cm \left({n\over 1 {\rm cm}^{-3}}\right)^{-1/3}\hskip -0.1cm{\rm days}$. Therefore, for typical values of $E$, $\Gamma_i$ and $n$, $t_{obs}$ is of order a second and a day, respectively, after the fireball explosion for the two cases shown in (1a) and (1b). These two cases may respectively be relevant for bursts in gamma-ray and afterglows at lower energy bands. It should be noted that, although the external shock model is used to illustrate the smoothing magnitude, the results should be applicable to the internal shock model as well. Recall that in all cases a delta function spectrum at $\nu$ is assumed in the comoving frame at a given time. One sees that this delta function spectrum is smoothed out to appear as a broad spectrum with a half-width-half-maximum (HWHM) of $\sim (0.6-2.0)\nu$ for all the cases considered, except for Case 1 where HWHM is $\sim 0.1\nu$. An immediate implication from Figure 1 (see the dot-long-dashed curves on the upper right corner in the two panels) is that the observed spectra of GRBs or their afterglows should roughly have $\nu^{-1}$, if the electron distribution function power index $p$ is equal to greater than $3$. In other words, {\it the observed spectra cannot be steeper than $\nu^{-1}$ regardless the value of $p$}. This slope  It is of interest to understand where the different radiation that correspond to different frequencies in Figure 1 comes from. Figure 2 shows the frequency seen by {\it O} at a given time as a function of the emission shell radius from the burster, $r$, expressed in units of $c t_{dec}$. It is seen that, for realistic cases (ii,iii,iv,v) that correspond to the continuum radiation of the GRBs and afterglows, higher frequency radiation comes from earlier time $t$ (in the burster frame) with larger angle $\theta$ up to $t_{dec}$ after which there is a downturn to lower frequency at still earlier times due to assumed constancy of $\Gamma$ thus constancy of $\nu_0(t)$. For case (i) lower frequency radiation comes from earlier time $t$ with larger $\theta$. In both panels (a,b) also shown along the solid curves using solid dots are the corresponding $\theta$ values in degrees. Dotted and dashed curves are also punctuated by open circles and open squares, corresponding to the same $\theta$ values. \section{Discussion} It is shown that differentially varying Doppler boost of different patches of the fireball front provides an intrinsic, unavoidable smoothing mechanism for the spectra of gamma-ray bursts and their afterglows. The detailed smoothing patterns are complicated, depending upon various factors such as the evolution of the Lorentz factor and the evolution of the intrinsic (i.e., comoving frame) radiation spectrum. Nonetheless, for plausible ranges of model parameters of interest, a comoving frame delta function spectrum at $\nu$ is smoothed to have a HWHM of $\sim (0.6-2.0)\nu$, assuming that the time evolution of the characteristic frequency $\nu(t)$ is proportional to some positive power of the shock front Lorentz factor, $\Gamma^\psi$ (where $\psi>0$; see equation 8). This type of smoothing may be applicable to continuous spectra such as from synchrotron mechanism. In the case $\psi=0$ (appropriate for atomic line features which are independent of the blastwave dynamics), the spectral smoothing is smaller, with a HWHM of $\sim 0.1\nu$. Thus, a sharp linelike emission feature (in the comoving frame) would be smoothed out to have an equivalent width of about $0.1\nu$. Furthermore, the spectral profile of such a feature will be asymmetrical with a sharp cutoff at the high end (see the two solid curves in Figure 1). Two interesting and natural consequences arise due to the differential Doppler smoothing. First, the observed spectra of GRBs or their afterglows cannot be steeper than $\nu^{-\alpha}$ with $\alpha=0.75-1.25$ (Figure 1), even though the intrinsic spectra in the comoving shock frame may be much steeper, if the circumburster medium is uniform and electron and magnetic energies are fixed fractions of the total post-shock energy with time. Second, a generic fast-rise-slow-decay type temporal profile of the GRB bursts is expected, not necessarily reflecting the intrinsic temporal profiles of the bursts in the comoving frame. This can be easily seen by considering the case where the intrinsic (blastwave frame) spectrum is a bivariate delta function in both time and frequency. In this case the fast rise occurs when the radiation from the region around $\theta=0$ enters the observer's finite band. Subsequently, radiation from regions with gradually increasing $\theta$ is received at decreasing amplitudes in the same observer's finite band with a decaying time scale of $\sim t_{dec}/2\Gamma^2$ (in observer's frame). \acknowledgments The work is supported in part by grants AST9318185 and ASC9740300. I thank Bohdan Paczy\'nski for discussion.
\section{Introduction} The abundances of the galactic globular clusters (henceforth GCs) are an important datum for study of the chemical evolution and the halo of the Galaxy, among many other areas. The determination of these abundances rests on spectroscopic observations of the brightest globular cluster stars, the red giants at the tip of the giant branch. This subject has a very long history, with observations of the brightest globular cluster stars attempted as soon as was technically feasible, beginning with Helfer, Wallerstein \& Greenstein (1959) using the coud\'e of the Hale Telescope. In the late 1970s, the commissioning of a pair of identical echelle spectrographs attached to the newly completed $4-$m telescopes at KPNO and at CTIO made it possible to obtain high signal-to-noise ratio (SNR) spectra at suitably high spectral resolution for the brightest giants in the nearer GCs, and several groups undertook programs to provide the required abundance determinations (Cohen 1979, Pilachowski, Sneden \& Wallerstein 1982, and references therein). These programs produced excellent results for the more metal poor GCs, but faltered at the metallicity of M71 or 47 Tuc for reasons described in more detail below. As no very metal rich GC is close enough for the giants to be reachable with instrumentation available at that time on a 4-m telescope, there was no attempt to observe stars in any GC expected to be more metal rich than 47 Tuc. Cohen (1983) tried a few of the most metal rich GCs on the Palomar 5-m Hale telescope, but at somewhat lower spectral resolution than ideal, to get preliminary abundance values. Since there are roughly 150 galactic GCs, detailed abundance analyses will only be available at best for a limited number of bright red giants in the nearer GCs. Over the years, many schemes have been devised to extend the abundance determinations for individual red giants in these few GCs to the full sample of galactic GCs. Some approaches, for example those of Zinn \& West (1984) and of Armandroff \& Zinn (1988), use narrow band photometry or spectra of the integrated light of GCs, some (see Armandroff \& Da Costa 1991) use the very strongest lines in individual GC red giants measured from low dispersion spectra, while other abundance schemes rely on characteristics of the c-m diagram, such as the color of the giant branch (Frogel, Cohen \& Persson 1983) or the slope of the giant branch in the infrared (Kuchinski \& Frogel 1995). But the fundamental calibration of all these schemes rests on the high dispersion spectroscopic abundance determinations for a limited number of individual GC red giants. Independent of these is the $\Delta(S)$ method for RR Lyrae stars developed by Preston (1959), with a recent calibration from high dispersion abundance analyses of field RR Lyrae stars by Clementini {\it et al.\/}\ (1995). Unfortunately, there are no RR Lyrae stars in the most metal rich GCs, hence this is only useful for the metal poor GCs. Up to the present the calibration of all these schemes at the extremely metal rich end has been unsatisfactory, a situation we hope to remedy with the present work. NGC 6553 is the best GC to use as a calibrator for the very metal rich GCs. It is reasonably close, populous, very metal rich with an abundance that is believed to lie near the maximum achieved among the galactic GCs, and with a reddening which is small compared to that of most such clusters. The purpose of this paper is to present a detailed abundance analysis for stars, avoiding the red giants, in NGC 6553. \section {The Characteristics of NGC 6553} The first photometric study of NGC 6553 was that of Hartwick (1975). Ortolani, Barbuy \& Bica (1990) provide a modern ground based CCD study. Their key finding was that the giant branch in NGC 6553 as viewed in a color-magnitude diagram using a blue magnitude such as $V, B-I$ or $V, V-I$, is not monotonic, but forms an arc such that the reddest of the presumed red giants in NGC 6553 are fainter than the somewhat bluer red giants. This anomalous behavior is also seen in their later study of NGC 6528 (Ortolani, Barbuy \& Bica 1991) and in the galactic bulge field giants (Rich {\it et al.\/}\ 1998). It is less prominent in c-m diagrams involving redder colors, i.e. $I, V-I$. It is not seen at all in c-m diagrams involving only infrared wavelengths where the giant branch behaves as expected, with the stars becoming brighter as they become redder (Cohen \& Sleeper 1995, Kuchinski \& Frogel 1995). This anomalous behavior is ascribed to the extreme line/molecular band absorption found at blue wavelengths in such metal rich cool stars. Guarnieri {\it et al.\/}\ (1998) provide a definitive photometric study of NGC 6553 which combines $V$ and $I$ photometry from HST with ground based infrared photometry at $J$ and $K$. They determine a distance based on the magnitude of the horizontal branch (henceforth HB) stars for this GC of 5.2 kpc with a reddening of $E(B-V)$ = 0.7 mag, in good agreement with that determined by Zinn (1980) ($E(B-V)$ = 0.78 mag). Guarnieri {\it et al.\/}\ (1998) obtain the reddening by forcing the red giant branch of NGC 6553 to overlay that of 47 Tuc, and then adopting a correction for the difference in metallicity between the two GCs; to deduce this correction they adopt [Fe/H] = $-0.22 {\pm}0.05$ dex for NGC 6553. \footnote{We adopt the usual spectroscopic notation that [A/B] $\equiv$ log($N_A/N_B)_{star} - {\rm log}(N_A/N_B)$\mbox{$_{\odot}$}. Also, unless otherwise specified, metallicity is arbitrarily defined as the stellar [Fe/H] value.} Barbuy {\it et al.\/}\ (1992) compiled the abundance estimates for NGC 6553 over the period 1983 through 1992 from the literature. The values range from +0.5 to $-0.7$ dex, with none of them being particularly well calibrated or having adequate spectral dispersion to produce a highly accurate result. Barbuy {\it et al.\/}\ (1992) obtained a CCD echelle spectrum from ESO covering the spectral region 475 -- 580 nm with a spectral resolution of 0.20\AA\ and a SNR of 40 of a single red giant in NGC 6553 (star III--17 as identified by Hartwick 1975), which is among the brightest red giants in the cluster (at $V$). They attempted an abundance analysis for this star. The available photometry for this star could not be used to determine \teff\ as different colors gave inconsistent results. The line crowding in their spectrum was so severe that they were forced to use spectral synthesis techniques throughout. They obtained a preliminary result of [M/H] = $-0.2$ (+0.2,$-0.4$) dex for this red giant in NGC 6553. A detailed abundance analysis of two giants in NGC 6553 (stars II--85 and III--3, both slightly fainter than star III--17, but still among the brightest stars in the cluster at $V$) was carried out by Barbuy {\it et al.\/}\ (1999). This time the spectral region was shifted somewhat to the red, 500 -- 750 nm, and was covered at a spectral resolution of R = 20,000 with a SNR $\approx$50. The effective temperatures for the two red giant stars were derived from the photometry of Guarnieri {\it et al.\/}\ (1998) and are 4000 K for both stars with log(g) = 0.8. They obtained [Fe/H] = $-0.55 {\pm}0.2$ dex with [$\alpha$/Fe] $\approx +0.5$ dex. \section{The Choice of Stars for Our Sample} \subsection{Lessons From the Past} In the early 1980s, the initial surge of enthusiasm for abundance determination of galactic GCs with the new generation of telescopes and instrumentation led to detailed results that have withstood the test of time for metal poor GCs. But when efforts were made to extend this upward in metallicity to 47 Tuc and to M71, very controversial abundances were initially derived by both of the two groups involved that were considerably lower than anyone anticipated ($-$1.1 dex for 47 Tuc by Pilachowski, Canterna \& Wallerstein 1980 and Pilachowski, Sneden \& Wallerstein 1983, and $-$1.3 dex for M71 by Cohen 1980). Subsequently Cohen (1983) attempted to determine whether these low abundances for 47 Tuc and for M71 were real or resulted from some problem in the analysis. Cohen (1983) derived [Fe/H] = $-$0.7 dex for M71 using (for perhaps the first time for a GC star) a CCD spectrum and demonstrated that the abundances initially deduced for stars in M71 (and presumably 47 Tuc as well) were too low. This was ascribed to the difficulty of locating the proper continuum level in such cool high metallicity stars and the tendency towards underestimating the continuum in such heavily blended spectra. Gratton, Quarta \& Ortolani (1986) also obtained [Fe/H] = $-0.8$ for M71 and for 47 Tuc, observing stars considerably warmer (and fainter) than the tip of the red giant branch, demonstrating yet again that there were problems with the initial abundances for these clusters obtained from the photographic 4$-$m echelle spectra. Note that the red giants near the tip of the giant branch in M71 and in 47 Tuc have \teff $\sim 4100$ K with log(g) ${\approx}0.7$ dex, somewhat hotter than the two red giants in NGC 6553 analyzed by Barbuy {\it et al.\/}\ (1999). After removing the reddening, the giant branch of NGC 6553 is even redder than that of 47 Tuc or M71 in a color-magnitude diagram. Hence presumably the giants near the tip are even cooler and more metal rich than those of M71. Issues of line crowding and continuum determination must be taken very seriously. We must take every precaution to avoid stumbling over the same problem again. \subsection{Our Choice of Stars in NGC 6553} We are fortunate to have available the collecting power of the Keck Telescope and the efficient HIRES spectrograph (Vogt {\it et al.\/}\ 1994). This means that we can get reasonable spectral resolution and SNR in a one hour exposure for stars considerably fainter than the tip of the RGB in NGC 6553. Instead of using red giants, even giants substantially fainter than the tip of the giant branch, we have chosen stars on the red HB in NGC 6553 for our sample. There are a number of major advantages of such a choice. The first is that the red HB stars are considerably hotter than those near the top of the RGB, with \teff\ comparable to the luminous giants in more metal poor globular clusters. We can use the same set of absorption features as were used for the lower metallicity GCs and they will be of roughly comparable strength in a red HB star in NGC 6553 as in a RGB star in a lower metallicity GC. We will thus be more able to maintain consistency of results over the full range of metallicity of the galactic GCs. Moving into the near IR (630 - 870 nm with incomplete coverage due to gaps between the orders of HIRES) at high resolution (R = 34,000, corresponding to a 1.1 arcsc entrance slit) guarantees the minimum line crowding possible, while still avoiding excessive contamination by terrestrial atmospheric absorption and emission lines. We will see later that the resulting red HB star HIRES spectra, even in a GC as metal rich as NGC 6553, are sufficiently clean that we do not need spectral synthesis techniques and can instead determine abundances using individual absorption features as is normally done for the RG stars in the metal poor GCs. Furthermore with the powerful equipment of the Keck Telescope + HIRES we can achieve a SNR of ${\approx}70$ per 4 pixel resolution element with exposure times of under 1 hour for a HB star in NGC 6553 in that spectral region. Other advantages of working with red HB stars include a well defined mass (assumed to be 0.8 M\mbox{$_{\odot}$}) and luminosity so that once \teff\ is specified, the surface gravity is determined as well. Furthermore, the use of a well populated evolutionary state that occupies only a small area on a cluster c-m diagram such as the red HB maximizes the probability that a star in that location in the c-m diagram is a member of the GC and minimizes contamination by field stars, a non-trivial issue for a cluster with ($l^{II}, b^{II}$) = (5.3, $-3.0$). In addition, the non-LTE corrections are smaller for these hotter and higher surface gravity stars. With these points in mind, we chose the least crowded red HB candidates in NGC 6553 from Guarnieri {\it et al.\/}\ (1998) to form the spectroscopic sample. At the request of A. Renzini, two of the extremely red stars at the extreme end of the arc in the cm diagram for NGC 6553 found by Ortolani, Barbuy \& Bica (1990) were also observed in order to check that they are indeed members of the cluster. Since the spectra of these stars show very strong TiO bands, they were not used for the abundance analysis. Table 1 lists the five red HB stars in NGC 6553 observed spectroscopically during a night in the summer of 1995 and one in the summer of 1997, as well as the two extremely red giants. The identification numbers are those assigned by Guarnieri {\it et al.\/}\ (1998), from which the photometry is taken. Star 30297, one of the extremely red stars, is star II--3 of Hartwick (1975). The exposure times and dates of observation are also given. The heliocentric radial velocity for this cluster was not well determined, hence measurements from our spectra are also given in Table 1. We find that all of the stars observed, including the two extremely red stars, are probably members of NGC 6553. The seven heliocentric radial velocities indicate that the cluster mean velocity is 4 km~s$^{-1}$, with $\sigma = 7.1$ km~s$^{-1}$, in good agreement with the recent determinations of the mean cluster velocity from two stars of Barbuy {\it et al.\/}\ (1999) of 6 km~s$^{-1}$\ and (from more stars, but at lower dispersion) of Rutledge {\it et al.\/}\ (1997a) of 8.4 km~s$^{-1}$. The HB stars are all slow rotators; this will be discussed in B. Behr's thesis. There is one possible disadvantage of using horizontal branch stars. Since they are in an even later stage of evolution than RGB stars, one might worry about the effects of mixing, gravitational settling, etc. In the stable atmosphere of a hot star, diffusion effects -- both gravitational settling of He, and radiative levitation of Fe and other metals -- can significantly alter the photospheric composition, as predicted by Michaud, Vauclair \& Vauclair (1983), and observed by Heber (1987) and Glaspey {\it et al.\/}\ (1989). These effects, however, appear to be limited to \teff $\ge 10000$ K (Behr {\it et al.\/}\ 1999) and are thus not of concern for the RHB. Deep mixing on the RGB can also produce surface abundance anomalies in CNO (Kraft {\it et al.\/}\ 1998 and many prior reports), but these same variations will appear in any evolutionary stage beginning with the RGB and including any more advanced stage such as the HB. Such abundance anomalies are not expected to extend to the heavier elements from Ca to the Fe peak. There have been a small number of high dispersion abundance analyses of RR Lyrae stars (Clementini {\it et al.\/}\ 1995 for M4) and other HB stars in galactic GCs (Cohen \& McCarthy 1998 for M92) which demonstrate that, at least if one avoids the bluest HB stars, good consistency with abundances from giant branch stars in the same GC is obtained. \section{Determination of Atmospheric Parameters} The initial values of \teff\ to be used in the high dispersion abundance analysis are determined for each of the red HB stars from the observed $V-K$ colors of Guarnieri {\it et al.\/}\ (1998) using the latest version of the of model stellar atmospheres from Kurucz (1992) and the latest update of the original calibration of Cohen, Frogel \& Persson (1978) which is given by Gratton, Carretta \& Castelli (1997). As has been discussed extensively in the past, the $V-K$ color offers maximum leverage in assigning \teff. A reddening of $E(B-V) = 0.78$ mag corresponding to $E(V-K)$ = 2.12 mag is assumed; we use the reddening curve of Cohen {\it et al.\/}\ (1981). We further assume that the mass of a HB star in NGC 6553 is 0.8 M\mbox{$_{\odot}$}\ and that the luminosity of a star on the red horizontal branch in NGC 6553 is log(L/L\mbox{$_{\odot}$}) = 1.62. These then determine the surface gravity. The resulting atmospheric parameters are listed in Table 2. \section{Equivalent Width Measurements} The spectra were reduced using the suite of routines for analyzing echelle spectra written by McCarthy (1988) within the Figaro image processing package (Shortridge 1988). The stellar data are flat fielded with quartz lamp spectra, thereby removing most of the blaze profile, and the results are normalized to unity by fitting a 10th-order polynomial to line-free regions of the spectrum in each order. We are trying to measure equivalent widths for weak lines from spectra which are of adequate resolution ($\lambda/\Delta(\lambda)$ = 34,000) but of only marginally adequate SNR (70/4 pixel resolution element). Even in the near infrared there will be some adjacent weak lines contaminating any attempt to measure equivalent widths. We therefore adopt the following procedure to make the $W_{\lambda}$\ more stable to the presence of weak spectral features near the desired one. We smooth the spectra with a Gaussian whose FWHM is 0.3 \AA, which is slightly larger than the instrumental resolution near 8000 \AA. The definition of the continuum was then updated slightly as seemed appropriate. Equivalent widths were measured in the convolved spectra of HB stars 40071 and 40123 in NGC 6553 for about 40 lines spanning the full range in strength to be used and judged to be unblended. A linear relationship was established between $W_{\lambda}$\ and central depth of an absorption line, and this relationship was used to determine $W_{\lambda}$\ for the rest of the features used in the abundance analysis. This process was repeated for the three HB stars in NGC 6553 which were observed in 1995. The fit between $W_{\lambda}$\ and FWHM determined for the spectra from 1995 is very slightly different from that of the 1997 data, as might be expect due to differing instrumental and/or telescope focus. In total, $W_{\lambda}$\ measurements were obtained for $\sim$80 lines in each of the five HB stars in NGC 6553, including $\sim$40 Fe I lines. The $W_{\lambda}$\ we obtained for the five RHB stars in NGC 6553 are listed in Table 3. Figure 1 shows a montage of a section of the convolved spectra covering part of a single echelle order for the five NGC 6553 RHB stars and for also the convolved spectrum for the comparison star $\zeta$ Cyg. Our use of convolved spectra to measure $W_{\lambda}$\ and the fact that the stars involved have the low surface gravities characteristic of giants means that any small variations in the intrinsic line profile due, for example, to variations in thermal width or to a small amount of hyperfine structure, will not significantly affect our measurements of $W_{\lambda}$. Several checks have been made to verify that our procedure produces accurate measurements of $W_{\lambda}$. All of the five red HB stars in NGC 6553 have have very similar atmospheric parameters (see Table 2). We compare the measured $W_{\lambda}$\ for the pair of stars 30180 and 30242 and find $W_{\lambda}(30242) = 0.96\times W_{\lambda}(30180) -3.0$ m\AA\, with a rms scatter about this relationship of 14.2 m\AA. Attributing equal errors to both sets of $W_{\lambda}$\ suggests that typical errors in $W_{\lambda}$\ are $\pm$10 m\AA\ which arise largely from uncertainties in the location of the continuum. A second check can be made using spectra of much brighter comparison stars taken with the identical configuration of HIRES on the same night as the NGC 6553 spectra. These were chosen to have well known abundances and at least some of them are believed to be similar to the RHB stars in NGC 6553. The one we utilize in this paper is $\zeta$ Cyg\ (BS 8115). The spectrum of this star (which is also included in Figure 1) has the same resolution as the RHB stars in NGC 6553, but much higher SNR. Thus the set of $W_{\lambda}$\ could be measured on both the original spectrum and the convolved spectrum; the $W_{\lambda}$\ for the latter are given in the last column of Table 3 and in Figure 2 these two sets of $W_{\lambda}$\ are compared. The regression line between $W_{\lambda}$\ measured in the original and convolved spectra is (for 47 lines) $W_{\lambda}$(conv) = $(1.033 \pm0.026) \times$ $W_{\lambda}$(orig) $- (2.1 \pm8.1)$ m\AA, where the error on the constant term is the rms scatter of individual points along this regression line. If we consider the set of $W_{\lambda}$\ measured on the original spectra as ``perfect'', there is no clear systematic error in the set of $W_{\lambda}$\ measured in the convolved spectra, and typical errors are $\pm$8 m\AA. This estimate agrees quite well with the error of $\pm$10 m\AA\ estimated for the stars in NGC 6553 (where the SNR is lower). \section {Abundance Analysis} The guiding principle we have followed in the abundance analysis of the RHB stars in NGC 6553 is to adhere as closely as possible to the procedures used in earlier analyses of the less metal rich GCs by Carretta \& Gratton (1997) (henceforth CG) so as to make the results for the NGC 6553 sample as consistent and comparable as possible with the earlier analyses. This means we use the same set of model atmospheres (those of Kurucz 1992 with convective overshooting), the same relationship between color and \teff\ (that of Gratton, Carretta \& Castelli 1997), the same set of gf values whose sources are given in CG, the same line-by-line abundance code, etc. The solar abundances are those defined in CG, and are obtained in a manner as consistently as possible as are those of the program stars, using a solar model from Kurucz (1992), etc. The metal abundance for the model atmosphere used for each RHB star was determined by looking at the results of a first iteration for [Fe/H], then repeating the analysis with models whose abundance is close to that inferred from the first iteration. The microturbulent velocity was determined in the usual manner by forcing the slope of the $W_{\lambda}$\ versus deduced abundance to be zero. No hyperfine structure corrections were used; none are necessary except perhaps for Sc II. Clementini {\it et al.\/}\ (1995) demonstrate that the non-LTE corrections for HB stars are expected to be small. We assume LTE, except for oxygen. The O abundance is from the infrared triplet near 7770 \AA; a small non-LTE correction ($-0.07$ dex) has been applied for oxygen based on the calculations of Gratton {\it et al.\/}\ (1999). We now consider the validity of our initial atmospheric parameters as judged by the results of the abundance analysis. For three of the five RHB stars in NGC 6553, adoption of the initial parameters led no detectable trend of abundance versus excitation potential for the Fe I lines. For stars 40123 and 30257, this was not true; to achieve this condition, a small adjustment was made in their \teff, raising it by 150 to 200K. This change could easily arise if the reddening were slightly larger than the mean for these two RHB stars. An increase in $E(B-V)$ by 0.08 mag above our adopted mean value for NGC 6553 of 0.78 mag is required for star 40123, while a somewhat smaller increase is needed for star 30257. With these two small adjustments of \teff, the derived abundances are independent of the excitation potentials of the FeI lines and the ionization equilibrium for Fe is quite good (see Table 4). The resulting abundances deduced for the five RHB stars in NGC 6553 are listed in Table 4. First the deduced [Fe/H] value is given as determined from the ions Fe I and Fe II. Then the element ratios [A/Fe] are given as determined from various ions. These are calculated using Fe I, except for oxygen and singly ionized scandium, where Fe II is used. The final column gives the solar abundance we use. When there is more than one line of a given ion, the rms dispersion about the mean is given in parentheses. The mean abundance of each element together with the rms dispersion about the mean for the cluster NGC 6553 as determined from the analysis of the five red HB stars is given in Table 5. Figure 3 shows the deduced abundances for Fe I lines as a function of excitation potential in the five red HB stars in NGC 6553. The 1$\sigma$ rms uncertainty in the slope of the best linear fit is 0.026 dex/eV, implying an uncertainty in \teff\ of $\sim$100 K. \subsection{Verification using $\zeta$ Cyg} We observed several bright stars as possible comparison stars for this analysis of RHB stars in the extremely metal rich GC NGC 6553. Of those, $\zeta$ Cyg\ (BS 8115) is the one actually used as its atmospheric parameters are closest to those of the program stars. Just as for the GC stars, we obtain \teff\ for $\zeta$ Cyg\ from its colors, including the $V-K$ color, given by Johnson (1964) and by Neugebauer \& Leighton (1969) using the same calibration as for the RHB stars. This star is so nearby that interstellar reddening can be ignored. The Hipparcos parallax (ESA 1997) is 21.62 mas, implying $M_V = -0.13$ mag, somewhat brighter than the HB for old stellar systems. However, we do not know the mass of $\zeta$ Cyg, and hence must determine its surface gravity by forcing ionization equilibrium for Fe using the set of $W_{\lambda}$\ measured from the original non-convolved spectra, and checking that the implied mass is reasonable. The abundance for the model atmosphere was set in a manner similar to that used for the RHB stars, and an abundance analysis was carried out both for the set of $W_{\lambda}$\ measured from the original spectra and for those from the convolved spectra. The resulting abundances are listed in Table 6. The organization of Table 6 is similar to that of Table 4 described in detail above. $\zeta$ Cyg\ has been the subject of at least five relatively recent high dispersion abundance analyses according to the compilation of Cayrel de Strobel {\it et al.\/}\ (1997). The \teff\ used in these five abundance analyses range from 4890 to 4990 K and log(g) ranges from 2.0 to 2.9. The deduced [Fe/H] ranges from $-$0.17 to +0.10 dex, with typical errors quoted as $\pm$0.10 dex. The mean of the five abundance determinations for $\zeta$ Cyg\ from the literature is $-$0.04 dex. Our adopted \teff\ is in the middle of the range of these values, as is our log(g); our value of [Fe/H] from Fe I is +0.05 dex. This is in some sense the expected abundance for a disk star. Thus we feel that this comparison demonstrates the validity of our procedure for measuring $W_{\lambda}$\ and of our abundance analysis procedure. \subsection{Verification Using Spectral Synthesis} To demonstrate how well the observed spectra can be fit by spectral synthesis with the model atmosphere parameters we have derived and with the abundances obtained from our analysis, we display in Figure 4 a synthesis for the spectral region shown in Figure 1 for the five RHB stars in NGC 6553. In considering the different abundances shown ($-0.4, -0.2$ and 0.0 dex) all metals (except Li) were scaled appropriately and a Ca overabundance of a factor of two ([Ca/Fe] = +0.3 dex) was used throughout. Figure 5 shows the same for the comparison star $\zeta$ Cyg, except that here the spectral syntheses shown utilize the abundances $-0.2, 0.0$ and +0.2 dex. The convolved spectra are shown for the NGC 6553 stars while the original spectrum is shown for $\zeta$ Cyg. The comparison with the predictions of spectral synthesis roughly confirm an abundance of about [Fe/H] = $-0.2$ for NGC 6553 ([Fe/H] = $-0.4$ appears definitely too low, while a solar Fe abundance is clearly too high), in agreement with the abundance given in Table 4 from the line-by-line analysis. For $\zeta$ Cyg, a higher abundance is indicated, as we expect from the discussion above. \subsection{Discussion of Errors} The dominant source of errors is in the $W_{\lambda}$\ because of the limited SNR of our spectra of such relatively faint metal rich cool stars, which gives rise to an uncertainty in the continuum location. Also contributing are problems of cosmic ray hits and accurate subtraction of night sky emission features, which may perturb individual lines. Another important factor is the uncertainty in \teff\ due to possible reddening variations across NGC 6553. As mentioned earlier, a change in $E(B-V)$ from 0.78 mag to 0.86 mag will produce a increase in \teff\ of 200 K. Figure 1 of Hartwick (1975) demonstrates that the outer isophotes of NGC 6553 (i.e. the stellar distribution) are highly non-circular, presumably the result of non-uniform interstellar absorption. The NE quadrant in particular shows evidence for absorption that is higher than the mean. Cohen \& Sleeper (1995) present evidence for a significant dispersion of the interstellar absorption in several other highly reddened GCs by analyzing the dispersion in color of the red giant branch measured at a fixed luminosity for a variety of colors. While NGC 6553 is not in their sample, NGC 6528, with a comparable $E(B-V)$, is included. The full width in $V-K$ of the upper giant branch in this cluster is $\sim$0.6 mag. Ascribing this to reddening variations across the cluster implies a range in $E(B-V)$ of $\sim$0.2 mag. Ortolani, Barbuy \& Bica (1991) and Ortolani, Bica \& Barbuy (1991) also noted reddening variations across the face of this GC of $\Delta(E(B-V) \sim 0.3$ mag. Since these are full amplitudes, not rms dispersions, for the variation in reddening, the fact that two stars in NGC 6553 appear to have a somewhat larger reddening than the mean by about $\Delta[E(B-V)] \sim 0.08$ mag should not be surprising. However, it does indicate that considerably larger variations in $E(B-V)$ should not be seen within our relatively small sample. For that change in \teff\ of +200K, without changing any other parameters, the Fe ionization equilibrium changes by $-0.33$ dex, and [Fe/H] (from the more numerous FeI lines) increases by +0.10 dex. [O/Fe I] is quite sensitive to \teff, but when it is compared with the dominant ionization stage Fe II, which unfortunately has far fewer measurable lines in this spectral region for these stars than does Fe I, the O abundance is relatively insensitive to a 200 K change in \teff. However, in practice, part of any increase in \teff\ will be taken up by adjustments of other parameters, particularly the microturbulent velocity. This is because there is a (loose) correlation between line strength and excitation in our list of measured spectral lines, with the lowest excitation lines in general being stronger. For this reason the impact of a change of \teff\ on the derived abundances is much less than the value quoted above. We verified this by repeating our analysis with \teff\ lower by 200 K, and adjusting $v_t$ again to avoid trends of derived abundances with $W_{\lambda}$. The average abundance we found for NGC 6553 was again [Fe/H] = $-0.16$. Furthermore, the ionization equilibria we have obtained just with our initial choice of model atmosphere parameters for three of the five red HB stars in NGC 6553 are extremely good. Model atmospheres whose detailed abundances element-by-element are identical to the results we have found for NGC 6553 are not available. We have checked that the error introduced by using models with the solar abundance ratios is small, less than 0.1 dex for all ions. In summary, we feel that an uncertainty of ${\pm}0.1$ dex is reasonable for [Fe/H] and ${\pm}0.15$ dex for the element-to-element abundance ratios in each of the five RHB stars in NGC 6553, with the caveat that the O abundance be compared to that of the dominant singly-ionized ionization stage of Fe. \subsection{Element Ratios} In our discussion to this point we have concentrated on Fe I as it is the only ion for which our red spectra contain large numbers of measured lines. For all other ions the number of measured lines is small. Typical errors in abundances from individual lines are $\pm0.2$ dex, and are dominated by uncertainties in the location of the local continuum and by the presence of blends. However, these error sources may represent systematic errors that behave in a similar way for a given line in different stars, so that the star-to-star scatter likely underestimates the real error bars. While the O abundance is derived from the IR triplet at 7770\AA\, whose lines are clean (i.e. not blended at this spectral resolution) and are a well known abundance indicator, the lines are of very high excitation from the dominant ion. Thus, as discussed above, O/Fe ratios are very sensitive to the adopted temperature and gravity when O abundances are compared to that given by Fe I lines. O/Fe ratios are much more robust when Fe II lines are used; unfortunately, our Fe II abundances are only based on one or two lines. Oxygen is clearly overabundant in NGC 6553 stars, with an abundance ratio similar to that measured in other globular clusters and metal-poor stars (Gratton \& Ortolani 1986, 1989; Barbuy 1988). Our O abundance for $\zeta$ Cyg\ is also quite high; however, this star is known to be a mild Ba star (Sneden, Lambert \& Pilachowski 1981), and its surface abundances may have been modified by accretion of material from an evolved, originally more massive, companion. Mg abundances are derived from the line at 8718\AA\ that appears clean both in the Sun and in all our spectra. Parameters for this line were taken from Lambert \& Luck (1978) solar analysis, and give a solar Mg abundance in good agreement with that given by Anders \& Grevesse (1989). Mg is then clearly overabundant in NGC 6553 stars. Si abundances are obtained from a few (1--4) lines; these are clean solar lines used in the analysis by Lambert \& Luck (1978). Results given by different lines are generally internally consistent with each other for each star. However, Si abundances are quite sensitive to the adopted effective temperatures: if \teff\ is lowered by 100 K (i.e. within uncertainties of the present analysis), [Si/Fe] increases by 0.12 dex. This sensitivity might explain the rather large star-to-star scatter. We conclude that our Si abundance for NGC 6553 may have an error bar as large as $\pm 0.2$ dex. Our Ca abundances are much more reliable than those for Si, since they are derived from at least 5 lines in each star giving very consistent results. Highly accurate laboratory oscillator strengths are used for the Ca lines (Smith \& Raggett 1981). Also, the sensitivity of the Ca abundances to atmospheric parameters is quite similar to that of Fe. Ca is then clearly overabundant in NGC 6553: the average value of [Ca/Fe]=$+0.26\pm 0.1$\ is very similar to that obtained by Gratton \& Sneden (1991) for both field metal-poor stars ([Ca/Fe]=$+0.29\pm 0.06$) and globular clusters ([Ca/Fe]=$+0.33\pm 0.15$). Finally, Ti abundances are obtained from only two Ti I lines, with solar oscillator strengths. However, they give very consistent star-to-star results, and the sensitivity to atmospheric parameters of the [Ti/Fe] ratios is not large (although a little larger than for Ca). On the whole our [Ti/Fe] ratio should have an error bar of about $\pm 0.15$~dex. Again, our average [Ti/Fe] ratio for NGC 6553 ([Ti/Fe]=$+0.19\pm 0.15$) agrees well with the results for both field metal-poor stars ([Ti/Fe]=$+0.28\pm 0.11$) and globular clusters ([Ti/Fe]=$+0.21\pm 0.16$) obtained by Gratton \& Sneden (1991). Summarizing, our analysis consistently gives overabundances for all the classical $\alpha-$elements in NGC 6553: for those elements having the most reliable results (O, Mg, Ti, and Ca), the abundance ratios are very similar to those obtained in the analysis of field metal-poor stars and other globular clusters. The error bar for Si is larger, so that we do not attribute much weight to the lower value of the overabundance we find for this element. (It still agrees with the value for field stars of [Si/Fe]=$+0.30\pm 0.08$\ obtained by Gratton \& Sneden 1991.) Our abundances agree also quite well with the pattern observed by McWilliam \& Rich (1994) for field giants in the Baade's Window (although these authors do not find an O overabundance). The most common interpretation of the overabundance of O and $\alpha-$elements in metal-poor stars is that only massive stars exploding as type II SN contributed initially to nucleosynthesis. This seems to be the case also for NGC 6553, pointing toward a fast formation of at least part of the galactic bulge. \section{Discussion of Results} We have carried out a detailed abundance analysis for five RHB stars in the globular cluster NGC 6553. We obtain a mean [Fe/H] = $-$0.16 dex, with the total range of the values for the five stars being only 0.23 dex. The relative abundance for the best determined $\alpha$ process element [Ca/Fe] is +0.26 dex. Because we have made a major effort to carry out this abundance analysis in a manner consistent with that adopted by Carretta \& Gratton (1997), our results should be directly comparable with theirs. Their work represents the largest collection of galactic globular cluster abundances based on a uniform analysis using both their own spectra and reanalyzing $W_{\lambda}$\ from high dispersion analyses in the literature. We have thus extended their abundance calibration up to the highest values of metallicity found among the galactic GCs. Zinn \& West (1984) assigned [Fe/H] = $-0.29 {\pm}0.11$ dex to NGC 6553. It is the GC with the highest metallicity among their calibrating clusters (Table 5 of their paper). The highest metallicity they assign to any galactic GC is +0.24 dex to Terzan 1 and to Terzan 5, both heavily obscured clusters in the galactic bulge. Their abundance for NGC 6553 is remarkably close (0.13 dex smaller) to the results of our abundance analysis. It is equal to the systematic difference of +0.12 ${\pm}$0.01 dex found by CG between the results of their abundance analysis and those of previous investigators using the same set of $W_{\lambda}$. CG compare their uniform well calibrated scale for a limited number of GCs with that of Zinn \& West (1984), which is the metallicity scale most commonly used for the family of galactic GCs, over the full metallicity range spanned by each. They find a curvature such that at both low and high metallicities (``high'' here is M71, 47 Tuc and NGC 6352), the Zinn \& West scale appears to be somewhat high (by about 0.1 dex). The point we have added for NGC 6553 at the extremely metal rich end appears to indicate that the comparison of the two metallicity scales for galactic GCs is somewhat more complex. Rutledge {\it et al.\/}\ (1997b) have introduced the parameter $W(CaII)$ which is determined from observations of the strength of the infrared Ca triplet in individual giants from low resolution spectra as the basis for a new metallicity ranking scheme for the family of galactic globular clusters. As is shown in Figure 6, with the addition of the point for NGC 6553, the relationship of the CG abundance scale with this parameter appears to be non-linear. A detailed abundance analysis of a second extremely high metallicity GC seems desirable to confirm this. The average metallicity for the five RHB stars in NGC 6553 is slightly higher than the average value of approximately $-0.2$ dex found for giants in Baade's Window in the galactic bulge by Rich (1988) as recalibrated by McWilliam \& Rich (1994) (see also Castro {\it et al.\/}\ 1995) and by Sadler, Rich \& Terndrop (1996). Castro {\it et al.\/}\ (1996) establish that the most metal rich star known in Baade's window from this sample and the most metal rich star known in the solar neighborhood ($\mu$ Leo) both have [Fe/H] = +0.45 dex, so there may well be galactic GCs that are even more metal rich than NGC 6553 as suggested by the Zinn \& West compilation. \subsection {Comparison with Barbuy {\it et al.\/}\ (1999)} We have found an abundance for NGC 6553 which is considerably higher than that of Barbuy {\it et al.\/}\ (1999). When comparing our abundances with those of Barbuy {\it et al.\/}, we should recall that there are differences in procedure which we believe will produce a systematic difference in the derived abundances. In particular these lie in the choice of model atmospheres and in the calculation of the solar abundances. Barbuy {\it et al.\/}\ adopt the model atmospheres of Plez {\it et al.\/}\ (1992) for their two giants in NGC 6553 while we use those of Kurucz (1992). They adopt the solar model of Holweger \& Muller (1974). For these reasons, our Fe abundances are expected to be 0.1 to 0.15 dex higher. This explains part of the difference between our result and that by Barbuy {\it et al.\/}, but not all. Some perhaps must be ascribed to our use of hotter stars and the higher SNR and dispersion of our spectra, leading to better continuum definition. \section{Summary} We provide a high dispersion line-by-line abundance analysis of five red HB stars in the extremely metal rich galactic globular cluster NGC 6553. In such a metal rich cluster, these red HB stars are significantly hotter than the red giants near the tip of the giant branch, and hence their spectra will be much less crowded. Since accurate location of the continuum is the critical key to success for abundance analyses of such metal rich objects, our approach offers the potential for a more reliable abundance determination. We find that the mean [Fe/H] for NGC 6553 is $-0.16$ dex, comparable to the mean abundance in the galactic bulge found McWilliam \& Rich (1994) and considerably higher than that obtained from an analysis of two red giants in this cluster by Barbuy {\it et al.\/}\ (1999). The relative abundance for the best determined $\alpha$ process element indicates an excess of $\alpha$ process elements of a factor of two as [Ca/Fe] = +0.26 dex. The metallicity of NGC 6553 reaches the average of the Galactic bulge and of the solar neighborhood. It is likely that there are even more metal rich galactic globular clusters that have solar metallicity. Our analysis of the abundance of NGC 6553 provides an important calibration point for determining the metallicity of the more distant/more heavily reddened extremely metal rich globular clusters found exclusively in the nuclear bulge. It also provides an important clue regarding the mean abundance of the stellar population in the galactic bulge and in luminous elliptical galaxies in general. \acknowledgements We are grateful to Maria Donata Guarnieri and her collaborators for providing access to their HST and IRAC photometry of NGC 6553 prior to publication and for providing finding charts for candidate HB stars from their images and to Jim McCarthy for help during the 1997 observing run. We also thank Beatriz Barbuy and Sergio Ortolani for having provided a copy of their recent paper in advance of publication. The entire Keck/HIRES user community owes a huge debt to Jerry Nelson, Gerry Smith, Steve Vogt, and many other people who have worked to make the Keck Telescope and HIRES a reality and to operate and maintain the Keck Observatory. We are grateful to the W. M. Keck Foundation, and particularly its late president, Howard Keck, for the vision to fund the construction of the W. M. Keck Observatory. This research has made use of the SIMBAD data base, operated at CDS, Strasbourg, France. \clearpage \begin{deluxetable}{lrrrrrrr} \tablenum{1} \tablewidth{0pt} \scriptsize \tablecaption{The Sample of Stars Observed in NGC 6553} \label{tab1} \tablehead{\colhead{ID} & \colhead{$V$\tablenotemark{a}} & \colhead{$I$\tablenotemark{a}} & \colhead{$J$\tablenotemark{a}} & \colhead{$K$\tablenotemark{a}} & \colhead{$v_r$} & \colhead{Date of Obs.} & \colhead{Exp. Time} \nl \colhead{} & \colhead{(mag)} & \colhead{(mag)} & \colhead{(mag)} & \colhead{(mag)} & \colhead{(km~s$^{-1}$)} & \colhead{} & \colhead{(sec)} } \startdata Red HB Stars \nl 30180 & 16.85 & 14.86 & 13.39 & 12.46 & $-$4.9 & 950805 & 3000 \nl 30242 & 16.90 & 14.97 & 13.38 & 12.42 & +2.2 & 950805 & 3000 \nl 30257 & 16.88 & 14.98 & 13.32 & 12.38 & +1.6 & 950805 & 3000 \nl 40071 & 16.77 & 14.82 & 13.38 & 12.40 & $-$4.5 & 970803 & 3000 \nl 40123 & 16.94 & 14.95 & 13.43 & 12.46 & +12.6 & 970803 & 3000 \nl Ext. Red Giants \nl 30291 & 16.92 & 12.15 & 8.90 & 7.17 & +9.1 & 950806 & 2000 \nl 30297& 16.62 & 12.04 & 8.81 & 7.31 & +10.3 & 950806 & 2000 \nl Comparison Star \nl $\zeta$ Cyg (BS 8115) & 3.19 & & 1.68 & 1.07 & +14.4 & 950805 & 2 \nl \enddata \tablenotetext{a}{From Guarnieri {\it et al.\/}\ (1998), except for $\zeta$ Cyg, whose photometry is from Johnson (1964) and with $K$ from Neugebauer \& Leighton (1969).} \end{deluxetable} \begin{deluxetable}{lrrr} \tablenum{2} \tablewidth{0pt} \scriptsize \tablecaption{Adopted Model Atmosphere Parameters} \label{tab2} \tablehead{\colhead{Star ID} & \colhead{\teff} & \colhead{log(g)} & \colhead{$v_t$} \nl \colhead{} & \colhead{(K)} & \colhead{(dex)} & \colhead{(km~s$^{-1}$)} } \startdata Red HB Stars \nl 30180 & 4700 & 2.3 & 1.8 \nl 30242 & 4630 & 2.3 & 1.8 \nl 30257 & 4750\tablenotemark{a} & 2.3 & 1.6 \nl 40071 & 4725 & 2.3 & 2.5 \nl 40123 & 4830\tablenotemark{b} & 2.3 & 1.4 \nl Comparison Star \nl $\zeta$ Cyg (BS 8115) & 4950 & 2.7 & 1.6 \nl \enddata \tablenotetext{a}{Initial \teff = 4600 K.} \tablenotetext{b}{Initial \teff = 4625 K.} \end{deluxetable} \begin{deluxetable}{lrrrrrrrrrrrrrrrrrr} \tablenum{3} \tablewidth{0pt} \tabcolsep 2pt \scriptsize \tablecaption{Equivalent Widths For 5 Red HB Stars in NGC 6553 and for $\zeta$ Cyg\tablenotemark{a}} \label{tab3} \tablehead{ \colhead{Ion} & \colhead{~~~~$\lambda$ (\AA)} &\colhead{$\chi$ (eV)} &$\log(gf)$ &\mc{400071} &\mc{40123} &\mc{30180} &\mc{30242} &\mc{30257} &\mc{$\zeta$ Cyg} \nl & & & & \mc{(m\AA)} & \mc{(m\AA)} & \mc{(m\AA)} & \mc{(m\AA)} & \mc{(m\AA)} & \mc{(m\AA)} } \startdata O I &7771.95 &9.11 &0.33 &&56.2 && &&30.2 &&39.7 &&56.8 &&49.0 \nl O I &7774.18 &9.11 &0.19 &&53.3 &&33.3 &&42.7 &&44.2 && &&55.9 \nl O I &7775.40 &9.11 &$-0.03$ &&45.9 &&31.7 &&39.7 &&42.5 &&43.3 &&47.8 \nl Mg I &8717.80 &5.93 &$-1.09$ &&122.4 &&86.9 &&94.9 &&101.0 && &&97.9 \nl Si I &6848.57 &5.86 &$-1.75$ &&24.5 &&20.6 && &&24.1 &&34.8 &&38.7 \nl Si I &7034.90 &5.87 &$-0.88$ &&81.2 &&63.8 &&90.3 && &&60.4 && \nl Si I &7405.79 &5.61 &$-0.82$ &&114.6 && && && && && \nl Si I &7932.40 &5.96 &$-0.47$ &&74.8 &&72.8 && &&92.4 &&80.0 &&95.7 \nl Ca I &6455.60 &2.52 &$-1.29$ && &&108.6 &&102.8 && && && \nl Ca I &6462.57 &2.52 &0.26 &&334.5 &&230.6 &&315.4 &&311.1 &&311.0 && \nl Ca I &6471.67 &2.52 &$-0.69$ &&173.9 &&133.0 &&165.2 &&142.3 &&161.0 && \nl Ca I &6493.79 &2.52 &$-0.11$ &&251.8 && &&199.0 &&176.9 &&170.4 &&182.8 \nl Ca I &6499.65 &2.52 &$-0.82$ &&158.2 &&116.0 &&137.3 &&122.9 &&134.2 &&131.9 \nl Ca I &6572.80 &0.00 &$-4.32$ &&186.1 &&108.5 &&131.9 &&129.7 &&137.6 &&105.9 \nl Sc II &6604.60 &1.36 &$-1.14$ &&91.4 &&74.5 &&89.6 &&91.3 &&68.4 &&90.5 \nl Ti I &6508.12 &1.43 &$-2.05$ &&41.7 &&41.3 &&42.5 && &&42.7 && \nl Ti I &6743.13 &0.90 &$-1.63$ &&126.9 &&76.3 &&78.5 &&112.7 && &&69.7 \nl Cr I &6979.80 &3.46 &$-0.22$ &&111.2 && &&90.6 && && && \nl Cr I &7400.19 &2.90 &$-0.11$ &&156.3 &&105.1 &&147.0 &&137.4 &&132.8 &&131.7 \nl Fe I &6380.75 &4.19 &$-1.34$ &&98.2 &&80.8 &&91.3 &&86.0 &&97.4 &&100.8 \nl Fe I &6392.54 &2.28 &$-3.97$ &&93.3 &&51.3 && && && && \nl Fe I &6393.61 &2.43 &$-1.43$ &&271.4 &&167.0 &&218.1 &&235.4 &&223.7 &&208.5 \nl Fe I &6400.32 &3.60 &$-0.23$ &&345.9 && && && && && \nl Fe I &6411.66 &3.65 &$-0.60$ &&236.4 &&125.2 &&156.8 && &&161.3 &&168.2 \nl Fe I &6421.36 &2.28 &$-1.98$ &&216.4 &&162.0 &&222.1 &&188.2 &&204.9 && \nl Fe I &6481.88 &2.28 &$-2.94$ &&158.7 &&106.3 &&142.8 &&124.4 &&124.9 &&132.5 \nl Fe I &6498.95 &0.96 &$-4.66$ &&153.2 &&113.5 &&143.7 &&124.5 &&121.1 &&134.5 \nl Fe I &6518.37 &2.83 &$-2.56$ &&133.6 &&95.0 && && && && \nl Fe I &6533.94 &4.56 &$-1.28$ &&92.8 && && && && && \nl Fe I &6574.25 &0.99 &$-4.96$ &&142.6 &&100.0 &&122.6 &&121.9 &&117.2 && \nl Fe I &6581.22 &1.48 &$-4.68$ &&133.0 &&81.8 &&93.6 &&108.5 &&102.5 &&102.7 \nl Fe I &6593.88 &2.43 &$-2.30$ &&183.8 &&146.0 &&170.9 &&170.9 &&163.7 &&157.0 \nl Fe I &6608.04 &2.28 &$-3.96$ &&92.5 &&56.5 &&73.8 &&55.9 &&63.8 && \nl Fe I &6609.12 &2.56 &$-2.65$ &&161.5 &&121.4 &&137.5 &&136.4 &&122.9 && \nl Fe I &6625.04 &1.01 &$-5.32$ && && && && &&114.3 && \nl Fe I &6627.56 &4.55 &$-1.50$ &&69.3 &&65.2 && &&52.0 &&63.7 &&64.7 \nl Fe I &6703.58 &2.76 &$-3.00$ &&128.2 &&88.7 &&102.4 &&92.2 &&97.4 && \nl Fe I &6713.74 &4.79 &$-1.41$ && &&32.2 && &&47.4 && && \nl Fe I &6725.36 &4.10 &$-2.21$ &&43.3 && &&48.8 && &&43.9 && \nl Fe I &6726.67 &4.61 &$-1.05$ &&76.3 &&59.0 &&63.3 && &&77.1 &&86.0 \nl Fe I &6733.15 &4.64 &$-1.44$ &&52.2 &&42.9 &&57.0 &&46.5 &&61.8 &&57.8 \nl Fe I &6750.16 &2.42 &$-2.58$ && && &&155.0 &&147.7 &&145.6 &&128.1 \nl Fe I &6786.86 &4.19 &$-1.90$ &&66.0 && &&64.9 && && && \nl Fe I &6839.84 &2.56 &$-3.35$ &&96.6 &&83.3 &&100.4 && &&102.3 && \nl Fe I &6843.66 &4.55 &$-0.86$ &&109.9 &&84.3 &&111.3 &&74.9 &&103.1 &&98.9 \nl Fe I &6857.25 &4.07 &$-2.07$ &&51.6 &&36.0 &&54.6 &&31.1 && &&51.7 \nl Fe I &6858.16 &4.61 &$-0.95$ &&86.7 &&60.2 &&76.5 &&69.1 &&82.5 &&92.3 \nl Fe I &6898.29 &4.22 &$-2.08$ &&45.5 && && &&43.7 && &&48.5 \nl Fe I &6916.69 &4.15 &$-1.35$ &&126.9 &&88.0 && && && &&109.7 \nl Fe I &6988.53 &2.40 &$-3.42$ &&126.6 &&106.0 && &&107.3 &&115.1 &&106.2 \nl Fe I &7007.97 &4.18 &$-1.80$ && && && && && &&78.6 \nl Fe I &7022.96 &4.19 &$-1.11$ &&121.0 && && &&122.9 &&120.9 && \nl Fe I &7130.93 &4.22 &$-0.76$ &&146.5 &&142.2 && && && && \nl Fe I &7132.99 &4.07 &$-1.66$ &&83.9 &&63.6 && && && && \nl Fe I &7142.52 &4.95 &$-0.93$ &&99.7 &&74.3 && &&79.9 &&80.7 && \nl Fe I &7151.47 &2.48 &$-3.58$ &&103.2 && && &&107.5 && && \nl Fe I &7180.00 &1.48 &$-4.71$ &&147.7 && &&96.5 &&118.1 &&111.9 && \nl Fe I &7189.16 &3.07 &$-2.77$ &&96.5 &&64.0 && && && && \nl Fe I &7190.13 &3.11 &$-3.28$ &&94.4 && &&50.1 &&49.3 && && \nl Fe I &7306.57 &4.18 &$-1.55$ &&63.9 &&48.9 &&70.2 &&63.2 &&59.5 &&69.0 \nl Fe I &7401.69 &4.19 &$-1.60$ &&94.4 && && && && && \nl Fe I &7411.16 &4.28 &$-0.48$ && &&134.9 &&161.0 &&143.6 &&172.8 &&168.3 \nl Fe I &7418.67 &4.14 &$-1.44$ &&121.8 &&74.3 &&91.6 &&86.4 &&97.3 &&94.3 \nl Fe I &7421.56 &4.64 &$-1.69$ &&30.2 && && &&27.2 && &&33.4 \nl Fe I &7447.40 &4.95 &$-0.95$ &&64.8 && && && && && \nl Fe I &7461.53 &2.56 &$-3.45$ &&127.4 &&82.2 &&97.4 &&90.0 &&82.8 && \nl Fe I &7491.66 &4.30 &$-1.01$ &&106.0 &&78.5 &&113.4 &&91.3 &&100.5 &&82.1 \nl Fe I &7568.91 &4.28 &$-0.90$ &&138.8 &&116.5 && && &&137.4 && \nl Fe I &7583.80 &3.02 &$-1.93$ &&151.3 &&124.6 && &&169.9 &&151.5 && \nl Fe I &7719.05 &5.03 &$-0.96$ &&71.6 &&51.3 &&66.7 && &&57.7 &&54.6 \nl Fe I &7723.21 &2.28 &$-3.62$ && &&72.5 &&123.0 && &&98.8 && \nl Fe I &7751.11 &4.99 &$-0.74$ &&104.2 &&71.1 &&81.4 &&84.2 && &&87.2 \nl Fe I &7807.91 &4.99 &$-0.51$ &&104.5 && &&96.1 &&94.8 &&89.8 &&108.6 \nl Fe I &7912.87 &0.86 &$-4.85$ &&168.1 && && && && && \nl Fe I &7941.10 &3.27 &$-2.29$ &&109.8 &&86.0 && &&111.7 &&107.7 &&82.1 \nl Fe II &6416.93 &3.89 &$-2.70$ &&55.0 &&45.9 &&52.5 &&52.7 &&50.0 &&53.9 \nl Fe II &6456.39 &3.90 &$-2.10$ && && &&80.8 && && && \nl Fe II &7449.34 &3.89 &$-3.10$ &&50.9 && && && && && \nl Ni I &6378.26 &4.15 &$-0.82$ &&45.9 &&61.5 &&63.5 &&51.8 && && \nl Ni I &6384.67 &4.15 &$-1.00$ && &&41.0 && &&45.9 && && \nl Ni I &6482.81 &1.93 &$-2.78$ &&145.1 &&95.6 &&120.4 &&121.8 &&105.0 &&104.6 \nl Ni I &6532.88 &1.93 &$-3.42$ &&94.3 && && &&73.2 && && \nl Ni I &6586.32 &1.95 &$-2.78$ &&136.4 &&77.8 &&105.4 &&110.0 &&93.3 &&97.3 \nl Ni I &6635.14 &4.42 &$-0.75$ &&58.4 &&49.9 && &&45.0 && && \nl Ni I &6767.78 &1.83 &$-2.06$ &&161.6 &&128.1 &&165.1 &&142.2 &&151.9 &&132.7 \nl Ni I &6772.32 &3.66 &$-0.96$ &&81.4 &&70.9 &&99.4 &&77.6 &&80.9 &&90.5 \nl Ni I &7030.02 &3.54 &$-1.70$ &&63.0 &&46.4 && &&54.1 &&47.6 &&53.9 \nl Ni I &7110.91 &1.93 &$-2.91$ &&143.4 &&102.0 && &&113.7 && && \nl Ni I &7422.29 &3.63 &$-0.29$ &&153.1 &&127.0 &&155.2 &&149.0 &&147.8 &&147.3 \nl Ni I &7555.61 &3.85 &$-0.12$ &&174.0 &&125.7 &&174.9 &&148.1 &&142.7 &&141.9 \nl Ni I &7574.05 &3.83 &$-0.61$ && &&92.0 && && && && \nl Ni I &7715.58 &3.70 &$-0.98$ &&86.4 && && && && && \nl Ni I &7727.62 &3.68 &$-0.30$ &&164.7 &&128.2 &&143.7 &&125.7 &&148.5 &&133.8 \nl Ni I &7797.59 &3.30 &$-0.82$ &&155.4 &&103.5 &&141.3 &&101.3 &&129.3 &&112.4 \nl \enddata \tablenotetext{a}{As described in the text, all $W_{\lambda}$ are measured from convolved spectra.} \end{deluxetable} \begin{deluxetable}{llcllcllcllcllcllc} \tablenum{4} \tablewidth{0pt} \tabcolsep 3pt \scriptsize \tablecaption{Abundances for Five Red HB Stars In NGC 6553} \label{tab4} \tablehead{\colhead{\bf Ion} && \multicolumn{2}{c}{\bf Star 40071} && \multicolumn{2}{c}{\bf Star 40123} && \multicolumn{2}{c}{\bf Star 30180} && \multicolumn{2}{c}{\bf Star 30242} && \multicolumn{2}{c}{\bf Star 30257} && \colhead{\bf Sun} \nl \colhead{} && \colhead{\# of} & \colhead{abundance} && \colhead{\# of} & \colhead{abundance} && \colhead{\# of} & \colhead{abundance} && \colhead{\# of} & \colhead{abundance} && \colhead{\# of} & \colhead{abundance} && \colhead{abund.} \nl \colhead{} && \colhead{lines} & \colhead{(dex)} && \colhead{lines} & \colhead{(dex)} && \colhead{lines} & \colhead{(dex)} && \colhead{lines} & \colhead{(dex)} && \colhead{lines} & \colhead{(dex)} && \colhead{(dex)} } \startdata [Fe/H] \nl Fe I &&50 & $-$0.18 (0.20) &&39 & $-$0.18 (0.15) &&32 & $-$0.14 (0.32) &&35 & $-$0.26 (0.22) &&36 & $-$0.03 (0.18) && $-$4.50 \nl Fe II &&2 & $-$0.11 (0.22) && 1 & $-$0.33 && 2 & $-$0.18 (0.03) && 1 & $-$0.07 &&1 & $-$0.20 && $-$4.56 \nl & \nl [A/Fe] \nl O I\tablenotemark{a} &&3 & +0.53 (0.07) &&2 & +0.34 (0.13) &&3 & +0.41 (0.32) &&3 & +0.56 (0.22) &&2 & +0.68 (0.03) && +1.18 \nl Mg I &&1 & +0.50 &&1 & +0.34 &&1 & +0.31 &&1 & +0.48 && & && +0.04 \nl Si I &&4 & +0.07 (0.22) &&3 & $-$0.01 (0.13) &&1 & +0.39 &&2 & +0.27 (0.05) &&3 & $-$0.03 (0.24) && +0.04 \nl Ca I &&5 & +0.36 (0.26) &&5 & +0.33 (0.12) &&6 & +0.23 (0.19) &&5 & +0.13 (0.12) &&5 & +0.26 (0.24) && $-$1.34 \nl Sc II\tablenotemark{a} &&1 & $-$0.16 &&1 & $-$0.13 &&1 & $-$0.04 &&1 & +0.11 &&1 & $-$0.38 && $-$4.48 \nl Ti I &&2 & +0.18 (0.12) &&2 & +0.23 (0.3)7 &&1 & +0.24 &&1 & +0.10 &&1 & +0.19 && $-$2.52 \nl Cr I &&2 & +0.10 (0.22) &&1 & $-$0.09 &&2 & +0.14 (0.01) &&1 & +0.04 &&1 & +0.01 && $-$1.88 \nl Ni I &&14 & $-$0.04 (0.23) &&14 & +0.02 (0.16) &&9 & +0.11 (0.17) &&14 & +0.02 (0.21) &&9 & $-$0.06 (0.17) && $-$1.25 \nl \enddata \tablenotetext{a}{O I and Sc II are calculated with respect to Fe II, all other ions are with respect to Fe I.} \end{deluxetable} \begin{deluxetable}{lclrr} \tablenum{5} \tablewidth{0pt} \scriptsize \tablecaption{Mean Abundances for NGC 6553 and Comparison with Baade's Window Results} \label{tab5} \tablehead{ Ion & \colhead{NGC 6553 Mean} & \colhead{$\sigma$} & \colhead{Mean BW \tablenotemark{a}} & \colhead{ {$\sigma$}\tablenotemark{a}} \nl & \colhead{(dex)} & \colhead{(dex)} & \colhead{(dex)} & \colhead{(dex)} } \startdata [Fe/H] \nl Fe I & $-$0.16 & 0.08 & $-$0.33 \nl Fe II & $-$0.18 & 0.10 \nl & \nl [A/Fe] & \nl O I & +0.50 & 0.13 & +0.03 & 0.18 \nl Mg I & +0.41 & 0.10 & +0.35 & 0.14 \nl Si I & +0.14 & 0.18 & +0.18 & 0.24 \nl Ca I & +0.26 & 0.09 & +0.14 & 0.17 \nl Sc II & $-$0.12 & 0.18 & +0.29 & 0.20 \nl Ti I & +0.19 & 0.06 & +0.34 & 0.10 \nl Cr I & +0.04 & 0.09 & $-$0.04 & 0.19 \nl Ni I & +0.01 & 0.07 & $-$0.04 & 0.08 \nl \enddata \tablenotetext{a}{Abundances for Baade's Window are from the 11 giants studied by McWilliam \& Rich (1994).} \end{deluxetable} \begin{deluxetable}{lrlrlr} \tablenum{6} \tablewidth{0pt} \scriptsize \tablecaption{Abundances for $\zeta$ Cyg} \label{tab6} \tablehead{\colhead{Ion} & \colhead{\# of} & \colhead{Original Spectrum} & \colhead{\# of} & \colhead{Convolved Spectrum} & \colhead{Solar Abund.} \nl \colhead{} & \colhead{Lines} & \colhead{Abundance} & \colhead{Lines} & \colhead{Abundance} \nl \colhead{} & \colhead{} & \colhead{(dex)} & \colhead{} & \colhead{(dex)} & \colhead{(dex)} } \startdata [Fe/H] \nl Fe I & 43 & +0.08 (0.12) & 27 & +0.05 (0.20) & $-$4.50 \cr Fe II & 3 & +0.09 (0.17) & 1 & $-$0.06 & $-$4.56 \cr & \nl [A/Fe] \nl O I & 4 & +0.35 (0.14) & 3 & +0.38 (0.18) & +1.18 \nl Mg I & 1 & +0.15 & 1 & +0.20 & +0.04 \nl Si I & 3 & +0.10 (0.11) & 2 & +0.11 (0.26) & +0.04 \nl Ca I & 5 & +0.08 (0.14) & 3 & +0.07 (0.04) & $-$1.34 \nl Sc II & 1 & +0.30 & 1 & +0.12 & $-$4.48 \nl Ti I & 2 & $-$0.08 (0.16) & 1 & $-$0.28 & $-$2.52 \nl Cr I & 4 & +0.03 (0.06) & 1 & +0.02 & $-$1.88 \nl Ni I & 16 & +0.02 (0.14) & 9 & $-$0.08 (0.17) & $-$1.25 \nl \enddata \end{deluxetable} \clearpage
\section{Introduction} In the preceding paper (Rajagopal, Srinivasan and Dwarakanath, 1998; paper I) we presented the results of a program to obtain the HI absorption profiles towards a selected sample of bright stars. As mentioned there, detailed optical absorption studies exist in the direction of these stars in the lines of NaI and CaII. The motivation for such an observational program was also described in the previous paper. In the present paper, we wish to discuss the results obtained by us. There are two major issues that we wish to address and discuss later: \begin{itemize} \item[1] Are the properties of interstellar clouds seen in optical absorption the same as those seen in HI emission and absorption in general? \item[2] What is the origin and nature of the faster clouds seen in optical and UV absorption ? \end{itemize} Let us elaborate a bit on the first issue. Although the general picture of the ISM that emerged from optical and radio observations, respectively, is the same, viz., clouds in pressure equilibrium with an intercloud medium, it has not been possible to directly compare the inferred properties. Whereas both the column densities and the spin temperatures of the clouds have been estimated from HI observations, there have only been indirect and often unreliable estimates for the clouds seen in optical absorption. Our HI absorption measurements, combined with HI emission measurements in the same directions will enable us to directly estimate for the first time the column densities and spin temperatures of the clouds seen in optical absorption. The second question mentioned above arises as follows. The existence of a {\it high velocity tail} in the distribution of random velocities of clouds was firmly established by Blaauw (1952) from the data obtained by Adams (1949). As mentioned in paper I, it was noticed quite early on by Routly and Spitzer (1952) that the faster clouds have a smaller NaI to CaII ratio than the lower velocity clouds. Early HI emission measurements (see paper I for references) in the direction of the bright O and B stars provided an added twist. Whereas the lower velocity clouds clearly manifested themselves in the HI emission measurements, the higher velocity clouds were not detected. To illustrate this point, we show in Fig 1 the optical absorption features and the HI emission profiles towards the star HD 219188. It may be seen that there is no counterpart in HI emission to the higher velocity optical absorption feature. \begin{figure} \begin{center} \epsfig{file=hi_opt.ps,height=10cm,width=8.8cm} \caption{Comparison of optical absorption (top) from CaII towards HD 219188 (Sembach, Danks and Savage, 1993) and HI emission (Habing 1968) in the same direction. The absorption feature at higher velocity is missing in the HI profile.} \end{center} \end{figure} In the next two sections (sections 2 and 3) we will discuss the results of our absorption survey presented in paper I. We shall classify the HI absorption features into two broad classes, viz., ``low velocity'' and ``high velocity''. It is of course difficult to define a sharp dividing line between ``low'' and ``high'' velocities. But based upon earlier analyses of NaI to CaII ratios, as well as HI emission measurements, we adopt a velocity of the order of 10 \kms as the dividing line. The main conclusions from our study are summarized in section 4. Section 5 is devoted to a detailed discussion of the nature and origin of the high velocity clouds. \section{The low velocity clouds} In this section we discuss the low velocity absorption features (i.e, $ v \lsim 10$\kms). To recall from paper I, we detected HI absorption in all but 4 of the 24 fields we looked at. We will discuss these four unusual fields at the end of the next section. \subsection{Coincident absorption features} All the lines of sight in our sample show optical absorption from CaII at both low and high velocities. In most of the fields where we have detected HI absorption, they occur at roughly the same velocities as the optical absorption lines for $ v \lsim 10$\kms. Not surprisingly, the HI absorption features have one to one correspondence with the HI emission features in the fields for which earlier emission measurements exist. In Table 1 we have listed all the fields with ``matching'' velocities in optical absorption and HI emission and absorption. \begin{table} \begin{center} \begin{tabular}{|l|c|c|l|c|l|}\hline Field & V$_{lsr}$(opt) & V$_{lsr}$(HI)& $\tau$ & $\Delta V$ & T$_s$ \\ \hline\hline {} & \kms & \kms & & \kms & K \\ \hline 14143-& -10.3& $-$11.2& 0.18& 11.8 & 283 \\ 14134 & & & & & \\ 14818& -6.6& -3.7 & 0.43 & 13.7 & 143 \\ 21278& -0.2& 2.8& 0.18 & 5.1 & 250 \\ 21291& -7.5& 7.0 & 0.60 & 6.2 & 180 \\ 24912& 4.7& 4.3& 0.80 & 14.9 & 73 \\ 25558& 10.1& 8.1& 1.13 & 5.6 & 73 \\ 34816& 4.1& 6.0& 1.50 & 3.2 & 45 \\ 41335& 0.2& 1.0& 0.22 & 7.2 & 342 \\ 42087& 10.2& 12.4& 1.26 & 3.5 & 211 \\ 141637& 0.0& 0.5 & 1.60 & 8.0 & NA \\ 148184& 2.2& 3.4& 4.90 & 3.4 & 50 \\ 156110& 0.4& 2.2& 0.28 & 7.5 & 102 \\ 159176& -22.5& $-$20.8&1.15 & 15.3 & NA \\ 166937& 5.9 & 5.4 & 1.70 & 5.7 & NA \\ 175754& 5.9& 6.8& 0.35 & 17.7 & 135 \\ 199478& -2.1, 8.7& 3.8$^+$ & 0.60 & 11.4 & 121 \\ 212978& 0.6& 0.3& 0.20 & 9.0 & NA \\ 214680& 0.1& 1.4& * & 2.1 & NA \\ \hline \end{tabular} \vspace{.5cm} \caption[Summary of coincident low velocities.]{Summary of coincident low velocities: Column 1 lists the HD number for the field. Columns 2 and 3 give the LSR velocities for the ``matching'' optical and HI absorption features respectively. Column 4 and 5 lists the optical depth and width for the HI absorption derived from the fitted gaussian. Column 6 lists the spin temperature T$_s$. These values have been derived by using the optical depth from our observations along with the emission brightness temperatures from Habing (1968). Where this was not available we used brightness temperatures from the Leiden-Green Bank survey (Burton, 1985). Those fields where emission measurements were not available are marked NA in column 6.\\+ The feature is a blend\\$*$ The feature is saturated} \end{center} \end{table} For illustration we have shown in the upper panels of Fig. 2 the HI absorption spectra (optical depth) in 3 fields; HI optical depth is plotted as a function of V$_{\rm LSR}$. The arrows indicate the velocities of the optical absorption lines. For comparison, we have shown in the lower panels the HI emission in these 3 fields: the data obtained by Habing (1968) has been digitized and re-plotted. The first field contains the star HD 34816. The optical spectrum obtained by Adams (1949) towards this star shows CaII absorption at $-$14.0 and $+$ 4.1 \kms. The HI absorption profile shows a prominent feature at 6 \kms (we have obtained spectra towards 2 radio sources in this field but only one of them is shown here). As may be seen from the lower panel, there is a corresponding emission feature at this velocity. As mentioned in paper I, absorption features in the optical and HI spectra may be taken to be at ``matching'' velocities provided they are within $\sim$3 \kms of one another (this window is to account for blending effects in the optical spectra and different corrections adopted for solar motion). In view of this one may conclude that the optical absorption at $+$ 4.1 \kms and HI emission and absorption at 6 \kms arise in the same cloud, even though the radio source is 20' away from the star. The second panel pertains to the field containing the star HD 42087. In this field also we have two radio sources within the primary beam, and the spectrum towards one of them is shown. The absorption features are clearly seen at 4.4 \kms and 12.4 \kms, with the latter being much stronger. The HI emission spectrum (shown in the panel below) shows a broad peak centred at $\sim$ 15 \kms. The absorption feature at 12.4 \kms may be taken to be the counterpart of the optical absorption at 10.2 \kms. There is no HI absorption at negative velocities corresponding to the other optical absorption lines indicated by the arrows. This may be due to the fact that the 2 radio sources in the field are 32' and 42' respectively from the star in question. Given that the star is at a distance of 1.3 kpc (paper I, Table 1) it is conceivable that we are not sampling all the gas seen in optical absorption. The spectra towards HD 148184 is shown in panel 3. Again there is good agreement between the HI spectrum and the optical spectrum as far as the lower velocity optical absorption is concerned. As in the previous two cases there is no counterpart of the higher velocity optical absorption in the HI spectrum. These two examples will suffice to illustrate the general trend in Table 1 viz., there is reasonably good agreement at low velocities ($v \lsim 10$ \kms) between the optical absorption features and the HI spectra. \begin{figure} \begin{center} \vspace{6.0cm} \mbox{ \epsfig{file=fig2.ps,width=13cm,angle=90}} \end{center} \end{figure} Returning to Table 1, we have listed the derived optical depths in column 4 and the velocity width of the HI absorption in column 5. In the last column we have given the derived {\it spin temperatures}. These are obtained by supplementing our HI absorption measurements with brightness temperatures derived from emission measurements (mostly from Habing, 1968; and in some cases from Burton, 1985). To the best of our knowledge, this is the first direct determination of the temperatures of the interstellar clouds seen in optical absorption. To be precise, the temperature derived by us is the spin temperature which may be taken to be an approximate measure of the kinetic temperature. The correspondence between the optical absorption features and the HI absorption suggest that one is sampling the same clouds in both cases. The derived spin temperatures and velocity widths are consistent with these clouds belonging to the same population as the standard cold diffuse clouds in the raisin-pudding model of the ISM. While it is conceivable that at low velocities one may merely be sampling the local gas (regardless of direction), statistical tests carried out by Habing (1969) suggest that this is unlikely. \subsection{Non-coincident absorption features} As we have already encountered in the case of HD 42087 (see Fig. 2), sometimes there is a mismatch between optical and radio spectrum even at low velocities. We mention two specific cases here. \subsubsection*{HD 37742} The HI absorption spectrum obtained by us (Fig. 3) shows a deep absorption feature at 9.5 \kms. There are no other absorption features down to an optical depth limit of 0.03. The optical absorption features are at 3.6 and $-$21 \kms. Since the radio source is only 12' away from the star, given the distance estimate of 500 pc to the star the discrepancy between the optical spectrum and the HI absorption spectrum is significant and intriguing. While the gas seen strongly absorbing in HI could be located beyond the star, one is left wondering as to why one does not see the low velocity cloud seen in optical absorption. \begin{figure} \addtocounter{figure}{1} \begin{center} \mbox{ \epsfig{file=37742n1.ps,width=8.8cm,height=6cm}} \end{center} \caption{Spectrum towards HD 37742. The arrows on the velocity axis mark the velocities at which optical absorption is seen.} \end{figure} \subsubsection*{HD 119608} The optical absorption spectrum towards this star obtained by M\"{u}nch and Zirrin (1961) shows 2 minima at 1.3 and 22.4 \kms. We have obtained HI absorption towards 2 strong radio sources in this field both within 15' of the star. Both show a deep absorption feature at $-$5.4 \kms (Fig. 4). This agrees with the HI emission feature shown in the lower panel. However the HI emission spectrum also shows a broad feature peaking at $\sim$ 20 \kms. This is also seen in the more recent measurement of Danly etal (1992). If this feature is indeed to be identified with the optical absorption at 22.4 \kms, then this represents an interesting case where the gas in the line of sight to the star causing optical absorption manifests itself in HI emission but not absorption. This could happen for example if the spin temperature of this gas is sufficiently high as to make the HI optical depth below our detection limit. HD 119608 is a high latitude star and one is presumably sampling the halo gas, and warrants a deeper absorption study. \begin{figure} \begin{center} \epsfig{file=119608.ps,width=8.8cm,height=10cm} \end{center} \caption{HI absorption (top) and emission (bottom) spectra towards HD 119608. The arrows mark the velocities at which optical absorption is seen. The emission spectrum is from Habing (1968).} \end{figure} \section{The high velocity clouds} Although in the previous section, we were primarily concerned with establishing the correspondence between the optical absorption features and the HI absorption spectra at low velocities, we did have occasion to comment on the {\it absence of HI absorption from the high velocity clouds} [The high velocity clouds we are discussing are those that populate the tail of the velocity distribution obtained by Blaauw (1952) and not those that are commonly referred to as HVCs in the literature]. It turns out that in all but 4 cases, we fail to detect HI absorption at velocities corresponding to the high velocity \mbox{($\gsim$ 10 \kms)} optical absorption lines. To illustrate this generic trend, we have shown some additional examples in Fig. 5. {\it It may be seen in the figure that the high velocity optical absorption features (indicated by arrows) are not seen in the HI emission spectra either}. A discussion of this will form the major part of section 5. \begin{figure} \begin{center} \vspace{6.0cm} \epsfig{file=fig3.ps,width=13cm,angle=90} \end{center} \end{figure} In only 4 out of the 24 lines of sight have we detected HI absorption at velocities $\gsim$ 10 \kms and which clearly correspond to the optical absorption lines. We discuss these below. \subsection{Coincident absorption features at high velocities} \subsubsection*{HD 14134, HD 14143} These two stars (in the same field) are members of the h and $chi$ Persei clusters (M\"{u}nch 1957). There were 3 radio sources within our primary beam (all within 10' of the star). The HI absorption features towards one of them is shown in Fig. 6. The prominent high velocity absorption features towards the 3 sources are at $-$52.8, $-$50.3 and $-$46.1 \kms, respectively (Table 1 of paper I). These should be compared with the optical absorption features towards the 2 stars in question which are at $-$46.8 and $-$50.8 \kms. Thus there is reasonable coincidence between the optical and HI data. Nevertheless we wish to now point out that the high velocity may not represent random motion but rather systematic motion. \begin{figure*} \addtocounter{figure}{1} \begin{center} \epsfig{file=14143n2.ps,height=6cm,width=8.8cm} \end{center} \caption{HI absorption spectra towards HD 14143. The arrows mark the velocities at which optical absorption is seen.} \end{figure*} While interpreting his pioneering observations M\"{u}nch (1957) attributed the high velocity features in the optical spectra to anomalous motions in the Perseus arm. Since then it has generally been accepted that there are streaming motions in the Perseus arm with velocities ranging from $-$10 to $-$30 \kms (Blaauw and Tolbert 1966; Brand and Blitz 1993). For the sake of completeness we have listed in Table 2 the spin temperature of the gas derived by us by combining our measurements with existing HI emission measurements. \begin{table}[t] \begin{center} \begin{tabular}{|l|l|l|l|l|l|}\hline Field & V$_{lsr}$(optical) & V$_{lsr}$(HI)& $\tau$ & $\Delta V$ & T$_s$ \\ \hline\hline {} & \kms & \kms & & \kms & K \\ \hline 14143-& $-$50.8& $-$50.3 & 0.30 & 10.0 & 96 \\ 14134 & & & & & \\ 21291& $-$34.0& $-$31.2& 0.40 & 3.3 & 181 \\ 159176& $-$22.5& $-$20.8& 1.15 & 15.2 & NA \\ \hline \end{tabular} \vspace{.5cm} \caption[Summary of coincident high velocities.]{Summary of coincident high velocities: Column 1 gives the HD number of the field. Column 2 lists the {\em high} velocity optical absorption seen towards the star. Column 3 is the ``matching'' HI absorption. Columns 4 and 5 give the fitted optical depth and width of the HI absorption features. The spin temperature is listed in column 6. As in the for low velocity features, we have used Habing(1968) and the Leiden-Green Bank survey for the emission temperatures needed to compute the spin temperature from the optical depth.} \end{center} \end{table} \subsubsection*{HD 21291} The spectrum towards this star near the Perseus arm has a prominent Na D line at a velocity of $-$34 \kms (M\"{u}nch, 1957). The HI absorption spectrum shows a feature at $-$31.2 \kms. The contribution to radial velocity from Galactic rotation can only be $\sim$ 10 \kms, thus indicating significant peculiar motion of the gas. In our opinion one must attribute this to streaming motion of the gas as in the case discussed above. HI emission clearly shows spatially extended gas covering the longitude range from $\sim$136$^{\circ}$ to 141$^{\circ}$ at the velocity of interest. This strengthens the conclusion that one must not attribute the observed velocity to random motions. \subsubsection*{HD 159176} There is pronounced optical absorption at $-$ 22.5 \kms which might be identified with the HI absorption seen by us at $-$20.8 \kms. Given the longitude of 356$^{\circ}$, it is difficult to attribute this to Galactic rotation. The measured velocity must correspond either to random velocity or systematic motion. \subsubsection*{HD 166937} In the case of this star, there is no strict coincidence (within 3 \kms) between the HI and optical absorption at high velocities. However we see two HI absorption features at velocities close to and straddling the optical feature at 41.1 \kms, so we include this field in our list of high velocity coincidences. It may be noted that this star is also close to the Galactic center direction. \begin{figure} \begin{center} \mbox{ \epsfig{file=93521.ps,width=8cm}} \end{center} \caption{HI absorption (top) and emission (bottom) spectra towards HD 93521. The arrows plot mark the velocities at which optical absorption is seen. The emission spectrum is from Habing (1968).} \end{figure} \section{Summary of results} In the preceding section we described the first attempt to directly compare HI absorption with optical absorption features arising in the ISM. We summarize below the main results: \begin{itemize} \item[1.] HI absorption measurements were carried out towards 24 fields. Each field has existing optical absorption spectra towards a bright star. In 20 of these fields we detected HI absorption features. \item[2.] In all but 4 of these 20 fields, the HI absorption features at low velocities ($<$ 10 km s$^{-1}$) correspond to the optical absorption lines. \item[3.] In most cases there is also corresponding HI emission. \item[4.] The spin temperatures derived by us for the low velocity gas is consistent with the standard values for the cold diffuse HI clouds. \item[5.] This lends strong support to the hypothesis that (at least) the low velocity clouds seen in optical absorption belong to the same population as those sampled in extensive HI studies. \item[6.] In 20 out of 24 fields surveyed, we did not detect any HI absorption corresponding to optical absorption at high velocities ($v > 10$\kms). It is unlikely that in all cases this is due to our line of sight not sampling the gas seen in optical absorption; in several cases the line of sight to the radio sources would have sampled this gas even if its linear size was of the order of 1 pc. \item[7.] Curiously, the early emission measurements also failed to detect HI gas at high velocities (in these same lines of sight). Given the size of the telescopes used, beam dilution could have accounted for the non-detection if the gas was ``clumpy''. Our absorption measurements rule this out as a generic explanation. More recent and sensitive measurements indicate that the high velocity gas has much smaller column density than the low velocity gas (N(HI) $<$10$^{\rm 18}$ cm$^{-2}$). If this is the case, then it is not difficult to reconcile why we do not see it in absorption since our sensitivity limit was $\tau \gsim$0.1. But then the correlation between high velocity and low column density would have to be explained. We venture to offer some suggestions in the next section. \item[8.] {\it Fields with no HI absorption}: We wish to record that in the fields containing the stars HD 38666, HD 93521, HD 205637 and HD 220172 {\it we did not detect any HI absorption - even at low velocities.} These are all high latitude stars. HI emission spectra also show only weak features. For illustration we show in Fig. 7 the HI absorption and emission spectrum towards HD 93521. From very detailed investigations - the case of HD 93521 is a good example - it has been concluded that most of the optical absorption arises from warm gas in the halo (detailed references may be found in Spitzer and Fitzpatrick 1993 and Welty, Morton and Hobbs 1996). The fact that we do not see HI absorption is consistent with the interpretation of this gas being warm. Weak HI emission indicates low column density also. \end{itemize} \section{Discussion} As we have already argued, our absorption measurements, taken together with earlier emission measurements, establishes that the low velocity clouds seen in optical absorption are to be identified with the standard HI clouds - their column densities and spin temperatures match. But the true nature of the high velocity clouds seen in optical absorption is still unclear. There are two questions to be addressed: (1) Do the high velocity clouds belong to a different population, and (2) is there a causal connection between their higher velocities and lower column densities? We wish to address these two questions below. An unambiguos indication that the high velocity clouds may have very different properties compared to their low velocity counterparts comes from an HI absorption study towards the Galactic center (Radhakrishnan and Sarma 1980). Given the statistics of clouds derived from optical studies (8 to 12 per kpc), if the high velocity clouds had optical depths comparable to the low velocity clouds, then an absorption experimant towards the Galactic center should straightaway reveal a velocity distribution similar to the one derived by Blaauw (1952) from Adams' data. The velocity distribution derived by Radhakrishnan and Sarma from precisely such a study did not reveal a pronounced high velocity tail. The velocity dispersion of 5 \kms derived by them was in good agreement with the low velocity component of Blaauw's distribution. Instead of a pronounced high velocity tail seen in optical and UV studies, there was at best a hint of a high velocity population of very weakly absorbing clouds. Even this conclusion has remained controversial (Schwarz, Ekers and Goss 1982). As for the possible correlation between higher velocities of clouds and lower column densities, fairly conclusive evidence comes from UV absorption studies. Since the UV absorption lines have larger oscillator strengths, they can be used to probe smaller column densities than is possible with optical absorption lines. The analysis of Hobbs (1984) seems to confirm this expectation - in several lines of sight there is more high velocity UV absorption features than in the optical. A more direct inference can be drawn from the work of Martin and York (1982). For the two lines of sight they studied, there is a clear indication of lower column density (N(HI)) at higher velocities. Over the years, three broad suggestions have been put forward in an attempt to elucidate the nature of the high velocity clouds. \subsubsection*{Circumstellar clouds} According to an early suggestion due to Schl\"{u}ter, Schmidt and Stumpf (1953), the high velocity clouds seen in optical absorption are to be identified with circumstellar clouds. This was an attempt to explain the predominance of {\it negative} velocities in the high velocity absorption features. If the clouds in the vicinity of massive stars are accelerated by the combined effect of stellar winds and radiation from the stars, then in an absorption study against the stars one would detect only those clouds accelerated towards us. A few years later, Oort and Spitzer (1955) developed the well known ``rocket mechanism`` in which the UV radiation from the star ionizes the near side of the cloud resulting in ablation and consequent acceleration of the cloud. This mechanism will naturally result in the higher velocity clouds having smaller mass and therefore smaller column density. The difficulty with this mechanism, however, is that one will have to invoke another mechanism to explain the large {\it positive} velocities which are also seen in absorption studies. In view of this we will not dwell any further on this scenario. \subsubsection*{Relic SNRs} An alternative scenario was advanced by Siluk and Silk (1974). Their suggestion was that the high velocity optical absorption features arise in very old supernova remnants (SNRs) which have lost their identity in the ISM. Their primary objective in advancing this scenario was to explain the high velocity tail of the velocity distribution of optical absorption features. The point was that if the absorption features arise not in interstellar clouds but in SNRs in their very late stages of evolution, then it would result in a power law distribution of velocities; such a distribution according to them provided a good fit to the observations. While this suggestion is quite attractive, it suffers from two drawbacks: (1) The early studies on the evolution of SNRs predicted the formation of very dense shells beyond the radiative phase. Such compressed shells were essential to explain the observed absorption features and the derived column densities. However, more recent studies which take into account the effects of the compressed magnetic field and cosmic ray pressure in the shells suggest that either dense shells do not form or if they do, do not last long enough (Spitzer 1990; Slavin and Cox 1993). (2) Given a supernova rate of one per $\sim$ 50 years in the Galaxy, the statistics of absorption features requires that the SNRs enter the radiative phase (and as a consequence develop dense shells) when they are still sufficiently small so as not to overlap with one another. This would indeed be the case if the intercloud medium into which the SNRs expand is dense enough (n $\sim$ 0.1 cm$^{-3}$). But if a substantial fraction of the ISM is occupied by low density hot gas (n$\sim$0.003 cm$^{-3}$; T $\sim$ 5$\times$10$^5$K) such as indicated by UV and soft X-ray observations then the supernova bubbles are likely to intersect with one another and perhaps even burst out of the disk of the Galaxy before developing dense shells (Cowie and York 1978). In view of these two drawbacks, we do not favour this suggestion. \subsubsection*{Shocked clouds} The third possibility is that the high velocity absorption features do arise in interstellar clouds but which have been engulfed and shocked by supernova blast waves. Indeed we feel that this is the most plausible explanation for it has support from several quarters. The earliest evidence that the high velocity gas may be ``shocked'' came from the Routly-Spitzer effect. The NaI/CaII ratio in the fast clouds was lower (sometimes by several orders of magnitude) than in the slow clouds. The variation in NaI/CaII ratio was primarily attributed to the variable {\it gas phase abundance} of calcium in these clouds. Due to its relatively high condensation temperature calcium is likely to be trapped in grains. Spitzer has argued that the observed trend in NaI/CaII ratio could be understood if the calcium is released back into the gas phase in the high velocity clouds due to sputtering. This is indeed what one would expect if the interstellar cloud is hit by an external shock, which in turn drives a shock into the clouds (Spitzer 1978). Supernova blast waves are the most likely candidates. Earlier in this section we referred to an HI absorption study by Radhakrishnan and Sarma towards the Galactic center. While they did not find strong absorption at high velocities they did conclude that there must be a population of weakly absorbing high velocity clouds. Radhakrishnan and Srinivasan (1980) examined this more closely and advanced the view that in order to explain the optical depth profile centered at zero velocity one had to invoke {\it two distinct velocity distributions}: a standard narrow distribution with a velocity dispersion of $\sim$ 5 \kms, and a second one with a much higher velocity dispersion of $\sim$35 \kms. While arguing strongly for a high velocity tail, they stressed that the latter distribution must consist of a population of very weakly absorbing clouds. They went on to suggest that this population of weakly absorbing clouds might be those that have been shocked by expanding SNRs; the very process of acceleration by SNRs might have resulted in significant loss of material and heating of the clouds, leading to low HI optical depths. To conclude this discussion we wish to briefly summarize the expected life history of a cloud hit by a supernova blast wave. The first consequence of a cloud being engulfed by an expanding SNR is that a shock will be driven into the cloud itself resulting in an eventual acceleration of the cloud. The effect of this shock and a secondary shock propogating in the reverse direction after the cloud has been overtaken by the blast wave, is to compress and flatten the cloud. Eventually various instabilities are likely to set in which will fragment the cloud. The detailed history of the cloud depends upon two important timescales: the time taken for the cloud shock to cross the cloud and the evolutionary timescale of the SNR. If the former is much smaller than the latter, the cloud is likely to be destroyed. However if the reverse is true, then the shocked cloud will survive and be further accelerated as a consequence of the viscous drag of the expanding hot interior. {\it Clouds accelerated in such a manner will however suffer substantial evaporation due to heat conduction from the hot gas inside the SNR.} Partial fragmentation could further reduce the size of the cloud. For detailed calculation and discussion we refer to McKee and Cowie (1975), Woodward (1976), McKee, Cowie and Ostriker (1978), Cowie, McKee and Ostriker (1981) and a more recent paper by Klein, McKee and Woods (1995). To summarize the above discussion, in our opinion the shocked cloud scenario has all the ingredients needed to explain the observational trends. In particular it would explain why the high velocity clouds seen so clearly in optical and UV absorption lines do not manifest themselves in HI observations. But this observation is predicated on the conjecture that the higher velocity clouds are not only warmer, but have smaller column densities. There is certainly an indication of this from optical and UV absorption studies. To recall, in UV observations which are sensitive to much smaller column densities than optical studies, the higher velocity absorption features are more pronounced. But it would be desirable to quantify the correlation between velocity and column densities. Reliable column densities are difficult to obtain from optical observations because of blending of lines and also depletion onto grains. The column densities derived from UV observations are also uncertain because of the effects of saturation of the lines. In view of these difficulties it would be rewarding to do a more systematic and much more sensitive HI absorption study, supplemented by emission studies.
\section{Introduction } WFPC2 pure parallel images from the HST Medium Deep Survey key project (\cite{1994ApJ...437...67G,1994ApJ...435L..19G} hereafter MDS ) cover a very wide range of signal-to-noise. For the few brightest galaxies observed, detailed structures such as spiral arms and bright regions of star-formation are well exposed and the morphology can be easily classified by eye and measured by traditional interactive one-dimensional profile fitting procedures. At these brighter magnitudes the two-dimensional light distributions of galaxies are not well fitted by simple parameterized models which are necessarily crude fits to the broad continuum using smooth image profiles. However, as the images get fainter and smaller (undersampled), the morphology is less apparent and requires a model-based two-dimensional image analysis to derive quantitative estimates. For the extreme faint and small objects there is very little morphological information in the observations. The MDS procedure described in this paper has been optimized for the intermediate (medium deep) galaxies, in the rough magnitude range between V $\approx 21$ to 24 mag., as imaged in exposures of about one hour. This has yielded a significantly large catalog of quantitative morphological and structural parameter estimates. This magnitude range is now accessible for spectroscopic determination of redshifts via the new generation of 8-10 meter class ground based telescopes. Decomposition of the images into disk and bulge has been a difficult task even at bright magnitudes with well sampled images (\cite{1977ApJ...217..406K,1981ApJS...46..177B,1984ApJS...56..105K,1985ApJS...59..115K}). Interactive procedures (\cite{1991PASP..103..396Y}) are also impractical for a large survey and in any case they do not generate an uniform catalog suitable for statistical analysis. The image analysis adopted is similar to that in stellar photometry programs like DAOphot (\cite{1987PASP...99..191S}). But unlike stellar photometry where the image can be characterized by the centroid, magnitude and the Point Spread Function (PSF), there is no simple model which will intrinsically fit all of the galaxy images. We adopt axisymmetric scale-free models which have been shown to fit the image continuum of normal galaxies (\cite{1959Hbphy..53..275V,1970ApJ...160..811F}). The procedure will average over any bright regions, as typically occurs in the data themselves at fainter magnitudes where the objects are smaller and less resolved. The residuals to these simple galaxy model fits at brighter magnitudes are the subject of a separate study (\cite{1997ApJ...476..510N}). To limit the complexity of the analysis, we assume that an image pixel is associated with a single object or background sky as is typical of the MDS WFPC2 images. We do not deal with the problems of crowding or image overlap, which are the major issues in programs for stellar photometry. The number and choice of parameters fitted to an extended image is clearly important. Fitting too few parameters to a well exposed image could significantly bias the estimates of the parameters fitted, by the implicit choice of the parameters that are not fitted. However, fitting too many parameters to faint and/or compact unresolved images could cause the fit to converge to a false local minimum of a likelihood function which is very noisy in that multidimensional space. For practical reasons, and to ensure statistical uniformity of the resulting catalog, we require an automated procedure which will select and fit those (necessary and sufficient) parameters which are constrained by each particular image. We have developed two-dimensional ``maximum likelihood'' image analysis software that attempts to automatically optimize the model and the number of parameters fitted to each image. We apply the Ockham's razor: {\it non sunt multiplicanda entia praeter necessitatem; i.e., } entities are not to be multiplied beyond necessity (Ockham 1285-1348). The model varies from a simultaneous decomposition of disk and bulge components of galaxy images ( hereafter \rm D+B\ models) at the bright end to circularly symmetric sources at the faint end. However, this choice of parameters creates selection effects which depend on the signal-to-noise of the image and needs to be included explicitly in any statistical analysis of the MDS database. The success of the procedure depends on the ability to efficiently generate smooth subpixelated galaxy images which can be convolved with an adopted Point Spread Function (PSF), such that precise derivatives can be evaluated with respect to all the parameters which need to be estimated. We will outline the procedure here but will avoid giving all the details of the numerical algorithm, since they are probably not of interest to the general reader. The algorithm is documented by comments in the software and the interested reader should contact the first author. A brief outline of the MDS pipeline is given in the Appendix. This paper is also the primary reference to the Medium Deep Survey database which has been made available on the MDS website in the HST archive \footnote{at http://archive.stsci.edu/mds/} and also mirrored at the Canadian Astronomy Data Center (CADC) \footnote{at http://cadcwww.dao.nrc.ca/mds/}. We avoid duplicating extensive tables since those can only be a snapshot of the present MDS database and we wish to ensure that users will always refer to the latest version which will be maintained on the Internet. The MDS website has a cgi-interface written in f77 which allows the database to be searched using coordinates or galaxy parameters, or looked at interactively by clicking on objects on an image-map of each stack. Direct access is also provided to the MDS database which is on CDROMs in a `jukebox'. The database contains WFPC2 Pure Parallel observations taken for the Medium Deep Survey (MDS - HST GO program ids 5369, 5370, 5371, 5372, 5971, 6251, 6802, 7203) and for the GTO observers (HST program ids 5091, 5092, 5201, 6252, 6254, 6609, 6610, 7202 ) as well as HST archival observations of randomly selected WFPC2 fields like that of the Groth-Westphal strip (HST GTO program ids 5090 5109 - hereafter GWS ) and the Hubble Deep Field (HST DD program id 6337 - hereafter HDF ), and selected galaxy cluster fields (HST archival program id 7536) and will continue to be expanded as more fields are processed (HST archival program id 8384). \section{The Observations} The HST MDS and GTO pure parallel observations were taken with the WFPC2 after January 1994, following the SM93 repair mission, and continued for four years until January 1998. Before the SM97 second servicing mission in February 1997, the instruments used for the associated HST primary observations were the FGS, FOC and FOS; after this mission, the primary instruments were FGS, STIS and NICMOS. \begin{figure} \plotfiddle{ratnatunga.fig01.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Location of pure parallel fields relative to targets in primary observations. Instrument of HST primary as indicated by symbols. \label{fig1}} \end{figure} We illustrate in Figure~\ref{fig1} the difference in pointing between the parallel and primary observations for all pure parallel fields in the MDS database, using different symbols for each primary instrument. The WFPC2 field is on average 4.5, 8.2, 12.1, 7.1, and 5.3 arc min away from the FOC, FOS, FGS, STIS, and NICMOS primary target respectively. \begin{figure} \plotfiddle{ratnatunga.fig02.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Hours of pure parallel WFPC2 observations over the 4 years of the Medium Deep Survey. Program as indicated by shading. \label{fig2}} \end{figure} About 25 hours of pure parallel exposure was obtained each month, giving a steady flow of observations. The database was supplemented using the archival data from primary observations which satisfied the survey criteria. The observation history is illustrated in Figure~\ref{fig2} as an indication of the quantity of HST data that was available for the survey. There was a significant drop in the number of WFPC2 parallels after SM97 as a result of the dithered observing strategy of NICMOS and STIS primary observations. The pure parallel GTO data was available to REG as a WFPC2 Investigation Definition Team member and Windhorst's Blue Survey data (WBS) was available from the HST archive 3 months after observation. \begin{figure} \plotfiddle{ratnatunga.fig03.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Distribution of the MDS pure parallel WFPC2 fields over the sky in Galactic Coordinates. Symbol size proportional to total length of exposure and are filled for fields processed for MDS. \label{fig3}} \end{figure} The HST MDS (\cite{1994ApJ...435L..19G}), with over 400 random WFPC2 fields distributed over the full Sky, the GWS (\cite{1994AAS...185.5309G}) with 28 contiguous WFPC2 fields and the Hubble Deep Field (HDF \cite{1996AJ....112.1335W}) are datasets which give three very complementary samples of field galaxies at faint magnitude. The HDF gives depth in a single WFPC2 field, the GWS gives a larger area uniformly observed, and the MDS samples the whole sky as illustrated in Figure~\ref{fig3}. All three sets have been analyzed uniformly through the MDS pipeline analysis software system. \slugcomment{ } \begin{center} \begin{table}[H] \title{MDS Field Priority} \medskip \begin{tabular}{ c l } \tableline Priority & Description \\ \tableline \tableline \\ 1 & 3 or more images in each of 2 or more Filters \\ 2 & 2 or more images in each of 2 or more Filters \\ 3 & 3 or more images in 1 Filter \\ 4 & 2 images in one and 1 image in other Filter \\ 5 & 1 image each in 2 Filters \\ 6 & 2 images in 1 Filter \\ 7 & 1 image only \\ 8 & Bad exposure \\ 9 & Failed observation \\ \tableline \end{tabular} \caption{Assignment of Field Priority in the MDS pipeline\label{tbl-1}} \end{table} \end{center} MDS and GTO observations were primarily done with the F814W and F606W broadband filters. When more than 3 exposures could be taken with each of these filters, then F450W observations were taken in addition to those in the first two. In order to achieve a similar signal-to-noise ratio in the images taken in all three filters, the exposure times in F814W and F606W were requested to be about equal while in F450W the requested exposure was about twice as long. However, all WFPC2 observations in the MDS were taken in ``non-interference'' pure parallel mode (\cite{1994ApJ...437...67G}), with the result that exposure times were of varied duration, with a variable number of exposures in the stack used for cosmic ray removal. \begin{figure} \plotfiddle{ratnatunga.fig04.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ The total hours of exposure and number of fields at each priority (see Table~\ref{tbl-1}) in each of the 3 primary WFPC2 Filters. The filled region of histogram are of fields processed through MDS pipeline. The partial shaded region is of non-survey fields. \label{fig4}} \end{figure} Each field was given a priority based on the number of exposures available, as listed in Table~\ref{tbl-1}. The total hours of exposure and the number of fields at each priority in each of the 3 selected WFPC2 Filters is illustrated in Figure~\ref{fig4}. Practically all of the higher priority data has been processed through the MDS pipeline and made available in the MDS website. The single exposure, single filter fields were given lowest priority because of the inability to remove cosmic rays and the lack of any color information. After October 1995, the MDS used only pure parallel opportunities in which a minimum of two exposures with total exposure time longer than 20 min could be taken in each of two filters, or one exposure with a total exposure time longer than 30 min in each of two filters. Special data processing code was developed to perform cosmic ray rejection using exposures through different filters. This procedure, although better than attempting to clean cosmic rays from a single exposure, is performed at the cost of losing any objects of extreme color. The pure parallel observations {\it per se} have therefore been the biggest challenge in the task of building a database using a clean statistical analysis. We will use the GWS for many of the illustrated distributions, in order to avoid complicating the discussion with effects due to changes in data quality. \section{The signal-to-noise index $\Xi\ $. } To characterize the ability of our method to extract quantitative parameter estimates we define an information index based on the signal-to-noise in the image. Since we are dealing with mostly extended images, the definition of the index is different from the signal-to-noise ratio generally used for point sources. We first define a contour around an object by selecting the subset of contiguous pixels which each contain a signal that is at least $1\sigma$ above the estimated local sky (see appendix for details). The signal-to-noise ratio of each of these pixels is computed individually. We define the signal-to-noise index $\Xi\ $ as the decimal logarithm of the integral sum of these ratios, and we have found this dimensionless quantity to be a good measure of the information content of the image, and we have used it to define thresholds within the image analysis procedure. For any particular field, exposure time and WFPC2 filter, $\Xi\ $ is linearly correlated with the magnitude of an extended image. Furthermore, it has the expected slope of 1 magnitude per 0.4dex, as shown in Figure~\ref{fig5} for GWS observations through F606W. Point-like stars follow a different sequence at brighter magnitudes. \begin{figure} \plotfiddle{ratnatunga.fig05.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ The decimal logarithm of the integrated signal-to-noise ratios of pixels within $1\sigma$ isophote ($\Xi\ $) of extended images in GWS is linearly correlated with slope of 1 magnitudes per 0.4dex. Point-like stars follow a different sequence at brighter magnitudes. Symbols represent different models as indicated in Fig~22. \label{fig5}} \end{figure} In most of the discussion on image analysis we will refer to $\Xi\ $ rather than magnitude since it is a measure of image quality which can be used without reference to exposure time, sky background and filter used. The MDS detection limit is at $\Xi\ \approx 1.6$, but the sample does contain some images with a smaller index, {\it viz.} those objects detected in the image of a different filter of the same region of sky. The completeness limit is at $\Xi\ \approx 1.8$ which is a half magnitude brighter than the detection limit. The morphology (disk-like or bulge-like) of galaxies can be determined for $\Xi\ \gtorder 2.0$ and \rm D+B\ models can be done for $\Xi\ \gtorder 2.4$, which is 2 magnitudes brighter than the detection limit. To avoid any contamination by image noise the detection limit is set at a conservative level since that was already much fainter than the image quality needed to estimate morphology. \begin{figure} \plotfiddle{ratnatunga.fig06.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Limiting magnitudes for morphology classification ($\Xi\ \approx 1.8$) as a function of total exposure time for all fields processed in 3 primary WFPC2 Filters. The best fields with 3 or more images in stack are plotted as circles. The northern HDF field is the extreme point on each graph. \label{fig6}} \end{figure} In Figure~\ref{fig6} we illustrate the limiting magnitude ($\Xi\ \approx 1.8$) as a function of total exposure time for all the MDS fields processed from WFPC2 pure parallel observations from HST Cycles 4 through 6, the GWS, the HDF, and archival cluster fields. The GWS comprises 27 WFPC2 fields, each observed uniformly with 4 exposures in each of the I (F814W) and V (F606W) filters, with total integration times of 4400 and 2800 seconds respectively and one deep WFPC2 field with $\approx 25,000$ seconds in each filter. Our object catalog for the GWS has 12,800 objects in the 27 WFPC2 fields. The percentage of images with $\Xi\ \gtorder 4.0,\ 3.0,\ 2.0\ \rm and\ 1.5$ is $0.3\% , 4.6\% , 30\% \rm and\ 67\%\ $ respectively. In these survey images, $\Xi\ \approx 1.8$ corresponds to I=24.5 mag and to V=25.2 mag. From a catalog of $\approx 10,800 $ galaxy images with $\Xi\ \gtorder 1.8$ in both F814W and F606W, 11\% of the images are fitted with two-component \rm D+B\ models, 7\% are classified as stars, 61\% are classified as either disk-like or bulge-like, 20\% are classified as generic galaxies (of uncertain disk or bulge nature) and less than 1\% remain unclassified. \begin{figure} \plotfiddle{ratnatunga.fig07.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Average number of pixels in selected region and above $5\sigma$ and $1\sigma$ contours as a function of $\Xi\ $ of images in GWS. We also show the number of pixels rejected as suspected residual hot pixels. \label{fig7}} \end{figure} As illustrated in Figure~\ref{fig7}, we find empirically that on average there are $10^{0.8\ \Xi\ }$ pixels above the $1\sigma$ contour, and $0.025\ 10^{1.1\ \Xi\ }$ pixels above $5\sigma$. When $\Xi\ \approx 2.0$ we thus have typically an image with 40 pixels above $1\sigma$ and 4 pixels above $5\sigma$. At the detection limit of our object-finding algorithm ($\Xi\ \approx 1.6$), we have typically an image with 15 pixels above $1\sigma$ and 1 pixel above $5\sigma$. Most images with $\Xi\ \ltorder 1.6$ are of regions corresponding to objects which were detected in another filter and model fits on them are typically very poor. \begin{figure} \plotfiddle{ratnatunga.fig08.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Fraction of images in GWS fields classified as different galaxy morphologies as a function of $\Xi\ $ (upper panel) and as a function of half-light radius in pixels (lower panel) for images with $\Xi\ >2.0$. \label{fig8}} \end{figure} In Figure~\ref{fig8}a we show empirically that the fraction of images for which we can use the likelihood ratio (see sec~7) to determine whether the galaxy is more disk-like or bulge-like follows the relation $\min(\max(0,(0.82\ \Xi\ -1)),1)$. Of these galaxies, the fraction of images for which we can fit a significant two-component model follows the relation $\min(\max(0,(0.46\ \Xi\ -1)),1)$. Both these relations were derived by fitting a straight line to the slopes in this figure. We can classify about 60\% of the galaxies with a $\Xi\ \approx 2.0$, and all of them with $\Xi\ \gtorder 2.5$. Hardly any galaxy with $\Xi\ \ltorder 2.2$ has sufficient signal to fit a two-component model, while 70\% of them can be fitted at $\Xi\ \gtorder 3.5$, of which however there are only very few examples in the MDS database. At $\Xi\ \approx 3.0$, about 40\% of the galaxies are modeled as \rm D+B\ . The saturation of the fraction at about 70\% is probably the intrinsic percentage of galaxies which have a significant component in both disk and bulge. In Figure~\ref{fig8}b, we show a plot similar to Figure~\ref{fig8}a, as a function of the half-light radius in pixels for images with $\Xi\ > 2.0$. For over 90\% of the images with a half-light radius $\xi\ \gtorder 2$ pixels, the image can be classified statistically using the likelihood ratio (see sec~7) to determine whether the galaxy is more disk-like or bulge-like. We can fit a significant two-component \rm D+B\ model to 20\%, 25\%, and 40\% of the galaxy images with $\xi\ \gtorder $ 2, 5 and 10 pixels respectively. None of the galaxies with $\xi\ \ltorder 2$ pixels had sufficient sampling to fit a two-component \rm D+B\ model fit. The ability and success of fitting models to an observed galaxy depend of course on both the integrated signal-to-noise index $\Xi\ $, as well as the half-light radius of the galaxy in pixels $\xi\ $. These two quantities are related to each other and to the morphology of the image. Systematic (or evolutionary) changes in the mean size and morphology as a function of apparent magnitude could slightly change Figure~\ref{fig8} if it were to be drawn for WFPC2 fields at significantly different limiting magnitudes. Figure~\ref{fig8} is applicable to exposures of about 1-hour in F814W and F606W. The difference in zero point magnitude for these filters is about the same as the mean color of galaxies, and therefore we can expect similar $\Xi\ $ for the typical galaxy. We have excluded in this figure the galaxies imaged on the PC camera in order to keep the spatial resolution constant. Each WFC pixel is $0\farcs1$. \section{Maximum likelihood estimation} Estimates of the centroid, magnitude, size, orientation and axis ratio of the observed galaxy image are initially evaluated using simple moments of the flux above the mean estimated sky, using those pixels within the $1\sigma$ contour. We next select an elliptical region around the object, ensuring that there are sufficient pixels to define the mean sky background to $\approx$ 0.5\% accuracy (0.005 mag). Any pixels within the elliptical region which are associated with some other object and which are $1\sigma$ above the mean sky are cut out from the region analyzed, together with any pixels which have been flagged as ``bad'' in the calibration procedure (\cite{1994AJ....108.2362R}). The procedure for the estimation of parameters via ``maximum likelihood'' starts by initial estimates of the model parameters from the observed moments of the image. For a given set of model parameters, the software creates a model image of the object and compares this image with the observations within the selected region (including the error image). The ``likelihood function'' is defined as the product of the probabilities for each model pixel value with respect to the observed pixel value and its error distribution; this function is evaluated as the integral sum of the logarithm of these probabilities. The likelihood function is then maximized by using a modified IMSL minimization routine (see \cite{1991AJ....101.1075R}). The 2D-image analysis used an improved version of the software developed for pre refurbishment WF/PC data (\cite{1994STScI.RT2..333R}), the catalog of which was presented in \cite{1995ApJ...453..599C}. \section{Model fitting} There are many types of empirical models that have been suggested over the years to represent galaxy profiles. We have decided in particular to use scale-free axisymmetric models with an exponential power-law profile which have been shown to fit the broad continuum of normal galaxies (\cite{1959Hbphy..53..275V,1970ApJ...160..811F}). This choice has many numerical advantages which are desirable in leading towards the development of a practical maximum likelihood fitting algorithm. Elliptical galaxies are assumed to have a $e^{-r^{1/4}}$ (bulge-like) profile, and disk galaxies a $e^{-r}$ (disk-like) profile. Each profile is characterized by a major axis half-light radius and axis ratio. Some well exposed images need to be modeled as the sum of two elliptical components. For about 4\% of the galaxy images with no central concentration, the images are better fit by a $e^{-r^2}$ (Gaussian) or even $e^{-r^4}$ profile in which the light distribution is both less centrally peaked and has no extended tail. The isophotes of some ellipticals may be {\it Boxy-distorted} (\cite{1989A&A...217...35B}) rather than the elliptical models which have been currently adopted. We will explore these and alternative models for fitting the continuum of irregular galaxies in a future paper. For a point-like stellar image (star or QSO), we need four parameters: sky background, centroid (x,y), and magnitude. For the extended images of galaxies, we need at least one extra parameter which measures the size of the image. Taking into account the image jitter (see discussion above) and any errors in the PSF, we have found it useful to adopt a Gaussian profile and to estimate a size parameter even for the point-like images, to be used as a star-galaxy separation index. This procedure also takes the stellar image analysis through the same convolutions as those done for galaxy images, enabling the likelihood functions to be compared, with some caveats. Errors in the adopted PSF would appear as an extended residual image following the model fit. This could make a bright stellar image significantly better fitted with a model image which includes an extended component. This is a particularly important issue when attempting to detect underlying galaxies in QSO images (see \cite{1995ApJ...450..486B}). \begin{figure} \plotfiddle{ratnatunga.fig09.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Plot of half-light radius in seconds of arc as a function $\Xi\ $ for images in GWS. Symbols as in Fig~22. \label{fig9}} \end{figure} In Figure~\ref{fig9}, we show for the GWS dataset a plot of half-light radius in seconds of arc as a function $\Xi\ \gtorder 2.0$. For most stellar images the estimated $\approx 0\farcs02$ or $\xi\ \approx 0.2$ WFPC2 pixel. At brighter magnitudes with $\Xi\ \gtorder 3.5$ we notice some larger objects (but $\xi\ \ltorder $ 1 pixel) which are very well separated from the sizes of galaxies. The PSF approximation adopted in the analysis is insufficient at these bright magnitudes. They could also be cases of stellar binaries which are just resolved and which at fainter magnitudes could contaminate the sample of objects classified as galaxies. \section{The parameters} We describe here the full list of model parameters in the order in which they are introduced as we increase the number fitted to an image. These are the intrinsic galaxy model parameters which are introduced before any PSF convolution or allowance for other instrumental effects such as (the small amount of) photon scattering in the CCD before detection. (1) Sky Background. The sky background is a very important part of the model estimates. A bias in the sky estimate could translate to a bias in the estimated morphology. Unlike for example \cite{1995ApJ...448..563B,1996ApJ...464...79S} we have therefore chosen to derive a maximum likelihood estimate for the mean sky background {\it simultaneous} with the other image parameters. We use sufficient pixels to ensure that the mean background sky is determined to an accuracy of 0.5\%. Typical fluctuations of order 1\% are seen in a single WFC frame. Some of this variation may be caused by the extragalactic background light (EBL) from faint unresolved galaxies. Much larger fluctuations are occasionally caused by the faint halos of nearby images, or by charge transfer problems caused by bright stars. The estimated sky backgrounds are seen to follow these variations very well. Sky background is assumed to be flat over the small region selected for analysis of each object. In the procedure we have adopted, disk-like or bulge-like model fits could possibly converge with slightly different sky backgrounds within the measurement errors. By allowing the sky to vary, we are not imposing some prior choice of sky background. The error in the sky background is then properly reflected in the error estimates for the galaxy parameters and the likelihood ratio used for morphological classification. (2,3) Centroid The $(x,y)$ centroid of the model image is in most cases very close to the centroid of the observed image. The mean errors for $\Xi\ \approx $ 2.0 and 1.6 are 0.1 and 0.2 pixels respectively. The error becomes much larger for images fainter than the detection limit (i.e. $ \Xi\ \ltorder 1.6 $). For the \rm D+B\ models we assume the same centroid for both components. The software does allow an independent offset for the center of the bulge from that for the disk (parameters (12,13)), but this has not yet been fully investigated. The extra degree of freedom resulted in poor convergence in many more galaxies than in the few which justified it. (4) Total magnitude. The adopted magnitude is the analytical total magnitude of the galaxy model. This estimate has the advantage of not needing an aperture correction as is required for a fixed aperture or isophotal magnitude. However, since the magnitude integration is over a smooth galaxy image, small errors could arise from the fact that the model may not average properly over bright regions of star formation, for example. For \rm D+B\ models the magnitude is the total for both components, a quantity better defined than the magnitudes of the individual components. The magnitudes of the individual disk and bulge components can be derived using the flux ratio ( ${\rm B/T\ }$ see below). Note that the total magnitude is integrated theoretically out to infinity. For disk galaxies practically all (99\%) of the light falls within 4 half-light radii. However for bulge-like galaxies only 85\% of the light is within 4 half-light radii, and the model needs to extend out to 19 half-light radii to contain 99\% of it. The Total magnitude for an elliptical could therefore be $\approx 15$\% brighter than when calculated by integration out to a typical isophotal detection radius, and correspondingly for the ${\rm B/T\ }$. (5) Half-light radius. This is the radius within which half the light of the unconvolved model would be contained if it were radially symmetric (an axis ratio of unity). For axisymmetric galaxies, this definition is independent of the observed axis ratio of the galaxy, a parameter which depends on the intrinsic axis ratio and its inclination to the line-of-sight. For point-like sources we fit a Gaussian profile with an exponent of 2.0, and the half-light radius is then 0.69 times the scale length. For disk-like galaxies with a profile exponent of 1.0, it is 1.68 times the exponential scale length. For bulge-like galaxies with a profile exponent of 0.25, it is the effective radius or 7.67 times the scale length. For \rm D+B\ models it is by definition still the major axis radius within which half the light of the combined profile is contained. Like the total magnitude, this is a quantity better defined than the half-light radii of the individual components. As a direct consequence of allowing the sky background to be a free parameter, we need to impose a maximum half-light radius in order to avoid this parameter from becoming meaninglessly large when a galaxy with no central concentration is fitted with a disk-like or bulge-like model. This limit has been set conservatively to equal half the maximum radius of the region selected for analysis. For $\approx 4$\% of the galaxy images, the half-light radius converges on this limit, and those models need to be rejected and flagged for fitting with a less centrally concentrated model. From numerical considerations we impose a minimum half-light radius of a tenth of a pixel on both the major and minor axes of a galaxy. For \rm D+B\ models this minimum is imposed independently for each component. This assumption does not put any significant constraints on the axis ratio distribution of galaxies with a half-light radius larger than one pixel. The quantity fitted is the logarithm of the half-light radius in seconds of arc. The half-light radius of the individual disk and bulge components can be derived using the bulge/(disk+bulge) flux ratio ${\rm B/T\ }$ and bulge/disk half-light radius ratio (see ${\rm HLF }\ $ below). (6) Orientation. The adopted position angle is that of the axis of symmetry of the galaxy model. Measured in radians in the range $[-\pi/2,+\pi/2]$, this is set equal to zero when the source is assumed to be azimuthally symmetric with an axis ratio of unity. For pre-refurbishment data with a highly asymmetric PSF, the observed orientation of the image could be significantly different from the intrinsic orientation of the fitted model. During the minimization procedure, the angle is measured clockwise from positive Y to positive X of the relevant CCD. It is then translated into a position angle as measured clockwise from North towards East using ${\rm PA}_{\rm V_3}\ $ of the HST attitude (pointing) vectors and the WFPC2 CCD plate-scale distortion map. For \rm D+B\ models we generally assume that the orientations of the disk and bulge components are the same. Since the bulge axis ratio is expected to be close to unity, any difference in orientation could be expected to be insignificant except in the brightest galaxy images. The software does allow for a difference in the orientation of the bulge from that of the disk (parameter(11)), but this too has not yet been fully investigated. (7) Axis ratio This is the ratio of the minor axis half-light radius to that of the major axis. This parameter has no units and is constrained to be smaller than unity to ensure proper definition of the major axis. For \rm D+B\ models it is defined independently for each component. If the axis ratio cannot be shown to be significantly different from unity then it is held at unity; for the one-component case, the position angle can then also be dropped as a free parameter. The size of individual pixels also imposes limits on the ability to usefully constrain an axis ratio. Note that we adopt a minimum minor axis half-light radius of 0.1 pixel; i.e. for a Galaxy with a half-light radius of 0\farcs5 this imposes a lower limit on the axis ratio of 0.02 since WFPC2 has a pixel size of 0\farcs1 . In a few rare cases, this limit was useful for the prevention of the minimization procedure from converging on an unrealistically low axis ratio. This observationally imposed limit could be taken into consideration in an analysis of the axis ratio distribution, but can practically be ignored for galaxies with half-light radii larger than 1 pixel. (8) bulge/(disk+bulge) flux ratio This is the fractional flux contribution of the bulge-like component to the (disk+bulge) light ( ${\rm B/T\ }$) in the galaxy image. It has no units and ranges from zero for pure disk-like galaxies to one for pure bulge-like galaxies. The ability to estimate this quantity depends on the integrated signal-to-noise index $(\Xi\ )$ in the image. A larger $\Xi\ $ is needed to separate out a second component with smaller fractional contribution to the total light (see Figure~\ref{fig17}). A second component is only fitted when there is a significant improvement to the likelihood ratio to compensate for the increased number of parameters. The definition has used (disk+bulge) rather than Total to allow for the possible extension of the model parameter set to a third component such as a central point source (see \cite{1996ApJ...471L..15S}). (9) Bulge axis ratio This is the ratio of the half-light radius of the minor axis to that of the major axis of the bulge-like component. In \rm D+B\ models it is often a poorly defined quantity when the disk component dominates the galaxy image, and the ratio is then adopted to be unity. We could not determine any meaningful relation between the bulge axis ratio and the disk axis ratio. Such a relation might have been expected if most disks and bulges have a typical axis ratio and were related by the common inclination to the observed line of sight. The latter does not seem to be the case. (10) The ratio of the half-light radii (${\rm HLF }\ $) bulge/disk This is the ratio of the half-light radius of the bulge-like component to that of the disk-like component. We observed that the logarithm of this ratio has a weak correlation with the ${\rm B/T\ }$ flux ratio (see Figure~\ref{fig12}). This correlation has been reported also by \cite{1985ApJS...59..115K}. For disk-like galaxies this ratio is about 0.25 and for bulge-like galaxies the ratio is about 1.6, i.e. on average, disk dominated galaxies have a disk half-light radius which is larger than the bulge half-light radius. Such is the case for our own Galaxy where this ratio is estimated to be about 0.65. However, there is a factor of 2.5 rms (i.e. one magnitude cosmic scatter) about the mean relation. It will be interesting to understand this relation using galaxy structure formation theories like those published by \cite{1998MNRAS..2XX..XXXM}. (11) Orientation difference of bulge from disk See discussion above on Orientation. (12,13) Centroid difference of bulge from disk See discussion above on centroid. \section{Optimizing the model fitted} In brief outline the procedure is as follows: The initial guess is typically far removed in parameter space from the final maximum likelihood model fit. At this point it is not useful to make any judgment about the selection of the model or the parameters to be fitted. However, testing has shown us that for 70\% of a typical catalog with $\Xi\ \ltorder 2$, we are never able to fit a significant \rm D+B\ model. These images are analyzed only as stars or pure disks or pure bulge-like galaxies and the better model is selected. In Figure~\ref{fig8} we show a histogram of the number of galaxies as a function of $\Xi\ $. We have highlighted the fraction fitted as \rm D+B\ and the fraction for which we can classify the object as being significantly disk-like or bulge-like. We first start with a disk-like model, or if $\Xi\ >2$ we attempt a 10-parameter \rm D+B\ model fit. The first fit is a special quick mode of the minimization routine (modified IMSL 9.2 ZXMIN subroutine that uses a Quasi-Newton method). This mode of minimization is fairly fast since it does not attempt to check full convergence. It reaches a point in the multi-dimensional parameter space which is close enough to the final answer to investigate the likelihood function and make some intelligent decisions. These investigations are made after each minimization, and depend on the number of parameters that were fitted. The quick mode does not use a higher resolution center (see Appendix). If a default resolution image had been used for the models, we investigate whether a high-resolution center will change the likelihood function. In over 75\% of the tests in a typical catalog reaching down to the detection limit, the absolute change in the likelihood function is less than three, which can be considered as insignificant justification for the introduction of a higher resolution center. Since we are merging parts of two independently convolved images, the high-resolution center option is only used when needed. If the half-light radius is less than $10^{-2.8+0.5 max(\Xi\ ,3.6)}$ arc seconds, the program branches to test if the object is point-like. As discussed above, we fit a symmetric 5-parameter Gaussian model to allow for image jitter and any errors in the PSF. For most images the cut is at $0\farcs1$ or one WFC pixel with a small increase for the brightest images (see fig~9). This test is done for about 30\% of the objects in the sample, although only about 8\% of the sample are eventually classified as probable point-like stellar sources, either stars or quasars. The star-galaxy classification is based on both the likelihood ratio for the best-fit galaxy model as well as the evaluated half-light radius for the object, which is typically 0.2 pixels (equal to the resolution used for the sub-pixel definition of the PSF). The next check is to see if a two-component \rm D+B\ model, if being considered, is significantly better than a single-component model with less parameters. In 60\% of the cases (for $\Xi\ \gtorder 2.0$) the numerical difference is less than 6 and this is insufficient justification for the fitting of a \rm D+B\ model. If the half-light radius is less than two pixels we again select a single-component fit. In Figure~\ref{fig8}b we show the fraction of galaxies as a function of half-light radius for which we can fit \rm D+B\ models and the fraction for which we can classify the object as being significantly disk-like or bulge-like. The peak of the distribution for which we can fit \rm D+B\ models is at about 5 pixels, and for obvious reasons we are not able to do so for galaxies with a half-light radius of less than 2 pixels. Even if the minimization gave a significant fit for a few of the latter galaxies, these fits are unlikely to be realistic models of these extremely under-sampled galaxy images. For single component galaxies we next check if the axis ratio is significantly different from unity. If not, then it is set equal to unity, and a five-parameter symmetric model is fitted to the data. For all galaxies we fit both a pure disk as well as a pure bulge model, selecting the better fit model. If the absolute value of the likelihood ratio is smaller than four, then the classification as disk or bulge is not significant and these objects are classified as generic ``galaxy''. If the object had been classified at a longer wavelength as disk or bulge, then the model output is selected to be that of the nearest wavelength for which the image was definitively classified. Otherwise, the model output is based formally on the likelihood ratio, ignoring the significance of it. For images with a sub-pixel half-light radius for which the likelihood ratio does not give a preference between star and galaxy, such objects are classified merely as ``object''. The star-galaxy separation at sub-pixel half-light radii needs more detailed investigation, particularly for the purpose of attempting to isolate an uncontaminated sample of stars needed for modeling our own Milky-Way Galaxy. \begin{figure} \plotfiddle{ratnatunga.fig10.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Comparison of the classification for images in GWS with $\Xi\ > 2.0$ in the two observed WFPC2 filters F606W and F814W. \label{fig10}} \end{figure} The image in each filter is modeled independently since the parameters in each filter need not be the same. In Figure~\ref{fig10} we compare the classification of images in GWS with $\Xi\ > 2.0$ in both the filters F606W and F814W. Most of the objects received the same classification in the two filters. As expected from Figure~\ref{fig9} there is very little ambiguity in the star-galaxy classification. \begin{figure} \epsscale{.60} \plotone{ratnatunga.fig11.eps} \medskip \caption{ Comparison of the parameter estimates for about 150 galaxies in GWS which got full 10 parameter \rm D+B\ MLE fits in both F606W and F814W filters. \label{fig11}} \end{figure} In Figure~\ref{fig11} we compare the parameter estimates for about 150 galaxies in GWS for which there is a full 10 parameter \rm D+B\ fit in both filters and a rms error estimate smaller than 0.5 in $log_e({\rm HLF }\ )$. The orientation (PA) is clearly the best defined parameter and this has proven very useful for studies of weak lensing (\cite{1996MNRAS.282.1159G}). The deviation for the total magnitude (Mag) is the color of the galaxy. The half-light radius (HLR) is the equal in the two filters for most galaxies. The axis ratios [for the disk components] (DAR) and for the bulge components (BAR) show scatter mostly from measurement error. The scatter in ${\rm B/T\ }$ flux ratio is real, and is caused by the different colors of the bulge and disk components. For galaxies which demonstrably have two components, i.e. disk and bulge, the least well defined parameter is ${\rm HLF }\ $ the ratio of the half-light radii. After a lot of effort, we have optimized an automated procedure to identify those cases for which a significant \rm D+B\ model can be fitted. We are now able to select and converge (with over 90\% success) on an unbiased estimate of the ratio of half-light radii for about half of these cases. The program determines if this quantity is unconstrained, by searching for a change in the likelihood as a function of this parameter. If the fainter component contributes less than 10\% of the light, or if the axis ratio of both components is unity, then we have generally found this parameter to be poorly constrained. In Figure~\ref{fig12} we show that the logarithm of this parameter is a linear function of the ${\rm B/T\ }$ flux ratio with a correlation coefficient of about 0.5. Bulge dominated galaxies have a systematically larger Bulge/Disk half-light radius ratio (${\rm HLF }\ $) than disk dominated galaxies. However the surface brightness limit for detection of the fainter component (see Fig.~20) probably contributes most of the observed correlation. \begin{figure} \plotfiddle{ratnatunga.fig12.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ The observed correlation with 0.4 dex rms cosmic scatter, of the Bulge/Disk half-light radius ratio (${\rm HLF }\ $) as a function of ${\rm B/T\ }$ flux ratio for galaxies in GWS. The mean value from relation (solid line) was adopted if ${\rm HLF }\ $ is unconstrained. The dashed lines are at $\pm 0.7$ dex. \label{fig12}} \end{figure} The scatter of 0.4 dex rms about the adopted mean relation (solid line) is equivalent to a cosmic scatter of one magnitude. If in the preliminary convergence the flux ratio ${\rm B/T\ } < 0.1 $, ${\rm B/T\ } > 0.9 $ or if the likelihood function was evaluated at extremes ${\rm HLF }\ \pm 0.7$ dex showed that the ratio of the half-light radius ratio was not constrained by the data, it is held fixed at the nominal value derived from the empirical relationship $$ \log_{10} ( {\rm HLF }\ ) = -0.7 + {\rm B/T\ } $$ Such relationships, although needed to facilitate convergence of the model fits at fainter magnitudes, are at best a rough approximation. However, when a parameter is unconstrained and the errors become comparable to the expected range of parameter space, this assumption does not significantly change the estimates of better defined parameters. The justification for the application of such a relationship is that it helps the routine to converge on a better defined minimum. The program may also choose to fix the bulge axis ratio, or less frequently, the disk axis ratio at unity, if either of them is determined to be not different statistically from unity. \section{Estimated errors of parameters} The covariance matrix is the inverse of the Hessian i.e. the second-order derivatives evaluated at the peak of the likelihood function. When it is normalized to have unit diagonal elements, the cross-terms then give the correlation coefficients between the estimated model parameters. If the cross-correlation terms are not large, we can expect to derive reliable error estimates for the parameters from the diagonal elements. The parameters were selected to try to minimize the covariance between the fitted parameters described above. In MLE theory, if the image being modeled is the same as the simple model assumed, then the parameter estimates and associated errors will be unbiased. However, real galaxy images which are well resolved are more complex than the simple axisymmetric image models that are assumed for MLE. The effects of spiral arms and bars on the parameter estimates are complicated and difficult to quantify using simulations. In general, we can expect that, given a sufficiently large sample, the cosmic dispersion caused by image peculiarities will be averaged out. \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig13.eps} \medskip \caption{ The running mean of estimated rms errors for parameters as a function of $\Xi\ $ for single component and \rm D+B\ MLE model fits for galaxies in GWS. See text for details. \label{fig13}} \end{figure} In Figure~\ref{fig13} we illustrate a running mean of rms errors for parameters as a function of $\Xi\ $. To first order, the logarithm of the rms error appears to increase linearly with $\Xi\ $. The errors for single component and two component \rm D+B\ fits are illustrated independently: in general, the latter errors are larger. There are a few points to notice. Firstly the sky error, of order 0.005 magnitude, is practically independent of $\Xi\ $ and is defined by our choice of the number of sky pixels to include in the MLE. The orientation and centroid position, which were held the same for both components, show no significant increase in error than a single component fit at the same $\Xi\ $. The errors in the bulge axis ratio are much larger for the two component fits. Since the rms of a random distribution between 0.13 and 1.00 is 0.25, rms errors larger than $\approx 0.1$ convey little useful information about the axis ratio. This occurs at a $\Xi\ $ of 1.93 and 2.12 for single component disk-like and bulge-like galaxies and $\Xi\ \approx 2.25$ and 2.72 for the disk and bulge components in \rm D+B\ model fits. The ${\rm B/T\ }$ errors do not become larger than 0.1 since a two-component model would not be significant if they did. The error in half-light radius is given in $log_{10}$ units. The error is 0.1 dex or 26\% at $\Xi\ $ values of 2.15 and 2.37 for single component and two-component models respectively. The ${\rm HLF }\ $ ratio, given in $log_{10}$ units, is clearly the worst constrained parameter, requiring $\Xi\ \gtorder 3.37$ for the expected error to be less than 0.1 dex or 26\% rms. The HDF superstack consisted of eleven individual HDF field pointings, and we can therefore use these to test the MLE method. We compare the MLE results for the HDF super-stack with the MLE results of the independent fits to the images of the same galaxies in each of the 11 sub-stacks. We limit the comparison to those galaxy images where the output from the sub-stacks resulted in the same morphology classification as that from the super-stack in the same filter. In the fits to all of the sub-stacks, we used the same object definition mask (see appendix) as that derived from object detection in the super-stack, together with the appropriate shifts. \begin{figure} \epsscale{.60} \plotone{ratnatunga.fig14.eps} \medskip \caption{ Comparison of parameter derived from our HDF super-stack (X-axis) with the weighted mean of the individual sub-stacks (Y-axis). See text for discussion of small systematic biases observed. \label{fig14}} \end{figure} In Figure~\ref{fig14} we compare the MLE parameters derived from the super-stack with the weighted means from the individual sub-stacks. We notice a small systematic bias: the axis ratios in the super-stack are slightly rounder and the half-light radii slightly larger, with a slightly larger ${\rm HLF }\ $ ratio. Our adopted approach to stack after shifting by closest integer number of pixels modifies the appearance of the peaked bulges. It will be instructive to see if the process of ``drizzling'' (\cite{1997STScI.CT3..518F}) and stacking with sub-pixel shifts helps to remove this effect completely. The errors in flux values of pixels in drizzled images are not independent and to use our MLE approach, the covariance error matrix for each pixel needs to be included in the evaluation of the likelihood function. Software to do this has yet to be developed. For the brighter galaxy images in the HDF it is probably better to use a weighted mean estimate of the galaxy parameters from the individual HDF sub-stacks rather than the MLE values derived for the super-stack. Although the image bias in our HDF super-stack is disappointing, it does show the power of MLE estimates to be sensitive to the true nature of the images analyzed. Of course, all these problems can be avoided by not stacking the images at all and, instead, by summing the likelihood over the individual images (\cite{1994STScI.RT2..333R}). This latter approach, however, is computationally impractical as yet. Figure~\ref{fig13} allowed us to easily estimate an expected error for a given $\Xi\ $. If the error estimate from inverting the Hessian is significantly smaller then it is unlikely to be real. This could happen for many reasons. There could be a sufficient covariance between parameters to make the diagonal only a small part of the error. The non-axisymmetric features of the galaxy image could have made a sharper dip in the likelihood function. The expected error could in fact be built into the evaluation of the Hessian at the peak of the likelihood function in order to pass over any sharp dips in the function. However, these relationships had not been derived at the time of the 1996-98 MDS pipeline processing. We find that a reasonable compromise for the current (October 1998) version of the database is to adopt a nominal expected error of half a magnitude brighter object if that is larger than the MLE error estimate from the Hessian. We find this is appropriate for all parameters except the orientation parameter, for which the original error estimates appear to be good. The orientation is not correlated with any of the other image parameters. \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig15.eps} \medskip \caption{ A histogram of the normalized deviations of the parameter estimates derived from the individual HDF sub-stacks from the value derived from the super-stack is compared with the expected standard normal distribution. See text for details. \label{fig15}} \end{figure} In Figure~\ref{fig15} we show a histogram of the resulting normalized deviations of the parameter estimates evaluated in the individual sub-stacks from the value derived from the super-stack, and we compare the results with the expected standard normal distribution. The small bias caused by stacking the parameter estimates discussed above is clearly emphasized. We see a significant tail larger than normal for the ${\rm B/T\ }$ flux ratio and for the ${\rm HLF }\ $ ratio because of the residual covariance in these parameters. The overall accuracy of the MLE parameter error estimates seems reasonable if we recognize that the simple galaxy model fitted does not include the structural detail seen in the real galaxy images at brighter magnitudes. \section{Selection effects effects due to Ockham's razor} Since the adopted procedure fits the minimum number of parameters which are required to get a best MLE fit which is statistically significant, the parameter estimates reflect that decision. \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig16.eps} \medskip \caption{ The Distributions of estimated disk and bulge axis ratios as a function of $\Xi\ $ for galaxies in GWS. For illustration the face-on (axis ratio = unity) case has been distributed randomly in the finite range [1.00,1.05] outside the fitted range [0.01,1.00]. Symbols as in Fig~22. See the text for details. \label{fig16}} \end{figure} In Figure~\ref{fig16} we show the distributions of disk and bulge axis ratios as a function of $\Xi\ $. For illustration, the face-on case (axis ratio = unity) has been distributed randomly in the finite range [1.00,1.05] outside the fitted range [0.01,1.00]. The disk axis ratio appears to be randomly distributed within the range [0.10,1.00] for $\Xi\ $ brighter than $\approx 3.0$. As images get fainter the axis ratios close to unity are found to be insignificantly different from unity. For example at $\Xi\ \approx 2$, axis ratios in the range [0.8,1.0] get set equal to unity, thus removing two parameters from the MLE fit. The same increase in errors produces a scattering of the observed axis ratios below 0.10. The bulge axis ratios show a similar distribution except that they are expectedly larger than the disk axis ratios. We also notice a number of small bulge axis ratios which are spurious and caused by barred galaxies which have not been properly included in the current MLE models. \begin{figure} \plotfiddle{ratnatunga.fig17.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ The Distributions of ${\rm B/T\ }$ flux ratio as a function of $\Xi\ $ for galaxies in GWS. For illustration, single component fits as pure disks ${\rm B/T\ }=0$ and pure bulges ${\rm B/T\ }=1$ have respectively been distributed randomly in the finite ranges [-0.05,0.00] and [1.00,1.05] outside the fitted range [0.00,1.00]. See text for details. \label{fig17}} \end{figure} In Figure~\ref{fig17} we show the distributions of ${\rm B/T\ }$ flux ratio as a function of $\Xi\ $. For illustration, single component fits as pure disks ${\rm B/T\ }=0$ and pure bulges ${\rm B/T\ }=1$ have respectively been distributed randomly in the finite ranges [-0.05,0.00] and [1.00,1.05] outside the fitted range [0.00,1.00]. The ${\rm B/T\ }$ flux ratio is distributed within the fitted range [0.00,1.00] for $\Xi\ $ brighter than about 3.0, with the understood excess of disk like galaxies. As images get fainter, ratios close to zero and unity are not observed since these galaxies do not show a significant second component. The disk component in ellipticals is `lost' before small bulges are lost in disk-like galaxies. For example at $\Xi\ \approx 2$, the observed ${\rm B/T\ }$ flux ratios are in the approximate range [0.1,0.6] \begin{figure} \plotfiddle{ratnatunga.fig18.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Distributions of ${\rm B/T\ }$ flux ratio as a function of half-light radius for galaxies in GWS. The distribution for single component fits are as in fig~17. See text for details. \label{fig18}} \end{figure} In Figure~\ref{fig18} we show the distributions of ${\rm B/T\ }$ flux ratio as a function of half-light radius. The distribution for single component fits are as those in Figure~\ref{fig17}. As galaxy images get smaller, the ${\rm B/T\ }$ flux ratios close to zero and unity are not observed since for these galaxies a significant second component cannot be resolved. Not unexpectedly, small bulge components in spirals can be inferred to have been lost from the MLE models of galaxies with half-light radii of a few pixels. \begin{figure} \plotfiddle{ratnatunga.fig19.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Distributions of Bulge/Disk half-light radius ratio (${\rm HLF }\ $) as a function of $\Xi\ $ for galaxies in GWS. See text for details. \label{fig19}} \end{figure} In Figure~\ref{fig19} we show the distributions of the ${\rm HLF }\ $ ratios as a function of $\Xi\ $. For $\Xi\ $ brighter than about 3.0 the ratio is seen to be distributed over a wide range. As images get fainter than those corresponding to $\Xi\ \approx 2.4$, the MLE routine does not estimate ratios larger than unity. This is because, as seen in Figure~\ref{fig17}, the MLE program does not resolve disk-like components in faint galaxy images dominated by bulges. We now look at the mean surface brightness within the central half-light radius ellipse. This has a constant magnitude offset from the central surface brightness of $-1.12463$ mag for disks and $-6.18126$ mag for bulges. The central surface brightness is a commonly quoted quantity, independent of axis ratio and inclination for our simple galaxy models. The advantage of discussing mean surface brightness here is that we find that the limiting mean surface brightness for morphological classification is a quantity which is about the same for disk-like and bulge-like components of the galaxy. \begin{figure} \plotfiddle{ratnatunga.fig20.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Mean surface brightness within the half-light radius ellipse as a function of major-axis half-light radius (left) and total magnitude (right) plotted independently for bulge (top) and disk (bottom) components. Cross symbol used for galaxies fitted with single component Model. \label{fig20}} \end{figure} On the left-side of Figure~\ref{fig20} we show the mean surface brightness within the half-light radius ellipse as a function of the estimated major-axis half-light radius. In the case of \rm D+B\ models, each of the components are considered separately. There is clearly a limiting magnitude for morphological classification which appears to be the same in each case, i.e. independent of whether it was a component of a \rm D+B\ galaxy model, or a single component. We illustrate this for the GWS 4-stack images of 2800 seconds in F606W. A very similar graph is seen for F814W. There appears to be a slight numerical bias for MLE to converge on integer or half-integer half-light radii at the smaller values. This bias is presumably caused by our attempt to merge in a high-resolution center (see appendix). On the right-side of Figure~\ref{fig20} we show the same surface brightness estimates within the half-light radius ellipse as a function of the total magnitude of the galaxy. Within the half-light radius ellipse of a galaxy or component, morphological classification can be done to a limit in surface brightness which is independent of the total magnitude of the galaxy up to certain magnitude limits. These two magnitude limits will be very useful as simple selection criteria in future models used to interpret the observed ${\rm B/T\ }$ distribution of galaxies and surface brightness dimming for cosmology. \section{Results} Preliminary versions of the MDS catalog have been the source of many scientific investigations: see, for example the papers on the size - redshift relation (\cite{1994ApJ...434L..55M}); angular size evolution (\cite{1995ApJ...441..494I,1996MNRAS.282.1247R,1997MNRAS.288..200R,1998MNRAS.293..157R}) axis ratio distribution (\cite{1995ApJ...445L..15I}); weak gravitational lensing (\cite{1996MNRAS.282.1159G}); luminosity functions of elliptical galaxies (\cite{1996ApJ...461L..79I}); morphological classification (\cite{1996MNRAS.281..153O,1997ApJ...476..510N,1997ApJS..111..357N,1999ApJ...510...82I}); galaxy interactions and mergers (\cite{1997ApJ...480...59N}); compact nuclei (\cite{1996ApJ...471L..15S}) the HST MDS cluster sample (\cite{1998AJ....116.2644O}) and a study of high-redshift clusters (\cite{1998AJ....116..584L}). The catalog used in these analyses was mostly based on the star, disk or bulge model that best fit each object. Most of the previous analyses can be repeated on the new catalog and refined using the \rm D+B\ models for the brighter sample. We do not, however, expect any significant changes to these previously reported results. It is especially interesting to look at results on the two observables which have not been previously measured for large numbers of galaxies, especially in the magnitude range observed here, viz. the ${\rm B/T\ }$ flux ratio and the Bulge/Disk half-light radius ratio (${\rm HLF }\ $). In fact we need to apply the same procedure to a large sample of bright nearby galaxies like those from the SLOAN digital sky survey (\cite{1998AJ....116.3040G}) in order to establish the behavior of these parameters on galaxies in the local universe. \section{Surface Brightness} \begin{figure} \plotfiddle{ratnatunga.fig21.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ The running mean of surface brightness as a function of total magnitude for galaxies in GWS (left) compared with HDF (right). Components from \rm D+B\ models illustrated with thicker line follow same sequence as galaxies fitted with single component models. See text for discussion. \label{fig21}} \end{figure} We have made plots similar to Figure~\ref{fig20} for much deeper observations such as the Hubble Deep Field (HDF). In Figure~\ref{fig21} we show a running mean of surface brightness as a function of total magnitude for the GWS galaxies on the left-side and compare it with those estimated for the HDF on the right-side. This graph illustrates Freeman's result for disk galaxies (\cite{1970ApJ...160..811F}); the mean is the same, indicating that the observed distribution of surface brightness is intrinsic to the galaxies, with a cosmic dispersion of only about 1-magnitude. The expected trend of surface brightness dimming as the mean redshift increases for galaxies with fainter total magnitude is also seen. Correcting by $-1.12$ mag., we estimate that the mean central surface brightness of disk galaxies is 20.6, 21.4 mag in F814W, F606W for the GWS and 21.0, 21.8, 22.4 mag in F814W, F606W, F450W for the HDF. It is interesting that the mean surface brightness is the same for galaxies fitted as pure disk, as it is for the disk component of \rm D+B\ galaxies. For bulges the scatter appears to be very much larger and our observations in GWS do not reach the limiting mean surface brightness bulges. Consequently, the mean for bulges in the HDF is about 2-magnitudes fainter than for bulges in the GWS fields. \section{Galaxy Color} \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig22.eps} \medskip \caption{ Color of GWS galaxies as a function of the F606W apparent magnitude. The dotted vertical line is drawn at the adopted the limiting magnitude for morphology classification at $\Xi\ \approx 1.8$. \label{fig22}} \end{figure} In Figure~\ref{fig22} we look at the color of GWS galaxies as a function of the F606W apparent magnitude. We have shown all 6 classifications. The dotted vertical line is drawn at the observed completeness magnitude ( $\Xi\ \approx 1.8$ ). Most of the objects which were not classified morphologically are fainter than this limit. Furthermore, some of the images fainter than this limit and which have been classified as point-like are probably faint galaxies rather than stars. We have chosen not to filter the MDS catalogs brighter than this limit in order to avoid additional censorship of the sample in statistical analyses. All parameter estimates for galaxies fainter than this limit (i.e. magnitudes corresponding to $\Xi\ \ltorder 2.2$) can be used for statistical analyses only, i.e. studies should not be focused on individual galaxies, particularly any outliers of such a distribution. \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig23.eps} \medskip \caption{ Color of the bulge-components of galaxies with the corresponding disk-components for galaxies in GWS which were fitted with \rm D+B\ models in both F814W and F606W. \label{fig23}} \end{figure} In Figure~\ref{fig23} we compare the color of the bulge-components of galaxies with the corresponding disk-components for GWS galaxies which were fitted with \rm D+B\ models in both F814W and F606W. It appears that the colors of disk and bulge components of many galaxies are similar (the dotted line), although bulges are observed to be systematically redder, as expected, except for a few isolated cases. \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig24.eps} \medskip \caption{ The colors of the disk and bulge components as a function of the ${\rm B/T\ }$ flux ratio for galaxies in GWS \label{fig24}} \end{figure} In Figure~\ref{fig24} we look at the color of disk and bulge components of galaxies as a function of the ${\rm B/T\ }$ flux ratio. As may be expected, the disk components of galaxies appear to become $\approx 0.5$ mag redder as we follow the plot from disk-like to bulge-like galaxies, with a cosmic scatter of 0.45 mag. The colors of bulges remain practically the same $\approx 0.2$ mag., with a larger cosmic scatter of 0.6 magnitudes. \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig25.eps} \medskip \caption{ The colors of galaxies in GWS as a function of the Bulge/Disk half-light radius ratio (${\rm HLF }\ $). \label{fig25}} \end{figure} In Figure~\ref{fig25} we look at the color of galaxies as a function of the ${\rm HLF }\ $ ratio. It appears that redder galaxies have a smaller ratio. Careful statistical analysis is needed to ensure that this is ``real'' and is not caused by a selection effect in which the GWS galaxies were sufficiently bright that \rm D+B\ models could be fitted to images in both F814W and F606W filters. \section{Conclusions} An automated maximum Likelihood procedure has been developed to calibrate, detect and quantitatively measure objects in the HST WFPC2 fields. The procedure measures the parameters of faint galaxies, despite the potential difficulties related to the undersampling in WFPC2. \rm D+B\ models are now fitted routinely to the brighter galaxy images as a part of the MDS pipeline. A \rm D+B\ galaxy model, a pure disk, a pure bulge or a star model is chosen automatically using likelihood ratio tests. Classification is done for images with significant confidence. Most HST MDS fields observed in 1994-1997 have been processed, resulting in a catalog of over 200,000 objects which have been put on the MDS website with a searchable browser interface. Clicking on a stack image will pick out and display the maximum likelihood model fit and the parameters for that object. The statistical properties of the HST-MDS Catalog has resulted in many publications and comparisons with models of galaxy evolution will continue. \acknowledgments This paper is based on observations with the NASA/ESA {\it Hubble Space Telescope}, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. The Medium Deep Survey was funded by STScI grant G02684 {\it et seqq.} and by the HST WFPC2 Science Team under JPL subcontract 960772, under NASA contract NAS7-918. Some of the data was also processed under the STScI archival grants GO6951, GO7536, and GO8384. We acknowledge the multiple contributions of Dr. Stefano Casertano, Dr. Myungshin Im, Mr. Adam Knudson and Dr. Lyman Neuschaefer who were associated with the MDS pipeline processing and analysis. We also thank the rest of the original MDS Co-I team, including Dr. Richard Ellis, Dr. Gerard Gilmore, Dr. John Huchra, Dr. Garth Illingworth, Dr. David Koo, Dr. Antony Tyson and Dr. Rogier Windhorst for their contributions to the program. \centerline{\bf Appendix - MDS Pipeline} \bigskip \section{Association of WFPC data and MDS Field names} The MDS database was maintained and updated using Starview (\cite{1994STScI.AM5...23F}). Observations are assigned an alphanumeric 5-character name that is based on Galactic coordinates as described below such that fields which are from the same region of the sky are associated by name. The choise of the individual characters in the name is as follows: 1) The first letter of the name, 'u', is the HST instrument letter assigned to WFPC2 observations. It was 'w' for older WF/PC data. 2) Galactic Latitude from `a' in the south to `z' in the north in equal steps of sin(latitude), using numeric index [6-9,0-5] within 16\degpt1 from the Galactic plane. 3) Galactic Longitude using sequence `1-5,a-z,6-9,0' in steps of 10\degpt0 such that numeric indexed fields are towards the Galactic Center. The Galactic caps within 3\degpt9 from pole are assigned ``a-'' and ``z\_'' SGP for NGP respectively. 4) Chronological sequence of primary target within the 31 degree$^2$ cells defined above, based on coordinates. Observations within a 0\degpt5 radius are assumed to be of the same target. 5) Chronological sequence of Association around the same primary target set. These fields may overlap each other. The program using a list of all pure parallel WFPC2 GO and GTO observations assigns the names. We have not included the STScI UV-survey program (pid=6253) or the current archive program (pid=7909) for all parallel WFPC2 data since February 1998. Every dataset in an associated group is allowed to be a maximum of 8\farcs0 (10\% of WFC CCD width) from any other dataset in the same association. This range is sufficient to associate all WFPC data taken in parallel with a STIS or NICMOS observations, which are dithered, say within a 5\farcs6 square. For most cases the ${\rm PA}_{\rm V_3}\ $ orientation is identical. If it is not, then we ensure that the difference in rotation is less than 0\degpt03 . This ensures a 1-1 mapping of the pixels, keeping any effect caused by the small rotation or differential distortion to be under about 0.5 pixels, the maximum error made by adopting integer truncated pixel shifts between images in a stack. \begin{figure} \plotfiddle{ratnatunga.fig26.eps}{6.0in}{-90.}{75.}{75.}{-288}{432} \medskip \caption{ Repeat observations give a ring of pure parallel WFPC2 fields around FOS calibration star BD+28-4211. \label{figA1}} \end{figure} Around some objects such as the FOS calibration star BD+28-4211 MDS has many repeat observations as illustrated on Figure~\ref{figA1}. \section{Calibration procedure} Briefly, the calibration procedure is as follows. The WFPC2 images are calibrated using the best available calibration data. We adopt the STScI static mask, super-bias and super-dark and flat field calibration files created for the HDF. Tables of hot pixels from STScI are used to correct fluxes in fluctuating warm pixels for the period of observation. Correction is made to ensure that the noise from any residual warm current is smaller than the read noise. Hot pixels, which cannot be corrected to that accuracy, are rejected. Saturated pixels and pixels with large dark current are flagged as bad and ignored. No attempt is made to interpolate over them. The software has been specifically developed to recognize the existence of missing pixel values. In general we have more than one exposures in the same filter, in the same field, to reject the numerous cosmic rays by stacking exposures with a $3\sigma$ clip. We use a corrected version of the IRAF/STSDAS combine task. See \cite{1994AJ....108.2362R} for a detailed discussion of various aspects of the stacking procedure and the statistical errors which are corrected in the ``combine'' algorithm. A error image is also generated which computes the rms error from the noise model, taking proper account of pixels rejected by cosmic rays, the dark current, flat-field. Shifts between images were determined by cross-correlation of the images. The coordinates listed in HST WFPC2 image and/or jitter file headers are often found to be insufficiently precise for the process of image stacking (\cite{1997STScI.CT3..361R}). The shifts are determined to an estimated rms accuracy of 0.1 WFC pixels. To avoid interpolation (which spreads the charge from cosmic rays, charge that is otherwise well confined), exposures are stacked with shifts corresponding to the nearest integer number of pixels, without any rotation or drizzling. Drizzled images (\cite{1997STScI.CT3..518F}) are most useful for very deep exposures like the HDF, which do not occur in pure parallel observations. Drizzling causes the errors in adjacent pixels in the image to become correlated and significantly complicates a proper statistical analysis of the image. A mode offset is employed to allow for changes in the sky background in different exposures due to changes in the fluorescent glow and scattered Sun/Earth light. The calibration accuracy is partly limited by the fluorescent glow. This can contribute as much as 50\% of the dark current, and is strongly correlated with the cosmic ray activity during the WFPC2 exposure, which in turn depends on the particular orbit. However, except for very deep stacks like the HDF, the noise created by improper correction of this fluorescent glow results in a term which is small compared to other noise terms. We next remove any large-scale gradient from the faint outer regions of bright galaxies, for which the nucleus was probably the target of the primary observation. The four CCD images of the WFPC2 are first oriented and merged along the pyramid edge and a single 2nd order 6-parameter polynomial surface is fitted across all four. This surface is then subtracted from each of the individual images and this automated procedure is iterated 2 or 3 times until no gradient is visible. Only about 4\% of the processed MDS observations required this gradient removal. After stacking, the image is multiplied by a selected factor, which is a power of 2, followed by an integer truncation and division by the same factor. This makes the images compressible without any loss of useful information, since the differential values before and after this process are much smaller than either the accuracy of the calibration or the averaged read-noise. The selected power depends on NCOMB, the number of images stacked and we adopt the function ${\rm nint}(\log_{2}(8\ {\rm NCOMB}))$, i.e. $2^3$ for a single image and $2^6$ for a deep 6-stack. The estimated rms error has an expected dynamic range of 0 to 25 ADU and the accuracy is unlikely to be better than 0.01 ADU. Therefore the rms ADU error image is multiplied by 100 and truncated to the nearest integer, in order to generate a short Integer image which is half the size when uncompressed than the size of the corresponding image of real numbers. \section{Object detection} The MDS pipeline was used to process only the typical field in which crowding was not a problem. We selected sparse fields in which the number of pixels $4\sigma$ or more above sky is typically under 5\% of the total pixels in the field. We classified as non-survey and excluded from the MDS pipeline image analysis all low galactic latitude fields with lots of stars, and those fields close to Globular Clusters and local Group galaxies with, say, more than about 1600 objects detected. Non-Survey MDS fields were analyzed independently by other members of the team. \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig27.eps} \medskip \caption{ Number of objects detected as a function of of pixels above $4\sigma$. The fields selected to be part of MDS survey have less than about 5\% of pixels above $4\sigma$. WFPC2 images with no cosmic-ray split have been indicated with crosses to show the systematic smaller number of object detections because of the attempted cosmic ray cleaning. \label{figA2}} \end{figure} In Figure~\ref{figA2} we illustrate the number of objects detected as a function of the number of pixels above $4\sigma$. The figure shows that most of the fields selected as part of the MDS survey have a smaller fraction of pixels over $4\sigma$ and a smaller number of objects detected than in non-survey fields. In both cases, images with no cosmic-ray split have been indicated with crosses and show a systematic smaller number of object detections because of the attempted cosmic ray clean out. Objects are located independently on each image using a `find' algorithm developed for HST-WFPC data. This algorithm does not do any pre-convolution of the data, so that it is specifically designed to be insensitive to hot pixels and missing pixel values. It is based on finding local maxima and mapping nearby pixels to the central object, and then selecting those detections which are significantly above noise. The detection threshold algorithm originally developed for pre-refurbishment WF/PC data was optimized for WFPC2. This resulted in the location of a practically identical list of objects in the overlapping region of three WFPC2 MDS parallel fields USA0[1-3] observed in June 1994. To ensure that we do not break up bright galaxies into small regions of star formation, we adopted an object resolution of $0\farcs5$, and small regions within this radius were allowed to merge with a brighter center. A larger radius of $1\farcs0$ was used for WF/PC data. This algorithm has been observed to locate real objects with as much or better efficiency as the FOCAS algorithm (\cite{1979ApJ...230L.153T}), which was developed mainly for ground-based data. The MDS `find' algorithm generates both a catalog and a `mask' image, which associates each pixel with one object. This is a short integer file since the MDS pipeline assumes that there are less than 10K objects in a single WFPC2 field suitable for analysis. The stacking and initial object location procedure is a fully automated first step of the MDS pipeline. After the initial find, we have the only interactive part of the operation. We first look at the exposure, and confirm that it satisfies our requirements on inclusion into the MDS. A typical MDS survey field is uncrowded, with about 400-800 objects detected in the 5 arc-min$^2$ field. We also exclude from the MDS catalogs those objects with a centroid within 10 pixels ( $1\farcs0$ ) from the pyramid and CCD edge, thus reducing the area surveyed by about 5\% from 5.03 arc-min$^2$ to 4.77 arc-min$^2$ per WFPC2 field, and causing a $2\farcs0$ wide gap in the shape of a cross in the center of each field. Rapid changes in the image distortion and the PSF ( a residual consequence of the original HST Spherical Aberration ) make the edge a very difficult region for reliable quantitative analysis. The next operation is to fix up the mask for any bright objects which have been over resolved, or to delete any ghost images or extremities of bright stellar diffraction spikes which have been spuriously detected as objects. The detection algorithm has been optimized to work best at intermediate to faint magnitudes at the cost of over resolving a few bright objects. The numbers plotted in Figure~\ref{figA2} are raw counts before cleanup. The spurious detections are flagged with an interactive cursor for rejection or merger with the central image. This interactive operation takes about 30 min per stack and is done with a well defined set of guidelines which were originally developed for WF/PC data and modified appropriately for WFPC2. The object detections in the various filters are then matched by software, and a single catalog is created, together with a revised mask for each image, so that corresponding pixels in the different filters are associated with the same object. Looking at a grid of the individual object detections, the final masks are inspected and the procedure is iterated as required to ensure that the object definitions as encoded by the final masks are acceptable. This is the conclusion of the calibration and object detection phase of the MDS pipeline. The object detection algorithm and search thresholds were kept unchanged over the four years of the MDS. When new calibration data became available we recalibrated the data and created a new stack to obtain a slightly lower noise in the image. We however do not redetect objects, however. The masks remain constant, and after the field has been setup in the MDS database, model fits to any objects can be reprocessed with practically no human intervention. When there has been a significant improvement in the calibration or the fitting software, the whole database is reprocessed to obtain an improved version of the catalog, which is uniform over the whole period of observation. This has, however, become practical only after we obtained a SPARC Ultra-1, which on its own can do the reprocessing of the current database of over 400 fields in about two months. All of the MDS fields have been reprocessed with the last (July 1996) version of the MDS image analysis software. The shifted stacks were refitted after they were improved in July 1997 using inter-image shifts derived from cross-correlation analysis. \section{Definition of the object region for analysis} Most galaxies are analyzed by picking out a 64-pixel square region centered on the galaxy. The very few images (on average about 3 galaxies per WFPC2 field or 0.67\% of the catalog) which are larger are analyzed as 128-pixel square images and in an extreme case as 256-pixel square region. The integral power of two in the region size in pixels was chosen for efficient convolution of models by fast Fourier transforms (FFT). An initial guess of the local sky background is determined from an algorithm that determines the sky using an adaptation of the iterative, asymmetric clipping procedure as described by \cite{1984AJ.....89..176R}. In the very few cases for which it is detected that the local sky is poorly defined (large rms and skew in distribution) then the global sky is adopted as the initial guess. We next use the mask of detected objects generated by the MDS `find' program in order to define a $1\sigma$ contour around the object. This is done by selecting the subset of pixels which are next to each other and are $1\sigma$ above the estimated local sky. Despite careful `dark' calibration and correction for suspected hot and warm pixels, pixels with fluctuating dark current are seen to leave a few ``hot'' pixels in the image. Since they could contribute significant flux compared with the flux of some of the faint images, any isolated pixels in the region outside the $1\sigma$ contour and over $5\sigma$ above immediate neighbors were located and were assumed to be hot pixels and rejected. This algorithm detected hot pixels in only 25\% of the images and in these cases found on average only 5-pixels in a 64 pixel square region (See Figure~\ref{fig7}). These values are for the GWS taken with the WFPC2 before it was cooled down from $-78^o C$ in April 1994, and for which warm-pixel corrections are not available. There are many less hot pixels in the newer data taken at $-88^o C$. The initial guess of the local sky and the choice of pixels within the $1\sigma$ contour associated with the object are factors which influence only the region picked out for analysis. The pixels within the $1\sigma$ contour get no different treatment when the likelihood function is integrated. \section{The observational error distribution.} The presence of cosmic rays makes the observational error distribution of the raw observation non-Gaussian. We have found that the cosmic ray contamination can be represented by a Weibull distribution with index 0.25. In theory, the likelihood function can be defined by taking the model all the way back through the calibration procedure in order to make the comparison by summing over independent raw observations without any stacking. If this is done, one can take proper account of the effect of telescope ``breathing'' which results in slight changes to the observed PSF. One can also allow for contamination by faint cosmic rays and even any analog to digital conversion errors (a problem mainly for old WF/PC data) on the observational error distribution (\cite{1994AJ....108.2362R}). However, after extensive software development and investigation using simulations, we found that this analysis of raw observations and the use of a complex error distribution gained only about 0.15 magnitudes in quality of morphology classification over the very much simpler analysis of calibrating and stacking the image to remove cosmic rays and the assumption of a Gaussian error distribution. With the latter approximation the log likelihood function is equal to $$-0.5(1+\log(2\pi))\chi^2=-1.42\chi^2 .$$ Maximizing the likelihood function is then identical to minimization of $\chi^2$. We have currently chosen to use the simpler analysis since the very slight improvement in results does not justify the very large increase in computation. \section{Generation of model images for comparison with observation.} The creation of the model image is the most technical and computer intensive part of the procedure. On average, of order 700 model images are used by the minimization routine to converge on the best-fit model of a single object. Since our minimization routine uses derivatives, an efficient high precision algorithm is required. For under-sampled images like those from WFPC2, sub-pixelation is very important, particularly close to the central peak of the galaxy image. We have developed a procedure which is automatically optimized by the algorithm by testing the evaluated likelihood function on the image being analyzed. We find that for many images the central pixels of the model image convolution needs to be done in sub-pixel space, and then block averaged for comparison with observation. \section{The creation of the image.} In order to ensure that the evaluated likelihood is a smooth function of all the model parameters, require computation of the model image at much higher resolution than that observed, particularly in under-sampled regions close to the center. The image models we have adopted are scale free and have an axis of symmetry. In order to minimize computation and make use of this symmetry, we therefore first evaluate the image by adopting an origin at the middle of central pixel of the array. The outer regions of the image are evaluated without sub-pixelation. If the models are scale free, then the outer regions can be multiplied by a constant factor to obtain sub-pixeled values for the inner pixels. For example, we generate an 81 pixel square image by first computing pixels outside the inner 27 pixel square. Using the axis of symmetry, only half these pixels need to be computed. Then each pixel outside the inner 9 pixel square and within the 27 pixel square region is integrated with 3x3 subpixelation by integrating the 9 pixels at 3 times the radius and using a scale factor appropriate for the selected model. Following this step, each pixel outside the inner 3 pixel square and within the 9 pixel square can be integrated with an effective 9x9 subpixelation, and the region outside the central pixel and within the inner 3 pixel square can be integrated with an effective 27x27 subpixelation. Finally the central pixel with an 81x81 subpixelation is integrated from the rest of the whole image and the contribution for the very central subpixel. In this way the model image that is created has a very high degree of subpixelation for the inner pixels at practically no extra computation. In this example, the central 64-pixel square region used gets computed as a 280 pixel square image with increasing subpixelation towards the center. The image is effectively computed at 39200 points at the cost of 2917 evaluations, more than an order of magnitude increase in speed. This approach of sub-pixelation can be used on any scale free model, even if not elliptical. \section{The Point Spread Function} Selection of the Point Spread Function (PSF) is not easy. The choice is between, using observed PSF's of well-exposed stars, or model PSF's from programs such as tinytim (\cite{1992STScI.TTM....1K}). Observed stellar PSF's are under-sampled, have random observational errors, and are not always available close to the image being analyzed. A compiled grid of stellar PSF's from various observations has systematic errors comparable to the small systematic errors seen in model PSF's. tinytim PSF images have the added advantage of being able to be generated as a sub-sampled image without observational jitter or the scattering in the WFPC2 CCD photon detection (see below for details). Convolution of the WFPC2 model image is best done in sub-pixel space where it is less under-sampled. Tinytim (\cite{1992STScI.TTM....1K}) PSF's are evaluated with 3 and 5 times sub-sampling for the PC and WFC CCD chips respectively. The 267 square PSF images are stored in the same data file format as the observations in a 3 by 3 PSF grid for each chip, and centered on the image at the pixel for which they were evaluated. A PSF grid image data file is made for each filter used in the observations. In the image analysis we choose from the grid the PSF for which the center is nearest to the location of the object. A 3 by 3 grid is sufficient for the corrected optics of WFPC2. A 11 by 11 grid with no sub-sampling was used for pre-refurbishment WF/PC data for which under-sampling of the extended PSF due to the spherical aberration was relatively less of a problem than the rapidly changing PSF as a function of the location on the chip. We have so far ignored the changes to the PSF caused by the gradual shift of the mean focus between resets (maximum of 6 microns) and the telescope breathing which has a rms of ( 3 microns ). This in itself is a complicated issue when a stacked image is used in the analysis. The focus of every exposure in the stack cannot be assumed to be the same. Simulations have shown that for typical extended images with a half-light radius larger than say 2-pixels, model parameters derived using slight changes to the PSF are well within the parameter error estimates. For extended images, deviation of a galaxy from the simple model assumed could give larger errors to the parameter estimates. \section{Convolution} Convolution is done with IMSL 9.2 FFT routines. For most images a 64 pixel square array is used. We correct for any aliasing by generating a 16 pixel square image with a factor of 4 lower resolution and convolving it at the center of a 32 pixel square image where the region outside the central 16 pixel square is set equal to zero. The flux which gets convolved out to this border region is a sufficient estimate of the alias, and subtracted appropriately from the convolution of the original 64 pixel square array which is not surrounded by a zero border to prevent aliasing. This correction procedure takes only 1.25 times longer, rather than the 4 times increase needed to surround the 64 pixel square array with a zero buffer, and doing a 128 pixel square array convolution. The subpixelation chosen is such that the region of the image selected for analysis will fit within a 64-pixel square array at the subpixel resolution. However, for WFPC2 images this was found to be insufficient for many highly peaked images that also covered most of the 64-pixel square array without any subpixelation. In these cases, we generate a second image for just the central pixels at the sub-pixelation used for PSF and then convolve this image as a 32-pixel square array, and correct it for aliasing effects using a similar procedure as for the main image. The high-resolution image replaces the central 5-pixel square region of the model image. Since the center region after convolution is at 5-pixel subpixelation, the image can be shifted to the required center and block-averaged to the observed pixel scale using a simple algorithm that assumes a uniform flux distribution within each subpixel. Including such a higher-resolution center takes factor of 1.25 longer in CPU time, and used as required based on changes to the likelihood function. \section{Scattering at time of photon detection in WFPC2 CCD.} There is a non-negligible probability that a photon will be counted in a pixel adjacent to that in which it should have been detected. For a highly peaked source such as a stellar image, the location of the centroid within a pixel will govern the spill over to the adjacent pixels. The photon detection scattering of the model image therefore needs to be done at the observed image resolution and cannot be incorporated within a sub-sampled PSF. After shifting the center and block averaging down to the size of the observed pixels, the image is convolved by a 3-pixel square kernel to allow for the photon detection scattering in WFPC2 data. The symmetric kernel adopted has [center,side,corner] values of [0.75,0.05,0.0125] and is as recommended in tinytim 4.1. We have compared this kernel with an azimuthally averaged kernel like that which was recommended in tinytim 4.0 with elements [0.5628,0.0937,0.0156]. We find that the revised, more centrally peaked kernel yields better model fits to a sample of stars. However, the combination of PSF and scattering is still not perfect and leaves residuals of about 5\% to 10\% which are significant in the brighter images. \section{Image Jitter} Since the tinytim PSF models are generated without including any contribution from telescope jitter, the intrinsic half-light radius estimated for a point source is non-zero. For a WFPC2 primary pointing in fine lock we expect about 10 mas (milli-arc-seconds) of telescope jitter. However, for parallel WFPC2 observations it could be larger because aberration corrections made for the primary instrument are slightly different from those required for the WFPC2. Since jitter data is still not available for all WFPC2 observations, they have as yet not been incorporated them into the MDS pipeline ( see \cite{1995STScI.CT2..351R} for details). Shifts between images were determined by cross-correlation of the images. Any small sub-integer shifts between the images that are ignored in the stacking procedure would also increase the effective jitter in the stacked image. Pointing Errors of 10 mas are possible at the times of target reacquisition between consecutive orbits of the HST and 20 mas if the target is reacquired after some other observation. Any systematic radial errors between the actual PSF for the observation and the tinytim PSF model would also translate to a larger effective half-light radius. We typically estimate about 20 mas for unsaturated stellar images, growing to as large as 60 mas for the very bright saturated stars. The half-light radius computed by the program is not corrected as yet for jitter this correction which not have any significant effect, except on those images with a half-light radius smaller than a pixel. \section{The flow chart of the fitting and the output files} The input data for the program are four STSDAS/GEIS images for each filter, the calibrated image, corresponding error image, object definition mask image, the PSF grid image, and an ASCII data file with keyword information about the pointing and global noise characteristics of each calibrated stack. These files are identified in the header of the catalog that then identifies the object to be analyzed by the group and coordinate of the centroid, together with the mask number. In a special mode, it is possible if necessary to identify a small group of adjacent objects resolved with different mask numbers as a single object in the analysis. The program fits all available images of the object in the different filters and outputs an ASCII data file with the fitted parameters, covariance matrix and other information about the likelihood ratio, and the sequence of intermediate results and tests. Catalogs for all objects in a field or a number of fields can be obtained using a keyword search of these data files. Also created is a FITS data image for each object. This file has the format of a grid with a single row of 7 images for each filter, starting from the longest wavelength at the top and progressively shorter wavelengths below it. From left to right the images on a single row are (1) Full image area read from the stack as observed. (2) Selected region for analysis, with any adjacent objects masked out. (3) Maximum likelihood model image, following PSF convolution. (4) Maximum likelihood model image. (5) Residual image. (6) Error image. (7) Object mask image. To make the FITS files short integer and compressible, the sky subtracted stack images and the residuals are multiplied by 10 and transformed to short integer. The error image and mask are in the integer format used for these images (see sec~15). \begin{figure} \epsscale{.75} \plotone{ratnatunga.fig28.eps} \medskip \caption{ An example of a well exposed galaxy image from the MDS database as displayed via the MDS website. The dark image is in color and is generated from the available filters using fixed color transformation algorithm. \label{figA3}} \end{figure} We show on Figure~\ref{figA3} an example of a well exposed galaxy image from the MDS database as displayed on the MDS website. Note that even the fainter F606W image which has $\Xi\ \approx 2.5$ is in the range signal-to-noise index which gives reasonable \rm D+B\ decomposition. As of October 1998 similar output for over 200,000 galaxies and stars had been made available on 19 CDROMS installed on a `Jukebox' at STScI. Analysis of an image as a star, disk-like or bulge-like galaxy is fairly straightforward. The only parameter that may be dropped is the axis ratio and in that case the orientation parameter is also not needed and the number of parameters fitted drops from 7 to 5. The software can also use any profile index or even attempt to optimize its value as was done for images in the Uppsala galaxy catalog (\cite{1989ESO...XXX....1L}). However numerical simulations have shown that the profile index gives a measure of the ${\rm B/T\ }$ flux ratio only if the axis ratio of the two components are very similar. Else, the minimization procedure computes an index which is not within the range of one (for pure disk-like) and a quarter for (pure bulge-like). The value is often larger than two, as seen for the Uppsala galaxy catalog in which the index seems to have been constrained to be smaller than three for the same reason. The \rm D+B\ analysis is much more complicated. From the very choice and definition of the fitted parameters, to the automated selection of them to ensure convergence, has been a long investigation based on both fits to real data and to realistic simulations. Almost like the minimization process, getting close to the answer was much faster than checking and ensuring that the algorithm was optimum. Getting an algorithm that worked on a majority of the images has been tested and in use for sometime. The final optimization became practical with the aid of a SPARC Ultra-1 which is more than an order of magnitude faster than a SPARC-2 on which the programs were developed. Automation gives uniformity at the cost of a few complicated cases (particularly at bright magnitudes) where a decision made by human eye would probably be different. The program was improved constantly till about July 1996 to reduce the percentage (currently about 2\% ) of fits which are in error. All of the MDS database has been reprocessed with the improved version of the program logic.
\section{Introduction} Recently, several papers \citep[e.g.][]{JOstrikerPSteinhardt95a,MTurner96a,JBaglaPN96a,LKrauss98a, MWhite98a,MTegmarkEH98a,MTegmarkEHK98a,DEisensteinHT98a, DEisensteinHT98b,MWebsterBHLLR98a,SBridleELLHCFH99a,GEfstathiouBLHE99a} have pointed out the advantages of joint analyses of cosmological parameters, i.e.~combining the information from more than one cosmological test. Ideally, such tests would be complementary, i.e.~the degeneracy in the $\lambda_{0}$-$\Omega_{0}$ plane would be in orthogonal directions. However, even if this is not the case, indeed, even if the degeneracy is exactly the same, the combination of tests can tighten the constraints as well as serve as a consistency check. Here, I discuss constraints on cosmological parameters in the $\lambda_{0}$-$\Omega_{0}$ plane from a joint analysis of gravitational lensing statistics \citep[hereafter Paper~II]{PHelbigMQWBK99a} and the magnitude-redshift relation for Type~Ia~supernovae, using the results of the Supernova Cosmology Project and the High-Z Supernova Search Team \citep[hereafter SCP and HZSST, respectively]{SPerlmutteretal99a,ARiessetal98a}. Although both tests are preliminary in the sense that they will improve with more and better observational data, the time is already ripe for a joint analysis, to demonstrate both what already can be done and how each test can be improved to lead to tighter joint constraints. The plan of this paper is as follows. In Sect.~\ref{theory} I briefly review the basis of each of these two cosmological tests. In Sect.~\ref{results} I present and discuss the joint constraints. Sect.~\ref{conclusions} provides a summary and conclusions. \section{Theory review} \label{theory} I here use the notation of \citet{RKayserHS97a} with regard to cosmology and refer the reader there for the relevant definitions. In particular, $\Omega_{0}$ refers only to `ordinary matter' and $\lambda_{0}$ is the normalised cosmological constant, such that $\lambda_{0}+\Omega_{0}=1$ for a flat universe. Both gravitational lensing statistics and the magnitude-redshift relation are `classical' cosmological tests, i.e.~the theoretical dependence of an observable quantity on redshift is compared with observations. This is done straightforwardly in the case of the magnitude-redshift relation, and in a somewhat more roundabout way in the maximum-likelihood analysis of gravitational lens statistics used here. The redshift range probed by the magnitude-redshift relation extends at present out to $z \approx 1$. In the case of lensing statistics, the source population extends to quite large redshifts ($z \approx 4$) although the redshift range of significant optical depth is smaller. Thus, the two tests are both `global' rather than `local' cosmological tests and probe similar, though not identical, redshift ranges. Otherwise, the tests are completely independent. The $m$-$z$ relation is concerned essentially only with the luminosity distance $D^{\mathrm{L}}$ whereas lensing statistics deal with several different angular size distances (between observer and lens ($D_{\mathrm{d}}$), observer and source ($D_{\mathrm{s}}$) and lens and source ($D_{\mathrm{ds}}$)) \citep[see, e.g.,][for a discussion of the various cosmological distances]{RKayserHS97a} and the volume; they also depend on several other `astrophysical' parameters \citep[e.g.][hereafter Paper~I]{CKochanek96a,RQuastPHelbig99a}. \subsection{The $m$-$z$ relation for type Ia supernovae} The basic idea of the $m$-$z$ relation is simple: one has an object of known absolute magnitude $M$ and compares it to the observed magnitude $m$. The difference or distance modulus is \begin{equation} \label{eq:m} m - M = 5\log_{10} D^{\mathrm{L}} + K + 42.384 - 5\log_{10}h , \end{equation} where $D^{\mathrm{L}}$ is in units of the Hubble length, $K$ is the $K$-correction and $h$ is the Hubble constant in units of $100$ \mbox{km/s/Mpc} \citep[see, e.g.,][for a derivation]{RKayserHS97a}\footnote{Note that the second occurrence of the term `Hubble length' in \citet{RKayserHS97a} should actually be `Hubble length for $h=1$', although this is obvious from the context.}. This depends on the cosmological model since $D^{\mathrm{L}}$ depends on the cosmological parameters $\lambda_{0}$ and $\Omega_{0}$. Note that, as is often the case in practice, if $M$ is known modulo $h$, then Eq.~(\ref{eq:m}) does not depend on the Hubble constant at all. On the other hand, if $M$ is known absolutely, this is equivalent to knowing $h$, assuming one has at least one object at low redshift (where the dependence on $\lambda_{0}$ and $\Omega_{0}$ is negligible). In any case, our knowledge (or lack of it) about the value of the Hubble constant $H_{0}$ does not appreciably affect the ability of this cosmological test to measure the cosmological constant $\lambda_{0}$ and the density parameter $\Omega_{0}$. Thus, one has a number of objects with observed magnitudes $m_{i}$ and a way of calculating the absolute magnitudes $M_{i}$ (see, e.g., SCP and HZSST for a description of how this is done in practice) and fits for the parameters $\lambda_{0}$ and $\Omega_{0}$. If all objects are in a narrow redshift range, then confidence contours in the $\lambda_{0}$-$\Omega_{0}$ plane will only allow one to measure approximately $\lambda_{0}-\Omega_{0}$ whereas having objects at different redshifts breaks this degeneracy \citep[e.g.][]{AGoobarSPerlmutter95a}. \subsection{Gravitational lensing statistics} See Paper~I and references therein for a discussion of how constraints on $\lambda_{0}$-$\Omega_{0}$ are derived from gravitational lensing statistics. Gravitational lensing statistics, at least in the `interesting' part of parameter space, constrain approximately $\lambda_{0} - \Omega_{0}$ \citep[e.g.][]{ACooray99a}. Thus the degeneracy is approximately the same as that of the $m$-$z$ test. Thus, rather than reducing the allowed area of parameter space through orthogonal degeneracies, these two cosmological tests provide a consistency check on each other. Also, the $m$-$z$ relation provides a good \emph{lower} limit on $\lambda_{0}$ while lensings statistics provides an \emph{upper} limit; obviously, the former should be smaller than the latter. If this is the case, then the two cosmological tests are consistent with each other, and it is meaningful to construct joint constraints, which allow a region of parameter space smaller than that allowed by either test alone. \section{Data and results} \label{results} \subsection{Individual results} For the $m$-$z$ test I used the results presented in SCP and HZSST, which have kindly been made available by the respective collaborations, as well as our own results from the analysis of JVAS, the Jodrell Bank-VLA Astrometric Survey (Paper~II and references therein). Fig.~\ref{fi:supernovae} shows the likelihood ratio as a grey scale and the 68\%, 90\%, 95\% and 99\% confidence contours for the results from SCP and HZSST; Fig.~\ref{fi:lenses} does the same for the results from Paper~II. \begin{figure*} \noindent \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f1}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f2}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f3}} \vspace{1ex} \noindent \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f4}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f5}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f6}} \vspace{1ex} \noindent \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f7}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f8}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f9}} \caption[]{The likelihood function $p(D|\lambda_{0},\Omega_{0})$ (cf.~Paper~I) from \citet{SPerlmutteretal99a} (SCP, equivalent to their Fig.~7; hereafter data set $\cal A$) (left column), \citet{ARiessetal98a} (HZSST) ($\Delta m_{15}(B)$ method, equivalent to the dotted contours of their Fig.~7; hereafter data set $\cal B$) (middle column) and \citet{ARiessetal98a} (MLCS method, equivalent to the dotted contours of their Fig.~6; hereafter data set $\cal C$) (right column) in the original parameter space and resolution (top row), in the parameter space used for the calculations in this paper but in the original resolution (middle row) and in the parameter space used for calculations in this paper in the resolution used for calculations in this paper (bottom row). The pixel grey level is directly proportional to the likelihood ratio, darker pixels reflect higher ratios. The contours mark the boundaries of the minimum $0.68$, $0.90$, $0.95$ and $0.99$ confidence regions for the parameters $\lambda_{0}$ and $\Omega_{0}$} \label{fi:supernovae} \end{figure*} \begin{figure*} \noindent \resizebox{0.45\textwidth}{!}{\includegraphics{8785.f10}} \hfill \resizebox{0.45\textwidth}{!}{\includegraphics{8785.f11}} \caption[]{The likelihood function $p(D|\lambda_0,\Omega_0)$ from Paper~II (hereafter data set $\cal D)$ in the original parameter space (left) and in the parameter space used for the calculations in this paper (right). (The resolution in Paper~II is (due to the fact that the lens statistics calculations are numerically much more demanding) the worst of all the data sets considered here and is thus the one used for the calculations in this paper.) See Fig.~\ref{fi:supernovae} for a description of the plotting scheme} \label{fi:lenses} \end{figure*} See these references for discussions of these results individually. I use the JVAS results of Paper~II, rather than those of Paper~I, since the former seem more reliable, despite the remaining uncertainties (see Paper~II for a discussion). Also, using only one set of lensing statistics results, rather than a combination, is conservative, since the joint constraints are tighter than individual constraints.\footnote{Of course, if one is concerned with the consistency of the results, rather than in reducing the parameter space through joint constraints, then one should use as many results as possible. However, it only makes sense in this context to use reliable results, so this is a reason to neglect the results based on optical gravitational lens surveys discussed in Paper~I.} The basic format here is that of a probability density function, i.e.~a relative probability as a function of $\lambda_{0}$ and $\Omega_{0}$. Ideally, this would cover \emph{all} values of $\lambda_{0}$ and $\Omega_{0}$, or at least all for which there is a non-negligible probability. Alternatively, one can impose a prior constraint on $\lambda_{0}$, $\Omega_{0}$ or both, such that there is a non-negligible probability only in a comfortably small region of parameter space. The simplest way to do this is to use a top-hat function, such that the a posteriori likelihood is given by the a priori likelihood within some range and is exactly zero outside of this range. This is a conservative approach if the allowed range is large enough to include the correct cosmological model in any case and also since, within the allowed range, the likelihood depends only on the cosmological tests considered and not on the priors (which makes for easier interpretation). The first three rows of Table~\ref{ta:ranges} show the range of parameter space covered by the references which are used to provide input data for this work. \begin{table*} \caption[]{The range of $\lambda_{0}$ and $\Omega_{0}$ explored by the references used here} \label{ta:ranges} \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lrrcrrc} \hline Reference & \multicolumn{3}{c}{$\lambda_{0}$} & \multicolumn{3}{c}{$\Omega_{0}$} \\ & \multicolumn{2}{c}{range} & resolution & \multicolumn{2}{c}{range} & resolution \\ \hline \citet{SPerlmutteretal99a} (SCP) & -1.00 & 2.98 & 0.02 & 0.00 & 2.99 & 0.01 \\ \citet{ARiessetal98a} (HZSST) & -1.00 & 3.00 & 0.01 & 0.00 & 4.00 & 0.01 \\ \citet{PHelbigMQWBK99a} (Paper~II) & -5.00 & 3.00 & 0.10 & 0.00 & 2.00 & 0.10 \\ this work & -1.00 & 2.90 & 0.10 & 0.00 & 2.00 & 0.10 \\ \hline \end{tabular*} \end{table*} I take as the only prior that the likelihood is zero outside the overlap of the various ranges of the various cosmological tests used, as in the last row in Table~\ref{ta:ranges}. The lower limit on $\Omega_{0}=0$ is physical and the upper limits $\Omega_{0}=2$ and $\lambda_{0}=2.9$ are certainly large enough (see the discussion in Paper~I on these values and on the use of prior information in general). The lower limit on $\lambda_{0}$ comes mainly from the fact that $\lambda_{0}<-1$ is strongly excluded by the $m$-$z$ relation itself, although the analysis of current cosmic microwave background observations \citep[e.g.][]{CLineweaver98a,JPerezHQ99suba,PHelbigBBdBFJKMMMQRWX98_Ca} also suggests this. In any case, an analysis of joint constraints from other cosmological tests, even excluding the $m$-$z$ relation, suggests that $\lambda_{0}>0$ is a robust result \citep[e.g.][]{MRoosSHoR99a}. With an individual test, likelihood contours are found (in all of the cases discussed here; see the discussion in Paper~I for possible caveats when comparing the results of various cosmological tests as presented in the literature) by finding the highest contour at constant likelihood such that the corresponding fraction of the total likelihood is enclosed. Note that these contours depend not only on the likelihood ratio but also on the range of parameter space plotted. In this work, the parameter space plotted should be considered to have an a priori likelihood of 1, while the parameter space outside the plot should be considered to have an a priori likelihood of 0. Note that these are `real' likelihood contours, not approximations based on $\Delta\chi^{2}$, the assumption that the probability distribution is (a 2-dimensional) Gaussian etc such as one often finds in the literature. It should also be noted that I take the likelihood as presented in the references in question. In the case of the lens statistics (data set $\cal D$), all parameters except $\lambda_{0}$ and $\Omega_{0}$ were held constant. In the case of the $m$-$z$ relation (data sets $\cal A$--$\cal C$), the results are obtained by marginalising over the nuisance parameters (see, e.g., Paper~I for a discussion). However, this is of no concern at the level of accuracy I am concerned with here, especially since there are no nuisance parameters common to the $m$-$z$ test and the lensing statistics test.\footnote{The publicly available data from SCP are actually not the likelihood itself, but rather the value for each point in the parameter space is the (normalised) sum of all likelihood values for all points in the parameter space which are not less than the value for the point in question. This format, which allows one to immediately plot a given confidence contour by plotting a contour at that level, I have converted back to the original probability density function.} \subsection{Joint constraints} The simplest thing to do when building joint constraints would be to multiply the corresponding probability density functions (PDFs).\footnote{Of course, this must be done at the same resolution. Rather than interpolate the low resolution lens statistics results, I have reduced the resolution of the $m$-$z$ results to that of the lensing statistics results by using only those points in the $\lambda_{0}$-$\Omega_{0}$ plane which were examined in the lens statistics calculations, all of which were examined by both $m$-$z$ tests.} One can then plot confidence contours in the manner described above. However, it is obvious that this is not meaningful if the PDFs are not consistent with each other, i.e.~if the region of confidence for a `sensible' confidence level from the joint constraints does not overlap with the corresponding confidence level for all component tests. A necessary, though not sufficient, condition for this inconsistency to exist is that the corresponding confidence contours for the individual component tests do not overlap. Fig.~\ref{fi:overlap} shows the 68\%, 90\%, 95\% and 99\% confidence contours for the four data sets considered here. \begin{figure*} \noindent \resizebox{0.45\textwidth}{!}{\includegraphics{8785.f12}} \hfill \resizebox{0.45\textwidth}{!}{\includegraphics{8785.f13}} \vspace{1.5ex} \noindent \resizebox{0.45\textwidth}{!}{\includegraphics{8785.f14}} \hfill \resizebox{0.45\textwidth}{!}{\includegraphics{8785.f15}} \caption[]{The 68\% (top left), 90\% (top right), 95\% (bottom left) and 99\% (bottom right) confidence contours for each of the data sets. The thick curves are for the lensing statistics results (data set $\cal D$). In all plots, data set $\cal A$ has the contour with the lowest value of $\Omega_{0}$ at its maximum height. Starting from this point and moving left, towards smaller values of $\lambda_{0}$, in all plots one crosses first the contour of data set $\cal B$ then that of data set $\cal C$} \label{fi:overlap} \end{figure*} As the 90\% confidence contours from all supernovae data sets overlap with that of the lensing statistics, and even the 68\% confidence contours from two of three supernovae data sets overlap with that of the lensing statistics, the results from the two cosmological tests are consistent and one is justified in calculating joint constraints by multiplying the probability distributions of the individual tests. Interestingly, they are most consistent at small, but not too small, values of $\Omega_{0}.$ The results of this are shown in Fig.~\ref{fi:joint}. \begin{figure*} \noindent \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f16}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f17}} \hfill \resizebox{0.3\textwidth}{!}{\includegraphics{8785.f18}} \caption[]{Joint constraints from the lensing statistics calculations of \citet{PHelbigMQWBK99a} (Paper~II) and the $m$-$z$ relation for type Ia supernovae as in data set $\cal A$ (left), data set $\cal B$ (middle) and data set $\cal C$ (right). See Fig.~\ref{fi:supernovae} for an explanation of the data sets} \label{fi:joint} \end{figure*} Note that if there is some offset between the allowed regions from each individual test, as is the case here, then certain aspects of the joint constraints, such as in this case the upper limit on $\lambda_{0}$, will not necessarily be tighter than the corresponding aspect from each individual test. The joint constraints are nevertheless better in that the allowed region is smaller and that this allowed region should contain the correct cosmological model, assuming of course that the results of the individual tests are correct as far as they go. The tests are very different in nature and one should not expect the form of the probability density function to be the same in each case. In particular, as lensing statistics is especially sensitive to a large cosmological constant, the gradient in this area of parameter space is quite steep, thus it is not surprising that the lensing statistics upper limit on $\lambda_{0}$ is tighter. The fact that the confidence contours from the individual tests overlap shows that the tests are not inconsistent, and of course the allowed region from the joint constraints, which is consistent with each individual test, is approximated by this overlap. Since the two $m$-$z$ results are not completely independent, the question of the consistency of or joint constraints from the two supernovae data sets will not be discussed in this paper. Rather, the question is the consistency of and joint constraints from each of these data sets individually with the lensing statistics constraints. Fig.~\ref{fi:joint} is the main conclusion of this paper. Although lensing statistics and the $m$-$z$ relation individually allow, for appropriate values of $\lambda_{0}$, rather large values of $\Omega_{0}$, the joint constraints clearly indicate a lower $\Omega_{0}$, in accordance with observational evidence which measures $\Omega_{0}$ more `directly' (see the discussion in Paper~I). Compared to the supernovae results, the allowed region of parameter space is shifted somewhat towards lower values of $\Omega_{0}$ in the joint constraints. Although the actual best-fit value should not be taken too seriously, it is comfortably close to the current `standard cosmological model' with $\lambda_{0}\approx 0.7$ and $\Omega_{0}\approx 0.3$ The quantity measured by both the lens statistics and the $m$-$z$ relation for type Ia supernovae discussed here is approximately $\lambda_{0}-\Omega_{0}$. Table~\ref{ta:results} shows the 95\% confidence ranges for $\lambda_{0} - \Omega_{0}$ allowed by each of the four data sets individually and by the joint constraints of data set $\cal D$ with data sets $\cal A$, $\cal B$ and $\cal C$. \begin{table*} \caption[]{95\% confidence ranges for $\lambda_{0} - \Omega_{0}$ allowed by each of the four data sets individually as well as various joint constraints} \label{ta:results} \begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcc} \hline Reference & lower limit & upper limit \\ \hline \citet{SPerlmutteretal99a} (SCP) (data set $\cal A$) & $-0.05$ & $+0.65$ \\ \citet{ARiessetal98a} (HZSST, $\Delta m_{15}(B)$) (data set $\cal B$) & $+0.30$ & $+0.81$ \\ \citet{ARiessetal98a} (HZSST, MLCS) (data set $\cal C$) & $-0.12$ & $+0.79$ \\ \citet{PHelbigMQWBK99a} (Paper~II) (data set $\cal D$) & $-1.90$ & $+0.45$ \\ $\cal A$ + $\cal D$ & $-0.20$ & $+0.55$ \\ $\cal B$ + $\cal D$ & $\pm 0.00$ & $+0.55$ \\ $\cal C$ + $\cal D$ & $-0.25$ & $0.60$ \\ \hline \end{tabular*} \end{table*} At 95\% confidence, the upper limit on $\lambda_{0}-\Omega_{0}$ from lensing statistics alone is 0.45 and from supernovae alone is in the range 0.65--0.81 (depending on the data set).\footnote{Note that this is not the same as that quoted in Paper~II; this is because, as discussed above, the range of parameter space examined or, equivalently (in this case), the prior is different.} For joint constraints, the upper limit on $\lambda_{0}-\Omega_{0}$ is in the range 0.55--0.60 (again depending on the data set). For a flat universe with $\lambda_{0} + \Omega_{0} = 1$, this corresponds to upper limits on $\lambda_{0}$, taking the top of the range from different data sets, of 0.72, 0.90 and 0.80 for lensing statistics alone, supernovae alone and the joint analysis, respectively. Again, this is perfectly consistent with the current `standard cosmological model' with $\lambda_{0}\approx 0.7$ and $\Omega_{0}\approx 0.3$ \citep[e.g.][]{MRoosSHoR99a,MTurner99a} and is consistent with a flat universe but, neglecting other cosmological tests, does not require it. \subsection{Systematic errors} As far as the $m$-$z$ relation for type Ia supernovae goes, various possible sources of systematic errors have been discussed in detail by SCP and HZSST. See particularly Fig.~5 in SCP. Basically, there is no evidence that the purported effects could significantly bias the results or, as in the case of grey dust, while a modest amount cannot be ruled out, it seems physically rather implausible \citep[but see, however,][]{AAguirre99a, AAguirre99b,AAguirreZHaiman99a}. (Note that \citet{EFalcoIKLMcLRKMP99a} find no evidence for grey dust at optical wavelengths based on studies of extinction in gravitational lens galaxies.) It is interesting to note that while one can invoke grey dust to explain the dimming of supernovae relative to the expectation in for example a dust-free $\lambda_{0}=0$ model, instead of invoking for example a low-density model with a positive cosmological constant, this degeneracy can be broken by observing supernovae at higher redshift \citep[e.g.][]{AAguirre99a} than has been done up until now: the $m$-$z$ relation as a function of $\lambda_{0}$ and $\Omega_{0}$ is exactly known, so the larger the range in redshift for which the $m$-$z$ relation is observed, the more ad hoc alternative explanations become, provided of course that there is a cosmological model (which is not ruled out on other grounds) which provides an acceptable fit to the data. Grey dust is also something which can effect gravitational lensing statistics \emph{based on optical samples}, although the effects are not so straightforward. On the one hand, if the grey dust is concentrated in (lensing) galaxies, this could lead to lens systems being missed in the survey. To first order, this would lead to an underestimate of the optical depth and thus of the value of $\lambda_{0}$. On the other hand, again if dust is concentrated in (lensing) galaxies, the sources in the identified lens systems can suffer from extinction, which, depending on the details of the luminosity function, could lead to a wrong estimate of the magnification bias. In the `normal' case of a flattening of the luminosity function for fainter objects, this will lead to an underestimate of the amplification bias and hence an overestimate of the optical depth and thus the value of $\lambda_{0}$ \citep{EFalcoIKLMcLRKMP99a}. Radio surveys of course are not affected by dust, so in principle one could detect grey dust through a systematic difference in the results from optical and radio surveys. In practice, however, the presence of other systematic effects makes such a detailed comparison impractical. Radio surveys for gravitational lenses offer many advantages over optical surveys (see the discussion in Paper~II). However, at present, the main source of uncertainty, lack of knowledge about the source population, makes them worse than optical surveys in this respect. For the calculation of the amplification bias, one needs to know, at a given redshift, the luminosity function.\footnote{Due to the amplification of the gravitational lens effect, lensed sources near the lower flux-density limit of the survey will have an unlensed flux density lower than this, so the luminosity function thus needs to be known down to a flux-density level a factor of several below that of the survey.} On the other hand, much information is gained from the sources in a survey which are not lensed (see the discussion in Paper~I); to interpret this, one needs to know, at a given flux density, the redshift distribution of the sources. Of course, these two things---the redshift-dependent luminosity function and the flux-density dependent redshift distribution---are different sides of the same coin. The number counts of the Cosmic Lens All-Sky Survey, of which the Jodrell Bank-VLA Astrometric Survey, the results of which are used here, is a subset, suggest that amplification bias is not a big effect; hence the systematic error from lack of knowledge of the luminosity function is probably small, although it is conceivable that the number counts (integrated over redshift, as in general the CLASS sources have unknown redshifts) of CLASS are not representative of the luminosity function at \emph{all} redshifts. As the lensed sources are generally at higher redshifts, the luminosity function might be different here and thus the true amplification bias different from that which was used in the JVAS analysis of Paper~II (where it was assumed that the CLASS number counts are representative of all redshifts). In Paper~II, it was also assumed that the redshift distribution of JVAS is equal to that of CJF, independent of flux-density. There is some preliminary evidence that, as one moves toward lower flux-density levels, the typical redshift of flat-spectrum radio sources decreases. If this is the case, then our JVAS analysis will have underestimated the value of $\lambda_{0}$, as a higher value of $\lambda_{0}$ (all other things being equal) is needed to achieve the same optical depth for a low-redshift source than is needed for a high-redshift source. Although the results from the $m$-$z$ relation for type Ia supernovae and gravitational lensing statistics are not inconsistent, and although, due to the different dependence on the cosmological parameters, the lower limit on $\lambda_{0}$ will always be stronger from the former and the upper limit from the latter, if it does turn out to be true that the CJF redshift distribution is systematically higher than the true redshift distribution of JVAS, then the results from the $m$-$z$ relation for type Ia supernovae and gravitational lensing statistics will become even more consistent. \section{Summary and conclusions} \label{conclusions} I have presented the first detailed analysis of joint constraints between gravitational lensing statistics and the $m$-$z$ relation for type Ia supernovae, making use of data from \citet{PHelbigMQWBK99a}, \citet{SPerlmutteretal99a} and \citet{ARiessetal98a}, presenting the individual results and the new joint constraints in a uniform way. The two tests are not inconsistent, the joint constraints are tighter than those from either test individually and provide additional evidence in favour of the current `standard cosmological model' with $\lambda_{0}\approx 0.7$ and $\Omega_{0}\approx 0.3$, although (neglecting constraints from other sources such as the CMB) a reasonable range of other cosmological models is not excluded. In the near future, gravitational lensing statistics from CLASS, the Cosmic Lens All-Sky Survey \citep{SMyersetal99a} should reduce both the random and systematic errors. Should the results from lensing statistics and the $m$-$z$ relation for type Ia supernovae remain consistent, this should reduce the allowed parameter space even further. We are truly entering an era of precision cosmology, where the overlap of the allowed regions of parameter space from many different and independent cosmological tests is very small but not zero. The data for the figures shown in this paper are available at \begin{quote} \verb|http://multivac.jb.man.ac.uk:8000/ceres|\\ \verb|/data_from_papers/snlens/snlens.html| \end{quote} or \begin{quote} \verb|http://gladia.astro.rug.nl:8000/ceres|\\ \verb|/data_from_papers/snlens/snlens.html| \end{quote} \begin{acknowledgements} It is a pleasure to thank Saul Perlmutter, Brian Schmidt and Saurabh Jha for helpful discussions and the Supernova Cosmology Project and the High-Z Supernova Search Team for making their numerical results available. This research was supported in part by the European Commission, TMR Programme, Research Network Contract ERBFMRXCT96-0034 `CERES'. \end{acknowledgements} \bibliographystyle{aa}
\section{TeV$^{-1}$ dimensions, supersymmetry breaking and submm forces} String theories provide at present the only consistent framework of quantum gravity and unification of all fundamental interactions. Traditionally, it was believed that they become relevant only at Planckian energies due to the simple tree-level expression for the string scale, \begin{equation} M_H\simeq gM_P\, , \label{het} \end{equation} valid in the context of perturbative heterotic theory. Here, $M_P=1.2\times 10^{19}$ GeV is the Planck mass, and $g\simeq 1/5$ is the gauge coupling at the heterotic string (unification) scale $M_H\simeq 10^{18}$ GeV. Despite this relation, there are however physical motivations which indicate that large volume compactifications may be relevant for physics. One of them comes from the problem of supersymmetry breaking by the process of compactification which relates the breaking scale to the size of some internal compact dimension(s). The latter should therefore be in the TeV$^{-1}$ region in order to keep the scale of supersymmetry breaking close to electroweak energies~\cite{a}. An immediate defect of this scenario is that the ten-dimensional (10d) string coupling is huge invalidating the perturbative description. From the 4d point of view, the problem arises due to the infinite massive tower of Kaluza-Klein (KK) excitations that are produced at energies above the compactification scale and contribute to physical processes: \begin{equation} M^2=M_0^2+{n^2\over R^2}\ ;\qquad n=0,\pm 1,\dots\, , \label{KK} \end{equation} where $M_0$ is a 5d mass and $R$ is the radius of an extra dimension. If the theory makes sense for some special models, a possible way out consists of imposing a set of conditions to the low energy theory that prevent the effective couplings to diverge. An example of such conditions is that KK modes should be organized into multiplets of $N=4$ supersymmetry, containing for each spin-1 4 Weyl-fermions and 6 scalars with the same quantum numbers, so that their contribution to beta-functions vanish for every $n\neq 0$ and gauge couplings remain finite. The breaking of supersymmetry by compactification is realized through boundary conditions and is similar to the effects of finite temperature with the identification $T\equiv R^{-1}$. It follows that the breaking is extremely soft and insensitive to the ultraviolet (UV) physics above the compactification scale. The summation over the KK excitations amounts to inserting the Boltzmann factors $e^{-E/T}$ to all thermodynamic quantities --or equivalently to the soft breaking terms-- that suppress exponentially their UV behavior~\cite{a,add}. This is in contrast to the behavior of supersymmetric couplings that generally blow up, unless special conditions are imposed. The extreme softness of supersymmetry breaking by compactification has two important phenomenological consequences: (i) a particular spectroscopy of superparticles that differs drastically from other scenaria~\cite{a,adpq}, and (ii) a new scalar force in the (sub)mm range mediated by the scalar modulus whose vacuum expectation value (VEV) determines the radius $R$ of the TeV dimension used to break supersymmetry~\cite{add}. One can easily check by direct dimensional reduction of the higher-dimensional Einstein action that the scalar field associated to the radius modulus with canonical kinetic terms is $\phi=\ln R$ in Planck units. Its VEV corresponds to a flat direction of the scalar potential, which is lifted only after supersymmetry breaking, through a potential generated by radiative corrections. The softness of the breaking mechanism implies that the vacuum energy $V$ behaves for large $R$ as \begin{equation} V\sim T^4\equiv{1\over R^4}\, , \label{vacen} \end{equation} up to logarithmic higher loop corrections. Notice the absence of a ``quadratically divergent" $M_P^2 R^{-2}$ contribution. It follows that the mass of the radius modulus $\phi$ is of order~\cite{fkz,add} \begin{equation} m_\phi\sim R^{-2}/M_P\simeq 10^{-4} - 10^{-2}\ {\rm eV} \, , \label{mphi} \end{equation} for a compactification scale $R^{-1}\simeq 1 - 10$ TeV. This corresponds to Compton wavelengths of the modulus in the range of 1 mm to 10 $\mu$m. The coupling of this light scalar to matter arises dominantly through the dependence on $\phi$ of the QCD scale $\Lambda_{\rm QCD}$, or equivalently of the QCD gauge coupling $\alpha_{\rm QCD}$. The coupling of the radius modulus to nucleons is given by the derivative of the nucleon mass relative to $\phi$, ${\partial m_N\over\partial\phi}$. Since the graviton coupling to nucleons is $m_N/M_P$, the coupling $\alpha_\phi$ of the modulus relative to gravity is \begin{equation} \alpha_\phi = {1\over m_N}{\partial m_N\over\partial\phi}= {\partial\ln\Lambda_{\rm QCD}\over\partial\ln R}= -{1\over b_{\rm QCD}}{\partial\over\partial\ln R}{2\pi\over\alpha_{\rm QCD}}\, , \label{coupling} \end{equation} with $b_{\rm QCD}$ the one-loop QCD beta-function coefficient. Since in these models, in the absence of supersymmetry breaking all couplings of the effective field theory are well behaved in the large radius limit, while after supersymmetry breaking all soft masses are proportional to $1/R$, one has \begin{equation} {\partial\over\partial\ln R}{2\pi\over\alpha_{\rm QCD}}= b_{\rm SQCD}-b_{\rm QCD}\, , \label{deltaR} \end{equation} with $b_{\rm SQCD}$ the supersymmetric beta-function. Using $b_{\rm SQCD}=3$ and $b_{\rm QCD}=7$, one obtains $\alpha_\phi=1-b_{\rm SQCD}/b_{\rm QCD}=4/7$. Thus, the force between two pieces of matter mediated by the radius modulus is $(4/7)^2 \simeq 1/3$ times the force of gravity~\cite{add}. In principle there can be other light moduli which couple with larger strengths. For example the dilaton, whose VEV determines the (logarithm of the) string coupling constant, if it does not acquire large mass from some dynamical supersymmetric mechanism, leads to the strongest effect. Its coupling is $\alpha_{\rm dilaton}=\ln(M_P/\Lambda_{\rm QCD})\sim 44$~\cite{tv}, which corresponds to a strength $\sim 2000$ times bigger than gravity. \begin{figure} \centerline{\psfig{figure=forces.eps,height=7cm}} \caption{Strength of the modulus force relative to gravity ($\alpha^2$) versus its Compton wavelength ($\lambda$). \label{fig:forces}} \end{figure} In fig.~\ref{fig:forces} we depict the theoretical predictions together with information from previous, present and upcoming experiments. The vertical axis is the strength, $\alpha^2$, of the force relative to gravity; the horizontal axis is the Compton wavelength of the exchanged particle; the upper scale shows the corresponding value of the large radius in TeV. The solid lines indicate the present limits from the experiments indicated~\cite{oldexperiments}. The excluded regions lie above these solid lines. Our theoretical prediction for the radius modulus is the thin horizontal line at 1/3 times gravity. The dilaton-mediated force is the thin dashed horizontal line at $\sim$ 2000 times gravity. Other moduli may mediate forces between the above two extremes. Of course, measuring gravitational strength forces at such short distances is quite challenging. The most important background is the Van der Walls force. The Van der Walls and gravitational forces between two atoms are equal to each other when the atoms are about 100 $\mu$m apart. Since the Van der Walls force falls off as the 7th power of the distance, it rapidly becomes negligible compared to gravity at distances exceeding 100 $\mu$m. The most important part of the figure is the dashed thick line; it is the expected sensitivity of the present and upcoming experiments, which will improve the actual limits by almost 6 orders of magnitude and --at the very least-- they will, for the first time, measure gravity to a precision of 1\% at distances of $\sim$ 100 $\mu$m~\cite{price}. \section{TeV strings, submm dimensions and signals in particle colliders} As mentioned before, the main defect of the above heterotic string models is the strong 10d coupling due to the large internal dimension(s) used to break supersymmetry. Fortunately, this problem can be addressed using the recent results on string dualities, by considering the corresponding dual weakly coupled theory. In the minimal case of one or two large dimensions, the dual weakly coupled description is provided by type II string theory, whose tension appears as a non-perturbative threshold below the heterotic scale~\cite{ap}. For instance, in the case of one TeV dimension, the type II scale appears at intermediate energies of order $10^{11}$ GeV, and the theory above the TeV but below the intermediate scale becomes effectively six-dimensional and is described by a tensionless string~\cite{ts} with no gravity. This 6d theory possesses a non-trivial infrared fixed point which encodes the conditions needed to be imposed in the low energy couplings, in order to ensure a smooth UV behavior. These conditions generalize the requirement of $N=4$ supersymmetry for the KK excitations, that keep the gauge couplings well behaved, to the Yukawa couplings of the theory. A generic property of these models is that chiral matter is localized at particular points of the large internal dimensions. As a result, quarks and leptons have not KK TeV excitations, which is welcome also for phenomenological reasons in order to avoid fast proton decay~\cite{a,ab}. The main prediction of these theories in particle accelerators is the existence of KK excitations~(\ref{KK}) for all Standard Model gauge bosons that can be produced for instance in hadron colliders, such as the Tevatron and LHC, through Drell-Yan processes~\cite{abq}. The corresponding KK resonances are narrow with a ratio of their width to their mass $\Gamma/m\sim g^2\sim$ a few per cent, and thus the typical expected signal is the production of a double resonance corresponding the first KK mode of the photon and $Z$, very nearly spaced the one from the other. The current limits on the size of large dimensions arise from the bounds of compositeness or from other indirect effects, such as in the Fermi constant and LEP2 data, and lie in the range of $1-2$ TeV~\cite{ab,limits}. On the other hand, direct production of KK states in LHC is possible for compactification scales of less than about 5 TeV. In the case of two TeV dimensions, the type II string scale appears also at the TeV, while its coupling is infinitesimally small $\sim 10^{-14}$~\cite{ap}. As a result, despite the fact that the string scale of the dual theory is so low, it has no observable effects in particle accelerators and the main experimental signal continues to be the existence of KK excitations along the two TeV dimensions with gauge interactions. The predictions are thus very similar to the previous case described above. It is amazing that the main features of these models were captured already in the context of heterotic string despite its strong coupling. When there are more than two large dimensions, a new phenomenon appears. The dual weakly coupled theory, that can now be either type II or type I, contains also some extra large dimensions of size much larger than TeV$^{-1}\simeq 10^{-16}$ cm, where only gravity propagates. It is now clear from the above discussion that in other than the heterotic perturbative descriptions of string theory, the string scale is not tied to the Planck mass by a simple relation like~(\ref{het}), but it corresponds to an arbitrary parameter that can be everywhere above the TeV scale~\cite{wl}. In particular, if it is at the TeV, the gauge hierarchy problem would be automatically solved without the need of low energy supersymmetry or technicolor. The price to pay is to introduce an extremely small string coupling and/or, alternatively, extra large dimensions in order to account for the hierarchy $M_{\rm str}/M_P$. For instance, it was recently proposed that the fundamental scale of gravity could be at the TeV if there were such extra large dimensions where only gravity propagates, while the Standard Model should be localized on a 3d wall~\cite{add1,aadd}. The observed weakness of 4d gravity would then be a consequence of the higher dimensional Gauss-law: \begin{equation} G_N={G_{N(4+p)}\over R_{\perp}^p}\, , \label{gl} \end{equation} where $G_{N(4+p)}$ is the $(4+p)$-dimensional Newton's constant and $R_{\perp}$ denotes the (common) size of $p$ extra large spacial dimensions. This scenario can be realized naturally in the context of type I theory of closed and open strings~\cite{aadd,stk}. Gravity is described by closed strings, while gauge interactions are described by open strings whose ends are confined to propagate on subspaces, called D-branes. The theory is weakly coupled when the size of the internal dimensions parallel to the D-brane where the Standard Model is localized is of order of the string scale $M_I^{-1}$, while there is no restriction on the size of transverse dimensions. One then obtains that the type I string scale is simply related to the higher-dimensional Newton's constant \begin{equation} G_{N(4+p)}\simeq g^2 M_I^{-2-p}\, , \label{typeI} \end{equation} and is therefore a free parameter, while the string coupling is given by the square of the gauge coupling. This expression should be compared with the corresponding relation~(\ref{het}) in the heterotic string, where the string scale is fixed by the 4d Planck mass.\footnote{Note that in the context of type II theories, the string tension is decoupled from the scale where gravity becomes strong and can be lowered independently by decreasing the string coupling which is a free parameter.} Using eqs.~(\ref{gl}) and~(\ref{typeI}) and taking the type I string scale to be at the TeV, one finds a size for the transverse dimensions varying from $10^8$ km, .1 mm, up to .1 fermi for $p=1,2$, or 6 large dimensions. With the exception of $p=1$ which is obviously experimentally excluded, all other values are consistent with observations~\cite{add1,add2,ns}. The strongest bound applies for $p=2$ and comes from graviton emission on cooling of supernovae which restricts the 6d Planck scale to be larger than about 50 TeV~\cite{sn}, implying $M_I\stackrel{>}{{}_\sim} 7$ TeV. The main distinct experimental signals of these theories in particle accelerators are~\cite{aadd}:\hfil\\ (i) Production of higher spin Regge-excitations for all Standard Model particles with the same quantum numbers and mass-squared increasing linearly with the spin. For instance, the excitations of the gluon could show up as a series of peaks in jet production at LHC. However, the corresponding resonances are very narrow, with a ratio of their width to their mass $\Gamma/m\sim g^4\sim$ a few per thousand, and thus are difficult to detect.\hfil\\ (ii) Graviton emission into the extra dimensions leading to events of jets + missing energy. LHC is sensitive to higher-dimensional gravity scales in the range of 3 to 5 TeV, when the number of transverse dimensions varies from $p=6$ at the subfermi to $p=2$ at the submm, where the effect becomes stronger~\cite{ns,grav}. This is comparable to the sensitivity on possible TeV dimensions with gauge interactions, as discussed previously. The most exciting possibility is of course when the available energy is bigger than the gravity scale, in which case gravitational interactions are strong and particle colliders become the best tools for studying quantum gravity. When the string scale is of order of a few TeV, the question of gauge hierarchy is to understand why some dimensions transverse to our brane-world are so large or/and why the string coupling is so small. Thus, the large number to explain varies from $10^{15}$ up to $10^5$ in the case of six transverse dimensions. This problem has a difficult part which is its dynamical aspect and a {\it technical} part that consists to guarantee its stability under quantum corrections. Supersymmetry solves this technical aspect by allowing only logarithmic dependence on the UV cutoff --usually taken at $M_P$-- to the masses and other parameters of the Standard Model. Although naively one would expect that low string scale would also solve the problem by nullification~\cite{add1}, it turns out that this is not true in general~\cite{aba}. The reason is due to the appearance of local tadpoles in various processes on the world-brane, associated to the emission of a massless closed string in the transverse space. These tadpoles diverge linearly when the closed string propagates ``effectively" in one preferred large transverse direction ($d_\perp=1$) and logarithmically when $d_\perp=2$. Local tadpole cancellation implies severe constraints for model building while their presence implies a set of conditions for the compactification manifold that forbids linear divergences. The remaining divergences are at most logarithmic and the technical aspect of gauge hierarchy is solved at the same level as with supersymmetry. Moreover, the case of $d_\perp=2$ is singled out as the only one in which the origin of the hierarchy would not be attributed to `out of this world' bulk physics, since the presence of logarithmic sensitivity could allow for a dynamical determination of the hierarchy by minimizing the effective potential.\footnote{A similar conclusion appears to hold in low-scale type II models with infinitesimal coupling~\cite{ap}.} Although the initial motivation of low energy supersymmetry is in principle lost, there may be still a reason to be introduced in the bulk, in connection to the cosmological constant and stability problem. In fact, the bulk energy density of a generic non-supersymmetric string model is $V_{\rm bulk}\sim M_{\rm str}^{4+p}$, giving rise to an energy density on our world-brane~\cite{adpq}: \begin{equation} V_{\rm brane}\simeq V_{\rm bulk}R_{\perp}^p\sim M_{\rm str}^2 M_P^2\, , \label{rho} \end{equation} by virtue of eqs.~(\ref{gl}) and~(\ref{typeI}). This is analogous to the quadratically divergent contribution to the vacuum energy in softly broken supersymmetry, which in general destabilizes the hierarchy since it induces a large potential for $R_\perp$. On the other hand, its absence is guaranteed if the bulk remains approximately supersymmetric, which is the case when for instance supersymmetry is broken spontaneously primordially only on the brane (even maximally)~\cite{aadd}. \section{Gravity modification at short distances} TeV scale strings with extra large transverse dimensions predict important modifications of gravitation at short distances that could be measured in experiments which test gravity in the submillimeter range (see fig.~\ref{fig:forces}). These ``tabletop" experiments, that cost orders of magnitude less than high energy particle accelerators, might provide an independent verification of the above ideas much before LHC. In fact, there are two classes of general predictions:\hfil\\ (i) Deviations from the Newton's law $1/r^2$ behavior to $1/r^{2+p}$ which can be observable for $p=2$ transverse dimensions of submm size. This case is particularly favored for theoretical reasons because of the logarithmic sensitivity of Standard Model couplings on the size of transverse space, as we discussed above, but also for phenomenological reasons, since the effects to particle colliders are maximally enhanced. Notice also the coincidence of this scale with the possible value of the cosmological constant in the universe that recent observations seem to support.\hfil\\ (ii) New scalar forces in the submm range mediated by light bulk scalar fields with masses $M_{\rm str}^2/M_P$, motivated by the problem of supersymmetry breaking. Such a universal scalar corresponds to the transverse radius modulus $\phi_\perp\equiv\ln R_\perp$ in Planck units. Following our previous discussion, if supersymmetry is broken primordially only on the brane, its transmission to the bulk is gravitationally suppressed and $\phi_\perp$ acquires a mass proportional to the gravitino mass, of the order of~\cite{aadd} \begin{equation} m^2_{\phi_\perp}\sim G_{N(4+p)}{V_{\rm brane}\over R_{\perp}^p}\simeq {M_{\rm str}^4\over M_P^2}\, , \label{mphiperp} \end{equation} using eq.~(\ref{gl}). Unlike to the case of the TeV radius modulus $\phi$~(\ref{mphi}), its coupling to matter is in general model dependent. Following the same steps as before, one finds that its coupling to nucleons relative to gravity $\alpha_{\phi_\perp}$ is given by eq.~(\ref{coupling}) replacing $R$ by $R_\perp$. It follows that $\alpha_{\phi_\perp}$ is strongly suppressed, except again in the case of $d_\perp=2$, when gauge couplings acquire logarithmic sensitivity on $R_\perp$. In this case, the resulting scalar force is of order ${\cal O}(1)$ relative to gravity --although not exactly computable-- and therefore in principle within the expected sensitivity of present and future experiments displayed in fig.~\ref{fig:forces}. \section*{References}
\section{Introduction} The Einstein equations present difficulties of size and complexity somewhat greater than those normally encountered in scientific computation. A numerical simulation must confront a range of theoretical and numerical challenges, starting with the theoretical problem of finding a well-posed and geometrically natural reduction of the full system of Einstein equations. Numerical algorithms are then needed to control the various facets of the reduced system, efficiently, since very large data structures are inevitable when modelling fully 3+1-dimensional spacetimes. Finally there are the twin theoretical and practical problems of understanding the nature of the solution represented by the data, and of certifying its reliability. In \cite{Bartnik97a} we presented a new coordinate formulation for the vacuum Einstein equations, based on a characteristic (null) coordinate and a quasi-spherical foliation \cite{Bartnik93}. This null quasi-spherical (NQS) formulation is well-adapted to modelling spacetimes containing a single black hole, extending from the black hole to null infinity. The purpose of this paper is to describe the numerical algorithms we have used in our implementation of the NQS Einstein equations, and to present results of some accuracy tests of the code. Interactive access to the data sets described here, and many other simulations, is available online at \texttt{http://gular.canberra.edu.au/relativity.html}. More detailed discussions of the physical and geometric significance of the results of the code will be presented elsewhere. {}From the numerical programming viewpoint, the most significant features of the code are: \begin{enumerate} \item a characteristic coordinate $z\sim t-r$ (cf.~\cite{Stellmacher38}) plays the role of ``time'', with the numerical evolution in the direction of increasing $z$; \item the numerical grid is based on spherical polar coordinates on the $z$-level sets $\mathcal{N}_{z}$; \item $S^{2}$ dependencies are handled by a combination of spectral coefficients with respect to a basis of spin-weighted spherical harmonics (for spins $0,1,2$, corresponding to scalar, vector and tensor harmonics); values of the field on a uniform grid in the spherical polar coordinates $(\vartheta,\varphi)$; and Fourier coefficients of the values on the polar coordinate grid; \item a non-uniform radial grid along the outgoing null hypersurfaces $\mathcal{N}_{z},$ which compactifies future null infinity $\mathcal{I}^{+}$ and is adjusted dynamically so radial grid points approximately follow the inward null geodesics; \item reformulations of the hypersurface (radial) Einstein equations which enable the numerical modelling of the asymptotic expansions of fields near null infinity; \item 4th order Runge-Kutta time evolution; \item the characteristic transport (hypersurface, radial) equations are treated as a system of ordinary differential equations and integrated using an 8th order Runge-Kutta method \cite{DormandPrince81}; \item an 8th order convolution spline is used for interpolation, for computing radial derivatives, to realign the fields with the dynamically varying radial grid, and for suppressing high frequency modes; \item a first order elliptic system on $S^{2}$ is solved at each radius and time step by the conjugate gradient method, accelerated by a geometrically natural spectral preconditioner. \end{enumerate} A number of numerical consistency checks suggest that most quantities of interest which are calculated by the code (eg., the NSQ metric functions, the connection coefficients and the Weyl curvature spinors) have relative errors of about $0.001\%$, for simulations where the gravitational waves carry no more than 20\% of of the total spacetime mass. Of course, greater accuracy is found for more nearly linear, weak field, simulations. The major limiting factor in determining the accuracy appears to be the spherical harmonic resolution, currently at $L\le15$. Although we have implemented routines which extend this to $L\le31$, a full implementation is not possible due to limitations in our present hardware. The algorithm was initially developed and tested on a 300MHz DEC Alpha with 512Mb memory, and presently runs on a 300MHz Sun Ultra 2 with 784Mb, with a typical (L=15) run taking between 2 and 4 days. Preliminary descriptions of these results are given in \cite{Bartnik96,BartnikNorton98,Bartnik99b,Norton96,Norton98}. The Cauchy and characteristic initial value problems differ significantly in the nature of their appropriate initial conditions. The Cauchy problem initial data consists of the initial 3-metric and extrinsic curvature \cite{SmarrYork78}, whereas the appropriate initial data for a characteristic initial surface is just the null metric. In the NQS case the null metric is \begin{equation} ds^{2}_{\mathcal{N}_{z}} = (r\,d\vartheta+\beta^{1}\,dr)^{2} + (r\sin\vartheta\,d\varphi+\beta^{2}\,dr)^{2}, \label{ds2Nz} \end{equation} parameterised by the angular shear vector $\beta = \tfrac{1}{\sqrt{2}}(\beta^{1}-{\rm i}\,\beta^{2}) = \beta(z,r,\vartheta\,\varphi)$. Unlike the Cauchy problem, the NQS Einstein equations \cite{Bartnik97a} do not impose any additional algebraic or differential constraints on $\beta$, so the initial data $\beta(z=0)$ is an arbitrary function of the spherical polar coordinates $(r,\vartheta\,\varphi)$, except for a mild size constraint \bref{bsize}. Heuristically $\beta$ represents the in-going gravitational radiation of the spacetime; an interpretation which is consistent with the initial ($z=0$) values of $\beta$ being freely specifiable. Note that the geometric invariant $\sigma_{NP}/\rho_{NP}$ \cite{Penrose72} becomes $-2\eth\beta/(2-\div\beta)$ in the NQS coordinates \cite{Bartnik97a}. The NQS geometric gauge provides a formulation of the Einstein equations as an explicit \emph{characteristic transport} system, coupled to a time evolution equation. This type of structure is also found in characteristic formulations of other hyperbolic equations such as the wave equation, and it is well-known for the Einstein equations in other characteristic coordinate based gauges \cite{Bondi62, Sachs62, Stellmacher38, NewmanUnti63}. Although there are existence results for characteristic initial value formulations of hyperbolic equations, these rely on reduction to the Cauchy problem \cite{Rendall90, MullerzumHagen90} rather than directly on the transport form. The only exception of which we are aware is the analysis of the linear wave equation in \cite{Balean96, Balean97}. It would be valuable to have theoretical existence results for systems of characteristic transport equations, which could justify the numerical formulation described here. The characteristic-based approach has recently been strongly advocated by Winicour and his coworkers \cite{Bishop92,GomezEt92,GomezWinicour92b}, who have developed codes for solving the scalar wave equation \cite{GomezEt92}, axially symmetric spacetimes \cite{GomezEt94a} and for the full Einstein equations \cite{BishopEt97b}, based on the Bondi coordinate system \cite{Bondi62}. These works have been fundamental in establishing the feasibility of numerical formulations based on a characteristic coordinate and in motivating the present implementation. However, the stability analyses and experience of \cite{GomezEt92,GomezEt94a} are not directly applicable to our evolution procedure, since the formulations and numerical methods used have several significant differences. Like the Bondi parameterisation, the gauge conditions are directly implemented in the metric form; however, the NQS Einstein equations are considerably simpler than the Bondi coordinate form of the equations \cite{Bartnik97a,BishopEt97b}. The treatment of angular derivatives is greatly simplified by the quasi-spherical condition, which encourages the use of spherical harmonic expansions. These in turn will simplify comparisons with theoretical results on perturbations of the Schwarzschild black hole \cite{ReggeWheeler57,Zerilli70,Moncrief74,Chandrasekhar84}. We note that the power of spectral methods in practical applications is well-known, in fields such as meteorology \cite{Swarztrauber89}, astrophysics \cite{BonazzolaEt98}, and fluid dynamics \cite{CanutoEt88}. Finally, the combination of the characteristic-transport and method of lines techniques, considerably simplifies the use of high order algorithms such as RK4 for time evolution. The method of lines approach to evolution equations is common in fluid dynamics \cite{CanutoEt88}, but has not previously been attempted in numerical relativity. The situation we consider here, of smooth variations in a single black hole geometry, is well-suited to the method, since high frequency modes are less likely to be of physical interest and thus may be treated by smoothing or artificial viscosity. The use of higher order methods, together with spherical harmonics and radial smoothing, leads to considerably more accurate results than would be possible with the 2nd order methods more commonly employed. However, this suggests that our code is restricted to very smooth spacetimes, and cannot reliably treat spacetimes with strong localised features (such as planets, or gravitational shocks). Because we are concerned primarily with the gravitational wave perturbation modes of a single black hole, this does not present an important restriction, since the dominant modes are known to occur only for low angular momentum $l$ -- in particular, the $l=2,3,4$ modes are expected to carry practically all the radiated gravitational energy. The code models evolution in an exterior domain, and consequently boundary data must also be prescribed at an interior boundary surface $r=r_{0}$. In general only two pieces of the boundary data can be freely specified, thereby fixing the null hypersurfaces $\mathcal{N}_{z}$ and the outgoing radiation flux at the boundary \cite{Bartnik97a}. The rest of the boundary data is constrained by boundary evolution equations (see \bref{eq:Gnn},\bref{eq:Gnm}). For simplicity, the version of the code reported here assumes fixed boundary conditions, corresponding to the past horizon of a Schwarzschild black hole. This assumption considerably simplifies the treatment of the inner boundary, and is completely consistent with the Einstein evolution. It follows that the code models the interaction of gravitational waves with a single Schwarzschild-like black hole. Future versions of the code will incorporate dynamical inner boundary conditions. The paper is organised as follows. Section 2 gives the NQS form of the equations, and describes the structure of the equations and the formal steps in the solution algorithm. The geometric significance of the resulting equations is described in \cite{Bartnik97a}. Section 3 describes the numerical techniques used, including the representation of spin-weighted spherical harmonics used to encode the angular variation of the various fields, and the high order convolution splines used for interpolation and differentiation in the radial direction. Two aspects of the treatment of spherical fields appear to be non-standard \cite{Orszag74,Swarztrauber89}: the use of FFT's in both the $\varphi$ and $\vartheta$ directions, based on the ``torus'' model of $S^{2}$ \cite{Norton96}; and the use of \emph{spin-weighted} spherical harmonics to handle, in a unified and frame-invariant manner, vector and higher rank tensor harmonics as well as invariant derivative operators. Section 4 describes the various stages in the evolution algorithm --- solving the hypersurface equations out to null infinity $\mathcal{I}^{+}$; reconstructing the metric from the connection variables (which includes the solution of a 1st order elliptic system on the 2-sphere at each radial grid position); and the evolution of the primary field $\beta$. In section \ref{sec:accuracy} we describe various techniques for estimating the accuracy of the code, testing both numerical and geometric properties of the numerical solution. Numerical convergence tests estimate the effects of separately increasing the resolution in the radial, time and angular directions. The two geometric tests described here demonstrate the consistency of the numerical metric by verifying the constraint equations for the Einstein tensor components $G_{nn},G_{nm}$ (\ref{eq:Gnn},\ref{eq:Gnm}), and by testing the accuracy of the solution at infinity using the Trautman-Bondi mass decay formula \cite{Trautman58b,Trautman62,Bondi60,Bondi62}. These provide highly nontrivial tests of the consistency of the numerical solution. Other tests of the code, based on geometric properties of vacuum spacetimes, and comparisons with known exact and approximate solutions of the Einstein equations, are envisaged for future work. \section{Einstein equations and NQS metric functions} \label{sec2} \subsection{Spacetime metric} We consider spacetimes admitting global null-polar coordinates $(z,r,\vartheta,\varphi)$ in which the metric takes the null quasi-spherical form \cite{Bartnik97a} \begin{equation} ds^{2}_{NQS} = -2u\,dz(dr+v\,dz) + (r\,d\vartheta+\beta^{1}\,dr+\gamma^{1}\,dz)^{2} + (r\sin\vartheta\,d\varphi+\beta^{2}\,dr+\gamma^{2}\,dz)^{2}, \label{ds2:nqs} \end{equation} where $u>0,v$, and $\beta=\beta^{1}\partial_{\vartheta}+\beta^{2}\csc\vartheta\partial_{\varphi}$, $\gamma=\gamma^{1}\partial_{\vartheta} + \gamma^{2}\csc\vartheta\partial_{\varphi}$ are the unknown NQS metric functions, to be determined by numerical solution of the Einstein equations. Note that we use $\partial_{\vartheta}, \partial_{\varphi}$ to denote equivalently the coordinate tangent vectors and the coordinate partial differential operators $\frac{\partial}{\partial \vartheta}$, $\frac{\partial}{\partial \varphi}$. We may consider $\beta,\gamma$ either as vector fields on $S^{2}$ or as spin~1 quantities, defined by the complex combinations \cite{Goldberg67} \begin{equation} \label{vector} \beta = \tfrac{1}{\sqrt{2}} (\beta^{1}-{\rm i}\,\beta^{2}),\qquad \gamma = \tfrac{1}{\sqrt{2}} (\gamma^{1}-{\rm i}\,\gamma^{2}). \end{equation} The canonical example of a spacetime with metric in NQS form is Schwarzschild spacetime in Eddington-Finkelstein retarded coordinates \cite{dInverno92,HawkingEllis73} \begin{equation} ds^{2}_{Schw} = - 2 dz\,(dr+\tfrac{1}{2}(1-2\textsc{M}/r)\,dz) + r^{2}(d\vartheta^{2}+\sin^{2}\vartheta\,d\varphi^{2}) \label{ds2:schw} \end{equation} with $u=1$, $v=\tfrac{1}{2}(1-2\textsc{M}/r)$, $\beta^{A}=\gamma^{A}=0$ and $\textsc{M}=const$. This includes Minkowski space $\mathbb{R}^{3,1}$ as the case $\textsc{M}=0$ and $z=t-|x|$. \subsection{Edth} Using the complex notation \bref{vector}, we have the canonical angular covariant derivative operator ``edth'' \cite{Goldberg67,PenroseRindler84,EastwoodTod82}, \begin{equation}\label{edth} \eth \eta = \frac{1}{\sqrt{2}}\sin^s \vartheta \left( \frac{\partial } {\partial \vartheta} - \frac{{\rm i}\,}{\sin \vartheta} \frac{\partial }{\partial \varphi} \right)\left(\eta\,\sin^{-s}\vartheta\right) = \frac{1}{\sqrt{2}}\left( \frac{\partial }{\partial \vartheta} - s\cot\vartheta - \frac{{\rm i}\,}{\sin \vartheta} \frac{\partial }{\partial \varphi} \right)\,\eta , \end{equation} acting on a spin $s$ field $\eta$, and its ``conjugate'' operator $\bar{\eth}$, \begin{equation}\label{edthbar} \bar{\eth} \eta =\frac{1}{\sqrt{2}}\left( \frac{\partial }{\partial \vartheta} + s\cot\vartheta + \frac{{\rm i}\,}{\sin \vartheta} \frac{\partial }{\partial \varphi} \right)\,\eta . \end{equation} All geometric angular derivative operators may be defined in terms of $\eth$, $\bar{\eth}$. For example, the covariant directional derivative of a spin $s$ field $\eta$ in the direction $\beta$ is \[ \nabla_{\beta}\eta = \beta\bar{\eth} \eta + \bar{\beta}\eth \eta\ ; \] the divergence and curl (of a vector) are \begin{eqnarray} \div\beta & = & \bar{\eth}\beta + \eth\bar{\beta} = \nabla_{1}\beta^{1}+\nabla_{2}\beta^{2}, \label{div} \\ {\mbox{curl}} \beta & = & {\rm i}\, (\bar{\eth}\beta-\eth\bar{\beta}) = \nabla_{2}\beta^{1}-\nabla_{1}\beta^{2} \ ; \label{curl} \end{eqnarray} and the spherical Laplacian is \begin{equation}\label{laplace} \Delta \eta = (\eth\bar{\eth} + \bar{\eth}\eth)\eta \ . \end{equation} Further properties of edth are described in section \ref{sec3} and in \cite{EastwoodTod82,PenroseRindler84,Bartnik97a}. \subsection{Connection variables} \label{sec2.3} In addition to the metric functions $(u,v,\beta,\gamma)$ we introduce the connection fields $H,J,K,Q,Q^{\pm}$ \begin{eqnarray} H &=& \frac{1}{u}(2- \div\beta), \label{def:H} \\[5pt] J &=& v(2- \div\beta) + \div \gamma, \label{def:J} \\[5pt] K &=& v\eth \beta - \eth \gamma, \label{def:K} \\[5pt] Q &=& r\frac{\partial \beta}{\partial z} - r \frac{\partial \gamma}{\partial r} + \gamma + \nabla_\beta \gamma-\nabla_{\gamma}\beta, \label{def:Q}\\[5pt] Q^{\pm} &=& \frac{1}{u}(Q \pm \eth u). \label{def:Qpm} \end{eqnarray} Observe that $u,v,H,J$ are real and have spin~0, whereas $\beta,\gamma,Q,Q^{+},Q^{-}$ have spin~1 and $K$ has spin~2. Given the metric functions $u,v,\beta,\gamma$ on a $z$-level set $\mathcal{N}_{z}$, we may construct $H,J,K$ on $\mathcal{N}_z$ directly, and $Q$ (and $Q^{\pm}$) may be reconstructed if in addition, $\partial\beta/\partial z$ is also known on $\mathcal{N}_{z}$. It is clear from (\ref{def:H}--\ref{def:Qpm}) that this construction does not require any compatibility conditions on the data $u,v,\beta,\gamma,\frac{\partial}{\partial z}\beta$. Rather remarkably, there is a converse construction for the metric functions $u,v,\gamma$ and $\partial \beta/\partial z$, which also involves totally free and unconstrained data, namely the connection variables $H,J,K,Q$. This contrasts sharply with the description of the connection via the Newman-Penrose spin coefficients \cite{NP62,PenroseRindler84}, which requires numerous differential constraint equations, expressing the property that the connection is {}torsion-free. The converse construction works as follows. Given $\beta$ and the connection variables ($H,J,K$) on $\mathcal{N}_{z}$, we reconstruct $u$ via the relation \begin{equation} u = \frac{2-\div\beta}{H}, \label{u:def} \end{equation} and we find $v,\gamma$ by solving an elliptic system for $\gamma$, \begin{equation} \mathcal{L}_{\beta}\gamma := \eth \gamma + \frac{\eth \beta}{2 - \div \beta}\,\div \gamma = J \frac{\eth \beta}{2 - \div \beta} - K, \label{Lbeta:def} \end{equation} and setting \begin{equation} v = \frac{J-\div\gamma}{2-\div\beta}. \label{v:def} \end{equation} The system (\ref{Lbeta:def}) is $\mathbb{R}$-linear and elliptic with 6-dimensional kernel, provided $\eth\beta$ is not too large ($|\eth\beta| < (2-\div\beta)/\sqrt{3}$ is sufficient). Prescribing the $l=1$ spherical harmonic coefficients of $\gamma$ (for example, by requiring $\gamma_{l=1}=0$) suffices to determine the solution $\gamma$ uniquely. The remaining connection parameter $Q$, together with the now known values of $\beta,\gamma$ on $\mathcal{N}_{z}$, determines the \emph{evolution equation} \begin{equation} \label{dbdz:eq} \frac{\partial \beta}{ \partial z } = \frac{\partial \gamma}{ \partial r } + \frac{1}{r}( Q+ \nabla_\gamma \beta -\nabla_\beta \gamma -\gamma). \end{equation} {}To summarise, given the field $\beta$ on a single level set $\mathcal{N}_{z}$, satisfying the size constraint \begin{equation} |\eth\beta| < (2-\div\beta)/\sqrt{3}, \label{bsize} \end{equation} the map $(u,v,\gamma,\beta_{z})\mapsto (H,J,K,Q,\gamma_{l=1})$ is invertible, assuming all fields are sufficiently smooth. In section \ref{sec:elliptic} we will describe the numerical implementation of the inverse map. \subsection{NQS Einstein equations} {}To compute the curvature of $ds^{2}_{NQS}$, and thereby to determine the NQS form of the Einstein equations, we introduce the complex null vector frame $(\ell,n,m,\bar{m})$, \begin{eqnarray} \ell &=& \partial_{r}-r^{-1}\beta, \nonumber\\ n &=& u^{-1}(\partial_{z}-r^{-1}\gamma -v(\partial_{r}-r^{-1}\beta)), \label{tetrad:def} \\ m&=&\frac{1}{r\sqrt{2}} (\partial_{\vartheta}-{\rm i}\,\csc\vartheta\,\partial_{\varphi}), \nonumber \end{eqnarray} and the directional derivative operators \begin{equation} \mathcal{D}_r = \partial_{r}- r^{-1}\nabla_{\beta}\,, \quad \quad \mathcal{D}_z = \partial_{z}- r^{-1}\nabla_{\gamma}. \label{DrDz} \end{equation} Expressions for the Newman-Penrose spin coefficients \cite{NP62} with respect to the frame $(\ell,n,m,\bar{m})$ are given in terms of $H,J,K,Q$ in \cite{Bartnik97a}. The frame components of the Einstein tensor $G_{ab}$, $a,b=\ell,n,m,\bar{m}$, may be written in terms of the NQS metric functions and NQS connection variables. These expressions may be grouped into {\em hypersurface equations} (or \emph{main equations} \cite{Bondi62,Sachs62a}): \begin{eqnarray} r \mathcal{D}_{r}H&=& \left(\tfrac{1}{2}\div\beta - \frac{2|\eth\beta|^2+r^{2}G_{\ell\ell}}{2-\div\beta} \right)H, \label{eq:Gll} \\[3pt] r\mathcal{D}_{r}Q^{-} &=& (\eth\bar{\beta}-uH)Q^{-} + \bar{Q}^{-}\eth\beta + 2\bar{\eth}\eth\beta {}+ u\eth H - H\eth u + 2r^{2} G_{\ell m} , \label{eq:Glm} \\[3pt] r \mathcal{D}_{r}J &=& -( 1-\div\beta)J + u - \tfrac{1}{2} u |Q^{+} |^2 {}-\tfrac{1}{2} u\,\div(Q^{+} ) - u r^2 G_{\ell n} , \label{eq:Gln} \\[3pt] r \mathcal{D}_{r}K &=& \left(\tfrac{1}{2} \div\beta + {\rm i}\,{\mbox{curl}}\beta \right) K -\tfrac{1}{2} J \eth\beta {}+ \tfrac{1}{2} u \eth Q^{+} +\tfrac{1}{4} u (Q^{+} )^2 +\tfrac{1}{2} ur^2G_{mm}, \label{eq:Gmm} \end{eqnarray} the {\em boundary equations} (or \emph{subsidiary equations}) \begin{eqnarray} r\,\mathcal{D}_z\left({J/u}\right) &=& {v^2}\,r\mathcal{D}_r\left(J/(uv)\right) +\tfrac{1}{2}(\div\gamma - v\,\div\beta)J/u + 2u^{-1}|K|^{2} \nonumber\\ && {} - \nabla_{Q^{+}}v - \Delta v + u {r^2} G_{nn} , \label{eq:Gnn} \\[3pt] r\,\mathcal{D}_z Q^{+} &=& (v\,r\mathcal{D}_r + J -v\eth\bar{\beta}+\eth\bar{\gamma})Q^{+} - K\bar{Q}^{+} + 2\bar{\eth} K + 2u^{-1}r\mathcal{D}_{r}(u\eth v) \nonumber\\ && {} -(2 + {\rm i}\, {\mbox{curl}}\beta)\eth v +\eth J -2u^{-1}J\eth u - 2u r^{2} G_{nm} , \label{eq:Gnm} \end{eqnarray} and the {\em trivial equation} \begin{eqnarray} \label{eq:Gmb} u r^{2}G_{m\bar{m}} & = & r\mathcal{D}_{r}J -\tfrac{1}{2} \div \beta\,J - u|Q^{+}|^{2} +\tfrac{1}{2} u\,\div Q^{+} + \bar{Q}^{+}\eth u + Q^{+}\bar{\eth} u \nonumber \\[1pt] & & {} + \bar{K}\eth\beta + K \bar{\eth}\bar{\beta} + r^{2}(v \mathcal{D}_{r} - \mathcal{D}_{z})(u^{-1}\mathcal{D}_{r}u) + u^{-1}r^{2}\mathcal{D}_{r}(u\mathcal{D}_{r}v). \end{eqnarray} Observe that the hypersurface equations (\ref{eq:Gll}--\ref{eq:Gmm}) have no explicit $z$-derivatives, and they each contain only one radial ($r$) derivative. The form of the connection variables ($H,J,K,Q$) was determined by exactly these properties. Consequently, the hypersurface equations may be written schematically in terms of $U=(H,Q^{-},J,K)$ in the ``characteristic-transport'' form \begin{equation} r\frac{\partial U}{\partial r} = F(\beta(z,r),U(z,r)) , \label{transport} \end{equation} by treating angular derivatives such as $\eth U$ as determined by the set of values $U(z,r)$ on the full $S^{2}$. This system has the effect of transporting the fields $U$ along the characteristic curves with tangent vector $\ell$ which foliate the null hypersurfaces $\mathcal{N}_{z}$. Note that alternative reformulations of the hypersurface equations are possible, preserving the general characteristic transport structure. For reasons associated with reliably capturing the asymptotic behaviour of the fields, the present version of the code integrates the following radial equations, for the variables $\log (H/2)$ (instead of $H$), $j=\tfrac{1}{4} r(1-HJ)-\textsc{M}$ (instead of $J$) and $rQ^{+}$ (instead of $Q^{-}$): \begin{eqnarray} r\partial_{r}\log (H/2) & = & \nabla_{\beta}\log (H/2) + \tfrac{1}{2}\div\beta - \frac{2|\eth\beta|^2+r^{2}G_{\ell\ell}}{2-\div\beta} , \label{eq:lh} \\ r\partial_{r}(rQ^{+}) & = & \nabla_{\beta}(rQ^{+}) - (1-2\eth\bar{\beta})(rQ^{+}) +\nabla_{rQ^{+}}\beta +2r\eth(r\mathcal{D}_r\log u) \nonumber \\ &&{}+2r\beta-{\rm i}\, r\eth{\mbox{curl}}\beta +2r^{3}G_{\ell m} , \label{eq:rQp} \\ r\partial_{r}j & = & \nabla_{\beta}j + (j+\textsc{M}-\tfrac{1}{2} r)(\div\beta - r\mathcal{D}_r\log u) \nonumber \\ &&{} +\tfrac{1}{4} r (|Q^{+}|^{2}+\divQ^{+}) + \tfrac{1}{2} r^{3} G_{\ell n} ,\ \label{eq:j} \\ 2r\partial_{r}K & = & 2\nabla_{\beta}K + \left(\div\beta + 2{\rm i}\,{\mbox{curl}}\beta \right) K - J \eth\beta \nonumber \\ &&{} + u \eth Q^{+} +\tfrac{1}{2} u (Q^{+} )^2 + ur^2G_{mm}. \label{eq:K} \end{eqnarray} Here $\textsc{M}=1$ is a constant which fixes the bare mass of the background Schwarzschild black hole. Of course, the Einstein tensor components in these formulae are set to zero for the vacuum equations. It is remarkable that the boundary equations (\ref{eq:Gnn},\ref{eq:Gnm}) and the trivial equation \bref{eq:Gmb} may be regarded as \emph{compatibility} relations, by virtue of the conservation (contracted Bianchi) identity $G_{ab}^{\ \ ;b}=0$ \cite{Sachs62a,Bartnik97a}. This identity is valid for any Einstein tensor $G_{ab}$, regardless of the metric. Substituting the hypersurface equations $G_{\ell\ell}=G_{\ell m}=G_{\ell n}=G_{mm}=0$ into the conservation identity, yields equations $HG_{m\bar{m}}=0$ and a propagation system for $G_{nn},G_{nm}$ which has the unique solution $G_{nn}=G_{nm}=0$ if the boundary equations are satisfied on one hypersurface transverse to the outgoing null surfaces $\mathcal{N}_{z}$. Thus in order to construct a solution of the full vacuum Einstein equations, it suffices to satisfy the hypersurface equations everywhere, and the boundary equations just on the boundary surface $r=r_{0}$ (for example). \section{Numerical techniques} \label{sec3} In this section we describe the data representation and manipulation techniques. These consist mainly of techniques for handling angular fields and derivatives, and an unusual convolution spline used for interpolation, differentiation and high frequency filtering in the radial and time directions. \subsection{Fields on $S^{2}$} The evolution algorithm treats the angular derivatives $ \frac{\partial }{ \partial \vartheta}, \frac{\partial }{ \partial \varphi}$ as ``lower order'', compared to the radial and time derivatives $ \frac{\partial }{ \partial r}, \frac{\partial }{ \partial z}$. This attitude in a numerical computation can be justified only if it is possible to easily and accurately compute and manipulate angular derivatives. This is achieved by using spectral representations (both Fourier and spherical harmonic) for fields on the 2-sphere. This approach is widely used in geophysical and meteorology applications \cite{Merilees73b,Orszag74,Boyd78a,Swarztrauber79,Swarztrauber96} and is known to have significant advantages compared to finite difference approaches \cite{Swarztrauber89}, based on either angular coordinate grids or overlapping stereographic projection charts \cite{Starius77,Singleton90,GomezEt97a}. Nevertheless, spectral methods have rarely been used in numerical general relativity (however, see \cite{Prager96,BonazzolaEt98}) and they have not been used previously for solving the full Einstein equations. The basic manipulations required of $S^{2}$ fields are: \begin{itemize} \item computing non-linear algebraic terms such as $1/(2-\div\beta)$, $u|Q^{+}|^{2}$ etc; \item computing angular derivatives operators such as $\div\beta,\ \ethQ^{+}$ etc; \item inverting the linear elliptic operator $\mathcal{L}_{\beta}$ (which appears in equation (\ref{Lbeta:def})); and \item projecting aliased or noisy field value data onto certain subspaces of spin-weighted spherical harmonics. \end{itemize} {}To carry out these manipulations, three separate representations are used for fields on $S^2$: \begin{itemize} \item \emph{field values} $\eta_{jk}=\eta(\vartheta_{j},\varphi_{k})$ at the polar coordinate grid points \begin{equation} (\vartheta_{j},\varphi_{k})=((j-\tfrac{1}{2})\Delta\vartheta,(k-1)\Delta\varphi), \label{phik} \end{equation} where $\Delta\vartheta=2\pi/N$, $\Delta\varphi=2\pi/N$ with $1\le j\le N/2,1\le k\le N$ and (in our implementations) $N=16,\,32$ or $64$; \item \emph{Fourier coefficients} $\hat{\eta}_{mn}$ arising from FFT transforms in the $\vartheta$ or $\varphi$ directions, of the field values $\eta_{jk}$; \item spin-weighted \emph{spherical harmonic coefficients} $\eta^{lm}$, $|m|\le l$, $l=s,\ldots,L$, $L=N/2-1$ (for spin $s=0,1$ and 2). \end{itemize} The field value representation is used when computing non-linear algebraic terms such as $u|Q^{+}|^{2}$. The Fourier representation is used for computing $\vartheta$ and $\varphi$ angular derivatives, which are needed in the formulas for $\eth$, $\div$, for example. The spherical harmonic representation is used in solving the elliptic system (\ref{Lbeta:def}), {}to spectrally limit the field values by projection to spherical harmonic data, and to summarise the computation results (which are stored using spherical harmonic coefficients). For fields which do not alias on the $(\vartheta,\varphi)$-grid, the three representations are completely equivalent in the sense that conversion between them is essentially exact, depending on the machine precision and on algebraic details of the specific FFT algorithm used. The requirement that a field does not alias is satisfied when it can be represented by a finite expansion in spin-weighted spherical harmonics with angular momentum $l\le L=N/2-1$. Our implementations use spectral cutoffs $L=15$ (for $N=32$) and $L=31$ (for $N=64$). Transformation to the spherical harmonic representation involves a projection, because both the field value and Fourier representations have approximately twice as many degrees of freedom. For example, a (real) spin~0 field with $l \le L= N/2-1$ has $(L+1)^2 = N^2/4$ spherical harmonic coefficients, whereas it has $N^2/2$ values on the $(\vartheta,\varphi)$-grid. The space of non-aliasing spherical harmonics is a {\em linear} subspace of the space of functions represented by either Fourier coefficients or field values. For example, when the $(\vartheta,\varphi)$-grid field values of the product of two fields is calculated, the result, which will contain components up to $l \le 2L$, is aliased onto the grid in such a way that its field values no longer lie in the appropriate spin weighted spherical harmonic subspace. To clean up after such non-linear effects, we project the result back onto the correct subspace, as described in section \ref{sph:sec}. {}To minimise the possibility of unstable feedback of quadratic aliasing errors, we may invoke the Orszag 2/3 rule \cite{CanutoEt88, Orszag71} at various points within the code. The effective spectral resolution of the code is then $l_{\rm max} \approx 2L/3$ ($l_{\rm max} = 10$ for $N=32$ and $l_{\rm max} = 20$ for $N=64$). \subsection{Spherical harmonics} We first summarise the more important properties of $\eth$ (``edth'') and spin weighted spherical harmonics. The edth formalism provides a unified geometric approach to the treatment of angular derivatives on $S^{2}$ and vector and higher-rank tensor harmonics. Detailed descriptions of the properties of spin-weighted fields and spherical harmonics may be found in Penrose and Rindler \cite{PenroseRindler84} or \cite{Goldberg67,EastwoodTod82}. Here we describe only the basic formulae. We use a real-valued basis $Y_{lm}$, $l = 0,1,2,\ldots$, $m = -l,\ldots,l$ for the space of spin~0 spherical harmonic functions, defined by \begin{equation}\label{Ylm} Y_{lm} = \overline{P}_{lm}(\vartheta)F_m(\varphi)\,, \end{equation} where \begin{equation}\label{Fm} F_m(\varphi) = \left\{ \begin{array}{cr} 1 & \quad m = 0 \\ \sqrt{2}\cos m\varphi & m > 0 \\ \sqrt{2}\sin |m|\varphi & m < 0, \end{array}\right. \end{equation} and the $\overline{P}_{lm}(\vartheta) = \overline{P}_{l|m|}(\vartheta)$ are related to the associated Legendre functions $P_{lm}$ by \begin{eqnarray} \label{Pblm} \overline{P}_{lm}(\vartheta) &=& (-1)^m \sqrt{2l+1} \sqrt{\frac{(l - m)!}{(l + m)!} } P_{lm}(\cos \vartheta)\,, \\ \label{Plm} P_{lm}(\cos \vartheta) &=& \frac{(-1)^m}{2^l l !} \sin^m\vartheta \left[ \frac{d^{l+m}}{dx^{l+m}}(x^2-1)^l \right]_{x=\cos\vartheta}\,. \end{eqnarray} The spin $s$ spherical harmonics $Y^{s}_{lm}$ are then defined explicitly by \begin{eqnarray}\label{Yslm} Y^{s}_{lm}& = & \phantom{(-1)^{s}} \left[\frac{2^{s}(l-s)!}{(l+s)!}\right]^{1/2} \eth^{s}Y_{lm}\,,\qquad s>0, \label{Y+slm} \\ Y^{-s}_{lm}& = & (-1)^{s} \left[\frac{2^{s}(l-s)!}{(l+s)!}\right]^{1/2} \bar{\eth}^{s}Y_{lm}\,,\qquad -s<0, \label{Y-slm} \end{eqnarray} where necessarily $l\ge |s|$. Note that the differential operator $\eth$ is spin-raising, sending spin $s$ into spin $(s+1)$ fields, and $\bar{\eth}$ is spin-lowering, \begin{eqnarray} \eth Y^{s}_{lm} & = & \left[\tfrac{1}{2}(l+s+1)(l-s)\right]^{1/2} Y^{s+1}_{lm}, \label{edthYslm} \\ \bar{\eth} Y^{s}_{lm} & = & -\left[\tfrac{1}{2}(l+s)(l-s+1)\right]^{1/2} Y^{s-1}_{lm}, \label{edthbarYslm} \end{eqnarray} for all $s\in\mathbb{Z}$, and $Y^s_{lm}$ and $Y^{-s}_{lm}$ are related by complex conjugation, \begin{equation} \label{Y-s} Y^{-s}_{lm} = (-1)^{s} \bar{Y}^{s}_{lm}. \end{equation} Since $\Delta Y_{lm}= -l(l+1) Y_{lm}$, the fundamental commutation relation \begin{equation} [\bar{\eth},\eth]\eta = (\bar{\eth}\eth - \eth\bar{\eth})\eta = s \eta, \label{commutator} \end{equation} for any spin $s$ field $\eta$, may be used to show that \begin{equation} \Delta Y^{s}_{lm} = (s^{2}-l(l+1)) Y^{s}_{lm}, \label{lap-s} \end{equation} where $\Delta=\eth\bar{\eth}+\bar{\eth}\eth$. With these conventions we have the orthogonality relations \[ \frac{1}{4\pi}\int_{S^2} Y^{s}_{lm}\, \bar{Y}^{s}_{l'm'} \sin\vartheta\, d\vartheta d\varphi = \delta_{ll'} \delta_{mm'}, \] which show that the $Y^{s}_{lm}$ form a basis (over $\bC$) of the Hilbert space of square-integrable spin $s$ fields on $S^{2}$, which is orthonormal in the natural Hermitian inner product \begin{equation} \langle \phi,\psi\rangle = \frac{1}{4\pi}\oint_{S^{2}}\Re( \bar{\phi}\psi). \label{s2ip} \end{equation} {}From (\ref{Ylm}--\ref{Plm}) it is evident that the spin~0 harmonics $Y_{lm}$ are trigonometric polynomials in $\vartheta$ and $\varphi$ \cite{Merilees73b}. Using expression (\ref{edth}) for $\eth$, we also see that the $ Y^{s}_{l m}$ are trigonometric polynomials. The highest wave number Fourier modes which occur in the set of basis functions $\{Y^{s}_{l m} : s\leq l \leq L,\ |m|\le l \}$ are $\cos (L\vartheta),\ \sin (L\vartheta),\ \cos (L\varphi)$, and $\sin (L\varphi)$. Therefore, on a uniform $(\vartheta,\varphi)$-grid of size $N/2\times N,$ one can represent all of the spin weighted spherical harmonic functions up to $L = N/2 - 1$. \subsection{Even/Odd decomposition} Because $\eth$ is surjective onto the space of smooth spin $s$ fields for $s\ge1$ \cite{PenroseRindler84}, the decomposition of spin~0 fields or functions into real and imaginary parts may be propagated to higher spin. The resulting decomposition into \emph{even} and \emph{odd} components plays an important role in the analysis of the linearised Einstein equations about the Schwarzschild spacetime \cite{ReggeWheeler57}. Because the NQS geometry distinguishes the Schwarzschild metric and is also based on spherical harmonics, it is ideally suited to comparing nonlinear evolution to the comparatively well-understood black hole linearised Einstein equations \cite{Chandrasekhar84, FuttermanEt88}. It is thus not surprising that the even-odd decomposition proves to be very important in analysing the results of the NQS evolution. We say that the spin $s$ field $\eta$ is \emph{even} if $\eta =\eth^{s}f$ for some real-valued function $f$, and $\eta$ is \emph{odd} if $\eta ={\rm i}\,\eth^{s}g$ for some real-valued function $g$. (If $s<0$ then we interpret $\eth^s$ as $(-\bar{\eth})^{|s|}$). This matches the usage in \cite{ReggeWheeler57} --- note that for axially symmetric fields the terms \emph{polar} (for even) and \emph{axial} (for odd) are sometimes used \cite{Chandrasekhar84}. The surjectivity of $\eth$ onto spin $s\ge1$ ensures that every spin $s$ field may be uniquely decomposed into a sum of even and odd parts. For $s=1$ this decomposition corresponds exactly to the classical Hodge-Helmholtz decomposition of a vector field into the sum of a gradient and a dual gradient (or curl) --- see Table \ref{table:odd-even} for a summary of the various nomenclatures. \begin{table}[ht] \centering \begin{tabular}{c|c|c|c|c|c} Even: & polar & irrotational & $\eth f$ & $\textrm{grad}f$ & $(\nabla_{1}f)\,v_{1}+(\nabla_2f)v_2 $ \\ \hline \hline Odd: & axial & divergence-free & ${\rm i}\, \eth g$ & ${\mbox{curl}} g$ & $(\nabla_{2}g)\,v_{1}-(\nabla_{1}g)v_{2}$ \\ \end{tabular} \caption{Equivalent terminologies for vector fields on $S^{2}$} \label{table:odd-even} \end{table} The even/odd decomposition has a natural interpretation in terms of the spectral decomposition \begin{equation} \label{eq:shc} \eta = \sum_{l=s}^{\infty} \sum_{m=-l}^{l} \eta^{lm} Y^{s}_{lm} \end{equation} of a spin $s$ field $\eta$, because we are using a basis of real-valued $Y_{lm}$. Namely, $\eta$ is \emph{even} if the spectral coefficients $\eta^{lm}$ are real, and \emph{odd} if the coefficients are pure imaginary. We sometimes use $\textrm{Even}(\eta)$ and $\textrm{Odd}(\eta)$ {}to represent the respective projections, so $\eta = \textrm{Even}(\eta) + \textrm{Odd}(\eta)$ with \begin{eqnarray} \label{even} \textrm{Even}(\eta) & = & \sum\Re(\eta^{lm}) Y^{s}_{lm} \\ \label{odd} \textrm{Odd}(\eta) & = & {\rm i}\, \sum\Im(\eta^{lm}) Y^{s}_{lm}. \end{eqnarray} Observe that $\eth Y^{s}_{sm}=0$ for $s\ge0$ and thus $\eth$ acting on spin $s$ fields with $s\ge0$ has kernel having (complex) dimension $2s+1$. Likewise the formal adjoint $-\bar{\eth}$ acting on spin $s\le0$ fields has $(2|s|+1)$-dimensional kernel. In particular, $\eth$ acting on spin~1 fields has kernel consisting of the $\bC$-linear space spanned by the three $l=1$ spin~1 spherical harmonics $Y^{1}_{1m}$ --- the corresponding real vector fields are the dual gradients of functions linear in $\mathbb{R}^{3}$, which are just the infinitesimal rotations, and the gradients, which are the conformal dilation vector fields. The correspondence between vector fields on $S^2$ and spin~1 fields generalises to spin~2 fields, which correspond to symmetric traceless 2-tensors on $S^{2}$. If $\lambda$ is a symmetric traceless 2-tensor then with respect to the standard polar coordinate derived orthonormal frame \begin{equation} e_{1} = \partial_{\vartheta},\quad e_{2} = \csc\vartheta\,\partial_{\varphi}, \label{e12} \end{equation} we have the correspondence \begin{equation} \lambda \sim \tfrac{1}{2} (\lambda^{11}-\lambda^{22} - 2{\rm i}\,\lambda^{12})\,. \label{2tensor} \end{equation} This correspondence extends to higher integer spins with higher rank symmetric traceless tensors on $S^{2}$. The cases $s=0,1,2$ of most importance in our work correspond to the usual scalar, vector and tensor harmonics. \subsection{Fourier representation} The fast Fourier transform (FFT) is used to transform between Fourier coefficients and field values on the uniform $(\vartheta,\varphi)$-grid. Fourier convergence problems arising from discontinuities in coordinate derivatives and vector and tensor components at the poles, are sidestepped by an observation relating fields on $S^{2}$ to fields on the torus ${\mathbb{T}}^{2}=S^{1}\times S^{1}$ \cite{Norton96}. The torus method enables coordinate and covariant derivatives for all types of field to be computed using Fourier methods, so is particularly well-suited to handling the derivative operator $\eth$ \bref{edth}. This approach to handling component discontinuities at the poles is simpler than the techniques reviewed in \cite{Swarztrauber81} for manipulating vector fields, and readily extends to any rank $s\ge 0$. For integer $s$ the real and imaginary parts of a spin $s$ field on $S^2$ may be identified with the two independent frame components of a completely symmetric trace-free tensor of rank $|s|$ on $S^2$ \cite{PenroseRindler84}. Since the frame $e_{1},e_{2}$ (\ref{e12}) is not continuous at the poles, the tensor components will not be continuous at the poles, so are not obviously suited to Fourier expansion in the $\vartheta$ direction. However, along any smooth curve crossing through a pole, both basis vectors $e_1$ and $e_2$ reverse direction at the pole. Thus, for any smooth tensor field $T = T^{j_1 \ldots j_s} e_{j_1}\otimes\cdots\otimes e_{j_s}$, by continuity of $T$ the component functions $T^{j_1 \ldots j_s}$ will change by a factor $(-1)^{s}$ across the poles. Consequently, if we extend the domain of definition of $T^{j_1 \ldots j_s}$ to $\vartheta\in [-\pi,\,\pi]$ by \begin{equation}\label{T-torus} T^{j_1 \ldots j_s}(-\vartheta,\varphi) \;=\; (-1)^s T^{j_1 \ldots j_s}(\vartheta,\varphi+\pi)\,, \quad \mathrm{for }\ \vartheta \in [0,\pi] \end{equation} (using the $2\pi$ periodicity in $\varphi$), then the resulting extension is $2\pi$-periodic and continuous in $\vartheta$. This argument extends to higher (covariant) derivatives of $T$, showing that the extension is in fact \emph{smooth} and periodic in $\vartheta$. Derivatives of $T^{j_1 \ldots j_s}$ with respect to $\vartheta$ can then be calculated just as for $\varphi$ derivatives, provided that the direction of increasing $\vartheta$ is properly taken into account. In effect, the extension just described defines a (smooth) field $T^{j_1\ldots j_s}$ on the torus $\mathbb{T}^{2}=S^1 \times S^1$. This may be understood geometrically by noting that the map \begin{equation} \Upsilon:{\mathbb{T}}^{2}\to S^{2}, \quad (\vartheta,\varphi)\mapsto \left\{ \begin{array}{ll} (\vartheta,\varphi), &\vartheta\in (0,\pi], \ \phi\in (-\pi,\pi] \\ (-\vartheta,-\varphi), &\vartheta\in (-\pi,0],\ \phi\in (-\pi,\pi] \\ \end{array}\right. \label{torusmap} \end{equation} is in fact \emph{smooth}. This follows by noting that because $\vartheta$ is a radial coordinate near the north pole $\vartheta=0$, the differential structure near the pole is represented by the rectangular coordinates $(\xi,\eta) = (\vartheta\cos\varphi,\vartheta\sin\varphi)$ and the map $(\vartheta,\varphi)\mapsto (\xi,\eta)$ is manifestly $C^{\infty}$ for $\vartheta$ near $0$. Consequently any smooth tensor $T$ on $S^{2}$, when expressed in a cotangent basis, pulls back to a smooth tensor on ${\mathbb{T}}^{2}$ (ie.~$\Upsilon^{*}(T)\in C^{\infty}({\mathbb{T}}^{2})$), and thus admits a well-behaved Fourier representation on ${\mathbb{T}}^{2}$. The converse is of course false: a smooth tensor on ${\mathbb{T}}^{2}$ does not necessarily arise from a smooth tensor on $S^{2}$, even if it satisfies the parity condition \bref{T-torus} satisfied by pull-back tensors. The coordinate derivative form \bref{edth} of $\eth$ (and similarly any other covariant angular derivative operator) can be easily evaluated by transforming to the Fourier representation of the field, multiplying the Fourier coefficients by the appropriate wavenumber factors, transforming back to obtain the field value representation of the $\vartheta$ and $\varphi$ derivatives, then finally including the $\csc\vartheta$ factors. Thus, computing a derivative operator such as $\eth$ is an $O(N^{2}\log N)$ operation. Since $\csc(\tfrac{1}{2}\Delta\vartheta)\simeq 10$ for $N=32$, there is no significant loss of accuracy in calculations near the poles. It should be emphasised that because the $Y^{s}_{l m}$ are trigonometric polynomials in $(\vartheta,\varphi)$, the FFT computation of their numerical derivatives is {\em algebraically exact}. For example, the Laplacian relation \bref{lap-s} is numerically verified to 1 part in $10^{12}$ \cite{Norton96}. \subsection{The spherical harmonic representation} \label{sph:sec} The field value and Fourier representations suffice for most numerical calculations, such as computing the nonlinear terms in (\ref{eq:Gll}--\ref{eq:Gmm}). However, at some steps it is essential {}to use the representation by spherical harmonic coefficients \bref{eq:shc}: \begin{enumerate} \item to solve the elliptic system \bref{Lbeta:def} (see section \ref{sec:elliptic}); \item to implement spectral projections after each time step, with the aim of suppressing aliasing effects and rounding errors at the poles \cite{Orszag74}; \item to store results for later analysis and display, since the spherical harmonic representation is more compact and the coefficients may be readily interpreted physically, by comparison with well-studied solutions of the linearised black hole Einstein equations \cite{ReggeWheeler57,Zerilli70,Chandrasekhar84}. \end{enumerate} We next describe the methods used to transform fields from the field value and Fourier representations to the spherical harmonic representation. Transformations for fields of spin~0, 1 and 2 are required by the code; the spin~0 case which we describe here to illustrate the technique is slightly less complicated, since we may assume the field $f$ is real-valued. There are $(L+1)^2$ basis functions in the set $\{ Y_{l m}: 0\leq |m| \leq l \leq L\}$. However, to represent these functions as trigonometric polynomials on a regular $S^2$ grid we require a grid of size $(L+1)\times 2(L+1)$, and thus $2(L+1)^{2}$ real coefficients. The spin~0 functions of angular momentum at most $L$ therefore form a subspace of real dimension $(L+1)^2$ in the space of Fourier series representable on the grid, which has real dimension $2(L+1)^2$. We use a projection onto the spherical harmonic subspace which is orthogonal with respect to the natural inner product in the Fourier space, \begin{eqnarray} \langle f_1,\, f_2\rangle_{2} &=& \frac{1}{4\pi^2}\int_0^{2\pi}\int_0^{2\pi} f_1(\vartheta,\varphi)f_2(\vartheta,\varphi)d\vartheta d\varphi \nonumber \\ \label{f1f2} &=& \frac{1}{N^2}\sum_{i=1}^{N}\sum_{j=1}^{N} f_1(\vartheta_i,\varphi_j) f_2(\vartheta_i,\varphi_j)\,, \end{eqnarray} where $\{(\vartheta_i,\varphi_j): i,j=1,\ldots,N\}$ are grid points (cf.~\bref{phik}) and $N = 2(L+1)$. {}To make use of (\ref{f1f2}) we use (\ref{T-torus}) to extend functions defined on $S^2$ to functions defined on the torus $\mathbb{T}^2 = S^1\times S^1$. In particular, given any set of values $\{f_{ij} \in \mathbb{R}: i=1,\ldots,N/2,\ j=1,\ldots,N\}$ on the $S^2$ grid, we use (\ref{T-torus}) to construct grid values on $\mathbb{T}^2.$ There is then a unique interpolating trigonometric polynomial $f$ such that $f(\vartheta_i, \varphi_j) = f_{ij}$. We project $f$ to the $l\le L$ spherical harmonic subspace as follows. The $Y_{l m}$ are not orthonormal with respect to (\ref{f1f2}), but instead have Fourier inner product \begin{eqnarray} G_{l m\; l^\prime m^\prime} &=& \langle Y_{l m},\, Y_{l^\prime m^\prime} \rangle_{\textsc{F}} \nonumber \\ &=& \langle \overline{P}_{l m}(\vartheta),\, \overline{P}_{l^\prime m}(\vartheta) \rangle_{\textsc{F}} \, \delta_{m m^\prime}\,. \label{Gdown} \end{eqnarray} Here, the index pair $l m$ (and $l^\prime m^\prime$) is a combined index which takes $(L+1)^2$ values, and \bref{Gdown} is the matrix for the induced Fourier metric on the spin~0 subspace. The summation convention will be employed for raised and lowered repeated indices. For fixed $m$, the inner product (\ref{Gdown}) of the $\overline{P}$ functions forms a matrix, \begin{displaymath} A_{(m)ll^\prime} = \langle \overline{P}_{l m}(\vartheta) ,\, \overline{P}_{l^\prime m}(\vartheta) \rangle_{\textsc{F}} \,. \end{displaymath} These matrices are defined only for $|m| \leq l,l' \leq L$, so are square and of size $(L+1 - |m|)\times (L+1 - |m|)$. Denoting the inverse matrix by $A_{(m)}^{ll^\prime}$, we have \begin{equation} \label{Gdown2} G_{l m\; l^\prime m^\prime} = A_{(m)ll^\prime} \delta_{m m^\prime}\,, \end{equation} and the components of the inverse metric are given by \begin{equation}\label{Gup} \quad\quad\quad\quad\quad G^{l m\; l^\prime m^\prime} = A_{(m)}^{ll^\prime} \delta^{m m^\prime} \ \quad (\mbox{no sum on}\ m). \end{equation} The dual basis vectors for the spin~0 subspace are \begin{equation}\label{Yup} Y^{l m} = G^{l m\; l^\prime m^\prime} Y_{l^\prime m^\prime}\,, \end{equation} and satisfy $ \langle Y_{l m},\, Y^{l^\prime m^\prime} \rangle_{\textsc{F}} = \delta^{l^\prime}_l \delta^{m^\prime}_m\, $. The orthogonal projection of $f$ onto the subspace is given by \begin{displaymath} {\rm proj}(f) = \langle f,\, Y^{l m} \rangle_{\textsc{F}} Y_{l m} = f^{lm}Y_{lm}, \end{displaymath} where \begin{equation}\label{flm} f^{l m} = \langle f,\, Y^{l m} \rangle_{\textsc{F}} \end{equation} are the spherical harmonic coefficients of the function $f$. {}To calculate the inner product (\ref{flm}), first note that using (\ref{Ylm}), (\ref{Gup}) and (\ref{Yup}), the dual basis vectors can be written as \begin{equation}\label{Yup2} Y^{l m} = \overline{P}^{l m}(\vartheta) F^m(\varphi)\, \end{equation} (in analogy with (\ref{Ylm})), where we have set \begin{equation}\label{Pup} \overline{P}^{l m}(\vartheta) = A_{(m)}^{l l^\prime} \overline{P}_{l^\prime m}(\vartheta)\,, \quad F^m(\varphi) = F_{m}(\varphi) . \end{equation} By Fourier analysis of $f$ in the $\varphi$ direction we can write $f = \hat{f}^k(\vartheta)F_k(\varphi)$. In particular, by $\varphi$-FFT of $\{f_{ij}\}$ we get the numbers $\hat{f}^k(\vartheta_i)$. The spectral coefficients $f^{lm}$ can then be evaluated using (\ref{f1f2}) and (\ref{Yup2}) as \begin{eqnarray} f^{l m} &=& \langle \hat{f}^k(\vartheta)F_k(\varphi),\, \overline{P}^{l m}(\vartheta)F^m(\varphi) \rangle_{\textsc{F}} \nonumber \\ &=& \langle \hat{f}^m(\vartheta),\, \overline{P}^{l m}(\vartheta) \rangle_{\textsc{F}} \nonumber \\ &=& \frac{1}{N} \sum_{i=1}^{N} \hat{f}^m(\vartheta_i) \overline{P}^{l m}(\vartheta_i)\,. \label{fup} \end{eqnarray} The converse process of reconstructing the function values $f_{ij} = f(\vartheta_i,\varphi_j)$ from the spherical harmonic coefficients $f^{l m}$ follows from \begin{eqnarray*} f &=& f^{l m}\, Y_{l m}(\vartheta,\varphi) \\ &=& f^{l m} \overline{P}_{l m}(\vartheta)F_m(\varphi)\,. \end{eqnarray*} First we construct the quantities \begin{equation}\label{fhatm} \quad\quad\quad\quad\quad \hat{f}^m(\vartheta_i) = \sum_{l=|m|}^L f^{l m} \overline{P}_{l m}(\vartheta_i)\,, \ \quad (\mbox{no sum on}\ m), \end{equation} and then we use inverse FFTs in the $\varphi$ direction to reconstruct $f_{ij}$ via \begin{displaymath} f_{ij} =\sum_{m=-L}^{L} \hat{f}^m(\vartheta_i) F_m(\varphi_j)\,. \end{displaymath} Both constructions, of $f^{lm}$ from $f_{ij}$ and conversely, are $O(L^{3})$ operations, due to the matrix multiplications in \bref{fup}, \bref{fhatm}. Routines for transforming between grid values and spherical harmonic coefficients have been implemented for maximum angular momentum $L = 7,15\ {\rm and}\ 31$. The grid values $\overline{P}^{l m}(\vartheta_i)$ which appear in the sum (\ref{fup}) were pre-computed in multiple precision using REDUCE \cite{Hearn}. The functions $\overline{P}^{l m}(\vartheta)$ defined by (\ref{Pup}), were constructed symbolically using exact inversion of the matrices $A_{(m)\,ll^\prime}$. This symbolic approach was feasible because the metric $G_{l m\; l^\prime m^\prime}$ factorized as the tensor product (\ref{Gdown}), thus allowing exact inversion of $G$ using matrices of size at most $(L+1)\times(L+1)$ rather than $(L+1)^2\times(L+1)^2$. The analysis of spin~1 and spin~2 grid functions into spherical harmonic coefficients is similar, but complicated by the fact that the induced metric on the subspace factorizes as a tensor product only in a complex (mixed parity) basis. Separating the even and odd parity coefficients therefore requires some extra book keeping. Techniques for handling spherical harmonic spectral representations have been described by many authors \cite{Merilees73b,Orszag74,Swarztrauber79,Swarztrauber89,Jakob97}. Our method differs from the Muchenhauer and Daly projection (see \cite{Swarztrauber79}) in the choice of inner product (\ref{f1f2}) used to define the orthogonal subspace. More general spectral transform methods (eg. \cite{Swarztrauber89}) use other choices of weightings and node points to define the projection, and do not have such a simple underlying inner product. All these methods are also $O(L^{3})$. Jakob \cite{Jakob97} gives an $O(L^{2}\log L)$ spectral projection, which however bypasses the construction of the spherical harmonic coefficients. Since we need the spectral coefficients, and because we work with a relatively small value of $L$, the Jakob projection would not provide any improvement. The torus method described here and in \cite{Norton96} has the advantage that it applies also to higher rank tensors, in particular vectors and 2-tensors. Representations in terms of spin-weighted fields are more efficient for vectors (spin $s=1$) than 3-vector representations \cite{Swarztrauber81,Swarztrauber84}, and the operator $\eth$ gives a transparent derivation of all invariant derivative combinations \cite{Swarztrauber81}. \subsection{Convolution splines} At various stages it is necessary to interpolate and differentiate grid-based fields. For example, the radial integration of the hypersurface equations by the 8th order Runge-Kutta method requires values of the source field $\beta$ at $10$ intermediate points; the dynamic regridding of the radial grid requires interpolation to determine the field values of $\beta$ at the new grid points; and derivatives such as $\partial\gamma/\partial r$ and $\partial Q^{+}/\partial z$ must be computed from field values on grids. A convolution spline algorithm described in \cite{Norton92} provides a convenient technique. The method has the effect of fitting a spline curve to sample data, and is implemented by a convolution of the form \cite{Norton92} \begin{equation} \bar{f}(x) = \sum_{k\in\mathbb{Z}}f(k)\phi_{n}(x-k), \label{Cspline} \end{equation} where the $f(k)$ are the raw data (samples) and $\phi_{n}(x)$ is a $C^{n-2}$ sampling kernel. The sampling kernel $\phi_{n}$ is constructed as a certain sum of central B-splines $M_{n}$ of order $n,$ \begin{equation} \phi_{n}(x) = \sum_{i=1}^{n-1} a_{i}^{(n)}M_{n}(x-\tfrac{n}{2}+i), \label{phin} \end{equation} where the coefficients $a_{i}^{(n)}$, $i=1,\ldots,n-1$ are chosen so that the convolution \bref{Cspline} acts as the identity on polynomials $f(x)$ of degree $n-1$ ($n$ even) or $n-2$ ($n$ odd). Recall that the central B-spline $M_{n}(x)$ is a $C^{n-2}$ piecewise polynomial of degree $n-1$ normalised by $\sum_{k \in \mathbb{Z}}M_n(x-k) = 1$, with support on $|x|\le n/2$ \cite{Schoenberg73}. The support of the kernel $\phi_{n}(x)$ is therefore $|x|\le n-1.$ Algorithms for computing the $a_{i}^{(n)}$ and tabulations for $n\le11$ are given in \cite{Norton92}. Coefficients for the kernel $\phi_{9}$ used in the code are given in Table \ref{a9i}, and $\phi_{9}$ is plotted in Figure \ref{fig:phi}. \begin{table}[hb] \centering \caption{Convolution coefficients $a_{i}^{(9)}$.} \begin{tabular}{|c|c|c|c|c|} \hline $i$: & $1,8$ & $2,7$ & $3,6$ & $4,5$ \\ \hline $a_{i}^{(9)}$:\rule[-2mm]{0mm}{6mm} & $-\frac{67}{2520}$ & $\frac{1111}{5040}$ & $-\frac{421}{560}$ & $\frac{1333}{1260}$ \\ \hline \end{tabular} \protect\label{a9i} \end{table} \begin{figure}[ht] \centering \resizebox{8cm}{!}{\includegraphics{kernel.ps}} \caption{Comparison of the $C^{7}$ spline kernel $\phi_{9}$ (solid curve) and the sinc method \cite{Boyd89,Stenger93} kernel ${\rm sinc}(x)= \sin(\pi x)/\pi x$ (dotted curve). The kernel for a sampling method is the response to the delta-like discrete data (solid dots). The advantages of the convolution spline method are that the kernels have finite support ($[-8,8]$ in this case) and the data is automatically filtered. The convolution \bref{Cspline} using kernel $\phi_{9}$ exactly reproduces polynomials of degree $7$ or less. } \label{fig:phi} \end{figure} The expressions \bref{Cspline}, \bref{phin} may be rearranged into a form which is more efficient for numerical calculations, \begin{equation} \bar{f}(x) = \sum_{k\in\mathbb{Z}} \tilde{f}_{k}M_{n}(x-\tfrac{n}{2}-k), \label{tfk} \end{equation} where the modified sample values $\tilde{f}_{k}$ are given by \[ \tilde{f}_{k} = \sum_{i=1}^{n-1} f(k+i)a_{i}^{(n)}. \] The advantage of \bref{tfk} is that the $\tilde{f}_{k}$ can be computed once and then reused to evaluate $\bar{f}(x)$ at many different points $x$, using the explicitly known values of the B-spline $M_{n}(x).$ The same $\tilde{f}_{k}$ values may also be used to compute derivatives of the spline function $\bar{f}$, \begin{equation} \bar{f}^{(j)}(x) = \sum_{k\in\mathbb{Z}} \tilde{f}_{k}M_{n}^{(j)}(x-k-\tfrac{n}{2}), \label{dtfk} \end{equation} where again the derivatives $M_{n}^{(j)}(x)$ are known functions. These techniques are routinely used to supply intermediate values and derivatives of fields in the radial and time directions. Non-uniform distributions of sample points are handled by a mapping between the independent variable and the sample number variable ($x$ in the above). Numerical derivatives are then calculated using the chain rule. For example, the radial grid described in Section \ref{r_grid} is non-uniform, specified by some known relation of the general form $r = r(n)$ (where $n$ is now being used to denote the sample number variable, with radial grid points being at $n=0,\ldots,n_\infty$). The operator $\frac{\partial}{\partial r}$ is then implemented as $\left(\frac{dr}{dn}\right)^{-1}\frac{\partial}{\partial n}$, with a formula for the first factor being known explicitly. Similarly, one can transform to an independent variable $s = s_0 + hx$ which has grid spacing $h$, to examine the behaviour of the approximation \bref{Cspline} as $h \rightarrow 0.$ Let $g(s) := f((s-s_0)/h) = f(x)$, so $g^{(j)}(s) = h^{-j}f^{(j)}(x)$. It can be shown \cite{Norton92} that using the $\phi_9$ kernel, the Taylor series truncation errors for \bref{Cspline} at a grid point $s$ are \begin{equation} |\bar{g}^{(j)}(s) - g^{(j)}(s)| = c_{j} h^{8}|g^{(8+j)}(s)| + O(h^9),\quad j=0,1,2, \label{errj} \end{equation} where $c_{0}=\frac{2021}{134400}$, $c_{1}=\frac{4547}{302400}$, $c_{2}=\frac{4549}{302400}$. This reflects the fact that convolution with $\phi_{9}$ is exact on polynomials of degree $7$. The predicted $h^8$ convergence of the $\phi_9$ spline convolution is clearly evident in Figure \ref{v_approx}. Here the function $v(x) = e^{x}\sin 10x$ has been approximated at varying grid resolutions corresponding to $N=2^p$ grid points over the interval $[-1,1]$. \begin{figure}[ht] \centering \begin{tabular}{ll} \resizebox{0.47\textwidth}{!}{\includegraphics{esin.ps}} & \resizebox{0.47\textwidth}{!}{\includegraphics{esin_errors.ps}} \\ (a) & (b) \end{tabular} \caption{Convergence of the $\phi_9$ spline convolution: (a) samples of $v(x) = e^{x}\sin 10x$ at $N=16$ points over $[-1,1]$, and the corresponding convolution spline; (b) logarithmic plots of the absolute error $|v(x) - \bar{v}(x)|$ for grid resolutions of $N=2^p$ with $p=4,\ldots,7,$ showing a reduction in the error by a factor of $2^8$ on each doubling of the resolution.} \label{v_approx} \end{figure} Convolution splines do not generally preserve sample values, except for samples from polynomials of degree less than or equal to the degree of reproduction. This results in some damping of high frequency components of the data, which we expect helps to suppress numerical noise and algorithmic instabilities. Within the context of spectral methods for PDEs, the direct filtering of Fourier coefficients of a numerical solution is common practice and has been extensively studied (cf. \cite[\S 8.3]{CanutoEt88} and references). On the other hand, explicit use of a digital filter \cite{Hamming77} in conjunction with finite difference methods is comparatively rare. Nevertheless, from an algorithmic point of view, this is the effect of using a convolution spline. The filtering inherent in the method can be examined via the response function \begin{equation} \Phi_{n}(\theta) = \sum_{k\in\mathbb{Z}}\phi_{n}(k) \cos k\theta,\quad 0 \le\theta\le\pi, \label{Phi} \end{equation} which is equal to the factor by which the Fourier mode $e^{{\rm i}\,\theta x}$ is amplified by the approximation \bref{Cspline} at a grid point $x \in \mathbb{Z}.$ The value $\theta = \pi$ corresponds to the Nyquist frequency for the grid. Figure \ref{fig:resp} shows the response function $\Phi_{9}(\theta)$, compared to the well-known Lanczos and raised cosine (artificial viscosity) filters \cite{Hamming77}. The filtering characteristics of convolution splines and their derivatives are described in \cite{Norton92}. \begin{figure}[ht] \centering \resizebox{8cm}{!}{ \rotatebox{0}{ \includegraphics{filter_resp.ps}}} \caption{Filter response functions: (a) raised cosine filter $\sigma(\theta) = \tfrac{1}{2}(1+\cos \theta)$; (b) Lanczos filter $\sigma(\theta)=\theta^{-1}\sin\theta$; (c) sharpened raised cosine \cite{CanutoEt88}; (d) $\Phi_{9}(\theta)$.} \label{fig:resp} \end{figure} The limitations of convolution splines are illustrated in Figure \ref{fig:step}, which shows Gibbs-like effects associated with approximation of step-like data. The figures also give an indication of the number of grid points needed to resolve a sharp transition in field values. \begin{figure}[ht] \centering \resizebox{\textwidth}{!}{ $\begin{array}{cc} \includegraphics{tanh_N16.ps} & \includegraphics{tanh_N64.ps} \\ \includegraphics{tanh_N32.ps} & \includegraphics{tanh_errors.ps} \end{array}$ } \caption{Convolution spline approximation of step-like data. Here the function $u = \tanh(10x)$ is represented by $N$ samples over the interval $[-3,3]$. For $N=16$, the transition from $u\approx -1$ to $1$ takes just one grid interval and a Gibbs-like phenomenon is evident. The convolution splines are constructed using the $\phi_9(x)$ kernel. The logarithmic plot shows the absolute error, $|u(x) - \bar u(x)|$. } \label{fig:step} \end{figure} In order that the convolution spline smoothing should introduce only negligible errors, the grid resolution should be chosen sufficiently fine that typical field variations take place over enough grid points that the expected frequency $\theta$ of the field data lies well within the part of the Nyquist frequency interval where $\Phi_{n}(\theta) \approx 1$. Of course this requires some prior knowledge of the length scale of the field, and cannot be applied where shocks (or arbitrarily rapid variations) occur in the data. In such cases the spherical harmonic representation would become equally unsuitable. The choice of high order convolution splines ($h^8$ rather than say $h^4$) was motivated by the need to reduce storage requirements. Low order spline convolutions have a markedly reduced usable proportion of the Nyquist interval \cite{Norton92}, so to avoid over-smoothing and to achieve comparable accuracy with a low order method would require significantly higher grid point densities. Both storage costs and the cost of the radial integration increase linearly with the number of radial grid points. The $\phi_9$ kernel was chosen also because its accuracy matches that of the RK8 method used for the radial integration. {}To use convolution splines near endpoints of a data set, one can extend the data set, using a suitable mapping between the independent variable and the sample number variable \cite{Norton98}. The mapping is chosen so that when expressed in terms of the sample number variable $n$, $0\le n\le n_{\infty}$, the fields admit expansions in powers of $n^{2}$ near $n=0$ and $(n_{\infty}-n)^{2}$ near $n=n_{\infty}$. The sampling kernel convolution can then be applied to the even extension of the fields through $n=0$ or $n=n_{\infty}$. This technique is particularly important in extracting radiation data near null infinity (scri, $\mathcal{I}^{+}$), where the radial grid is chosen so that $n_{\infty}-n = O(r^{-1/2})$, as described in Section \ref{r_grid}. \section{Solution algorithm} \label{sec:4} The hypersurface equations (\ref{eq:Gll}-\ref{eq:Gmm}) suggest the following process for evolving the metric in the exterior region with interior boundary the cylinder $r=r_{0}$: \begin{enumerate}\label{algorithm} \item Choose boundary data $(H,Q^{-},J,K)$ on the cylinder $r=r_{0}$, consistent with the boundary equations (\ref{eq:Gnn},\ref{eq:Gnm}); \item Assume $\beta$ is given on a null hypersurface $\mathcal{N}_{z}$; \item Solve the $\mathcal{N}_{z}$ hypersurface equations $r\partial_{r}U=F(\beta,U)$, by integrating along the radial curves $(z,\vartheta,\varphi)=const.$ with initial conditions at $r=r_{0}$ determined in step 1; \item reconstruct the metric functions $u,v, \gamma$ from $H,J,K$ and $\beta$ using the converse construction (\ref{u:def}--\ref{v:def}); \item reconstruct $\partial\beta/\partial z$ from $Q$ and the now known values of $\beta,\gamma$ on $\mathcal{N}_{z}$, using \bref{dbdz:eq}; \item use $\partial\beta/\partial z$ from step 5 to evolve $\beta$ to the ``next'' null hypersurface $\mathcal{N}_{z+\Delta z}$ and repeat from step 3. \end{enumerate} In the following we will show how this heuristic algorithm is implemented numerically, using the techniques and data representations of the previous section. \subsection{Geometry and inner boundary conditions} The code models gravitational waves propagating on a black hole spacetime, with metric approximating that of the Schwarzschild solution in the Kruskal-Szekeres coordinates \cite{HawkingEllis73}. Introducing the double-null coordinates \[ z=t-r^{*},\qquad y=t+r^{*},\qquad r^{*}=r+2\textsc{M}\log\left(\frac{r}{2\textsc{M}} - 1\right), \] the Schwarzschild metric becomes \[ ds^{2}_{Schw} = -(1-2\textsc{M}/r)\,dy\,dz + r^{2}d\Omega^{2}\,, \] where $d\Omega^{2}= d\vartheta^{2}+\sin^{2}\vartheta\,d\varphi^{2}$. The coordinate singularities at the past and future horizons $t=\pm\infty$ are removed by defining $\tilde{y}=e^{y/4\textsc{M}}$, $\tilde{z}=e^{-z/4\textsc{M}}$, giving the metric \[ ds^{2}_{Schw} = \frac{32\textsc{M}^{3}}{r}e^{-r/2\textsc{M}}\,d\tilde{y}d\tilde{z} + r^{2}\,d\Omega^{2}\,, \] where $r=r(\tilde{y},\tilde{z})$ is defined implicitly by (see Figure \ref{KS:fig}) \begin{equation} e^{r/2\textsc{M}}\left(\frac{r}{2\textsc{M}} - 1\right) = \tilde{y}\tilde{z} . \label{tilde_yz} \end{equation} Because radial light rays are straight lines at $45^{\circ}$ and $r=0$ is singular, the surfaces $\tilde{z}=0$ and $\tilde{y}=0$ (ie. $r=2\textsc{M}$) form the past and future event horizons, and these are smooth hypersurfaces with bounded curvature. The approximate Minkowski structure of Schwarzschild spacetime is better illustrated by the radial null geodesics in $(r,t)$ coordinates, see Figure \ref{rtschw:fig}. Note however that the $(r,t)$ coordinates are singular along the event horizons $r=2\textsc{M}$. \begin{figure}[ht] \centerline{ \resizebox{7cm}{!}{\includegraphics{KS_coords.ps}} } \caption{Schwarzschild spacetime in Kruskal-Szekeres coordinates. Radial light rays are straight lines at $45^{\circ}$.} \protect\label{KS:fig} \end{figure} \begin{figure}[ht] \centering \begin{tabular}{ll} \resizebox{!}{7cm}{ \includegraphics{Schwarzschild_coords.ps}} & \resizebox{!}{7cm}{\includegraphics{rz_coords.ps}} \\ (a) & (b) \end{tabular} \caption{Exterior region $r>2\textsc{M}$ of Schwarzschild spacetime, with past and future radial null geodesics. (a) In Schwarzschild coordinates $(r,t)$; (b) in retarded Eddington-Finkelstein coordinates $(r,z).$} \label{rtschw:fig} \end{figure} Initial conditions for $\beta$ are imposed on $\{z=0,r\ge 2\textsc{M}\}$ (with $\textsc{M}=1$ usually), by specifying the spherical harmonic coefficient functions $\beta_{lm}(r)$. Since $\beta(z=0)$ is unconstrained, these coefficient functions may be freely chosen, subject only to the size condition \bref{bsize}. For simplicity the inner boundary conditions are set at $r_{0}=2\textsc{M}=2$ to agree with the Schwarzschild past horizon: $H_{0}=2$, $Q_{0}=J_{0}=K_{0}=0$. Since we choose $\beta(0,r)=0$ for $2\le r \le 5$, by causality the solution should agree with the Schwarzschild metric in a neighbourhood of $r=2$ for all time $z\ge0$, producing a ``white hole'' past horizon in the spacetime. This choice of inner boundary condition has the considerable advantage that the boundary equations (\ref{eq:Gnn}),(\ref{eq:Gnm}) are automatically satisfied, and it is not necessary for this class of simulations to separately ensure that the boundary data are numerically compatible with the boundary equations. The resulting spacetimes have the geometry of an isolated black hole with future event horizon at $z=\infty$, $r>2\textsc{M}$ and Schwarzschild-like white hole boundary along $r=2\textsc{M}, 0\le z <\infty$. Adding shear $\beta$ at the initial hypersurface $z=0$ results in spacetimes modelling the interaction of gravitational radiation with a single black hole. Although the fixed past horizon boundary conditions used in the present code allow many interesting issues to be addressed, it would be desirable to implement more general inner boundary conditions. Such conditions specify $H,J,K,Q^{+}$ at an inner surface ($r=1$, for example), subject to the dynamical ($\partial/\partial z$) constraints on the evolution of $J/u$ and $Q^{+}$ determined by the boundary equations \bref{eq:Gnn},\bref{eq:Gnm} \cite{Bartnik97a}. The free data on the inner boundary consist of $u_{0},K_{0}$, where $u_{0}$ represents a certain coordinate gauge freedom, whilst $K_{0}$ describes the gravitational radiation injected into the system through the inner boundary. Various exact solutions with such boundary conditions are described in \cite{Bartnik97b} (Robinson-Trautman \cite{RobinsonTrautman62}, boosted Schwarzschild, twisted Minkowski space), and would provide useful accuracy checks on the numerical methods. However, implementing general inner boundary conditions raises numerical and geometric difficulties --- the boundary data must be ``consistent'' with the RK4 solution evolution in order to maintain optimal accuracy (see \cite{AbarbanelEt96} for an analysis of similar but simpler situations), and constraining the radiation data $K_{0}$ such that the spacetime is still Schwarzschild near the past horizon is a difficult geometric problem. An arbitrary choice of $K_{0}$ (even $K_{0}=0$ if $u_{0}\ne 1$) will inject some additional energy into the spacetime. \subsection{Dynamic radial grid} \label{r_grid} There are two geometric features which the code should model accurately: future null infinity (``scri'' or $\mathcal{I}^{+}$, where $r\to\infty$, $z$ finite), and the future horizon $r\sim 2\textsc{M}$, $z\to\infty$. The field near null infinity $\mathcal{I}^{+}\cap \mathcal{N}_{z}=(r=\infty, z)$ determines the outgoing gravitational waves as seen by a distant observer and is consequently very important for applications to gravitational wave astronomy. Experience with 3+1 codes shows that it is not possible (as yet) to provide boundary conditions on an outer timelike boundary at a finite radius which do not either inject radiation or reflect radiation back into the grid. This deficiency has the effect of severely limiting the overall time duration of most $3+1$ simulations. We avoid all such reflection problems by using a radial grid coordinate $n$, which compactifies $r=\infty$ and leads to accurate modelling of gravitational radiation. The $(z,r)$ coordinates become singular near the future horizon $z\to\infty$ in the Schwarzschild spacetimes (see Figures \ref{KS:fig}, \ref{rtschw:fig}). In our case this picture is not exact, since the spacetime geometry is only approximately Schwarzschild, and thus the future horizon (for example) will not be located exactly at $z=\infty$. However, the NQS parameterisation must still become singular eventually, as the outgoing null hypersurfaces $\mathcal{N}_{z}$ approach the future event horizon. The effect of the nearly singular coordinates is that at late times, the in-falling gravitational components near the event horizon will be compressed into a region of small $r$-variation, and this compression will accelerate in time $z$, whilst retaining field structures from early times. Consequently no $r$-grid which is constant in time is able to accurately represent the in-falling radiation at late times. We have observed that numerical problems with a fixed radial grid arise as early as $z=10$. {}To overcome these problems, a dynamic and variable radial grid is used, based on double null coordinates $(z,\tilde{y})$. The time steps in the evolution direction are regular, with $\Delta z=0.1,0.05,0.025$ being typical. The grid in the radial direction is chosen to satisfy the criteria that it compactify null infinity and concentrate grid points in the region of greatest variation in the seed field $\beta$. Because the field features propagate along the inward and outward null characteristics, which correspond respectively to the curves $\tilde{y}=const.$ (approximately!) and $z=const.$ (exactly), the numerical grid is taken to be rectangular in the $(z,\tilde{y})$ coordinates, with radial grid point positions being determined by an initial distribution of grid points on the surface $z=0$. Introducing the radial grid coordinate $n$ with range $0\le n\le n_{\infty}$ (with typical values of $n_{\infty}$ being 128, 256, 512 and grid points at integer $n$), we specify an initial grid point distribution $r(z=0,n)=f(n)$, where $f$ is some monotone increasing function such that $f(n_{\infty}) = \infty$. The radial grid points on the initial ($z=0$, $\tilde{z}=1$) surface have $\tilde{y}$ ordinates given by \bref{tilde_yz}, \begin{equation} \tilde{y} = (f(n)/2\textsc{M} -1)\exp({f(n)/2\textsc{M}}) = \phi(f(n)/2\textsc{M}), \label{tyn} \end{equation} where $\phi(x) := (x-1)e^{x}$ is monotone and invertible for $x\ge0$. Since $\tilde{y},z,r$ are related by $\tilde{y}=e^{z/4\textsc{M}}\phi(r/2\textsc{M})$ and the grid points are required to inflow along the curves of constant $\tilde{y}$, we can determine the dynamic radial grid point distribution $r=r(z,n)$ in terms of the initial grid distribution function $f(n)$ by \begin{equation} r(z,n) = 2\textsc{M}\, \phi^{-1}\!\left(\exp({-z/4\textsc{M}})\phi(f(n)/2\textsc{M})\right), \label{eq:r(z,n)} \end{equation} where the inverse function $\phi^{-1}:[-1,\infty)\to[0,\infty)$ is evaluated numerically. With this definition, the surfaces $n=const.$ correspond to in-falling null hypersurfaces in the reference Schwarzschild metric. In order to express the hypersurface equations in terms of $n$ rather than $r$, we need to compute $\partial r/\partial n$ --- this expression follows immediately from \bref{eq:r(z,n)}: \begin{equation} \frac{\partial r}{\partial n} = e^{-z/4\textsc{M}} \frac{f(n)}{r(z,n)} \exp\left({\frac{f(n)-r(z,n)}{2\textsc{M}}}\right) \frac{df}{dn}. \label{eq:drdn} \end{equation} It remains to choose the initial grid distribution $f(n)$. The condition $n_{\infty}-n = O(r^{-1/2})$ is achieved by setting $f(n) = f_{1}(\nu)/(1-\nu)^{2}$, where $\nu=n/n_{\infty}$ and $f_{1}:[0,1]\to\mathbb{R}$ is any suitable smooth monotone bounded function. In the code, $f_{1}$ is a quadratic polynomial, with coefficients chosen to concentrate grid points across the support of the chosen initial data $\beta(z=0)$. Figure \ref{fig:rzn} shows sample curves $r(z,n)=const.$, illustrating the in-falling nature of the $(z,n)$ grid coordinates. \begin{figure}[ht] \centerline{ \resizebox{10cm}{!}{\includegraphics{zn_plane.ps}} } \caption{The map between radial grid point number $n$ and radius $r$ is dynamic, chosen so that grid points approximately follow inward null geodesics. At late null-time $z,$ grid points cluster near the black hole horizon at $r=2.$ } \protect\label{fig:rzn} \end{figure} This heuristic prescription for distributing the grid points works well in practice --- Figure \ref{fig:beta} shows the shear over the $(z,n)$ plane for \texttt{run\_160}, and clearly demonstrates the in-falling structure of this solution. The simulation eventually terminates at $z = 55$ because of some geometric effect associated with breakdown of the NQS gauge condition near the future event horizon. \begin{figure}[ht] \centerline{ \resizebox{8cm}{!}{\includegraphics{beta_160.ps}} } \caption{Evolution of $r\beta$ for $0\le z \le 55$, in radial in-falling coordinates. Observe that the in-falling grid tracks the dynamical evolution. This simulation has timestep $\Delta z=0.05$, $n=0$ is the past horizon $r=2\textsc{M}$ and $n=256$ represents future null infinity $\mathcal{I}^{+}$.} \protect\label{fig:beta} \end{figure} \subsection{Hypersurface equations} The hypersurface equations are solved by treating them as a large system of ordinary differential equations, with the radial grid coordinate $n$ playing the role of independent variable, and the dependent variables being the values taken by the fields $(H,J,Q,K)$ at the $N^2/2$ points of the $(\vartheta,\varphi)$-grid. The form (\ref{eq:lh}--\ref{eq:K}) of the hypersurface equations, for the variables $\log H$, $rQ^{+}$, $j$ and $K$, proves to be better behaved near $r=\infty$, since each of these variables has a finite (usually non-zero) limit. Integration of these radial ODEs is possible up to and including the final point $n=n_{\infty}$, with results whose numerical effectiveness may be seen by inspecting the field values in a neighbourhood of null infinity \cite{gular}. Tests described in the following section, in particular the consistency of the constraint equations and the accuracy of the Trautman-Bondi mass decay formula (Figure \ref{fig:massloss}) also confirm that asymptotic behaviour has been reliably calculated. Note that unlike methods based on Bondi-Sachs or Newman-Unti coordinates \cite{Bondi62,NewmanUnti63}, integration along the $r$-coordinate lines does not correspond to integrating along the radial null geodesics (the characteristics of the Einstein equations), since in general the NQS shear $\beta$ is non-zero, and the null direction is $\ell=\partial/\partial r -r^{-1}\beta$. The radial integration with respect to the coordinate $n$ is performed using an 8th order Runge-Kutta scheme \cite{DormandPrince81}, with RK step size $\Delta n = 1$. The RK8 method requires 13 derivative evaluations per RK step, of which 10 are at intermediate points not on the radial grid. Values of the field $\beta$ and its angular derivatives at these intermediate points are provided by convolution splines generated using the kernel $\phi_{9}$ with samples at integer $n$. \subsection{Reconstructing the metric} \label{sec:elliptic} Step 4 of the solution algorithm requires us to reconstruct the metric functions $(u,v,\gamma)$ from the solution $(H,J,K,Q)$ of the hypersurface system (\ref{eq:Gll}--\ref{eq:Gmm}) with seed $\beta$ and boundary data $(H,J,K,Q)|_{r=r_{0}}$. Note that the connection variables (\ref{def:H}--\ref{def:Qpm}) are determined by the values of the metric functions $(\beta,\gamma,u,v)$ and $\partial \beta/\partial z$ on the hypersurface $\mathcal{N}_{z}$. The reconstruction is carried out as described in section \ref{sec2.3}. This process requires solving the system (\ref{Lbeta:def}) on each $S^{2}$ of the radial grid. If $\beta$ is not too large, then (\ref{Lbeta:def}) is an elliptic system of partial differential equations on the sphere $S^{2}$, mapping surjectively to the space of spin~2 fields. We solve (\ref{Lbeta:def}) by first substituting \begin{equation} \gamma = \eth^{-1}\Gamma, \label{def:Gamma} \end{equation} where $\eth^{-1}$ is defined spectrally by \[ \eth^{-1} Y^{2}_{lm} = - \left[\tfrac{1}{2}(l+2)(l-1)\right]^{-1/2} Y^{1}_{lm},\quad l\ge 2, \] so (\ref{Lbeta:def}) becomes \begin{equation}\label{Gammeq} \mathcal{K}_{\beta}\Gamma := \Gamma + \frac{\eth \beta}{2-\div\beta}\div(\eth^{-1}\Gamma) \;=\; -K + J \frac{\eth \beta}{2-\div\beta} \,. \end{equation} Note that the choice \bref{def:Gamma} gauges the $l=1$ spherical harmonic components of $\gamma$ to zero --- a similar but more expensive construction may be used if nonzero $\gamma_{l=1}$ components are desired. The advantage of (\ref{Gammeq}) over (\ref{Lbeta:def}) is that the operator $\mathcal{K}_{\beta}$ in (\ref{Gammeq}) is close to the identity for small $B:= \eth\beta/(2-\div\beta)$. The corresponding discretized problem is therefore well suited to iterative matrix methods. We use the conjugate gradient (CG) method \cite{Johnson87}, which is an iterative method applicable to matrix problems of the form $Ax = b$ with $A$ symmetric positive definite. Accordingly we actually solve an associated self-adjoint equation obtained by applying to (\ref{Gammeq}) the operator adjoint to that in (\ref{Gammeq}) with respect to the $L^2$ norm on $S^2$. In order to compute the adjoint operator $\mathcal{K}_{\beta}^{T}$, notice first that $\mathcal{L}_{\beta}$ and $\mathcal{K}_{\beta}$ are {\em real}-linear but not complex-linear, so the adjoint must be computed with respect to the real form of the inner product (\ref{s2ip}). Expanding $\Gamma=\sum_{l\ge2}\Gamma^{lm}Y^{2}_{lm}$, $\phi=\sum_{l\ge1}\phi^{lm}Y^{1}_{lm}$ we have the spectral representations \begin{eqnarray} \label{edthinv} \eth^{-1}\Gamma & = & -\sum_{l\ge2,m}\left[\tfrac{1}{2}(l-1)(l+2)\right]^{-1/2}\Gamma^{lm}Y^{1}_{lm}, \\ \label{divpspec} \div\phi & = & \sum_{l\ge1,m}\left[\tfrac{1}{2} l(l+1)\right]^{1/2}(\phi^{lm}+\bar{\phi}^{lm})Y_{lm}, \end{eqnarray} and thus the real adjoints $(\eth^{-1})^{T}$, $\div^{T}$ are \begin{eqnarray} \label{edthinvT} \eth^{-1T}\phi & = & -\sum_{l\ge2,m} \left[\tfrac{1}{2}(l+2)(l-1)\right]^{-1/2}\phi^{lm}Y^{2}_{lm}, \\ \label{divT} \div^{T}(f+{\rm i}\, g) & = & 2 \sum_{l\ge1,m}\left[\tfrac{1}{2} l(l+1)\right]^{1/2} f^{lm}Y^{1}_{lm}, \end{eqnarray} where $f,g$ are real-valued functions. Consequently we may represent the adjoint $\mathcal{K}_{\beta}^{T}$ spectrally by \begin{equation} (\mathcal{K}_{\beta}^{T}\phi)^{lm} = \phi^{lm} - \sqrt{\frac{l(l+1)}{(l-1)(l+2)}} (\bar{B}\phi + B\bar{\phi})^{lm}, \label{KT} \end{equation} where $B=\eth\beta/(2-\div\beta)$ and $l\ge2$. The equation (\ref{Lbeta:def}) transformed into (\ref{Gammeq}) gives an invertible equation \begin{equation} \mathcal{A}_{\beta}\Gamma := \mathcal{K}_{\beta}^{T}\mathcal{K}_{\beta}\Gamma = \mathcal{K}_{\beta}^{T}(JB-K). \label{AGamma} \end{equation} The right hand side of \bref{AGamma} may be computed explicitly using the spectral representation and \bref{KT}. The operator $\mathcal{A}_{\beta}$ is symmetric and positive definite and close to the identity when considered in the spectral representation. Thus the conjugate gradient algorithm may be applied to \bref{AGamma} and we see that the parameterisation of \bref{Lbeta:def} in terms of $\Gamma$ \bref{def:Gamma} amounts to a preconditioner. Note CG requires not that the \emph{matrix} of $\mathcal{A}_{\beta}$ be given explicitly, but only that $\mathcal{A}_{\beta}\Phi$ can be evaluated for any spin~2 field $\Phi$. We carry out this evaluation by a sequential process which computes the actions of $\mathcal{K}_{\beta}$ and $\mathcal{K}_{\beta}^{T}$ using the spectral representations of $\eth^{-1},\eth^{-1T}$ and $\div,\div^{T}$ (which are simple diagonal operators) combined with transformations to the point representation to evaluate multiplication terms like $\bar{B}\Phi$ followed by projections back to the spectral representation. This scheme requires several transformations between representations of fields by their spin-weighted spherical harmonic coefficients and by their values on the $S^2$ grid. For example, the operator $\eth^{-1}$ is a trivial multiplicative operator on spectral coefficients, whereas the products in the source terms are best calculated in the grid representation. Although evaluating the action of $\mathcal{A}_{\beta}$ is thus numerically expensive, the expense is more than compensated for by the rapid convergence of the CG algorithm with this spectral preconditioning. The spectral representation has the further advantages that the solution is represented fewer unknowns ($2(L+1)^{2}-8$ compared to $4(L+1)^{2}$ for the grid value representation), and gauge conditions which specify the $l=1$ components of $\gamma$ (eg.~$\gamma_{l=1}=0$) can be directly implemented. It is possible to adapt the algorithm to allow for other NQS gauges (eg.~$\beta_{l=1}=0$), but this is numerically more expensive since (\ref{Lbeta:def}) must be solved 4 times at each sphere rather than once. Although such gauges have some geometric advantages \cite{Bartnik97b}, their numerical implementation has not yet been considered. Using CG to solve for the spin~2 spherical harmonic coefficients of $\Gamma$ turns out to be quite efficient, typically requiring fewer than 10 iterations for an $S^2$ grid of size $N/2\times N = 16\times 32$. On this size grid we resolve all components of $\Gamma$ up to angular momentum $L = N/2 - 1= 15,$ so in this case we are solving for $2((L + 1)^2 - 4) = 504$ spectral coefficients. The scheme's effectiveness is due in part to having a good initial guess for $\Gamma$ to use as the starting point of the CG iterations, namely the solution found for $\Gamma$ on the 2-sphere at the previous radial position. The CG iterations finish when the error, measured by the sum of squares of spherical harmonic coefficients of the difference of the two sides of \bref{Gammeq}, is $10^{-2}$ times the size of the aliasing error in the source term. This aliasing error is the difference between the raw field values of the source term (which is necessarily calculated in the field value representation because it involves products and quotients) and its field values after projection into the subspace spanned by spin~2 spherical harmonics. It provides an estimate of the error in the source term, and hence (because the operator $\mathcal{K}_{\beta}$ is close to the identity) it is reasonable to accept a solution of comparable accuracy. {}To ensure termination of the CG algorithm, other stopping criteria are also checked, but the relative error test is the usual termination cause and is found to work well in practice. For example, it can result in a 10-fold improvement over letting the CG iterations run until the solution is determined to machine precision. \subsection{Evolution} \label{sec:evolution} Given $\beta$ on a null hypersurface $\mathcal{N}_{z}$, we construct the time derivative $\partial\beta/\partial z$ by solving the hypersurface equations with seed $\beta$, determining $\gamma,v$ as outlined in the previous section, and then using formula (\ref{dbdz:eq}) to evaluate $\partial \beta/\partial z$. Let us write the result of this process as \begin{equation} \frac{\partial \beta}{\partial z} = \mathcal{B}(\beta,U_{0}) \label{BbetaU0} \end{equation} where the operator $\mathcal{B}$ is determined by the value of $\beta$ on the hypersurface $\mathcal{N}$ and the initial conditions $U_{0}=(H_{0},Q_{0},J_{0},K_{0})$ at $r=r_{0}$ for the hypersurface equations. The evolution formula (\ref{BbetaU0}) provides the basis of the spacetime evolution algorithm, which simply incorporates (\ref{BbetaU0}) into a standard 4th order Runge-Kutta algorithm. This approach is just the method of lines, treating the evolution equations as a very large system of ordinary differential equations for the (point) representation of the entire field $\beta(z) = \beta_{|\mathcal{N}_{z}}$. The method of lines, applied blindly in this manner, is generally prone to instabilities. Tests suggest the relative stability of the NQS code derives from the smoothing effects (a) of the convolution spline, and (b) of the spectral projection. The filtering implicit in the convolution spline is applied to $\beta$ during the radial integration of the hypersurface equations, at each of the 4 stages of the RK4 algorithm. It is not possible to turn off this radial filtering because the convolution splines for $\beta$ are an essential part of the algorithm for evaluating the right hand side of \bref{BbetaU0}. Smoothing of $\beta$ in the angular directions is done explicitly, by projecting $\beta$ onto the spin~1 subspace with maximum angular momentum $L$ or $2L/3$ (the Orszag $2/3$ rule, to eliminate quadratic aliasing). This angular filtering is done after each of the 4 stages of the RK4 algorithm. Removal of the angular filtering results in very rapid disintegration of the evolution, which then typically lasts only a few RK4 steps. For simplicity, the 4 RK4 stages evolve $\beta$ in the $z$ direction in the $(z,r)$ coordinates, along $r=const$. At the end of each full RK4 time step the key field $\beta$ is interpolated onto the new radial grid \bref{eq:r(z,n)}, using a convolution spline based on values of $\beta$ on the old grid. The RK4 time integration of $\beta$ evolves field values on the $(\vartheta,\varphi)$-grid. Equivalently, we could have evolved its spherical harmonic coefficients, of which there are only half as many. However, the amount of computation saved by doing so is insignificant in comparison to that required {}to evaluate $\partial\beta/\partial z$, so this choice is made for convenience. Likewise, the RK8 radial integration of the system of hypersurface equations uses the field value representation. In this case, however, it is found that projecting the fields onto their appropriate spherical harmonic subspaces during the integration is not required for either stability or accuracy. There is a definite computational advantage in staying within the field value representation, since several relatively expensive $O(L^3)$ projections are avoided. The first radial derivative of $\gamma$ is needed to evaluate $\partial\beta/\partial z$ \bref{dbdz:eq}. Grid values of $\partial\gamma/\partial r$ are calculated numerically as derivatives of convolution splines (in the radial direction) for the $(\vartheta,\varphi)$-grid values of $\gamma$, making use of formula \bref{eq:drdn} and the chain rule for derivatives. The radial derivative term $\mathcal{D}_r\log u$ appears in the hypersurface equations \bref{eq:rQp} and \bref{eq:j}. Using equation \bref{eq:lh} and definition \bref{u:def}, this term can be written as an expression involving only the 1st radial and angular derivatives of $\beta$. The program is normally run until the solution ceases to be well behaved. Blowup is detected by monitoring $2-\div\beta$, which must remain everywhere positive. For the initial data that we have used, the blowup has always occurred in the $l=2$ modes of $\beta$, at low values of $n$ corresponding to $r-2\textsc{M}\approx 0$ (see Figures \ref{fig:rzn},\ref{fig:beta}). Although the precise cause of the blowup is not yet understood, it is not a numerical instability, since it is unaffected by changes in the radial or timestep resolutions, nor does it appear to be primarily geometric, since most curvature scalars remain bounded. This suggests the blowup is a coordinate effect, probably arising from proximity to the future event horizon. For smooth initial data of intermediate strength (\texttt{run\_160}), the evolution extends {}to $z\sim 55$. The final time varies with the strength of the initial data --- see Table \ref{table:datasize}. The evolution of an intermediate strength solution is shown in Figure \ref{fig:beta}, which plots the mean square or $L^{2}(S^{2})$ size of $\beta$ at each radial sphere, for time $0 \le z \le 55$. Termination is caused by the blowup feature at low radius, which grows steadily from time $z=40$ onwards. \section{Accuracy tests} \label{sec:accuracy} The complexity of the NQS Einstein equations and the variety of algorithms employed in the code, make it problematic to prove rigorously that the numerical simulation accurately models the physics and geometry of the spacetime. Instead we rely on a range of tests to justify the reliability of the code, probing the numerical accuracy of the solutions through their convergence and geometric consistency. We consider here tests based on the \emph{numerical convergence} of the solutions as algorithmic parameters are varied; and on the \emph{algebraic consistency} of the numerical solutions. The consistency tests measure the constraint identities and the Trautman-Bondi mass decay formula \cite{Trautman58b,Bondi62}. Work in progress considers other tests, including comparisons with linearised theory, and with better known solutions such as Robinson-Trautman, Schwarzschild and Minkowski spacetimes in twisted NQS coordinates \cite{Bartnik97b}. The resolution of the simulations is determined by three parameters: the spherical harmonic spectral limit $L$ (or effective limit $l_{\rm max}$); the number of radial zones $n_{\infty}$; and the time step $\Delta z$. We shall examine in turn how the accuracy of a solution depends on each of these parameters. It is clear that numerical convergence can be estimated from the convergence properties of the key field $\beta$. However, convergence of $\beta$ guarantees only that the (limit) solution satisfies \emph{some} system of equations, which may not coincide with the desired vacuum Einstein equations. (For example, the Einstein equations may have been incorrectly implemented). Thus, to assert that the correct equations have been solved, it is essential to provide independent tests of the correctness of the code. The most natural independent test is to compare the numerical solution with an explicitly known solution. Unfortunately the Schwarzschild metric \bref{ds2:schw} is trivial in the NQS gauge and does not provide a useful comparison test, whilst the twisted shear-free metrics \cite{Bartnik97b} require boundary conditions which are more general than those available in the present version of the code. Instead we consider here another class of independent tests based on constraint relations. Such relations are typical of geometric equations arising in geometry and physics, which admit gauge and coordinate freedoms. Thus, we check the geometric consistency of the solution by evaluating $r^2G_{nn}$ and $r^2G_{nm}$, using the constraint relations \bref{eq:Gnn} and \bref{eq:Gnm}. Neither of these relations is used in generating the numerical solutions, and in theory these components should evaluate to zero. In practice, since each is a sum of terms having magnitude approximately $|\beta|\sim1$, the extent to which $r^{2}G_{nn}$, $r^{2}G_{nm}$ evaluate to zero serves both to confirm the consistency of the numerical solution with the vacuum Einstein equations, and also to assess the accuracy of the solution. The Trautman-Bondi mass decay formula provides another such test of geometric consistency, and of the accuracy of the solution near $r=\infty$. This theoretical result leads to a relation between the asymptotic ($r=\infty$) values and $z$-derivatives of the fields $H$, $J$, and $K$, and may be readily tested for our numerical solutions. In the following we discuss numerical solutions using three reference initial $\beta(z=0)$ fields, which differ only by the scale factors given in Table \ref{table:datasize}. In each case the initial $\beta$ consists of pure $l=2,m=2$ spherical harmonics with equal strength odd and even parts, and radial profile given by a bump supported on $5\le r\le 40$. We use the terms \emph{weak}, \emph{intermediate} and \emph{strong} to describe solutions generated using the three sizes of initial data. Another convenient measure of the strength of the gravitational field is the initial relative mass difference $m_{B}(0)/\textsc{M} - 1$, between the initial Bondi mass of the numerical spacetime (cf.~\bref{def:mB}) and the background Schwarzschild mass ($\textsc{M}$). Table \ref{table:datasize} gives the initial relative mass differences for the three reference initial $\beta$ fields. \begin{table}[ht] \centering \caption{Size of initial data sets } \vskip 3mm \begin{tabular}{|c|c|c|c|} \hline Field strength: & weak & intermediate & strong \\ \hline $\beta(0)$ scale factor: & 1 & 4.48 & 10 \\ \hline $m_{B}(0)/\textsc{M} - 1$: & $0.9472\times 10^{-2}$ & $ 0.1915 $ & $ 0.9940 $ \\ \hline Last $z$: & 61 & 55 & 51 \\ \hline \end{tabular} \label{table:datasize} \end{table} The qualitative conclusions of the error analysis of this section may be summarised as follows: The major determining factor in the overall accuracy is the spectral limit $L$. Truncating spherical harmonic coefficients beyond $L$ has the effect of modifying the Einstein equations to a system for which the constraint identities are no longer valid, and thereby places a lower bound on the numerical accuracy. For $L=15$ the weak and intermediate field solutions can be adequately resolved, but this is not sufficient to obtain adequate (beyond $10^{-3}$) accuracy for the strong field simulation \texttt{run\_170}. To suppress unstable quadratic aliasing effects, it is essential to use an Orszag 2/3 rule truncation. Within the bounds governed by the spectral limit $L$, accuracy can be improved by increasing the radial resolution $n_{\infty}$. For the weak field solution, $n_{\infty}=1024$ reduces the radial error contribution to the level of the spectral truncation error (see Figure \ref{Gnn_dz}(b)). For given resolutions $L$ and $n_{\infty}$, there is a range of values $\Delta z$ for which the simulation remains stable. Outside this range, the simulation follows the standard solution for some time, then rapidly blows up. The simulation is largely insensitive to the value of $\Delta z$ within the stable range, so $\Delta z$ may be chosen as large as possible, consistent with stable evolution. \subsection{Dependence on spectral limit $L$} Using our current hardware it is not generally feasible to run the code at $L=31$, and $L=7$ is too low to be of interest. The code is normally run at $L=15$ resolution (giving a $16\times 32$ $(\vartheta,\varphi)$ grid) with an anti-aliasing cutoff at $l_{\rm max}=10$. Orszag \cite{CanutoEt88, Orszag71} observed that quadratic aliasing can be eliminated by periodically removing the upper $1/3$ of the spectral bandwidth of a numerical solution. If fields contain only modes for which $l \leq \frac{2}{3} L$, then a quadratic product is band limited to $l \leq \frac{4}{3} L$. With a working bandwidth $L$, the modes for which $L \leq l \leq \frac{4}{3} L$ become aliased onto the modes $\frac{2}{3}L \leq l \leq L$. Therefore, truncation at $l_{\rm max} = \frac{2}{3}L$ will remove quadratic aliasing contamination. If no cutoff is used (ie.~the full $L=15$ resolution is retained) then the high $l$-modes of the intermediate strength simulations blow up at $z \approx 8$. The onset of instability (time until blow up) of the $L=15$ simulations with no $l_{\rm max}$ cutoff is largely independent of the time step and the radial resolution. This suggests that the effect of the nonlinear aliasing contamination is best regarded as changing the system of equations into a system which has unstable solutions. Because the nonlinear interactions in the NQS equations are predominantly quadratic, it is not surprising that the $l_{\rm max}=10$ cutoff is sufficient for long term stability. The intermediate strength solution lasts until $z=55$, when the code terminates for other reasons. Figure \ref{aliasing}(a) shows blow up of \texttt{run\_453}, an $L=15$ simulation of the intermediate field strength solution. The $l=15$ modes show rapid growth beyond $z=6$, indicating the instability of the aliasing feedback. Figure \ref{aliasing}(b) shows the difference between \texttt{run\_453} and the stable simulation \texttt{run\_456}, which has an $l_{\rm max}=10$ cutoff. Until the onset of the high $l$-mode instability (ie.~for $z \leq 6$) there is good agreement between the two simulations, with approximately $10^{-10}$ relative difference for $l=2$ modes and $10^{-2}$ relative difference for $l=10$ modes. \begin{figure}[h!tb] \centering (a) \resizebox{0.4\textwidth}{!}{\includegraphics{aliasing_453.ps}} (b) \resizebox{0.4\textwidth}{!}{\includegraphics{aliasing_453_456.ps}} \caption{Orszag's $2/3$ rule is used to remove aliasing instability: (a) unstable evolution of the high $l$-modes of an $L=15$ simulation (no anti-aliasing cutoff); (b) difference between an unstable $L=15$ simulation (no cutoff) and a stable simulation with an $l_{\rm max}=10$ cutoff. Each $l$-bin contains a radial plot (linear in $n$, with $n = 0,\ldots,n_\infty$) of the square root of the sum of the squares of the $(l,m)$-components for fixed $l$ with $m=-l,\ldots,l$. These simulations have $n_\infty = 512$ and $\Delta z = 0.05$.} \label{aliasing} \end{figure} The spectral limit $L$ is critical in determining the relation between gravitational field strength and simulation accuracy. This can be appreciated by observing the decay rate of the $l$-spectrum of $\beta$, as in Figure \ref{beta_spec}. By extrapolation, the error introduced by the anti-aliasing cutoff at $l_{\rm max}=10$ should be no more than the $l=10$ coefficient, and a relative error estimate follows by comparing the $l=10$ and the $l=2$ coefficients. Figures \ref{beta_spec}(a) and \ref{beta_spec}(b) show the dramatic difference in decay rates of the $l$-modes of $\beta$ for weak and strong fields. Assuming an $l_{\rm max}=10$ cutoff, it is evident that the relative error is at most $10^{-8}$ for weak field simulations, and about $10^{-3}$ for strong field simulations. \begin{figure}[h!tb] \centering (a) \resizebox{0.4\textwidth}{!}{\includegraphics{beta_spec_150.ps}} (b) \resizebox{0.4\textwidth}{!}{\includegraphics{beta_spec_170.ps}} \caption{Spectral resolution and field strength: (a) well resolved weak field with fast $l$-mode decay; (b) poorly resolved strong field with slow decay of $l$-modes (see Table \ref{table:datasize} for field strength details).} \label{beta_spec} \end{figure} From the observed decay rate of the $l$-modes of $\beta$ for a given field strength, it is possible to estimate the resolution $L$ required to achieve a prescribed accuracy. Thus although we cannot directly investigate the behaviour of errors with varying spectral limit $L$ (due to hardware constraints), we can still investigate spectral resolution effects by altering the $\beta$ field strength. Figures \ref{Gxx}(a) and \ref{Gxx}(b) show the effect of field strength (weak, intermediate, strong) on the constraint quantities $r^2G_{nn}$ and $r^2G_{nm}$. The parameters for these simulations are $l_{\rm max}=10$, $n_\infty = 256$ and $\Delta z = 0.05$. The four curves in each band are times $z=10, 20, 30, 40$. There is no significant $z$ dependence of either $G_{nn}$ or $G_{nm}$ until within about $5\textsc{M}$ of the final blow up time. \begin{figure}[h!tb] \centering (a) \resizebox{0.4\textwidth}{!}{\includegraphics{Gnn_150_160_170.ps}} (b) \resizebox{0.4\textwidth}{!}{\includegraphics{Gnm_150_160_170.ps}} \caption{Effect of spectral resolution on constraint quantities (a) $|r^2G_{nn}|_{S^2}$; (b) $|r^2G_{nm}|_{S^2}$, at times $z = 10,20,30,40$, for strong (top 4 curves), intermediate (middle 4 curves) and weak (bottom 4 curves) fields.} \label{Gxx} \end{figure} The second Bianchi identity implies the conservation law $G_{ab}^{;b}=0$, which leads to a radial system of equations for $G_{nn}, G_{nm}$ with sources linear in the hypersurface Einstein tensor components $G_{\ell\ell}$, $G_{\ell m}$, $G_{\ell n}$, $G_{mm}$. Thus $G_{nn}, G_{nm}$ give a measure of the accumulated error in the hypersurface equations in the radial direction. This provides some explanation of the structure of the $G_{nn}, G_{nm}$ graphs, particularly for the strong field solution: the numerical solution of the hypersurface equations will have greatest error in the region where the fields are strongest, in this case the range $64 < n < 128$, and this is precisely the region of greatest increase in $G_{nn}, G_{nm}$. \subsection{Dependence on radial grid resolution $n_\infty$} The radial regridding and interpolation of $\beta$, the radial differentiation of $\gamma$, and the RK integration of the hypersurface equations are all formally 8th order accurate. Figure \ref{beta_n} shows that this is consistent with the observed convergence of $\beta$ on increasing the radial resolution. \begin{figure}[h!tb] \centering \resizebox{0.4\textwidth}{!}{\includegraphics{beta_error_352_452.ps}} \caption{Convergence of $\beta$ with increasing radial resolution: weak field solutions with $n_{\infty}=256,512$ compared to $n_{\infty}=1024$. The error decreases by approximately a factor of $2^{8}$ on doubling the radial resolution.} \label{beta_n} \end{figure} The constraint quantities $G_{nn}$ and $G_{nm}$ also exhibit some convergence effects. Figures \ref{Gxx_n}(a) and \ref{Gxx_n}(b) show show significant improvement between $n_{\infty}=256$ and $512$, but little between $512$ and $1024$. The form of the $n_{\infty}=1024$ curve indicates that errors at the highest radial resolution are dominated by errors associated with the spectral truncation. \begin{figure}[h!tb] \centering (a) \resizebox{0.4\textwidth}{!}{\includegraphics{Gnn_352_452_552.ps}} (b) \resizebox{0.4\textwidth}{!}{\includegraphics{Gnm_352_452_552.ps}} \caption{Effect of radial resolution on weak field constraint quantities (a) $|r^2G_{nn}|_{S^2}$, (b) $|r^2G_{nm}|_{S^2}$. In each case the three curves are $n_\infty = 256,512,1024$ (top, middle and bottom curves respectively).} \label{Gxx_n} \end{figure} \subsection{Dependence on time step $\Delta z$} Although the solution algorithm is formally 4th order accurate in the time direction, at the typical resolutions at which the code is run, the RK4 errors are completely dominated by errors arising from the spectral truncation $L$ and/or the radial discretisation $n_\infty$. This is illustrated by Figure \ref{Gnn_dz}(a), which shows no significant difference in the constraint quantity $G_{nn}$ between $\Delta z = 0.1$ and $\Delta z=0.05$, with $n_{\infty}=256$. However, when the solution is better resolved in the radial direction, a small effect can be observed, cf.~Figure \ref{Gnn_dz}(b), where $n_{\infty}=1024$. Figure \ref{beta_dz} compares $r\beta$ for runs with $\Delta z = 0.1,0.05,0.025$ and $n_{\infty}=512$ and again shows only minor improvements from decreasing $\Delta z$. \begin{figure}[h!tb] \centering (a) \resizebox{0.4\textwidth}{!}{\includegraphics{Gnn_dz_342_352.ps}} (b) \resizebox{0.4\textwidth}{!}{\includegraphics{Gnn_dz_542_552.ps}} \caption{Effect of time step resolution on the constraint quantity $|r^2G_{nn}|_{S^2}$ for the weak field solution: (a) $n_\infty = 256, \Delta z=0.1,0.05$: the error is dominated by the radial discretisation error for $n < 192$ and by the spectral truncation error for $n > 192$. Refining $\Delta z$ produces no appreciable improvement in the solution. (b) $n_\infty = 1024, \Delta z=0.1,0.05$: the radial discretisation error is small enough that the RK4 integration error can be observed. For $n > 700$ the error is dominated by the spectral truncation, resulting in the same tail as in (a), while for $n < 700$ the constraint improves in places, consistent with a factor of 16 decrease in the error.} \label{Gnn_dz} \end{figure} \begin{figure}[h!tb] \centering \resizebox{0.4\textwidth}{!}{\includegraphics{beta_dz_442_452.ps}} \caption{Convergence of $\beta$ with decreasing time step: weak field solutions for $\Delta z=0.1,0.05$, compared against $\Delta z=0.025$. Where the error is not dominated by the radial discretisation error, the curves show a decrease in error which is consistent with 4th order convergence.} \label{beta_dz} \end{figure} Consequently, $\Delta z$ is optimally chosen as large as possible, subject to resulting in stable evolution. For $n_\infty = 256$ and $L=15$ with an anti-aliasing $l_{\rm max}=10$ cutoff, the evolution is stable for $\Delta z= 0.1$ and unstable for $\Delta z= 0.2$, which blows up at time $z=25$, after 125 RK4 steps. \subsection{Energy and asymptotic decay tests} \label{sec:BTmassloss} The Hawking mass \begin{equation} \label{mH} m_{H}(\Sigma) = \sqrt{\frac{\textrm{area}(\Sigma)}{16\pi}} \left(1-\frac{1}{2\pi}\oint_{\Sigma}\rho_{NP}\mu_{NP}\,dv_{\Sigma}\right) \end{equation} of a 2-surface $\Sigma$ reduces in the NQS gauge to \begin{equation} m_H(z,r) = \tfrac{1}{2} r\left( 1 - \frac{1}{8\pi} \oint_{S^2} HJ\right). \label{def:mH} \end{equation} $m_{H}(z,r)$ provides an easily computed quantity representing the ``quasi-local'' mass contained within the sphere $(z,r)$, and has asymptotic limit equal to the Bondi mass \begin{equation} m_B(z) = \lim_{r\to\infty} m_H(r,z). \label{def:mB} \end{equation} The Bondi mass is easily computed numerically, by $m_{B}(z)=m_{H}(n=n_{\infty},z)$. Figure \ref{fig:mH}(a) shows the Hawking mass plotted against the radial coordinate, for times $z=0,1,\ldots,60$. There are several features of interest in this plot: the limit Bondi mass (Figure \ref{fig:mH}(b)) decays in time, reflecting the Trautman-Bondi mass loss formula \bref{eq:dzmB}; the energy is radiated in bursts, reflecting near-linear behaviour dominated by pure $l=2$ modes; the Hawking and Bondi masses decay to background black hole mass $\textsc{M}=1$ at late times, suggesting that in this example, almost all the gravitational radiation has been scattered to $\mathcal{I}^{+}$ and essentially none will be absorbed by the black hole; and finally, the rapidly growing feature about $n=20$ at late times in Figure \ref{fig:beta}, does not affect the Hawking mass. The Trautman-Bondi mass loss formula \cite{Trautman58a,Trautman58b,Bondi60,Bondi62} \begin{equation} \frac{d}{dz}m_{B}(z) = {}- \frac{1}{16\pi} \lim_{r\to\infty}\oint_{S^{2}(z,r)} H|K|^{2}. \label{eq:dzmB} \end{equation} provides another test of the geometric consistency of the solution, particularly near null infinity. By comparing the numerical derivative $dm_{B}/dz$ with the computed value of the right hand side (evaluated at $n=n_{\infty}$), we may construct the error $\frac{d}{dz}m_B-\textrm{RHS(\ref{eq:dzmB})}$. Figure \ref{fig:massloss}(b) plots this error against time $z$, suggesting that the asymptotic ($r=\infty$) fields of \texttt{run\_160} are accurate {}to about $0.00001\%$. \begin{figure}[h!tb] \centering (a) \resizebox{0.4\textwidth}{!}{\includegraphics{hawking_mass_160.ps}} (b) \resizebox{0.4\textwidth}{!}{\includegraphics{bondi_mass_160.ps}} \caption{Mass functions: (a) Hawking mass for times $z=0,1,\ldots,55$; (b) Bondi mass. } \label{fig:mH} \end{figure} \begin{figure}[h!tb] \centering (a) \resizebox{0.4\textwidth}{!}{\includegraphics{dmdz_160.ps}} (b) \resizebox{0.4\textwidth}{!}{\includegraphics{dmdz_error_160.ps}} \caption{The Trautman-Bondi mass loss formula as test of numerical accuracy at $r=\infty$: (a)~Bondi mass decay rate; (b)~error in the mass decay formula, given by ${\rm LHS\bref{eq:dzmB}}-{\rm RHS\bref{eq:dzmB}}$. } \label{fig:massloss} \end{figure}
\section{Introduction} One of the methods called for estimating the age of the Galaxy is based on the analysis of observed abundance of some long-lived radionuclides. The most studied cosmochronometries involve $^{187}{\rm Re}$ or the actinides $^{232}{\rm Th}$, $^{235}{\rm U}$ and $^{238}{\rm U}$. Though some promising results have recently been achieved in decreasing the uncertainties affecting the $^{187}{\rm Re}$ cosmochronometry, the predictions based on the trans-actinides can still be regarded as relatively poor (Arnould \& Takahashi 1999). Nevertheless, the recent observation of r-process elements, including Th, in ultra-metal-poor halo stars, such as CS 22892-052 or HD 115444 (Sneden et al. 1996, 1998) has brought some renewed excitement in the estimate of the age of the Galaxy on grounds of the Th cosmochronometry (Cowan et al. 1997; Pfeiffer et al. 1998). With a metallicity as low as [Fe/H] = --3 and a composition enriched in some pure r-elements, these stars provide strong evidence that the production of heavy elements by the r-process already took place early in the history of the Galaxy. Moreover, the abundance pattern of the 15 r-elements heavier than Ba at the surface of CS 22892-052 (or the 9 elements in HD 115444) shows a striking similarity with the solar system r-abundance distribution, leading to the tempting (though hazardous) conclusion that the r-process mechanism is "unique", i.e any astrophysical event producing r-elements gives rise to a solar-like abundance distribution. This conclusion has been critically analyzed by Goriely \& Arnould (1997) who showed that this assumption may be valid indeed, but is by far {\it not the only possible one}, as the observations in the limited $56\le Z \le 76$ range are equally compatible with an abundance distribution that does not fit the solar one outside the observed domain. This ambiguity is assigned to the fact that the observed CS 22892-052 pattern of abundances reflects primarily nuclear physics properties, and not one or another specificity of a blend of r-process events. This universality assumption is a fundamental prerequisite to build a Th cosmochronometry upon the abundance analysis of metal-poor stars at the present time. In principle, it could be possible to derive the abundance of Th ingested in these metal-poor stars from theoretical extrapolations based on direct fits to the observed abundances. However, in practice, this exercise is affected by uncertainties even greater than when basing the fits on the solar abundances, because of the restricted number of elements observed, the impossibility to distinguish isotopic ratios and the smaller precision in the abundance determination compared with the data available in the solar system. In particular, Goriely \& Arnould (1997) showed that the r-elements distribution at the surface of CS 22892-052 could be reproduced satisfactorily by a random superposition of canonical r-process events. In this case, the theoretical extrapolation to the actinide region based on parametric r-process models is simply meaningless. Nevertheless, future accurate observations of r-elements in ultra-metal-poor star could change this situation. For this reason, we will consider in the present paper the universality assumption to be valid in order to analyze if, despite this difficulty, the recent accurate observation of Th at the surface of ultra-metal-poor stars can indeed provide a reliable estimate of the stellar age by comparing it with the universal r-abundance of Th. Such a procedure requires the estimate of the Th by r-process models, which are known to suffer from very many astrophysics and nuclear physics problems, in spite of much recent theoretical and experimental effort. In this respect, the Th problem is particularly acute, since with U, Th is the only naturally-occuring nuclide beyond $^{209}{\rm Bi}$, so that the estimate of Th production relies on extrapolation procedures based on fits to the solar (or stellar) r-abundance distribution. In Sect. 2, a brief description of the adopted r-process models is given in relation to the Th cosmochronometry. In Sect. 3, the various uncertainties affecting the Th production are studied and their impact on the estimate of the stellar age is analyzed. In Sect. 4, it is shown by comparing the solar fits to the stellar r-element distribution observed that future observations of Pb, Bi or U could put the Th cosmochronometry on safer grounds. \section{Th cosmochronometry and the r-process} Assuming that the whole r-abundance distribution observed in CS 22892-052 and HD115444 is essentially solar, it is straighforward to relate the star age $T_*$ to the Th abundances, \begin{equation} \Bigl({{\rm Th} \over {\rm Eu}} \Bigr)_{obs} = \Bigl({{\rm Th} \over {\rm Eu}} \Bigr)_{r} \exp\bigl[- T_* / \tau({\rm Th})\bigr] \label{eq1} \end{equation} \noindent where $\tau({\rm Th})=20.27~{\rm Gyr}$ is the characteristic $\alpha$-decay timescale of Th and the subscripts $obs$ and $r$ refer to the observed and universal r-process abundance ratios, respectively. As classically done, the Th abundance is here expressed relative to the spectroscopically relevant Eu r-dominant element. The recent accurate observation of Th at the surface of CS 22892-052 amounts to $\log ({\rm Th}/{\rm Eu})_{obs}= -0.70\pm 0.08$ (Sneden et al. 1996; Cowan et al. 1997). Th has also been observed at the surface of HD 115444, but its precise abundance remains to be determined (Pfeiffer et al. 1998). Assuming that a solar-like mix of the r-elements ingested in these halo stars originates from a small number of nucleosynthetic events that took place just before the formation of the stars, the age of the star can be estimated from Eq.~(\ref{eq1}) without calling for a complex model of the chemical evolution of the Galaxy. The only difficulty of the methodology is therefore related to the theoretical estimate of the r-production ratio $({\rm Th}/{\rm Eu})_{r}$. Unfortunately, the r-process remains the most complicated nucleosynthetic process to model from the astrophysics as well as nuclear physics point of view (for a review see Arnould \& Takahashi 1999). On the nuclear physics side, the nuclear structure properties (such as the nuclear masses, deformation, \dots) of thousands of nuclei located between the valley of $\beta$-stability and the neutron drip line have to be known, as well as their interaction properties, i.e the ($n,\gamma$) and ($\gamma, n$) rates, $\alpha$- and $\beta$-decay half-lives and the fission probabilities. Despite much recent experimental effort, those quantities for most of the nuclei involved in the r-process remain unknown, so that they have to be extracted on theoretical grounds and are subject to the associated uncertainties. On top of these nuclear difficulties, the question of the astrophysical conditions under which the r-process can develop is far from being settled. The site(s) of the r-process is (are) not identified yet, all the proposed scenarios facing serious problems. For this reason, only parametric approaches, such as the so-called canonical model (Seeger et al. 1965) can be used to estimate the Th production. We use in the present study the multi-event model\footnote{Note that, in the case of r-processes responsible for elements observed in ultra-metal-poor stars, the denomination "multi-event" does not refer to numerous astrophysical events, such as supernova explosions, but rather to numerous components of a given astrophysical event characterized by different thermodynamic conditions, for example in the different layers of a given supernova.} (Bouquelle et al. 1996; Goriely \& Arnould 1996) in which the best fit to the solar abundances is derived from a superposition of canonical events with the aid of an iterative inversion procedure. Compared with other treatments of the canonical model (e.g Pfeiffer et al. 1998), a major advantage of the multi-event approach is to provide an efficient tool for a systematic study of the various uncertainties affecting the model (Goriely 1999). The iterative inversion method works in such a way that the modification of a given (nuclear or astrophysics) input in the r-process model leads to an automatic renormalization of the thermodynamic conditions necessary to optimize the fit to the solar r-abundance distribution. Therefore, the uncertainties affecting the input data of the parametric model, as well as their impact on the Th production can be studied systematically within the multi-event approach, as shown in the next section. Our standard calculation is performed under the following thermodynamic conditions: $1.3 \le T[10^9{\rm K}] \le 1.7$, $10^{22} \le N_n [{\rm cm}^{-3}] \le10^{29}$ and $10 \le n_{cap} \le 200$ (where $T$ is the temperature, $N_n$ the neutron density and $n_{cap}$ the number of neutrons captured per seed nucleus). Note that the r-process calculations are performed making use of the waiting point approximation, since under the thermodynamic conditions considered here, an almost complete $(n,\gamma)-(\gamma,n)$ equilibrium is established (Goriely \& Arnould 1996). When not available experimentally, the nuclear data are taken from the ETFSI nuclear masses of Aboussir et al. (1995) and from the gross theory (GT2) of $\beta^-$ decay (and $\beta$-delayed neutron emission) of Tachibana et al. (1990). In addition, $\alpha$-decay and fission processes are also considered (before and after the neutron irradiation freeze-out). The fission processes include spontaneous, $\beta$-delayed and neutron-induced fission, the probabilities of which are calculated according to the prescriptions of Kodoma \& Takahashi (1975) with the ETFSI fission barriers (Mamdouh et al. 1998). The procedure used to fit the solar r-abundance distribution is similar to the one described in Bouquelle et al. (1996) though each isotope is now given a weight inversely proportional to the error affecting its solar r-abundance (Goriely, 1999). Since we are mainly concerned with Th cosmochronometry, no details are given for the representative thermodynamic conditions required to fit the solar system r-abundance distribution (such details can be found in Goriely and Arnould, 1996). \section{Uncertainties in the predicted Th abundance} The r-process production of Th is obviously model dependent. However, for cosmochronological purposes, it is of fundamental importance to know to what extent the remaining uncertainties in the r-process modelling can affect the Th synthesis. From Eq.~(\ref{eq1}), it can be seen that the Th abundance has to be determined within less than 16\% if we hope to predict the age of star within less than 3 Gyr. A high accuracy has already been achieved observationally with errors reduced to $\log \epsilon = 0.08$ affecting the stellar age by about 3.7 Gyr. Unfortunately, other uncertainties still need to be solved. These mainly concern the r-process modelling, but before focussing on this subject, it is of interest to stress that the normalization to the Eu abundance is not free from uncertainties. As shown by Goriely (1999), even if the r-process models would be able to reproduce exactly the Eu solar r-abundance, this value is still uncertain by about 20\% (the observed abundance in the ultra-metal-poor stars is known within 35\%) leading to an error in $T_*$ of about 6.6 Gyr. Note that with respect to normalization procedure, it might be safer to use the Ho abundance, since Ho is made of one stable isotope only and the s-contribution to its solar abundance is even smaller than for Eu, i.e the error bars on its solar abundance (about 13\%) are smaller than in the Eu case. The additional uncertainties related to the predicted Eu r-abundance are neglected in the present study by normalizing the calculated Th abundance to the solar r-abundance of Eu. The complete absence of correlation in the production of Th and Eu in the canonical approach of the r-process justifies this choice. The sensitivity of the calculated r-process abundances, and in particular of the Th abundance, to the different crucial inputs used in the multi-event model is now examined and the impact of such uncertainties on the age of the CS 22892-052 star ($T^{CS}_*$) is discussed. \subsection{Sensitivity to astrophysics conditions} Among the different thermodynamic parameters entering the canonical model, the most critical one affecting the Th synthesis is obviously the maximum number of neutrons captured by the initial seed nuclei, $n_{cap}^{max}$, which defines the strong component of the r-process. In analogy with the s-process nucleosynthesis, we can define a main r-process component responsible for the production of all the elements up to the $A=195$ peak and part of the Pb peak. This main component requires a value of $n_{cap}$ up to about 140. Till now, there is no constraint from realistic models on the largest value that $n_{cap}$ can take, i.e in analogy with the s-process, on a strong r-component responsible for the bulk production of Pb and Bi. Considering canonical events with values of $n_{cap}>140$ would lead to the production of the Pb-peak elements, as well as Th, without affecting the synthesis of the lower-mass elements. To illustrate such a sensitivity, multi-event calculations are performed considering canonical events with a maximum number of neutrons captured of $n_{cap}^{max}=$140, 145, 150 and 200 (Fig.~\ref{F1}). An excellent fit is obtained for all isotopes with $A \lsimeq 204$ and seen not to be affected at all by the change in the maximum value of $n_{cap}^{max}$ considered. On the contrary, increasing the $n_{cap}^{max}$ above 140 leads to an increase in the production of the Pb-peak, Th and U elements. The large uncertainties in the Pb-peak r-abundances cannot favour one or another fit. The complete absence of a strong r-component, i.e a maximum value of $n_{cap}=140$, leads to a negative age (when derived from Eq.~\ref{eq1}) of the CS 22892-052 star, and can obviously be rejected. Including a strong r-component with $n_{cap}=145$, 150 and 200 leads to a Th abundance that implies a star age $T^{CS}_*=12.2$, 22.9 and 28.9 Gyr, respectively. Since the fit is constrained by the upper value of the Pb and Bi abundances, increasing $n_{cap}^{max}$ above 200 does not affect the upper value of the Th abundance. More precisely, no event with a value of $n_{cap} > 170$ contributes to the fit to the solar system abundances with our adopted nuclear inputs. In summary, any age below about 29 Gyr can thus be obtained just by adjusting the strength of such a strong r-process component, unless the s- or r-origin of the Pb and Bi can be determined with a greater accuracy. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[scale=.7]{H1447F1.eps}} \caption{ Comparison between the solar system r-abundances (Goriely 1999) and the distribution predicted by our standard multi-event superposition of events characterized by maximum values of $n_{cap}^{max}=$140 (dashed), 145 (dot-dash), 150 (dotted) and 200 (full curve). The square corresponds to the Th abundance observed in CS 22892-052 (Sneden et al. 1996). The Bi and Th abundance are connected by a straight line to visualize the extrapolation predicted by the respective models. The vertical lines correspond to error bars in the solar system abundances taken from Goriely (1999).} \label{F1} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[scale=.7]{H1447F2.eps}} \caption{ Same as Fig.~\ref{F1} where the predictions are obtained with the ETFSI (full), ETFSI-Q (dashed), FRDM (dot-dash) and Duflo \& Zuker (dotted) mass models. } \label{F2} \end{figure*} \subsection{Sensitivity to nuclear physics input} The most fundamental nuclear input to r-process models is well known to be the nuclear masses. Various mass models are available, but for practical reasons we only consider here in addition to the ETFSI model, the ETFSI-Q model (Pearson et al. 1996) which takes into account the strong shell-quenching found in some microscopic calculations on highly neutron-rich nuclei, the popular FRDM model of M\"oller et al. (1995) and the recently-developed model of Duflo \& Zuker (1995), hereafter DZ, based on a very different approach than the previously cited models and which has proven its remarkable ability to predict experimentally known masses. Many studies have compared the quality of these models and their differences in the prediction of masses far away from the valley of $\beta$-stability. Their impact on the r-process nucleosynthesis has also been analyzed in various papers (e.g Goriely \& Arnould 1992), so that we will restrict ourselves to analyze their respective predictions of the Th abundance. Multi-event calculations are now performed making use of the 4 above-cited mass models. The resulting fits are shown in Fig.~\ref{F2}. The fits to the stable nuclei are of the same quality, and in particular in the Pb region, no major differences in the predicted r-abundances can be observed. In particular, it should be emphasized that no major deficiency in the fit is obtained in the pre-peak regions at $A\simeq 120$ and $A\simeq 180$ whatever mass model is used. Given the absence of realistic r-process models, there is obviously no reason to favour one or another mass formula on grounds of parametric fits to the solar r-abundance distribution, especially when dealing with the Th abundance predictions which exclusively depend on the r-process paths in the $A \ge 232$ region. The extrapolation to the Th abundance appears to be highly affected by the mass model used. The estimate of the star age amounts to $T_*^{CS}=28.9$, 23.8, 42.2 and 8.7 Gyr for the ETFSI, ETFSI-Q, FRDM and DZ models, respectively. Such differences are not surprising, since it is well known that the r-process paths for these mass models are significantly different, in particular in the $Z>82$ region where the strength of the shell correction energy around the $N=184$ shell closure can be very different. This is not the case for the ETFSI and ETFSI-Q models, because the shell quenching introduced in the ETFSI-Q model in the vicinity of the $N=184$ shell closure is small. Therefore, the abundance predictions in the Pb and actinide regions, and consequently the stellar age predictions, are globally similar when making use of ETFSI or ETFSI-Q. Compared with the other models, the DZ formula is characterized by a steep slope of the mass parabola and a weak shell effect around $N=184$, so that the progenitors responsible for the final Pb abundance are found in a lower mass region by-passing partially Th and U. The Th abundance obtained with FRDM model is higher than with the ETFSI mass models, because of a more widely spread shell effect in the vicinity of the $N\le 184$ shell closure affecting the r-process path down to $N=170$. The abundance peak around $N=184$ before freeze-out is consequently flattened to lower masses than in the ETFSI case and is less affected by fission processes after freeze-out. A higher Th abundance predicted with the FRDM model leads to a higher age estimate. The high sensitivity of the predicted Th abundance to the mass model will not be resolved before improving our mass predictions in the heavy ($Z>82$) neutron-rich region. Another important ingredient in the Th nucleosynthesis concerns the fission processes, i.e the spontaneous, neutron-induced and $\beta$-delayed fission. Most of the r-pro\-cess calculations do not include the fission processes at all or only partially. However, when dealing with Th cosmochronometry, fission processes must be included in the most careful way in order to describe the competing processes responsible for the final Th abundance (namely $\alpha$-decays, $\beta$-decays and fissions) correctly. The recent large scale calculation of ETFSI fission barriers (Mamdouh et al. 1998) up to $A=295$ is used to test the significance of fission on the Th cosmochronometry. The spontaneous, neutron-induced and $\beta$-delayed fission probabilities are determined in the same way as in Kodoma \& Takahashi (1975). It should, however, be stressed that when not available experimentally, the spontaneous fission rates are derived from a new regression fit to experimental data based on our fission barrier predictions. Figure~\ref{F3} shows the nuclear regions where the different fission modes influence the r-process flows. Because of the strong ETFSI shell effect on the fission barriers around $N=184$, no fission recycling is found during the neutron irradiation, at least before crossing the $N=184$ closure. Only r-process paths characterized by an astrophysical parameter $S_a\gsimeq 2~{\rm MeV}$ (for more details about the astrophysical parameter, see Goriely \& Arnould 1992) are stopped by neutron-induced fission. Spontaneous fission can also affect such r-process paths before neutron freeze-out. $\beta$-delayed fission is found to be of small importance compared with the other decaying modes, even after freeze-out. On the contrary, the spontaneous fission is found to be faster than the $\beta$-decay for almost all isobaric chains above $A=250$ and crucial in estimating the final r-abundances in the Pb and actinide region. In the specific fits studied in the present paper, the fission fragments do not affect the low-mass abundance distribution. Obviously, all the above conclusions should be taken with care, because of the uncertainties remaining in the determination of the fission barriers and fission probabilities, which are to be studied in a forthcoming paper. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[scale=.65]{H1447F3.eps}} \caption{ Representation in the nuclear chart of the dominating fission modes affecting the r-process flow, as given by the legend in the figure. Three r-process paths at $S_a=1$, 2 and 3 MeV (full line) are drawn for illustrative purposes, as well as the neutron drip line (dashed line).} \label{F3} \end{figure*} In order to quantify the impact of the fission processes on the Th cosmochronometry, a multi-event calculation is reiterated switching off all the fission processes. This numerical test just aims at illustrating the largest error possibly made when neglecting fission, but should not be regarded as a sensible test case for the Th prediction. It is found that the neglect of fission gives rise to an increase of the age of CS 22892-052 by 12.3 Gyr on grounds of the abundance distribution shown in Fig.~\ref{F4}. It can also be seen that when including fission processes, the fission fragments do not modify the global abundance distribution. Obviously, a complete and consistent treatment of the fission processes (especially spontaneous fission) is required to build a reliable cosmochronometry on the actinides. Figure~\ref{F4} also presents uncertainties associated with $\beta$-decays and $\beta$-delayed neutron emission by comparing the solar fits obtained with the GT2 model and the QRPA model of M\"oller et al. (1997). Both models have been extensively used in previous works dedicated to the r-process nucleosynthesis, so that it is of interest to study their influence on the Th cosmochronometry. If use is made of the QRPA model instead of the GT2 model, a reduction from 28.9 Gyr down to 15.1 Gyr is obtained for the age of CS 22892-052.Once again, it should be added that although the fit to the solar distribution obtained with the QRPA model is slightly worse than the one obtained with the GT2 approach, it cannot be rejected {\it a priori}, since other nuclear or astrophysics shortcomings of the model can be responsible for the observed discrepancies (for example in the $A=180$ region). As stated previously, given our poor understanding of the r-process nucleosynthesis (especially of the astrophysical site) the quality of nuclear models should not be tested on astrophysics arguments like fits to the solar abundance distribution. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[scale=.7]{H1447F4.eps}} \caption{ Same as Fig.~\ref{F1} where the predictions are obtained using the GT2 $\beta$-decay and $\beta$-delayed neutron emission rates with (full) or without (dash-dot) fission processes. The dotted curve corresponds to the use of the QRPA $\beta$-decay and $\beta$-delayed neutron emission rates including fission.} \label{F4} \end{figure*} \subsection{Uncertainties in the solar r-abundance} The uncertainties still affecting the s-process model are responsible for non-negligible imprecisions in the determination of the residual r-abun\-dan\-ces of the solar system content (Goriely 1999), in particular for the so-called s-dominant isotopes and Pb-Bi isotopes (see the large error bars in Fig.~\ref{F1}). The principal source of uncertainty in the solar r-abundance of Pb and Bi lies in our ignorance of the relative s- and r-contribution to their production. Both processes can produce Pb-Bi almost entirely, so that the r-contribution cannot be estimated in a reliable way on grounds of s-process calculations. This problematic aspect of the solar abundance splitting in the Pb region can only be resolved through realistic modelling of the s-process (or accurate abundance determination of Pb at the surface of ultra-metal-poor stars provided the assumption of the r-process universality be confirmed). It should be kept in mind that the prediction of the Th and Pb-Bi abundances are strongly correlated, so that any uncertainty in the solar r-abundances of the Pb and Bi elements is translated into the exponentially dependent uncertainty in the star age. Among the $A>200$ r-isotopes, $^{204}{\rm Hg}$, $^{206}{\rm Pb}$ and $^{209}{\rm Bi}$ play an important role since their r-abundances are better determined than for their neighbours. The solar r-abundance of $^{204}{\rm Hg}$ is indeed well determined, so that it seems logical to constrain the fit in such a way as to reproduce the $^{204}{\rm Hg}$ solar r-abundance. As regards $^{206}{\rm Pb}$ and $^{209}{\rm Bi}$, they are characterized by a relatively well-determined abundance to which the Th production is directly correlated. However, reproducing the recommended solar abundance of $^{204}{\rm Hg}$, $^{206}{\rm Pb}$ and $^{209}{\rm Bi}$ simultaneously appears to be impossible without strongly deteriorating the fit to the $A=195$ peak. For this reason, we reiterate multi-event calculations in which the fitting procedure is constrained (with an extra statistical weight) to the solar abundance of $A=204$ and 206 in one case and $A=204$ and 209 in the other case, using two different mass models, namely the ETFSI and DZ models (Fig.~\ref{F5}). Although a negative age is obtained with the DZ masses when constraining the fit to $^{209}{\rm Bi}$, the other predicted ages are 30.4 Gyr for the ETFSI calculation constrained to $^{206}{\rm Pb}$ and 10.5 Gyr in the two remaining cases. So, in addition to the large sensitivity of the stellar age to the mass models as studied in the previous section, the uncertainties in the solar r-abundances appear to affect the age determination by about 20 Gyr. As long as the s- or r-origin of the Pb and Bi solar abundance is not determined with high accuracy, the Th cosmochronometry will not provide any reliable age estimate. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[scale=.7]{H1447F5.eps}} \caption{ Same as Fig.~\ref{F1} when the fit is constrained on $^{204}{\rm Hg}$ and $^{206}{\rm Pb}$ solar abundances with the ETFSI (full) or DZ (dashed) mass models. The dash-dot and dotted lines correspond to the constrain on $^{204}{\rm Hg}$ and $^{209}{\rm Bi}$ abundances with ETFSI and DZ mass models, respectively. } \label{F5} \end{figure*} \section{The age of the stars} So far, we have been evaluating the stellar ages on the basis of a fit to the solar r-abundance distribution. Such a procedure is obvioulsy based on our initial fundamental assumption that the r-process site is unique in the Galaxy. This basic assumption is a fundamental prerequisite to build a Th cosmochronometry upon the abundance analysis of metal-poor stars at the present time, since, as explained above, a direct fit to the abundance distribution observed at the surface of metal-poor stars would present even larger uncertainties, because of the restricted number of elements observed, the impossibility to distinguish isotopic ratios and the much smaller accuracy in the abundance data as compared with our solar system. However, in the specific case of Pb, the uncertainties still affecting the solar abundance might be greater than the ones found in the observation at the surface of HD 115444 or HD 126238 (Sneden et al. 1998), so that the Pb abundance in such stars could probably be more constraining on the Th predictions than the solar value. Unfortunately, the Th abundance has not been determined yet in such stars, so that in this case, we limit the discussion on the relative stellar age. Compared with CS 22892-052, only a small number of r-elements are observed in HD 115444, but it is at the moment the only metal-poor star in which Pb and Th lines are detected simultaneously. All the physically sound (i.e leading to a positive age $T^{CS}_*$) calculations presented in the previous section are compared, in Fig.~\ref{F6}, with the elemental abundance distribution observed in HD 115444 and CS 22892-052. We only retain calculations which provide a good fit to the observed $Z \ge 55$ abundances, and in particular to the Pb abundance in the case of HD 115444. In order to achieve a good fit to all elements observed and to optimize the extrapolation to the Th region, a normalization of the abundance curves is done on the heaviest elements accurately observed, i.e Pt and Os for HD 115444 and CS 22892-052, respectively. In this case, the age of CS 22892-052 is found to lie in the $7 \le T^{CS}_* {\rm[Gyr]} \le 39$ range, and a similar error range of about 32 Gyr is predicted for the age of HD 115444. The different predictions of the stellar age, as well as the elemental abundances of Eu, Pb, Bi, Th and U are summarized in Table~\ref{tab1}. The stellar age can be determined in two different ways. If normalized to the calculated Eu abundance, the age $T_*^1$ has the advantage of beeing free from uncertainties in the normalization procedures on the observed abundances, but is sensitive to the theoretical uncertainties made in the r-process predictions. In particular, it is well known that the origin of the r-nuclides in the $A\simeq160$ region is not easily explainable (e.g Meyer \& Mullenax 1998). The impact of such errors remains to be estimated. On the contrary, normalizing the calculated Th abundance on the observed Eu avoids theoretical complications inherent in the origin of the $A\simeq160$ nuclides, but gives rise to additional errors of the order of $\pm 0.1$ dex on the abundances (as seen in Fig.~\ref{F6}), i.e of $\pm 4.7$ Gyr on the age $T_*^2$. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[angle=-90]{H1447F6.eps}} \caption{Elemental abundance distributions observed in HD 115444 and CS 22892-052 (open squares) compared with multi-event r-abundance predictions (see text).} \label{F6} \end{figure*} \begin{table*} \centering \caption{Ages $T_*$ (in Gyr) of CS 22892-052, elemental abundances (in log$\epsilon$) for Eu, Pb, Bi, Th, U and log(Th/U) predicted by the different theoretical calculations shown in Fig.~\ref{F6} and analyzed in Sect.~3. $T_*^1$ ($T_*^2$) corresponds to the age calculated with a Th abundance normalized to the calculated (observed) Eu abundance. } \begin{tabular}{|l|cccccccc|} \hline Comment (see Sect.~3)& Eu & Pb & Bi & Th & U & Th/U & $T_*^1$ & $T_*^2$ \\ \hline Standard ($n_{cap}^{max} = 200$, ETFSI, GT2) & -0.79 & -0.20 & -0.76 & -1.00 & -1.16 & 0.16 & 22.9 & 25.5 \\ $n_{cap}^{max} = 145$ & -0.82 & -0.43 & -1.03 & -1.38 & -1.49 & 0.11 &6.6 & 8.2 \\ $n_{cap}^{max} = 150$ & -0.80 & -0.27 & -0.85 & -1.13 & -1.23 & 0.10 &17.3 & 19.5 \\ ETFSI-Q masses & -0.79 & -0.26 & -0.86 & -1.10 & -1.32 & 0.22 &18.2 & 21.0 \\ FRDM masses & -0.91 & -0.14 & -0.64 & -0.72 & -0.67 & -0.05 & 41.6 & 38.9 \\ DZ masses & -0.75 & -0.19 & -0.85 & -1.40 & -1.44 & 0.04 & 2.4 & 6.9 \\ ETFSI + constraint on $A=206$ & -0.86 & -0.22 & -0.81 & -1.04 & -1.18 & 0.14 & 24.3 & 23.8 \\ DZ + constraint on $A=206$ & -0.75 & -0.20 & -0.85 & -1.36 & -1.38 & 0.02 & 4.3 & 8.9 \\ ETFSI + constraint on$A=209$ & -0.74 & -0.48 & -1.11 & -1.34 & -1.36 & 0.02 & 4.7 & 9.6 \\ \hline \end{tabular} \label{tab1} \end{table*} Although the present study suggests that at the moment great care should be taken in estimating the age of the stars on the basis of observed Th abundance, new accurate observations of heavy r-elements could put the Th cosmochronometry on safer grounds, especially if Th and U lines could be observed accurately and simultaneously in metal-poor stars, as already stressed by Arnould \& Takahashi (1999). As a matter of fact, if the Th and U lines are available, an age estimate could be derived from the expression \begin{equation} \log \Bigl({{\rm Th} \over {\rm U}} \Bigr)_{obs} = \log \Bigl({{\rm Th} \over {\rm U}} \Bigr)_{r} +\log {\rm e} ~ \Bigl( {1 \over \tau({\rm U})} - {1\over \tau({\rm Th})}\Bigr)~T_* \label{eq2} \end{equation} \noindent where $\tau({\rm U})=6.41~{\rm Gyr}$ is the characteristic $\alpha$-decay time\-scale of U. As seen in Fig.~\ref{F6} and Table 1, $\log({\rm Th/U})_r$ is found to lie within an 0.1 range, whatever theoretical inputs are used in the r-process model. Such an accurate estimate is principally bound to the fact that Th and U are neighbour nuclei, and consequently their production ratio is not strongly affected by unreliable extrapolation procedures, but rather by local nuclear uncertainties, such as nuclear masses or fission processes in the actinide region. From Eq.~(\ref{eq2}), it is found that a $\pm 0.1$ error on the observed or predicted production ratio of Th/U gives rise to a $\pm 2.1$ Gyr error on $T_*$. A future simultaneous observation of Th and U lines in ultra-metal-poor stars could therefore open the way to an accurate age determination in contrast to the still complicated Th cosmochronometry based on Th lines only. Equation~(\ref{eq2}) is shown in a graphical form in Fig.~\ref{F7} for 3 possible values of the Th/U production ratio. Note, however, that a deeper analysis of the Th/U production ratio would be required before rushing into an age determination. In particular, uncertainties in the fission processes have not been included in the present study, but could possibly affect the predicted Th/U ratio. \begin{figure} \resizebox{\hsize}{!}{\includegraphics[scale=.32]{H1447F7.eps}} \caption{ Relation between the observed Th/U ratio and the age of the star $T_*$ for 3 different estimates of the production ratio.} \label{F7} \end{figure} \section{Conclusions} The present paper analyzes critically the impact of the remaining uncertainties in the existing r-process models on the determination of the age of the Galaxy derived on grounds of the Th cosmochrometry. Although the direct and accurate observation of the Th abundance in ultra-metal-poor stars provides a possible method to date the age of the star without calling for complex model of the chemical evolution of the Galaxy, the Th cosmochronometry remains affected by all the difficulties associated with our poor understanding of the r-process nucleosynthesis. Even if we disregard the difficulty associated with the fundamental assumption relative to the unicity of the r-process site in the Galaxy, it is impossible at the present stage reliably to determine the age of a ultra-metal-poor star from parametric models of the r-process. The major difficulties lie in the very s- or r-origin of the Pb and Bi isotopes, in addition to the unavoidable problem related to the prediction of nuclear data (principally masses) and the still unknown thermodynamic conditions in which the r-process takes place (in particular concerning the maximum neutron irradiations that can be reached). New accurate observations including Pb and Th lines in ultra-metal poor stars could shed light on the ability of the r-process to produce the Pb-peak elements, and consequently constrain the Th cosmochronometry. In particular, the Th cosmochronometry could be put on safer grounds, especially if Th and U lines could be observed accurately and simultaneously in ultra-metal-poor stars. As regards nuclear uncertainties, only theoretical as well as experimental effort to come can help us achieve the required accuracy to pretend to derive the age of the Galaxy on the basis of the actinides cosmochronometry. \begin{acknowledgements} The authors are grateful to M. Arnould for helpful discussions \end{acknowledgements}
\section{\bf Introduction} Several experiments probing the effect of impurities or lattice defects on superconductivity in the cuprates have been carried out \cite{I1,I2,I3,I4,I5,I6,I7,I8,I9,I10,I11,I12,I13,I14,I15,I16,I17,I18,I19,I20,I21,I21a} in order to get more insight into the symmetry of the superconducting state. The most thoroughly studied defects are Zn and Ni substitutions \cite{I1,I2,I3,I4,I5,I6,I7,I8,I9,I10,I11,I12,I13,I14,I15,I16,I17,I18} on the planar Cu sites and irradiation induced oxygen vacancies \cite{I19,I20,I21,I21a} in the copper oxygen planes. Yet the results are difficult to explain within a standard Abrikosov-Gorkov type theory of impurity scattering \cite{II1} for the scenario of the $d$-wave superconductivity, which is predicted to be extremely suppressed by the impurities. \cite{I22,I22a,I22b,I22c} This issue was critically examined by Radtke et al. \cite{I22} who considered isotropic nonmagnetic impurity scattering in the Born approximation and obtained the critical temperature in both weak- and strong-coupling approach close to the Abrikosov-Gorkov scaling function. A comparison to the electron irradiation data \cite{I19} in $YBa_2Cu_3O_{7-\delta}$ ($Y\!-\!123$) showed that the theoretically predicted $T_c$ was about twice as much reduced by the impurities than observed. This inconsistency can be settled down within the weak impurity scattering model if the impurity scattering rate a factor of 3 less than the one deduced from the transport measurements is assumed \cite{I20,I21} which is equivalent to an introduction of two separate relaxation time scales - one defining the scattering time in pair-breaking processes and the other representing the transport scattering time. Such a distinction occurs naturally as a consequence of an impurity momentum-dependent scattering probability. \cite{II7} The issue of a possible anisotropy in the impurity scattering potential was suggested in a discussion of the irradiation data by Giapintzakis et al. \cite{I19} where the authors evoked a model by Millis et al. \cite{II4} A more general formulation of the problem was given in our previous paper, \cite{I24} where the effective correlation between two impurity vertex functions was assumed in the form $|w_0|^2+|w_1|^2f\left({\bf k}\right)f\left({\bf k'}\right)$, with $|w_0|$ and $|w_1|$ representing isotropic and anisotropic impurity scattering amplitude respectively and $f\left({\bf k}\right)$ determining the symmetry of the impurity potential. \cite{I24a} Our analysis showed that the symmetry of the anisotropic potential is an important factor and a significant reduction in the pair-breaking strength appears for large values of $\left<ef\right>^2= \left[\int_{FS}dS_{k}n\left({\bf k}\right)e\left({\bf k}\right) f\left({\bf k}\right)\right]^2$, where $\int_{FS}dS_{k}$ denotes integration over the Fermi surface (FS) and $n\left({\bf k}\right)$ is the normalized ($\int_{FS}dS_{k}n\left({\bf k}\right)=1$) angle-resolved FS density of states. Particularly low $T_c$ suppression is predicted for $\left<ef\right>^2=1$ that is for the anisotropy of the scattering potential in phase with the order parameter, which in the case of the $d_{x^2-y^2}$-wave superconductor corresponds to $f\left({\bf k}\right)\sim k_x^2-k_y^2$. Within this approach we were also able to understand quantitatively the irradiation data \cite{I19} in $Y-123$. Through the postulation of the analytic form of the square value of the impurity potential this approach has been designed for the second-order Born approximation and cannot be efficiently extended to include multiple impurity scattering processes. Such a generalization is important from the theoretical and experimental points of view, especially in the light of the recent experiments suggesting a possible strong (close to unitary) scattering by Zn atoms in $Y\!-\!123$ and $La_{2-x}Sr_xCuO_4$ ($La\!-\!214$) compounds. \cite{I2} For that purpose a model based on the assumption of the impurity potential and not its square value is needed. In this paper we study in the t-matrix approximation the pair-breaking effect of the nonmagnetic impurities with the anisotropic momentum-dependent factorizable potential. The influence on the superconducting transition temperature is analyzed quantitatively in the Born and unitary scattering limits. In particular, we find that the scattering from impurity potential in phase with the order parameter leads to a stronger suppression of the critical temperature than from the impurity potential orthogonal to the superconducting state. On the other hand, the superconducting state appears very robust to the potential scattering with its maxima in the region of the nodes of the order parameter. Finally, we find that the Zn and electron irradiation $T_c$ suppression data in $Y\!-\!123$ and $La\!-\!214$ are in the range predicted by our model. This work is organized as follows. In Sec. {\bf II}, we introduce the anisotropic momentum-dependent impurity potential. In Sec. {\bf III}, we derive the expressions necessary for the analysis of the scattering process of arbitrary strength. In Sec. {\bf IV}, we obtain the self-energies due to impurity scattering at the superconducting-normal state phase transition which allows the evaluation of the critical temperature. Then we examine analytically Born and unitary scattering limits. In Sec. {\bf V}, we calculate numerically the critical temperature for $d$- and $(d+s)$-wave superconducting states in the presence of impurities accounting for their anisotropic potential of $p$-, $d$-, $f$-, and $g$-wave symmetry in Born and unitary limit. In Sec. {\bf VI}, we compare the results with data on Zn substituted and irradiated overdoped $La\!-\!214$ and optimally doped $Y\!-\!123$ samples. In Sec. {\bf VII}, we present our conclusions. Except for comparing to the experimental data we assume $\hbar=k_{B}=1$ in the calculations. \section{\bf Impurity scattering potential} As the impurity scattering strength is rather impossible to be determined from first principles, the information about the scattering process is usually deduced from a comparison of the experimental data with the theoretical models. These models assume a certain phenomenological impurity potential \cite{II1,I22,II4,I24,I25,II2,II3,I26,I28} which is verified by a fit to the available data. We proceed in the same way by analyzing the impurity scattering potential of the factorizable form \begin{equation} \label{eI1} v\left({\bf k},{\bf k'}\right)=v_i+ v_af\left({\bf k}\right)f\left({\bf k'}\right) \end{equation} \noindent The above interaction consists of two channels - the isotropic scattering channel with the scattering amplitude $v_i$, and the anisotropic one determined by the scattering strength $v_a$ and momentum-dependent function $f\left({\bf k}\right)$. We assume that the $f\left({\bf k}\right)$ average value over the Fermi surface $\left<f\right>\!=\!\int_{FS}dS_k n\left({\bf k}\right)f\left({\bf k}\right) \!=\!0$. Therefore, $f\left({\bf k}\right)$ is orthogonal to the isotropic s-wave term in a sense of a scalar product defined as the FS integral and in consequence a symmetry other than the identity is introduced into the impurity potential. An additional normalization $\left<f^2\right>\!=\!1$ gives $v_a$ the meaning of the scattering strength magnitude in the anisotropic channel. The chosen potential depends on the absolute orientations in the crystal of the incoming and outgoing (scattered) particles momenta, not only on the angle between ${\bf k}$ and ${\bf k'}$. The usually assumed isotropic conditions \cite{II5} of a spherical (cylindrical in two dimensions) constant energy surface and the scattering probability dependent only on the angle of deflexion are broken here. \cite{II5a} This leads to a momentum-dependence of the time between scattering events determined by the imaginary part of the self-energy and a momentum-dependent relaxation time in the Boltzmann equation. \cite{II6,II7} This second quantity is worth mentioning as it provides the information about the normal state transport properties and can be used as an additional physical assessment of the phenomenological model. \cite{II7} A constraint on the potential (1) which follows immediately from the analysis of the normal state properties is a nonzero value of the scattering amplitude in the isotropic channel $v_i$. A lack of the s-wave scattering may lead in some cases to an infinite value of some elements in the dc conductivity tensor, \cite{II7} that is to nonphysical transport properties. The anisotropic potential from Eq. (\ref{eI1}) distinguishes a certain set of coordinates which we think should coincide with the main directions in the crystal. This assumption seems rather plausible because the impurity potential is determined by the dopant atom itself as well as its substitution site in the crystal. The potential produced by a given sort of impurity (defect) may be considered unique since the impurities tend to selectively substitute at characteristic sites in the crystal. The Zn and Ni atoms occupy the in-plane Cu sites \cite{I1,I2,I3,I4,I5,I6,I7,I8,I9,I10,I11,I12,I13,I14,I15,I16,I17,I18} and the electron irradiation displaces the oxygen atoms from the $CuO_2$ planes. \cite{I19,I20,I21,I21a} Therefore on a short length scale of the order of magnitude of the lattice constant a given impurity dopant produces the same potential of the same orientation throughout the crystal. On the other hand, on a large length scale the impurity distribution in the system is random and the Abrikosov-Gorkov's method \cite{II1} of averaging the Green's functions is applicable. We have not made any assumption about the electron energy band yet and we use a general formalism deriving the equations valid for an arbitrary Fermi surface. For the computational simplicity, however, the numerical results are obtained for a cylindrical FS. This approximation allows to study the effect of the anisotropy of the impurity potential alone. The anisotropy of the Fermi surface may equally enhance or suppress the pair-breaking effect of the impurities. The detailed quantitative calculations are needed to answer this question which is beyond the scope of the present paper. For the numerical calculations we take the function $f\left({\bf k}\right)$ in the scattering potential (\ref{eI1}) proportional to the harmonic functions which in a polar angle notation read $sin(l\phi)$ and $cos(l\phi)$, where $l$ is an integer number. Therefore we study the effect of the basis elements in a space of the functions determined by a two-dimensional momentum vector. \section{\bf T-matrix approximation for the self-energy} We study the effect of potential scattering by spinless, noninteracting impurities on the single-particle propagator for superconducting electrons \begin{equation} \label{eII1} \hat{G}\left({\bf k},\omega\right)=\left[i\omega\hat{\tau}_0\hat{\sigma}_0 -\xi_k\hat{\tau}_3\hat{\sigma}_0-\Delta\left({\bf k}\right)i\hat{\tau}_2 \hat{\sigma}_2-\hat{\Sigma}\left({\bf k},\omega\right)\right]^{-1} \end{equation} \noindent Here $\xi_{k}$ is the quasiparticle energy, $\omega=\pi T(2m+1)$, where T is the temperature and $m$ is an integer. $\hat{\tau}_j$, $\hat{\sigma}_j$ ($j=1,2,3$) are the Pauli matrices and $\hat{\tau}_0$, $\hat{\sigma}_0$ are the unit matrices in particle-hole (Nambu) and spin space respectively. The order parameter $\Delta\left({\bf k}\right)$ is defined as \begin{equation} \label{eII2} \Delta\left({\bf k}\right)=\Delta e\left({\bf k}\right) \end{equation} \noindent where $e\left({\bf k}\right)$ is a momentum-dependent real function which may belong to a one-dimensional (1D) irreducible representation of the crystal point group or may be given by a linear combination of the basis functions of different 1D representations. We normalize $e\left({\bf k}\right)$ by taking its average value over the Fermi surface $\left<e^{2}\right>\!=\!1$. This normalization gives $\Delta$ the meaning of the magnitude of the order parameter. The self-energy $\hat{\Sigma}\left({\bf k},\omega\right)$ and consequently the Green's function (Eq. (\ref{eII1})) have been obtained by applying Abrikosov-Gorkov's technique \cite{II1} of averaging over the coordinates of the impurities and depend on only one momentum vector ${\bf k}$. In this approximation \begin{equation} \label{eII3} \hat{\Sigma}\left({\bf k},\omega\right)=n\hat{T} \left({\bf k},{\bf k},\omega\right) \end{equation} \noindent where $n$ is the impurity concentration and $\hat{T}$ obeys the Lippmann-Schwinger equation \cite{II2,II3,III1,III2,III3,III4} \begin{equation} \label{eII4} \hat{T}\left({\bf k},{\bf k'},\omega\right)=\hat{v}\left({\bf k},{\bf k'}\right) +\sum_{{\bf k''}}\hat{v}\left({\bf k},{\bf k''}\right) \hat{G}\left({\bf k''},\omega\right)\hat{T}\left({\bf k''},{\bf k'},\omega\right) \end{equation} \noindent Since the scattering potential for a single electron \begin{equation} \label{eII5} \hat{v}\left({\bf k},{\bf k'}\right)= v\left({\bf k},{\bf k'}\right)\hat{\tau}_3\hat{\sigma}_0 \end{equation} \noindent is momentum-dependent (Eq. (\ref{eI1})), the vertex part $\hat{T}\left({\bf k},{\bf k'},\omega\right)$ is a function of two momenta ${\bf k}$ and ${\bf k'}$. Therefore it should be evaluated from Eq. (\ref{eII4}) using the explicit form of $v\left({\bf k},{\bf k'}\right)$ first, and then the self-energy can be obtained according to Eq. (\ref{eII3}) by taking ${\bf k'}={\bf k}$. We proceed to a solution by defining \begin{equation} \label{eII6} \hat{g}_i\left(\omega\right)=\sum_{{\bf k}}f^{i}\left({\bf k}\right) \hat{G}\left({\bf k},\omega\right)\;,\;\;\;i=0,1,2 \end{equation} \noindent and expanding all matrix quantities as \begin{equation} \label{eII7} \hat{\Sigma}\left({\bf k},\omega\right)=\Sigma_0\left({\bf k},\omega\right) \hat{\tau}_0\hat{\sigma}_0+\Sigma_1\left({\bf k},\omega\right) \hat{\tau}_1\hat{\sigma}_2+\Sigma_2\left({\bf k},\omega\right) \hat{\tau}_2\hat{\sigma}_2+\Sigma_3\left({\bf k},\omega\right) \hat{\tau}_3\hat{\sigma}_0 \end{equation} \begin{equation} \label{eII8} \hat{g}_i\left(\omega\right)=g_{i0}\left(\omega\right) \hat{\tau}_0\hat{\sigma}_0+g_{i1}\left(\omega\right) \hat{\tau}_1\hat{\sigma}_2+g_{i2}\left(\omega\right) \hat{\tau}_2\hat{\sigma}_2+g_{i3}\left(\omega\right) \hat{\tau}_3\hat{\sigma}_0 \end{equation} \noindent The expression for the one-particle Green's function is then \begin{equation} \label{eII8a} \hat{G}\left({\bf k},\omega\right)= -\frac{i\tilde{\omega}\hat{\tau}_0\hat{\sigma}_0 +\tilde{\xi_k}\hat{\tau}_3 \hat{\sigma}_0+\tilde{\Delta}\left({\bf k}\right)i\hat{\tau}_2 \hat{\sigma}_2+\tilde{\Delta}'\left({\bf k}\right)\hat{\tau}_1\hat{\sigma}_2} {\tilde{\omega}^2+\tilde{\xi_k}^2-\tilde{\Delta}^2\left({\bf k}\right) +\tilde{\Delta}'^2\left({\bf k}\right)} \end{equation} \noindent with $\tilde{\omega}=\omega+i\Sigma_0\left({\bf k},\omega\right)$, $\tilde{\xi_k}=\xi_k+\Sigma_3\left({\bf k},\omega\right)$, $\tilde{\Delta}\left({\bf k}\right)=\Delta\left({\bf k}\right)-i \Sigma_2\left({\bf k},\omega\right)$ and $\tilde{\Delta}'\left({\bf k}\right)= \Sigma_1\left({\bf k},\omega\right)$. The formal solutions for the self-energies $\Sigma_j\left({\bf k},\omega\right)$ $(j=0,1,2,3)$ obtained through Eqs. (\ref{eII3}) and (\ref{eII4}) read \begin{equation} \label{eII9} \begin{array}{l} \Sigma_j\left({\bf k},\omega\right)= s_j\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right) u\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right)\\ +s_j\left(v_a,v_i,\hat{g}_2,\hat{g}_1,\hat{g}_0\right) u\left(v_a,v_i,\hat{g}_2,\hat{g}_1,\hat{g}_0\right) f^{2}\left({\bf k}\right)\\ +\left[t_{3-j}\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right) +t_{3-j}\left(v_a,v_i,\hat{g}_2,\hat{g}_1,\hat{g}_0\right) \right]f\left({\bf k}\right) \end{array} \end{equation} \noindent Note a permutation: $v_i\leftrightarrow v_a$, $\hat{g}_0\leftrightarrow\hat{g}_2$, in the arguments of the anisotropic parts (proportional to $f\left({\bf k}\right)$ and $f^2\left({\bf k}\right)$) of $\Sigma_j\left({\bf k},\omega\right)$ in Eq. (\ref{eII9}). The functions $s_j$, $t_j$ and $u$ are given by a series of equations \begin{equation} \label{eII10} \begin{array}{l} s_0\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right)= a_0c_3+a_1c_2+a_2c_1+a_3c_0\\ s_1\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right)= a_0c_2+a_1c_3+ia_2c_0-ia_3c_1\\ s_2\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right)= a_0c_1-ia_1c_0+a_2c_3+ia_3c_2\\ s_3\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right)= a_0c_0+ia_1c_1-ia_2c_2+a_3c_3 \end{array} \end{equation} \begin{equation} \label{eII11} t_j\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right)= v_in c_j\left(c^2_0+c^2_1+c^2_2-c^2_3\right)^{-1} \end{equation} \begin{equation} \label{eII12} u\left(v_i,v_a,\hat{g}_0,\hat{g}_1,\hat{g}_2\right)=v_iv^{-1}_an d^{-1}\left(c^2_0+c^2_1+c^2_2-c^2_3\right)^{-1} \end{equation} \noindent with $d=-g^2_{10}+g^2_{11}+g^2_{12}+g^2_{13}$ and the coefficients $a_j$, $c_j$ $(j=0,1,2,3)$ determined by twelve $g_{ik}$ elements (Eqs. (\ref{eII6}) and (\ref{eII8})) given by the integrals of the products of the Green's function $\hat{G}\left({\bf k},\omega\right)$ and appropriate powers of the impurity potential anisotropy function $f\left({\bf k}\right)$. These coefficients are introduced in order to shorten and simplify the notation of the self-energy functions. We define them in the sequential formulas with a use of additional $b_j$ parameters \begin{equation} \label{eII16} \begin{array}{l} a_0=g_{13}\left(1-v_ag_{23}\right)-v_a\left(g_{12}g_{22}+g_{11}g_{21} -g_{10}g_{20}\right)\\ a_1=ig_{12}\left(1-v_ag_{23}\right)+v_a\left(ig_{13}g_{22}-g_{11}g_{20} +g_{10}g_{21}\right)\\ a_2=-ig_{11}\left(1-v_ag_{23}\right)-v_a\left(ig_{13}g_{21}+g_{12}g_{20} -g_{10}g_{22}\right)\\ a_3=-g_{10}\left(1-v_ag_{23}\right)-v_a\left(g_{13}g_{20}-ig_{12}g_{21} +ig_{11}g_{22}\right) \end{array} \end{equation} \begin{equation} \label{eII15} \begin{array}{l} b_0=\left(1-v_ig_{03}\right)a_0+v_i\left(ig_{02}a_1-ig_{01}a_2 -g_{00}a_3\right)\\ b_1=\left(1-v_ig_{03}\right)a_1+v_i\left(ig_{02}a_0+ig_{00}a_2 +g_{01}a_3\right)\\ b_2=\left(1-v_ig_{03}\right)a_2+v_i\left(g_{02}a_3-ig_{01}a_0 -ig_{00}a_1\right)\\ b_3=\left(1-v_ig_{03}\right)a_3-v_i\left(g_{02}a_2+g_{01}a_1 +g_{00}a_0\right) \end{array} \end{equation} \begin{equation} \label{eII14} \begin{array}{l} c_0=v_a^{-1}d^{-1}b_0-v_ig_{13}\\ c_1=iv_a^{-1}d^{-1}b_1-v_ig_{12}\\ c_2=-iv_a^{-1}d^{-1}b_2-v_ig_{11}\\ c_3=-v_a^{-1}d^{-1}b_3+v_ig_{10} \end{array} \end{equation} \noindent The self-energies can be evaluated with a simultaneous solution of the gap equation \begin{equation} \label{eII17} \Delta\left({\bf k}\right)i\hat{\sigma}_2=-T\sum_{\omega}\sum_{{\bf k'}} \frac{1}{2}V_{{\bf k},{\bf k'}}tr\left[\left(\hat{\tau}_1+\hat{\tau}_2\right)\hat{G} \left({\bf k'},\omega\right)\right] \end{equation} \noindent where $V_{{\bf k},{\bf k'}}=-V_0e\left({\bf k}\right)e\left({\bf k'}\right)$, $V_0>0$, is the pair potential. The above lengthy expressions have been derived without any additional constraints and are fundamental to the considerations within the model. This generally complicated problem simplifies with the assumption of particle-hole symmetry of the excitation spectrum. It has been shown by Hirschfeld et al. \cite{III2} that in this case the $\hat{\tau}_3$ component of the integrated Green's function ($g_{03}$) may be neglected in the presence of $s$-wave scatterers. Inclusion of higher-order angular momentum waves in the scattering potential makes this analysis considerably more difficult. In this paper, however, we calculate the self-energies at the phase transition and for that purpose we need to show only the consistency of the assumption $\Sigma_3=0$ at the critical temperature $T_c$. In the following we assume a particle-hole symmetry of the energy spectrum, take $g_{i3}=0$ $(i=0,1,2)$ and check if this condition leads to a vanishing self-energy $\Sigma_3$. \section{Self-energy at phase transition} We consider the effect of anisotropic impurity scattering on the critical temperature. At the superconducting-normal state phase transition the gap equation (\ref{eII17}) transforms into \begin{equation} \label{eIII1} 1=V_0T_c\sum_{\omega}\sum_{{\bf k}}e\left({\bf k}\right) \frac{\displaystyle e\left({\bf k}\right)+\frac{1}{i} \left(\Sigma_1\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0} +\frac{1}{i} \left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}} {\displaystyle\left(\omega+i\Sigma_0\left({\bf k},\omega\right)_{\Delta=0}\right)^2 +\left(\xi_k+\Sigma_3\left({\bf k},\omega\right)_{\Delta=0}\right)^2} \end{equation} \noindent In order to find $T_c$ the self-energies in $\Delta\rightarrow 0$ limit are to be obtained. For the sake of convenience we introduce new parameters $c_i=1/(\pi N_0 v_i)$ and $c_a=1/(\pi N_0 v_a)$ describing the scattering strength in the isotropic ($c_i$) and anisotropic ($c_a$) channel respectively. $N_0$ represents the overall single-spin density of states at the Fermi level. A new measure of impurity concentration $\Gamma=n/(\pi N_0)$ is also used. Taking into account that $\left(g_{00}\right)_{\Delta=0}=-i\pi N_0 sgn(\omega)$ and $\left(g_{20}\right)_{\Delta=0}=\left(g_{00}\right)_{\Delta=0}$ (because of normalization $\left<f^2\right>=1$) we obtain \begin{equation} \label{eIII2} \Sigma_0\left({\bf k},\omega\right)_{\Delta=0}= -i\Gamma\left[\frac{1}{c^2_i+1}+\frac{1}{c^2_a+1} f^{2}\left({\bf k}\right)\right]sgn\left(\omega\right) \end{equation} \begin{equation} \label{eIII4} \Sigma_3\left({\bf k},\omega\right)_{\Delta=0}= \Gamma\left[\frac{c_i}{c^2_i+1}+\frac{c_a}{c^2_a+1} f^{2}\left({\bf k}\right)\right] \end{equation} \noindent Calculation of $\left(\Sigma_1\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}$ and $\left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}$ quantities is more tedious and requires a solution of two sets of two linear equations which determine four unknowns $\left(g_{j1}/\Delta\right) _{\Delta=0}$, $\left(g_{j2}/\Delta\right)_{\Delta=0}$ $(j\!=\!0,2)$ (note that $g_{11}\!=\!g_{12}\!=\!0$ because of $\left<f\right>\!=\!0$). This procedure leads to $\left(\Sigma_1\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}=0$ and \begin{equation} \label{eIII5} \begin{array}{l} \displaystyle\left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}= -\frac{\Gamma}{\pi N_0}\left[\frac{1}{c^2_i+1}\left(g_{02}/\Delta\right) _{\Delta=0} +\frac{1}{c^2_a+1}\left(g_{22}/\Delta\right) _{\Delta=0}f^{2}\left({\bf k}\right)\right]sgn\left(\omega\right) \end{array} \end{equation} \noindent where \begin{equation} \label{eIII6} \begin{array}{l} \displaystyle\frac{1}{\pi N_0}\left(g_{02}/\Delta\right)_{\Delta=0}= \left(c^2_i+1\right)\left[c^2_i+1-\Gamma\left<\frac{1} {\tilde{\omega}_0}\right>\right]^{-1}\\ \\ \displaystyle\times\left[-i\left<\frac{e} {\tilde{\omega}_0}\right> +\Gamma\left(c^2_a+1\right)^{-1} \left<\frac{f^2}{\tilde{\omega}_0}\right> \frac{1}{\pi N_0}\left(g_{22}/\Delta\right)_{\Delta=0}\right] \end{array} \end{equation} \noindent and \begin{equation} \label{eIII7} \begin{array}{l} \displaystyle\frac{1}{\pi N_0}\left(g_{22}/\Delta\right)_{\Delta=0}= -i\left(c^2_a+1\right)\\ \\ \displaystyle\times\left[\left(c^2_a+1-\Gamma\left<\frac{f^4} {\tilde{\omega}_0}\right>\right)\left(c^2_i+1-\Gamma\left<\frac{1} {\tilde{\omega}_0}\right>\right)-\Gamma^2\left<\frac{f^2} {\tilde{\omega}_0}\right>^2\right]^{-1}\\ \\ \displaystyle\times\left[\left<\frac{ef^2} {\tilde{\omega}_0}\right>\left(c^2_i+1-\Gamma\left<\frac{1} {\tilde{\omega}_0}\right>\right)+\Gamma\left<\frac{f^2} {\tilde{\omega}_0}\right>\left<\frac{e}{\tilde{\omega}_0}\right>\right] \end{array} \end{equation} \noindent with $\tilde{\omega}_0=\omega+i\Sigma_0\left({\bf k},\omega\right)_{\Delta=0}$. In the following subsections we consider the above self-energies in the Born and unitary scattering limits. \subsection{Born scattering} When both isotropic and anisotropic impurity scattering channels are in the Born scattering regime i.e. $c_i\gg 1$ ($v_iN_0\ll 1$), $c_a\gg 1$ ($v_aN_0\ll 1$) only the lowest order terms in the impurity potential play role in the self-energies. Keeping up to the square terms in $v_i$ ($v_a$) we obtain from Eqs. (\ref{eIII2}) and (\ref{eIII4}) \begin{equation} \label{eIII11} \Sigma_0\left({\bf k},\omega\right)_{\Delta=0}= -i\pi nN_0\left[v^2_i+v^2_af^2\left({\bf k}\right)\right]sgn\left(\omega\right) \end{equation} \begin{equation} \label{eIII12} \Sigma_3\left({\bf k},\omega\right)_{\Delta=0}= n\left[v_i+v_af^2\left({\bf k}\right)\right] \end{equation} \noindent Though $\Sigma_3$ is nonzero, it may be absorbed into the chemical potential and its effect vanishes in the Born scattering. \cite{II1} For $\left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}$ we obtain \begin{equation} \label{eIII13} \left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}= i\pi nN_0\left[v^2_i\left<e\right>+v^2_a\left<ef^2\right> f^2\left({\bf k}\right)\right]\frac{sgn\left(\omega\right)}{\omega} \end{equation} \noindent It is noteworthy that apart from the average value of the order parameter symmetry function, $\left<e\right>$, a term reflecting the overlap between $e\left({\bf k}\right)$ and $f^2\left({\bf k}\right)$ influences the self-energy $\left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}$. The limit of $v_a=0$ gives the standard $s$-wave impurity scattering in the Abrikosov-Gorkov approximation \cite{II1} with $\Sigma_0\left({\bf k},\omega\right)_{\Delta=0}=-i\pi nN_0v^2_isgn\left(\omega\right)$, $\left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}= i\pi nN_0v^2_i\left<e\right>sgn\left(\omega\right)/\omega$, and the critical temperature determined by $ln\left(T_c/T_{c_{0}}\right)=\left(\left<e\right>^2-1\right) \left[\Psi\left(1/2+nN_0v^2_i/\left(2T_c\right)\right)-\Psi\left(1/2\right)\right]$. \subsection{Unitary scattering} The limit of the resonant impurity scattering in both isotropic and anisotropic channels i.e. $c_i\rightarrow 0$, $c_a\rightarrow 0$ in Eqs. (\ref{eIII2}) and (\ref{eIII4}) leads to \begin{equation} \label{eIII14} \Sigma_0\left({\bf k},\omega\right)_{\Delta=0}= -i\Gamma\left[1+f^2\left({\bf k}\right)\right]sgn\left(\omega\right) \end{equation} \noindent and $\Sigma_3\left({\bf k},\omega\right)_{\Delta=0}=0$ which is consistent with the assumption of particle-hole symmetry of the excitation spectrum. The unitarity limit in Eqs. (\ref{eIII5})-(\ref{eIII7}) gives \begin{equation} \label{eIII16} \begin{array}{l} \displaystyle\left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}= i\Gamma\left\{\left[1-\Gamma\left<\frac{1}{\tilde{\omega}_0}\right> \right]^{-1}\left<\frac{e}{\tilde{\omega}_0}\right>\right.\\ \\ \displaystyle+\left[\Gamma\left(1-\Gamma\left<\frac{1}{\tilde{\omega}_0}\right>\right)^{-1} \left<\frac{f^2}{\tilde{\omega}_0}\right>+f^2\left({\bf k}\right)\right]\\ \\ \displaystyle\times\left[\left(1-\Gamma\left<\frac{1}{\tilde{\omega}_0}\right>\right) \left(1-\Gamma\left<\frac{f^4}{\tilde{\omega}_0}\right>\right)-\Gamma^2 \left<\frac{f^2}{\tilde{\omega}_0}\right>^2\right]^{-1}\\ \\ \displaystyle\left.\times\left[\left(1-\Gamma\left<\frac{1}{\tilde{\omega}_0}\right>\right) \left<\frac{ef^2}{\tilde{\omega}_0}\right>+\Gamma \left<\frac{f^2}{\tilde{\omega}_0}\right> \left<\frac{e}{\tilde{\omega}_0}\right>\right]\right\}sgn\left(\omega\right) \end{array} \end{equation} \noindent The impurity potential may lead to the strong scattering in one channel and the weak scattering in another. If the strong scattering takes place in the isotropic channel and the scattering in the anisotropic one is in the Born limit, then we deal with a case of the isotropic unitary scattering. \cite{II2,II3,III1,III2,III3,III4} The opposite case with the strong scattering in the anisotropic channel and the weak scattering in the isotropic one is equivalent to a nonphysical situation of unitary scattering in the anisotropic channel alone as discussed in section II. \section{Critical temperature} The critical temperature in the weak-coupling BCS approximation is determined by the equation \begin{equation} \label{eIV1} \begin{array}{l} \displaystyle ln\left(\frac{T_c}{T_{c_{0}}}\right)= \pi T_c\sum_{\omega}\left[\frac{1}{\pi N_0}\sum_{{\bf k}}e\left({\bf k}\right) \frac{\displaystyle e\left({\bf k}\right)+\frac{1}{i} \left(\Sigma_1\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0} +\frac{1}{i} \left(\Sigma_2\left({\bf k},\omega\right)/\Delta\right)_{\Delta=0}} {\displaystyle\left(\omega+i\Sigma_0\left({\bf k},\omega\right)_{\Delta=0}\right)^2 +\left(\xi_k+\Sigma_3\left({\bf k},\omega\right)_{\Delta=0}\right)^2}\right.\\ \\ \displaystyle\left.-\frac{sgn\left(\omega\right)}{\omega}\right] \end{array} \end{equation} \noindent where because of the momentum-dependent self-energies the summations over the quasimomentum vector ${\bf k}$ restricted to the Fermi surface and Matsubara frequency $\omega$ need to be performed numerically. We do the calculations in two different scattering regimes - Born and unitary for the $d$-wave order parameter. Finally the results for the $(d+s)$-wave superconductor are discussed briefly. In order to proceed further one has to choose a function describing the anisotropy of the impurity potential. We study the functions given by the subsequent harmonics as they represent an orthonormal and complete set in the interval [0, 2$\pi$] and can be used for the Fourier expansion of any regular function. Harmonics up to the 4th order are analyzed and the results are extended to higher-order harmonic functions. It is important to realize that the impurity potential of $p$-wave anisotropy shows a particular property. Given by $sin\phi$ or $cos\phi$ function it represents a small-angle scattering which is relatively unimportant in contributing to the resistivity of the normal state. Taken even together with the $s$-wave scattering channel it results within the linear response approximation to the distribution function in a not well defined integrals determining the dc conductivity for some electric field orientations. \cite{II7} In other words, in the Boltzmann equation based analysis of the transport properties the $p$-wave scattering amplitude couples to the ''$1-cos(\theta)$" term in the collision integral. Therefore, in a fit to the real systems the $p$-wave scattering can be considered only as one of the components in the anisotropic channel of the impurity potential coexisting with some other higher-order harmonics. In order to emphasize this different feature of the $p$-wave potential among the other basis functions we present the momentum-dependent scattering with various amounts of the $p$-wave potential in a separate figure. The anisotropic impurity scattering is compared to the isotropic one of the same scattering strength. It means that we discuss the effect of substituting a part of a given isotropic impurity potential $v_0$ with an anisotropic term. The amplitude of the replaced isotropic potential, $v_0$, and the amplitudes of the isotropic and anisotropic scattering channels in the studied potential are related through the formulas $v_i=\alpha v_0$, and $v_a=\left(1-\alpha\right)v_0$, where the coefficient $\alpha$ ($0\leq\alpha\leq 1$) defines the partition of the potential. \subsection{Born scattering} The impurity scattering is analyzed as a function of the pair-breaking parameter $\Gamma'=\pi nN_0v^2_0$ ($v_0=v_i+v_a$) which includes the overall scattering strength that is, takes both isotropic and anisotropic channels into account. It is convenient to use the ratio $v_a/v_i=\left(1-\alpha\right)/\alpha$ to define a particular potential. \subsubsection{s-wave superconductor} We note, before discussing the unconventional superconductivity, that for the isotropic s-wave superconductor, given by $e\left({\bf k}\right)=1$, the $T_c$ equation (\ref{eIV1}) along with Eqs. (\ref{eIII11}) and (\ref{eIII13}) do not lead to a change of the critical temperature in agreement with the Anderson's theorem. \cite{IV1} \subsubsection{d-wave superconductor} The critical temperature for a $d_{x^2-y^2}$-wave superconductor that is for $e\left({\bf k}\right)\sim (k^2_x-k^2_y)$ is presented for a large range of $v_a/v_i$ ratio values in Fig. 1. The major common feature of these diagrams is a lower $T_c$ suppression by the anisotropic impurity scattering compared to the isotropic ($s$-wave) one. According to their effect on superconductivity the anisotropic potentials can be classified in three groups defined by the function $f\left({\bf k}\right)$: $cos2\phi$, $sin2\phi$, and higher-order harmonics. The first of them, determined by $f\left({\bf k}\right)\sim cos2\phi$, leads to the strongest $T_c$ suppression. In this case the directions of the maximum impurity scattering correspond to the maxima of the order parameter and most of the pair-breaking process takes place in this region. Therefore, the suppression of superconductivity is particularly strong. However, for the level of anisotropy up to $\left(1-\alpha\right)\approx 0.5$ ($v_a/v_i\approx 1$) the depairing effect of impurities is reduced with increasing contribution of $cos2\phi$ scattering compared to the isotropic scattering. Further increase of the anisotropic part in the impurity potential up to a level $\left(1-\alpha\right)\approx 0.6$ ($v_a/v_i\approx 1.5$) almost does not change the pair-breaking effect. When the contribution of the anisotropy exceeds this level the impurity effect on $T_c$ is enhanced and finally for $\left(1-\alpha\right)$ of the order of 0.83 ($v_a/v_i=5$) becomes comparable to the one of the $s$-wave scatterers as it is shown in Fig. 1e. The out of phase scattering takes place for $f\left({\bf k}\right)\sim sin2\phi$. This function is a $45^{\circ}$ rotation in the xy plane of the order parameter function $e\left({\bf k}\right)\sim cos2\phi$, that is, the impurity potential maxima correspond to the superconducting gap nodes and vice versa. In this way the impurity pair-breaking effect is minimized. The suppression of the critical temperature is reduced by an increasing amount of the anisotropic scattering in the impurity potential. For large $v_a/v_i$, however, the shape of the suppression lines changes slightly and a small enhanced suppression can be observed for some impurity concentration (Fig. 1e). In general, the anisotropic impurity potential given by a function orthogonal to the order parameter is less pair-breaking than $f\left({\bf k}\right)\sim e\left({\bf k}\right)$. The $T_c$ suppression by the 3rd and 4th order harmonics is less than the one of $cos2\phi$ but it exceeds that of $sin2\phi$. It is also decreasing for increasing level of anisotropy in the impurity potential up to $\left(1-\alpha\right)\approx 0.5$, then the pair-breaking effect of impurities practically saturates. The curvature of the graphs changes at some points, however, and the normalized critical pair-breaking parameter, $\Gamma'/\left(2\pi T_{c_{0}}\right)$, at which $T_c\approx 0$ may increase in some cases. It is worth observing that the critical temperature is almost equally suppressed by the impurity potentials given by the 3rd and 4th order harmonics for $v_a$ values up to $v_a/v_i=1$. This is particularly true for small anisotropic scattering levels. When the contribution of the anisotropic channel in the impurity potential is larger, the differences are more pronounced. Nevertheless, even for the amplitudes ratio as high as $v_a/v_i=2$\\ ( $1-\alpha\approx 0.67$ ) the curves differ only near a zero value of the critical temperature and the normalized critical pair-breaking parameters, $\Gamma'/\left(2\pi T_{c_{0}}\right)$, of these four harmonics are within an interval of the order of magnitude of $10^{-2}$. For the sake of transparency we present $cos3\phi$ scattering effect only in Fig. 1d. Though the curves of different harmonics overlap for the most of the critical temperature range for the amplitude in the anisotropic scattering channel five times as much as the one in the isotropic channel, $v_a/v_i=5$, they split distinctly at a low temperature of about $0.2\;T_c$ (Fig. 1e). A study of the higher order harmonics up to the 10th order for the anisotropic scattering strength equal to the isotropic one ($v_a/v_i=1$) leads to almost the same result as the 3rd and 4th order harmonics. This approximately universal behavior for harmonics from the 3rd to the 10th order is shown in Fig. 2. Concluding we may say that to a good accuracy the anisotropy of the impurity potential given by the harmonics of the order higher than two yields an approximately universal $T_c$ suppression. It is also worth mentioning that except for the isotropic scattering the critical temperature goes to a zero value asymptotically. The asymptotic tails start at very low temperatures of the order of magnitude of $10^{-4}\;T_{c_{0}}$ (not seen at the figure scale) where the fluctuation effects become important and may destroy superconductivity. Remembering that this potential cannot be considered on its own without any higher-order admixture, we show the pair-breaking effect of the $p$-wave scattering in the anisotropic channel for different $v_a/v_i$ values in Fig. 3. Compared to the other harmonics (Figs. 1-2) it yields the lowest $T_c$ suppression. This fact can be explained by an effectively small-angle scattering of the $p$-wave potential. If it is given by $cos\phi$ function then mostly the quasiparticles with their momenta parallel to the x-axis are affected, and for the $sin\phi$ function representing impurity potential the electrons moving along the y-axis are being scattered. Thus we deal with a weak practically one-dimensional scattering in a two-dimensional space. The pair-breaking effect of both basis functions $cos\phi$ and $sin\phi$ is the same within the accuracy of the numerical calculations. Worth observing is also a feature of a reduced $T_c$ suppression with an increasing amount of $p$-wave scattering in the impurity potential. However, at large amplitudes in the anisotropic channel ($v_a/v_i\sim 5$) the initial suppression of the critical temperature can be enhanced. Characteristic is also a clear asymptotic decrease of the critical temperature to its zero value. Any impurity potential given by Eq. (\ref{eI1}) can be represented as a combination of the potentials based on single harmonics. Therefore, we expect that the pair-breaking effect will be given by an appropriate superposition of the effects discussed above. \subsubsection{(d+s)-wave superconductor} As an example of a $(d_{x^2-y^2}+s)$-wave superconductor we consider the one with an $s$-wave admixture of $10\%$. This is the order of magnitude of the $s$-wave level which cannot be ruled out by the ARPES measurements in the cuprates. \cite{IV2,IV3} The normalized to unity order parameter is given then by $e\left({\bf k}\right)=\left(cos2\phi+s\right)/ \left<\left(cos2\phi+s\right)^2\right>^{1/2}$, where $s=\left(0.005/0.99\right)^{1/2}$ and its FS average, $\left<e\right>=0.1$, corresponds to $10\%$ of the $s$-wave fraction in the $(d_{x^2-y^2}+s)$-wave superconductor. The results for $v_a/v_i=1$ are shown in Fig. 4. Even a small amount of the $s$-wave component, which is not destroyed by the potential scattering, results in a robustness of $T_c$ compared to the pure $d$-wave superconductor (Fig. 1c) and its asymptotic reduction. A change in the symmetry of the order parameter also causes a split in the suppression of the critical temperature by the $p$-wave anisotropic channel scattering. \subsection{Unitary scattering} The relation determining the partition of the scattering amplitudes between the isotropic and anisotropic channels, $v_i+v_af\left({\bf k}\right)f\left({\bf k'}\right)= v_0\alpha+v_0\left(1-\alpha\right)f\left({\bf k}\right)f\left({\bf k'}\right)$, holds as long as these amplitudes are finite. In the case of the resonant scattering we make an approximation of $v_i$ and $v_a$ diverging to infinity. Therefore we cannot control the relative scattering strengths in both channels and these processes become independent. Each part of the impurity potential enters the self-energy functions (Eqs. \ref{eIII2}-\ref{eIII7}) through variables $c_i$ and $c_a$. It is important to note that the self-energies depend on $c_i$ and $c_a$ parameters separately so they are functions of $c_i$ and $c_a$ and not of any combination of them (like $c_a/c_i$ for instance). In the unitarity limit $c_i\rightarrow 0$, $c_a\rightarrow 0$ and the contributions from the isotropic and anisotropic channels are in fact equal. Similarly to the Born scattering limit, we want to discuss the effect of an anisotropic impurity potential that replaces the isotropic one. In order to do an appropriate comparison we need to refer to the isotropic scattering in two channels or equivalently to a regular isotropic scattering in one channel with a doubled impurity concentration $2n$. Therefore, to compare the anisotropic unitary scattering with a corresponding isotropic one, the impurity concentration for the anisotropic scattering model must be half of the concentration of the $s$-wave impurities. We take the impurity concentration equal to $n/2$ for the two-channel anisotropic potential and the results are plotted as a function of $\Gamma'=2\Gamma=n/\left(\pi N_0\right)$, where $n$ is the real impurity concentration the same as in the Abrikosov-Gorkov scaling function for the $s$-wave scattering. That is, the two-channel scattering potential is averaged over two impurities so there is only one scattering channel per impurity present. This procedure introduces the anisotropic scattering potential of the form given by Eq. (\ref{eI1}) and the same scattering strength as the isotropic one in the unitarity limit. \subsubsection{s-wave superconductor} The lack of influence of the potential anisotropic impurity scattering in the unitary limit on the critical temperature can be shown rigorously for a small concentration of the defects $n/(\pi N_0)\ll 1$, when the self-energies $\Sigma_0\left({\bf k}\right)_{\Delta=0}=-i\Gamma\left[1+f^2\left({\bf k}\right)\right] sgn\left(\omega\right)$ (Eq. (\ref{eIII14})) and $\left(\Sigma_2\left({\bf k}\right) /\Delta\right)_{\Delta=0}=-\Sigma_0\left({\bf k}\right)_{\Delta=0}/\omega$ (Eq. (\ref{eIII16})). For a larger impurity concentration a numerical analysis confirms this result with a very good accuracy. \subsubsection{d-wave superconductor} The effect of the anisotropic unitary scattering on a $d_{x^2-y^2}$-wave superconductor is shown in Fig. 5. The same general rule as for the Born scattering holds here. The weakest pair-breaking effect is caused by the impurity potential involving $sin2\phi$ and most suppression is seen for the potential containing $cos2\phi$, that is the anisotropic scattering channel in phase with the order parameter. All the other harmonics orthogonal to $cos2\phi$ yield a moderate suppression of the critical temperature comparable to the isotropic scattering. They also form an almost universal suppression curve. A study of the $T_c$ reduction by the harmonics of the order up to ten shows the same approximate universality as that in Fig. 2, with the differences in the normalized critical pair-breaking parameter, $\Gamma'/ \left(2\pi T_{c_{0}}\right)$, of the order of $10^{-3}$. The curves of the 3rd and 4th order harmonics overlap in Fig. 5. Particularly interesting feature of the resonant scattering is a strong pair-breaking effect of the $cos2\phi$ anisotropic potential. It destroys superconductivity even faster than the isotropic impurities. Another characteristic fact is practically the same critical temperature dependence on the impurity concentration (i.e. not distinguishable in Fig. 5) for the $p$-wave scattering in the anisotropic channel and $sin2\phi$ scattering potential. Similarly to the weak scattering limit, except for the $s$-wave scattering the suppression of superconductivity at low $T_c$ is asymptotic, however not seen at the scale of Fig. 5. \subsubsection{(d+s)-wave superconductor} A $(d_{x^2-y^2}+s)$-wave superconductor is more robust against the impurity scattering due to a nonzero $s$-wave component. Since $cos2\phi$ function is no longer in phase with the order parameter its pair-breaking effect is lowered and becomes even less than the isotropic one for low critical temperatures. The results for the same level of the $s$-wave part in a $(d_{x^2-y^2}+s)$-wave order parameter as the one discussed in the Born limit are shown in Fig. 6. It is worth mentioning that the $p$-wave scattering leads again to the same $T_c$ suppression as the $sin2\phi$ anisotropic potential. \section{Comparison to experiment} In the overdoped samples $T_c/T_{c_{0}}$ data points plotted vs. impurity concentration form a universal curve independent of the critical temperature in the absence of impurities. \cite{I10,V0} Such a universal scaling behavior is characteristic of the impurity limited superconductivity provided the pair-breaking parameter $\Gamma'/\left(2\pi T_{c_{0}}\right)$ does not change with $T_{c_{0}}$. This requirement is equivalent to a constraint $N_0 T_{c_{0}}=constant$ in the unitarity limit and may be obeyed in the overdoped systems where the critical temperature decreases with increasing hole concentration. \cite{I10} There are also indications that a large residual resistivity due to Zn atoms in the cuprates corresponds to an impurity potential scattering in the unitary limit. \cite{I2} Thus, we consider the case of the unitary scattering focusing on the effect of Zn dopant. Working in the Born scattering limit requires an estimation of the impurity scattering potential which can be obtained from the residual resistivity in the normal state. \cite{I20,I21,I22,II7,I24} In that analysis a clear distinction between the impurity scattering life time and the transport relaxation time, which may differ if the impurity potential is anisotropic, \cite{II7} is needed. In order to compare our unitary scattering results with the experiment we have to convert the pair-breaking parameter $\Gamma'/\left(2\pi T_{c_{0}}\right)$ into the impurity concentration. As $\Gamma'=n/\left(\pi N_0\right)$ only the values of the density of states on the Fermi surface and the critical temperature in the absence of impurities are needed for that purpose. The density of states at the Fermi level is estimated from the measurements of the specific heat jump at the phase transition $\Delta C$. \cite{I1,V1} We employ the BCS weak-coupling relation $\Delta C/\gamma T_c\approx 1.43$ to obtain the normal-state Sommerfeld constant $\gamma$ which gives the density of states through $\gamma =2\pi^2k_B^2N_0/3$. It is important to note that the strong-coupling corrections \cite{V2,V3} as well as the interaction with impurities \cite{V4,V5} may change this relation significantly and in consequence alter the overall density of states. It may result in a wide range of $T_c$ solutions. \cite{V6} Because of a difficulty in the separation of lattice and electron contributions to the specific heat the thermodynamic experiments provide $\Delta C$ values with the accuracy depending on the quality of a sample. For high purity, fully oxygenated $Y\!-\!123$ compound grown in $BaZrO_3$ the mean-field component of the electronic specific heat jump is estimated as $56 \pm 2 \left(mJ/K^2 mole\right)$. \cite{V1,V7,V8} In $La\!-\!214$ system this quantity is in the range of $14\pm 5 \left(mJ/K^2 mole\right)$. \cite{V1,V9} There is no specific heat jump observed at the phase transition in $Bi\!-\!2122$ compound \cite{V1} thus the present method of obtaining the density of states cannot be applied in this case. We discuss the experimental results in this system together with $Y\!-\!123$ and $La\!-\!214$ compounds more extensively in Ref. 53. The density of states calculated from the above $\Delta C$ values are used for the evaluation of the impurity concentration, which is represented as the per cent number of impurities (defects) per planar Cu site. We do the calculations with a fixed density of states as we do not have any quantitative data showing its change with doping or disorder. Therefore, the results are not really universal and depend on the critical temperature in the absence of impurities $T_{c_{0}}$. We group them according to the values of the critical temperature of a pure system. In the optimally doped $Y\!-\!123$ compound $T_{c_{0}}$ varies in a very narrow range of values and for the theoretical calculation we take $T_{c_{0}}=91K$. The phase diagrams for the $d$-wave superconductor together with the experimental data of Zn doped samples \cite{I13,I14,I15,I16,I17} are shown in Fig. 7. The scattering effect of a given potential corresponds to the area between two curves of the same style, for instance the solid lines stand for the $sin2\phi$ impurity potential. This broadened range of values stem from the uncertainty in the estimation of the FS density of states, that is, from the accuracy of the $\Delta C$ values. We note, that all the experimental points fall in the region of the theoretically predicted $T_c$ suppression by the unitary impurity scattering. Interestingly, the pair-breaking effect of Zn atoms in certain samples corresponds to the resonant scattering in different anisotropic channel of the impurity potential. We can clearly distinguish data that can be approximated by the scattering potentials given by $cos2\phi$ and $sin2\phi$ functions. The rest of the experimental points lie in the range of isotropic scattering or anisotropic scattering with higher order harmonics. Although the critical temperatures of the analyzed samples are very close, we do not see a universal suppression dependence as in $Bi\!-\!2122$ and $La\!-\!214$ compounds. \cite{I10,V0} This feature may suggest a possible sample-dependence of the effective impurity scattering caused for instance by the differences in a sample preparation or Zn substitution processes. Even small changes in the hole concentration alter the electronic density of states at the Fermi level and result in a modified impurity scattering rate. According to $T_{c_{0}}$ values, we gather the experimental data for Zn doped $La\!-\!214$ in two groups corresponding to $T_{c_{0}}$ equal to $30K$ (Fig. 8a) \cite{I1} and $36K$ (Fig. 8b). \cite{I1,I3} For the sample of $T_{c_{0}}\approx 30K$ the experimental suppression of the critical temperature is in the range of anisotropic scattering determined by $sin2\phi$ (Fig. 8a). The pair-breaking effect of Zn atoms in the samples characterized by $T_{c_{0}}\approx 36K$ is on the edge of anisotropic scattering in higher order harmonics but also right in the middle of $sin2\phi$ based anisotropic scattering. As we can see, the quantitative calculations for $La\!-\!214$ compound contain a large uncertainty margin which is caused by a lack of a precise value of the specific heat jump at the phase transition and consequently of the electron density of states on the Fermi surface. The error in the experimentally estimated magnitude of $\Delta C$ is of the order of $36\%$. A similar analysis of Ni substituted samples shows that the pair-breaking effect in the unitary limit is stronger than observed experimentally. For the sake of comparison we present Ni doped $Y\!-\!123$ compound data \cite{I12,I14,I15,I18} and our theoretical curves for the resonant scattering in Fig. 9. Less difference between the experimental and theoretical results is seen for $La\!-\!214$. \cite{V6} Although this comparison is suggestive for a weak potential scattering, the detailed calculations in the Born scattering limit giving the critical temperature dependence on the residual resistivity \cite{II7} are needed in order to draw more firm conclusions. Finally we discuss the electron irradiation experiments in $Y\!-\!123$. \cite{I19,I20,I21,I21a} The results for low-energy (60-120 keV) incident electrons read from Fig. 12 of Ref. 21 are shown in Fig. 10. The experimental points are in the range of the pair-breaking effect of the scattering potential with the anisotropy given by $sin2\phi$. The initial $T_c$ suppression, however, is more gradual than the one obtained from the theoretical calculation and cannot be explained by the standard impurity pair-breaking mechanism. Unfortunately, we do not have any other set of electron irradiation data representing the change of the critical temperature with defect concentration. The analysis by Tolpygo et. al. \cite{I21} shows that the initial $T_c$ suppression for low-energy (100 keV) electron irradiation reported by Legris et al. \cite{I21a} agrees with the one shown in Fig. 10. For higher-energy electrons \cite{I19} $T_c$ is reduced initially about twice as much as in Ref. 21 which becomes in the range of the $sin2\phi$ determined behavior for low defect concentration. \section{Conclusions} We have studied the pair-breaking effect of the anisotropic impurity scattering in the t-matrix approximation for $d_{x^2-y^2}$-wave and $(d_{x^2-y^2}+s)$-wave superconductors. The Born and the unitary limits have been discussed analytically and numerically. Although limited to the employed phenomenological model the conclusions we draw should be suggestive to a larger class of potentials capturing the feature of anisotropy. We have shown, that the effect of the anisotropic impurity scattering can be considered in four groups for the Born scattering and three groups in the case of unitary scattering which are determined by the form of the function $f\left({\bf k}\right)$ defining symmetry of the scattering potential. In both scattering limits the strongest suppression of the critical temperature is caused by the impurity potential given by $f\left({\bf k}\right)\sim cos2\phi$ which is in phase with the order parameter for the $d_{x^2-y^2}$ state and overlaps significantly with the $(d_{x^2-y^2}+s)$-wave state of a major $d$-wave component. The pair-breaking effect of this potential with a large amount of anisotropic amplitude can be comparable with the one of the isotropic scattering in the Born limit and exceeds the $s$-wave impurity effect in the unitary limit. This issue is particularly important as it shows that a weak reduction of $T_c$ due to anisotropy of the impurity potential proportional to the superconducting order parameter which follows from the weak-scattering model \cite{I24} is not a general feature of the anisotropic impurity scattering. Another class is defined by $f\left({\bf k}\right)\sim sin2\phi$. It leads to the lowest impurity pair-breaking effect in the case of the unitary scattering and second lowest for the Born limit. This kind of scattering is maximal in the direction of the nodes of the order parameter and it vanishes where the gap function has its maxima. Therefore, the effective scattering is minimized by the symmetry of the impurity potential. Any other function orthogonal to the $d$-wave order parameter results in a rather universal $T_c$ suppression and falls into the third group of the potentials. The pair-breaking in this case is less than that of the isotropic scattering but it exceeds the one of $sin2\phi$ based potential, and in the unitary limit is very close to the isotropic scattering. Resonant scatterers in the $p$-wave channel lead to the same $T_c$ suppression as the $d$-wave scattering given by $sin2\phi$ function. In the Born scattering limit, however, the $p$-wave anisotropic scattering results in the lowest $T_c$ suppression. We have compared our results for the $d_{x^2-y^2}$-wave superconductor with the experimental data assuming that the impurity potential is close to the unitary limit. Within the accuracy of our calculations the Zn atoms can be considered as the resonant scatterers in overdoped $La\!-\!214$ and $Y\!-\!123$ compounds. The pair-breaking effect of the structure defects produced by the electron irradiation is also in the range of magnitude of the unitary scattering with the anisotropy of the impurity potential given by $sin2\phi$ function. The theoretical results strongly depend on the value of the electronic density of states on the Fermi surface. Thus far, for some systems (like $La\!-\!214$) there is only an estimation of this value with a large margin of uncertainty available. It opens a wide range of possible $T_c$ solutions and makes a proper analysis very difficult. Therefore, before any final conclusion about the impurity suppression of the critical temperature in the cuprates can be made, a much more accurate determination of the electronic density of states on the Fermi surface has to be done. \section*{Acknowledgments} We would like to thank B. Nachumi for his help in analyzing the experimental data, as well as T. Kluge for the discussions on the effect of impurities in high-$T_c$ superconductors and pointing to us several important papers. We are also grateful to A. Junod for providing us with the specific heat data and for the enlightening explanations of the thermodynamic measurements in the cuprates. Finally, we would like to thank P. Hirschfeld for critical comments on the manuscript. This work was supported by the Natural Sciences and Engineering Research Council of Canada. \newpage
\section{Introduction} The 1720.53 MHz line from the hydroxyl radical (OH) was conclusively shown to be associated with the supernova remnant (SNR) W\thinspace{28} by Frail, Goss \& Slysh (1994). Since then several surveys have been made toward other Galactic SNRs (Frail et al. 1996, Yusef-Zadef et al. 1996, Green et al. 1997, Koralesky et al. 1998), clearly establishing that the OH(1720 MHz) line toward masers is a new class of OH maser, distinct from those in star-forming regions and evolved stars. Follow-up work (Claussen et al. 1997, Frail \& Mitchell 1998, Wardle, Yusef-Zadeh \& Geballe 1998) supports the hypothesis that the OH(1720 MHz) masers originate in C-type shocks, transverse to the line-of-sight, being driven into adjacent molecular clouds by the expanding SNR. The measured densities, temperatures and magnetic fields from these studies are consistent with collisional excitation of the OH by the H$_2$ molecules in the post-shock gas (Elitzur 1976, Pavlakis \& Kylafis 1996, Lockett, Gauthier, \& Elitzur 1999). Claussen et~al. (1997) imaged the masers toward the SNRs W\thinspace{28} and W\thinspace{44} and reported finding numerous maser features distributed across these SNRs. At their arcsecond resolution, some features were unresolved (``spots'') while other features appeared to be resolved with measured angular sizes ranging from 0.25'' to 2.5''. These measured sizes of the masers could, of course, be spots that appeared spatially blended due to insufficient angular resolution. Alternatively, the resolved features could be spots whose apparent sizes reflect scattering by interstellar turbulence along the line of sight (angular broadening). Many masers are known whose apparent sizes are dominated by angular broadening (e.g.,~ Diamond et~al. 1998; Frail et~al. 1994); most such masers are seen with similar sizes and elongations as noted by Claussen et~al. (1997). The current study was undertaken primarily to address the question of whether the measured size of the masers is due to scattering or to multiple maser components. In order to reach a definite conclusion on this question, we must investigate both the intrinsic size of the masers and the possible effects of interstellar scattering. The OH(1720 MHz) masers are quite rare in the interstellar medium as compared to main-line masers, and even more so toward late-type stars where the other satellite line at 1612 MHz is the dominant maser transition. So published sizes of OH(1720 MHz) masers, which necessitate the use of VLBI techniques are also very rare. Forster et~al. (1982) found upper limits of 20 mas to the sizes of OH(1720 MHz) masers toward the \hbox{H\hskip 2trueptII } region NGC~7538, while Masheder et~al. (1994) reported that the W3(OH) 1720 MHz masers are unresolved with sizes $<$1.2 mas. These measurements may not even be applicable in the case of supernova remnants where the physical conditions and pumping mechanisms could be very different from that in star-forming regions. A recent study of the pumping of the 1720 MHz masers toward supernova remnants (Lockett et~al. 1999) finds tight constraints on the physical conditions needed for their production (temperature in the range 50 --- 125 K, molecular hydrogen density $\sim$ 10$^5$ cm$^{-3}$, and OH column densities of order 10$^{16}$ cm$^{-2}$.) An upper limit for the size of the maser spots is the thickness of the shocked region over which such conditions exist. This thickness is estimated to be about 3$\times$10$^{15}$ cm. At the distance (3 kpc) of both W\thinspace{44} and W\thinspace{28} , this corresponds to an angular size of about 60 mas. The effects of interstellar scattering on the sizes of masers in W\thinspace{28} and W\thinspace{44} can be constrained by observations of nearby (in projection) pulsars and extragalactic sources. Interstellar scattering of an extragalactic source results in angular broadening while interstellar scattering of pulsars results in pulse broadening, an increase in the apparent width of a pulsar's average pulse profile beyond its intrinsic width. The degree of angular broadening for the masers and the extragalactic source and the degree of pulse broadening for the pulsar all depend in different ways upon the relative geometry of the observer, scattering material, and sources. Measurements of these three scattering effects can constrain the distribution of the scattering material and can be used to estimate, for example, the unscattered sizes of the masers. In this paper, we present VLBI observations of several bright OH(1720 MHz) masers in W\thinspace{28} and W\thinspace{44} using the VLBA of the NRAO. The large extent of the SNRs and associated maser emission (tens of arcminutes) precluded observations of all the OH masers for these remnants. In addition, we have used the MERLIN telescope of the Nuffield Radio Astronomy Laboratory at Jodrell Bank to observe the OH masers in W\thinspace{44} with intermediate angular resolution between that of the VLBA and the VLA. We present the results of these observations and discuss their impact on the question of the masers' intrinsic size vs. broadening due to interstellar scattering. \section{Observations and Data Reduction} \subsection{VLBA Observations} The 1720 MHz transition of the ground-state OH molecule was observed with the VLBA (Napier et~al. 1994) on 09 May 1997 toward one $\sim$ 25\char'023~ field in each of the two supernova remnants W\thinspace{28} and W\thinspace{44}. Table 1 lists the position of the pointing center for both sources. These positions were chosen to encompass both the ``OH E'' and ``OH F'' 1720 MHz masers in W\thinspace{28} and the ``OH E'' masers in W\thinspace{44}, following the nomenclature of Claussen et~al. (1997). A single antenna of the VLA was also used in conjunction with the antennas of the VLBA in order to provide short projected baselines ($\sim$60 km). The data were recorded with a 62.5 kHz bandwidth centered at velocities of 44.0~km~s$^{-1}$~ and 10.0~km~s$^{-1}$~ (LSR) for W\thinspace{44} and W\thinspace{28}, respectively. Both right and left circular polarizations were recorded with 2-bit sampling. The data were correlated with the NRAO VLBA correlator to provide 128 spectral channels for each polarization averaged every 4.1 seconds. All four polarization correlations were performed. This correlator mode provided a channel spacing of 0.09~km~s$^{-1}$~ per spectral channel. The velocity resolution is slightly large than this ($\sim$0.11~km~s$^{-1}$) due to the spectral weighting function applied. In order to provide manageable dataset sizes, the correlator averaging time was the limiting factor for the field of view. For averaging times of 4.1 seconds, the fringe-rate window for the longest baselines of the VLBA provides a field of view of approximately 24\char'175~. Thus for both SNRs we made two correlation passes near positions of strong OH(1720 MHz) masers in the primary field of view. Table 2 provides the positions of the two correlation positions for both remnants, and the peak flux density in the VLA {\bf A} configuration observations (Claussen et~al. 1997). Data reduction was performed using standard software contained in the NRAO AIPS package. Delays were measured via observations of nearby continuum sources (1748$-$253 for W\thinspace{28}, and 1904+013 for W\thinspace{44}) by performing fringe fitting. Bandpass calibration was determined using total-power observations of the strong sources 1921$-$293 and 1611+343. Residual fringe rates were determined by fringe fitting on a single strong spectral channel. For W\thinspace{28}, amplitude calibration was accomplished by fitting the bandpass-corrected, total-power, on-source spectra of each antenna in the array to a template total-power, on-source spectrum observed with a sensitive antenna (at the Mauna Kea, HI station) at a high elevation angle. The absolute amplitude calibration determined by this method is accurate to about 10\%. For W\thinspace{44}, the amplitude calibration was determined by using known antenna gains (provided by NRAO staff) and system temperatures measured during the observations (so-called {\it a priori} amplitude calibration). This procedure was carried out for the W\thinspace{44} data because the signal-to-noise ratio in the total-power spectra of the OH maser line was not high enough to attempt the template-fitting method. The absolute amplitude calibration determined by the {\it a priori} method is accurate to about 15\%. Even on the strongest spectral channels, no correlated signal was detected on the longer baselines (typically from one of the southwest locations to Saint Croix, VI; Hancock, NH; and Mauna Kea, HI), and so data to these stations were automatically deleted. The spectral channel with the greatest flux density was then used in an iterative self-calibration mapping procedure. These self-calibration solutions were then applied to all spectral channels. The rms noise ($\approx$ 100 mJy beam$^{-1}$) in these images is close to the expected theoretical noise limit. Images were then made of all spectral channels with maser emission. Naturally weighted maps for each velocity channel were then produced using the AIPS task IMAGR. The resultant synthesized beam was about 40 $\times$ 15 mas at a position angle of 0\ifmmode^\circ\else$^\circ$\fi\ for W\thinspace{28}, and about 40 $\times$ 30 mas at a position angle of $-$10\ifmmode^\circ\else$^\circ$\fi\ for W\thinspace{44}. \subsection{MERLIN Observations} The MERLIN radio telescope was used on 12 April 1998 to observe the position listed in Table 1 for W\thinspace{44} in the 1720 MHz transition of OH. Seven telescopes were used, including the 76-m Lovell telescope, the Mark II, the 32-m telescope at Cambridge, and the 25-m telescopes at Tabley, Darnhall, Knockin, and Defford. Both right and left-hand polarization was observed, and the correlator produced 512 spectral channels over a bandwidth of 500 kHz, for a channel spacing of 0.18 ~km~s$^{-1}$. The field of view for the MERLIN observations included both the W\thinspace{44} ``E'' and ``F'' sources (Claussen et~al. 1997). The synthesized beam from the MERLIN observations was 290 $\times$ 165 mas at a position angle of 25\ifmmode^\circ\else$^\circ$\fi\ . Initial phase and amplitude calibration was performed by MERLIN staff at the University of Manchester. Bandpass calibration was determined from observations of 3C~84, and phase calibration was determined from interleaved observations of 1904+013. The absolute amplitude calibration was determined by observations of 3C~286 on the shortest baselines. Similar to the VLBA reduction, AIPS was then used to apply self-calibration solutions, based on iterative self-calibration imaging carried out on the strongest spectral channel. The rms noise obtained after applying the self-calibration to a single channel was about 30 mJy beam$^{-1}$, also close to the theoretical noise limit. \section{Results} \subsection{W\thinspace{28} Masers} Figure 1 shows a contour image of the OH(1720 MHz) emission at 11.3~km~s$^{-1}$~, and Stokes {\it I} spectra at two emission peaks. This emission corresponds to feature F~39 in the nomenclature of Claussen et~al. (1997) (see their Table 2). The peak flux density in the image is 6.1 Jy beam$^{-1}$ with a total flux of $\sim$70 Jy, in close agreement with the peak flux density (73 Jy beam$^{-1}$) at this position in the VLA {\bf A} configuration maps. The OH(1720 MHz) emission is clearly resolved at this resolution. Several emission peaks can be seen. The size of the feature marked {\bf B} in Figure 1, determined by Gaussian fitting, is 75 $\times$ 34 mas, at a position angle of 9\ifmmode^\circ\else$^\circ$\fi\ . Thus the peak brightness temperature is $\sim$2 $\times$ 10$^9$ K. Other emission peaks in the contour image of Figure 1 have similar spectral profiles, differing primarily in their peak flux density. In addition to the masers shown in Figure 1, there is an additional maser emission peak $\sim$500 mas to the northeast, at a velocity of 9.6~km~s$^{-1}$~. Figure 2 shows a contour image of the peak velocity channel in this region and Stokes {\it I} spectra of two emission peaks. Again, the emission is quite extended (size $\sim$60 mas) compared with the beam. The two peaks {\bf C} and {\bf D} in the emission correspond to brightness temperatures of 1$\times$ 10$^9$ and 8 $\times$ 10$^8$ K, respectively. For the positions marked {\bf A} and {\bf B} in Figure 1, we have determined the Stokes parameters {\it I} and {\it V}. Assuming the {\it V} profile is due to the Zeeman effect, and that the splitting is small compared to the intrinsic (Doppler) line width, the Stokes {\it V} profile is proportional to the frequency derivative of the Stokes {\it I} profile. A fit of the {\it V} profile to the derivative of the {\it I} profile yields a measurement of the line-of-sight magnetic field (see Claussen et~al. 1997 for further discussion). Figure 3 shows the result of this fit for the two positions marked in Figure 1. The estimate of the line-of-sight magnetic field ($\sim$ 2 milliGauss) is larger by a factor of $\sim$10 than estimates based on the VLA observations. However, toward this specific position, Claussen et~al. (1997) were unable to estimate the line-of-sight magnetic field because the spectra were quite complex and did not show a clear signature (the classical S-shape in Stokes {\it V}) of Zeeman splitting. This relation used to derive the line-of-sight magnetic field is valid only for thermal absorption and emission lines. Nedoluha \& Watson (1992) conclude that the standard thermal relationship used here is a valid approximation of the line-of-sight field strength for observations of water masers, if they are not strongly saturated, despite the complications of the maser radiative transfer. Elitzur (1996, 1998) has derived a general polarization solution for maser emission and arbitrary Zeeman splitting. According to this solution, Elitzur (1998) concludes that masers require smaller fields to produce the same amount of circular polarization as thermal emission. Thus the estimate of the magnetic field given above may be overestimated by as much as factors of 2---4. At the other correlated position in W\thinspace{28} (W\thinspace{28} E), we did not detect OH emission. Based on the VLA observations, we expected that there should be several features detectable by the VLBA. Some of the flux densities measured in the VLA observations are 24 Jy/beam (E~30), 12.5 Jy/beam (E~31), and 10 Jy/beam (E~24). E~24 was {\it unresolved} with the VLA. If these non-detections are due to emission that is smooth, we can calculate a lower limit to the size of the emission features, based on the rms noise that we measured on the shortest projected baseline in the VLBA observations. Assuming a source size that is a circular Gaussian function, a maximum visibility of 0.1 (for E~24, for example) on the shortest baseline ($\sim$60 km), the lower limit must be about 400 mas. For E~30, the visibility must be correspondingly lower and thus the size must be larger than about 500 mas. \subsection{W\thinspace{44} Masers} The OH(1720 MHz) masers observed with MERLIN are shown in Figures 4 and 5. Figure 4 is a contour plot of the peak maser emission from the E~11 source along with the OH spectrum of the peak emission. The peak occurs at a velocity of 44.2~km~s$^{-1}$~. The stronger of the two features in the contour image has a peak flux density of 3.3 Jy beam$^{-1}$, and is slightly resolved with a fitted Gaussian size of 165 $\times$ 57 mas at a position angle of 147\ifmmode^\circ\else$^\circ$\fi\ . The total flux density over the emission region is about 6.1 Jy, which is comparable to the peak flux density measured with the VLA. The brightness temperature for this feature is 1.3 $\times$ 10$^8$ K. We convolved the MERLIN map of the E~11 source with the VLA {\bf A} configuration beam, and then made a Gaussian fit to the resulting image. The peak of the convolved image was 5.0 Jy beam$^{-1}$ with a fitted size of 715 $\times$ 250 mas at a position angle of 166\ifmmode^\circ\else$^\circ$\fi\ . This is comparable with the VLA observations which obtained a peak flux density of 6.6 Jy beam$^{-1}$, and a fitted size of 890 $\times$ 180 at a position angle of 137\ifmmode^\circ\else$^\circ$\fi\ . Figure 5 shows the OH(1720 MHz) maser emission from the W\thinspace{44} F~24 source which peaks at a velocity of 46.9~km~s$^{-1}$, with a peak flux density of 1.5 Jy beam$^{-1}$. The OH spectrum at the emission peak is also shown. The total flux density in the emission region is about 4.2 Jy, only about half of the VLA observed peak flux density. Figure 6 shows a contour plot of the VLBA image of the E~11 source. This maser source is unresolved at the resolution of the VLBA (40 $\times$ 30 mas). The peak flux density is 1.2 Jy beam$^{-1}$, and thus a lower limit to the brightness temperature is 8 $\times$ 10$^8$ K. This source is the core of the brightest MERLIN source shown in Figure 4. If a single Gaussian component with a large size were responsible for the difference in flux density between the VLBA and MERLIN measurements, then, based on the shortest projected spacing of the VLBA and the measured noise, the size of such feature would have to be larger than 270 mas. This is inconsistent with the MERLIN measurement. Thus we conclude that the 2.1 Jy of missing flux density must be in a few components whose peaks are each weaker than about 0.4 Jy beam$^{-1}$. \section{The OH Maser Sizes: Scattering Disks or Physical Size?} Table 3 summarizes the measured components and the estimated brightness temperatures of the OH(1720 MHz) masers in both W\thinspace{28} and W\thinspace{44}. The VLBA and MERLIN observations clearly resolve the masers seen by Claussen et al. (1997) into multiple components. Typical angular sizes in both SNRs are 50 to 100 mas with aspect ratios of order 2.5:1. While these measurements have shown that the Claussen et al. (1997) maser sizes were an artifact of the low resolution used, a major question remains: ``are these compact and resolved features the true size of the masers or are they due to interstellar scattering ?''. The observations and results of the present study cannot distinguish between these two options. This is due mainly to our ignorance of what the intrinsic sizes might be. In what follows we will interpret the results of these observations as they apply to both intrinsic structure and that due to scattering, and suggest further observational tests that should help to distinguish between these two possibilities. \subsection{Scattering Interpretation} If the maser sizes in W\thinspace{44} and W\thinspace{28} are dominated by scattering, other nearby (in projection) objects such as pulsars and extragalactic sources can also be affected. Pulse broadening and angular broadening depend in different ways upon the distribution of scattering material along the line of sight; additionally, angular broadening of an extragalactic source seen through the turbulence in our Galaxy's ISM is sensitive only to the strength of the turbulence, not to its distribution along the line of sight. The relative distances of the masers and nearby pulsars along with the distribution of scattering material can all be constrained using measurements of angular and pulse broadening. Adopting a distance of 3 kpc for both W\thinspace{28} and W\thinspace{44} , the models of Cordes et al. (1991) and Taylor \& Cordes (1993) predict angularly broadened sizes for the OH masers of order 1$-$3 mas. These angular broadening estimates are well below the sizes given in Table 3 but it should be noted that the angular broadening is underestimated for lines-of-sight subject to enhanced scattering. Examples of lines-of-sight with enhanced scattering include the Galactic Center region (van Langevelde et al. 1992), the Cygnus region (Fey, Spangler \& Cordes 1991), and towards 1849+005 (Fey, Spangler \& Cordes 1991). We discuss separately the implications of our maser size measurements for the scattering towards W\thinspace{28} and W\thinspace{44}. \subsubsection{Scattering in the Direction of W\thinspace{28} } The W\thinspace{28} SNR and its associated masers lie at a distance of approximately 3 kpc (e.g.,~ Kaspi et~al. 1993; Frail, Kulkarni, \& Vasisht 1993) in the direction $(l,b)=(6.8,-0.06)$. The 60,000 year old pulsar PSR~B1578$-$23 also lies in the same direction but is located outside the SNR, a few arcminutes to the north of its bright continuum edge. An extragalactic source, 1758$-$231, lies within two arcminutes of this pulsar. Frail et~al. (1993), using observations of the pulsar and the neighboring extragalactic source, argued in favor of the association of the pulsar and the SNR. Kaspi et al. (1993) disagreed, suggesting that the pulsar was much further away. The discussion of Frail et~al. was based upon a pulse broadening measurement of 70 milliseconds at 1~GHz for PSR~B1758$-$23 (Kaspi et~al. 1993) and upon an upper limit of $1$~arcsecond to the size of the extragalactic source 1758$-$231. Here we review the implications of our measurements of the $60$~mas maser size as it pertains to this disagreement. Under the usual assumption that masers and the extragalactic source are scattered by a turbulence confined to a single thin screen, angular broadening measurements completely constrain the location of the screen. The two angular broadening sizes are related to the location of the screen by $$d_s/d_m = 1-\theta_m/\theta_e ,$$ where $\theta_m$ and $\theta_e$ are the angular broadening sizes of the masers and the extragalactic source and $d_m$ and $d_s$ are the observer distances to the masers and the screen, respectively. If the masers and the extragalactic source have sizes of $\theta_e = 1$\arcsec and $\theta_m = 0.06$\arcsec, respectively, then for a maser distance of $d_m=3$~kpc, the scattering screen lies at $d_s = 2.8$~kpc. Note that the screen distance decreases if $\theta_m/\theta_e$ increases. The pulse broadening of PSR1758$-$231 and the angular broadening of 1758$-$231 provide a general constraint on the location of the pulsar. Frail et~al. (1993) assumed a distance of 3~kpc for PSR~B1758$-$23 and combined the pulse broadening of PSR~B1758$-$23 with the angular broadening of 1758$-$231 to constrain $f_p$, the ratio of the observer-screen distance to the screen-pulsar distance, to values between $0.3$ and $3.4$. A more general constraint on the location of the pulsar can be obtained if the equations presented by Frail et~al. are combined without assuming a distance to the pulsar to produce $$\theta_e = {0.42\over\sqrt{d_p/3}} \left(\sqrt{f_p}+{1\over\sqrt{f_p}}\right) ,$$ where $\theta_e$ is the angular size of the extragalactic source in arcseconds, and $d_p$ is the distance from the observer to the pulsar in kpc. It is easy to show that this implies that the distance to the pulsar is given by $$d_p = 2.12 / \theta_e^2 ,$$ if the screen is halfway to the pulsar and is further otherwise. In particular, if the measured extragalactic source size is smaller than $0.84$~arcseconds, the pulsar cannot be associated with the W\thinspace{28} SNR. \subsubsection{Scattering in the Direction of W\thinspace{44} } The W\thinspace{44} SNR and its associated masers lie at a distance of approximately $3$~kpc (Radhakrishnan et~al. 1972) in the direction $(l,b)=(34.7, -0.4)$. The association of pulsar PSR~B1853$+$01 with W\thinspace{44} is well established based on ages, dispersion measure distance, and positional coincidence (Wolszczan, Cordes, \& Dewey 1991). Assuming a $3$~kpc distance to PSR~B1853$+$01, the Taylor \& Cordes model predicts $30$~microseconds of pulse broadening and $1$~mas of angular broadening. Because the Taylor \& Cordes model prediction of $92$~mas severely underestimates the observed angular broadening size of $378$~mas for 1849+005, an extragalactic source situated only $3.7$ degrees away from W\thinspace{44}, we consider the possibility that the observed $100$~mas OH maser sizes are due to angular broadening. The observational limit to the pulse broadening is of order $1$ millisecond. Lines-of-sight with enhanced scattering are believed to intersect a clumped component of scattering material in the interstellar medium. All lines-of-sight passing within $30$~arcminutes of the Galactic Center are known to be heavily scattered (van Langevelde et~al. 1992; Lazio \& Cordes 1998). The scattering material has been constrained to lie within 50 parsecs of the Galactic Center. At a distance of $8.5$~kpc, a clump of size $30$~arcminutes corresponds to enhanced scattering in a region of about $40$~pc. Similar clump sizes have been proposed for scattering towards other lines-of-sight (e.g.,~ Dennison et~al. 1984). Because the OH masers in W\thinspace{44} and PSR~B1853$+$01 both lie at $3$~kpc, the distance $d_s$(in kpc) to an assumed scattering screen can be derived from the pulsar size formula presented by Frail et~al. : $$d_s = {3\over1+{\theta_m^2/2.52\tau}} ,$$ where $\theta_m$ is measured in arcseconds and $\tau$, the pulse broadening, is measured in seconds. If the angular broadening size for OH masers in W\thinspace{44} is assumed to be $80$~mas and the pulse broadening of order $0.5$~milliseconds, a clump of enhanced scattering would need to be placed only $0.5$~kpc away. At this distance, a $40$~pc clump would subtend an angle of about $4$ degrees and could also intercept the line of sight towards 1849+005. Since the line-of-sight towards 1849+005 is the second most heavily angularly broadened line of sight known, it is reasonable to suggest that other, neighboring, lines-of-sight should also be heavily scattered --- as is the case for the Galactic Center direction. The observed sizes of OH masers in W\thinspace{44}, if dominated by angular broadening, have sizes consistent with the scattering of 1849+005. This hypothesis predicts that other lines-of-sight close to 1849+005 should also be heavily scattered. In addition, improved better measurements of the pulse-broadening towards PSR~B1853$+$01 could help to prove or disprove this suggestion. \subsection{Intrinsic Structure of the Masers} If the structure that is observed in the OH(1720MHz) masers toward W\thinspace{44} and especially W\thinspace{28} are due to variations in emission intrinsic to the maser, then this data is the first demonstration of structure in 1720 MHz OH maser emission. The pumping requirements of these OH (1720 MHz) masers, modeled by Lockett et~al. (1999), strongly suggests an OH column density of 10$^{16}$ cm$^{-2}$ with molecular hydrogen density $\sim$ 10$^5$ cm$^{-3}$. This requires a linear dimension of ${10^{11}} \over {x_{OH}}$ cm, where $x_{OH}$ is the OH abundance. According to Lockett et~al. (1999), the highest OH abundance expected in C-shocks is $\sim 2\times10^{-5}$, so the expected thickness of the OH emitting region is about 5$\times10^{15}$ cm, similar to the maser sizes we observe. We could conclude that these masers appear to be similar to the stellar OH(1612 MHz) masers. The OH(1612 MHz) masers in circumstellar shells show a large range of size scales: 40 --- 1000 mas, as demonstrated by Bowers et~al. (1990). Both main-line and 1720 MHz OH found in the interstellar medium have been shown to be very compact and with little structure (e.g.,~ Reid et~al. 1980; Forster et~al. 1982; Masheder et~al. 1994). Bowers et~al. (1990) suggest that the OH(1612 MHz) maser structure is determined by a combination of density and velocity effects; our 1720 MHz observations could be indicative of a similar situation. A good test of whether or not the OH(1720 MHz) emission is really due to intrinsic structure would be a high-resolution observation of the OH(1720 MHz) emission toward a nearby SNR in the anti-center direction (to minimize possible scattering effects). A good candidate for this test observation would be the SNR IC~443. It lies at a Galactic longitude $\approx189$\ifmmode^\circ\else$^\circ$\fi\ , and is only 1.5 kpc distant. If a measurement of the size of the OH(1720 MHz) masers in IC~443 showed the masers to be smaller than those in W\thinspace{28} or W\thinspace{44}, then a good argument for scattering of the masers in W\thinspace{28} and W\thinspace{44} could be made. If the sizes of the masers in IC~443 were similar in size or larger than those in W\thinspace{28} or W\thinspace{44} , then the sizes of the masers in all three of the SNR would likely be intrinsic. \section {The OH Maser Polarization in W\thinspace{28}: Magnetic Fields} The measurement of the line-of-sight magnetic field of about 2 milliGauss toward the region of strong maser emission in W\thinspace{28} is stronger than the more widespread measurements reported by Claussen et~al. (1997) toward the SNR by about a factor of 10. This measurement is closer to that reported by Yusef-Zadeh et~al. (1996, 1998) for magnetic fields in the Galactic Center. It is interesting to note that the strongest field strength we measure is also in the region of strongest maser emission. This may be a selection effect, since our observations were limited to only a few maser regions. Since both the current observations and the VLA observations of {\it V} profiles show the classical S shapes of the Zeeman effect, we are confident that both sets of measurements are good estimates of the line-of-sight magnetic field. In this small region, the magnetic pressure must be 100 times the pressure estimated by Claussen et~al. (1997), or about 2$\times$10$^{-7}$ dyn cm$^{-2}$. Thus the magnetic pressure is very much larger than the thermal gas pressure of 6$\times$ 10$^{-10}$ dyn cm$^{-2}$ estimated from hot X-ray gas in the interior of the remnants (Rho et~al. 1996), and so the magnetic field is likely the dominant factor in the structure of the shock. As discussed by Lockett et~al. (1999) and Draine, Roberge, and Dalgarno (1983), the larger magnetic field estimated here is further strong evidence for a C-type shock in the OH maser region. \section{Conclusions} We have used the VLBA and MERLIN to observe some of the OH(1720 MHz) masers toward the two supernova remnants W\thinspace{44} and W\thinspace{28} at resolutions of 40 mas. We have resolved the masers in both SNRs. The range of observed sizes are 50$-$180 mas, and the derived apparent brightness temperatures are in the range 0.3$-$20$\times{10}^8$ K. Based on the present data, it is unclear if the observed structure of the masers is due to interstellar scattering or to intrinsic structure. If the OH(1720 MHz)structure is due to intrinsic maser emission, then we suggest that the OH(1720 MHz) masers in SNR may be similar to the OH(1612 MHz) maser emission from circumstellar shells. A possible test of whether or not the size and structures observed are intrinsic or due to scattering would be a high angular resolution observation of the remnant IC~443. If the sizes measured are considered to be dominated by the angular broadening effects of interstellar scattering, conclusions can be drawn about the location of the scattering material. In the case of W\thinspace{28}, the scattering material along the line of sight is most likely situated within $\sim100$~pc of the masers. Also, we would conclude that pulsar PSR~B1758$-$23 is definitely {\it not} associated with the SNR. In the case of W\thinspace{44}, we suggest that a 40~pc clump of enhanced scattering material located 500 pc from the Sun could explain the observed maser sizes as well as the scattering of 1849$+$005. This suggestion could be tested by searching for other heavily scattered sources within $\approx4$ degrees of 1849+005 and by improved pulse broadening measurements of the pulsar PSR~B1853$+$01. Finally, we measure a magnetic field in a small region of the SNR W\thinspace{28} ($\sim$2$\times$10$^{15}$ cm), which is a factor of about 10 higher than that which was measured using the VLA. The magnetic field clearly dominates the shock structure, and is further evidence for a C-type shock in the OH maser region. \acknowledgments The National Radio Astronomy Observatory is a facility of the National Science Foundation, operated under cooperative agreement by Associated Universities, Inc. MERLIN is a UK national facility operated by the University of Manchester on behalf of PPARC. We thank Peter Wilkinson for granting us time on MERLIN from the Director's discretion, Peter Thomasson for his invaluable assistance in scheduling MERLIN, and Anita Richards for performing the initial phase and amplitude calibration for the MERLIN data. Finally, we thank the referee, Moshe Elitzur, for useful comments and discussions which have improved the paper.
\section{Introduction} It was recently proposed that the large hierarchy between the weak and Planck scales arises because there exist $n$ extra compact spatial dimensions, within which only gravity, and not standard model particles and interactions, can propagate\cite{ADDone, ADDtwo, Nima}. In this framework, the Planck scale $M_{P}$ is not a fundamental scale of nature, but is rather an effective coupling related to $M$, the scale of $(4+n)$ dimensional gravity, by \begin{equation} M_{P}^{2}=4\pi r_{n}^{n}M^{2+n}, \label{eq:first} \end{equation} where $r_{n}$ is the radius of compactification of the $n$ extra dimensions\footnote{In this work we assume that the extra dimensions are compactified on an $n$ dimensional torus with a single radius. The scale $M$ defined in (\ref{eq:first}) is related to Newton's constant in $(4+n)$ dimensions according to $M^{2+n}=(2\pi)^{n}/S_{2+n} G^{-1}_{(4+n)}$, where $S_{k}$ is the surface area of a unit radius sphere in $k-1$ dimensions. This is the same definition of the gravitational scale used in several recent phenomenological studies \cite{{Nima},{peskin},{astro}}.}. Setting $M \sim {\rm TeV}$ transforms the hierarchy problem into the question of why the radii are large. The approximate values for $r_{n}$ obtained for $M \sim {\rm TeV}$ indicate that $n=1$ is ruled out immediately, while for the $n=2$ case, deviations from the standard force law may easily be detected by planned experiments sensitive to gravitational forces at distances of tens of microns \cite{submm}, depending on the precise value of $M$. The cases of higher $n$ can be tested instead at high energy colliders\footnote{In \cite{Nima} it is shown that if there exist gauge fields that propagate in the bulk, they can mediate long range forces relevant to sub-mm experiments, regardless of the number of extra dimensions. In this letter we restrict our attention to gravitational forces.}. Bringing the fundamental scale of gravity down near a TeV dramatically alters our view of the universe, and it is not a trivial matter that this picture is allowed experimentally. In \cite{Nima}, a diverse range of collider, astrophysical, and cosmological phenomena are examined to verify that the framework is in fact safe for all $n>1$. However, lower bounds on $M_{F}$ from rough estimates of both energy loss in stellar objects and cosmological constraints described in Section 2, cast uncertainty on whether these theories can be probed in future sub-mm gravity experiments, even for the $n=2$ case. In this letter we perform a detailed calculation of the most stringent cosmological constraints, and derive an upper bound on $r_{2}$ that is far below the anticipated range of these experiments. \section{Cosmology in Theories with Large Extra Dimensions} In standard cosmology, big bang nucleosynthesis (BBN) provides a detailed and accurate understanding of the observed light element abundances \cite{BBNrev}. In order not to lose this understanding in the context of theories with large extra dimensions, we must require that before the onset of BBN, the influence of the extra dimensions on the expansion of our 4D wall somehow becomes negligible. In particular we must imagine that starting at some ``normalcy temperature'' $T_{*}$, the extra dimensions are virtually empty of energy density and their radii are fixed. In \cite{Nima} it is suggested that the emptiness of the bulk can be explained if $T_{*}$ is the reheat temperature following inflation, and if the inflaton is localised on our 4D wall and decays only into wall states. What is the allowed range for $T_{*}$? We need $T_{*}>1 {\rm MeV}$ in order for ordinary BBN to be recovered. On the other hand, if $T_{*}$ is too large, then copious production of bulk gravitons by standard model particles can alter cosmology in unacceptable ways. The authors of \cite{Nima} perform rough estimates of several such effects and find that the most serious constraints come from overclosure of the universe by gravitons and contributions to the cosmic diffuse gamma (CDG) radiation from graviton decay. They estimate that these constraints require, for $n=2$, $M \lower.7ex\hbox{$\;\stackrel{\textstyle>}{\sim}\;$} 10{\rm TeV}$, even if the normalcy temperature is pushed down to $T_{*}\sim 1 {\rm MeV}$. As $M$ is raised to this level, it becomes unclear whether experiments probing macroscopic gravity at small distances will be sensitive to the extra dimensions, even if $n=2$. In light of the potential implications of cosmological constraints on planned experiments, it is worthwhile to calculate them more carefully. Detailed studies \cite{BBN1, BBN2} show that, in the early universe, the electron neutrinos decouple at 1.25 MeV, while the other flavors of neutrinos decouple at 2.15 MeV. From the results of \cite{BBN1}, one can deduce that at $T={\rm 1MeV}$, the relaxation time for muon and tau neutrinos is 10 times longer than the inverse Hubble rate of expansion. If the reheat temperature were less than an ${\rm MeV}$, the weak interactions would thus be unable to produce the thermal distribution of neutrinos required as an initial condition for standard BBN. For this reason we believe that by taking $T_{*}=1 {\rm MeV}$, we suppress the cosmological effects of the extra dimensions as much as is conceivably allowable, so that bounds we derive on $M$ by requiring $T_{*}>1 {\rm MeV}$ should be robust. We also present bounds obtained using the less conservative choice $T_{*}=2.15 {\rm MeV}$, which, given that this it is the decoupling temperature for two of the three neutrino species, may in fact be a more realistic value. We find that the strongest bounds on $M$ come from the CDG radiation, to which we dedicate the bulk of our analysis. \section{Calculation of the Diffuse Gamma Ray Background} To calculate the CDG background, we imagine that at the normalcy temperature $T_{*}$, the bulk is entirely empty, while standard model particles on our 4D wall assume thermal distributions. The KK excitations of the graviton are produced through the process $\nu \overline{\nu} \rightarrow G$, for example. The spin-summed amplitude squared for this process is\footnote{Feynman rules for the coupling of gravity to matter are derived in \cite{{giudice},{Han}}.} \begin{equation} \sum |{\cal M}|^2 = \frac{s^2}{4\overline{M}^{2}_{P}}, \end{equation} where $\overline{M}^{2}_{P}$ is the reduced Planck mass. The number density of mass $m$ KK states is then governed by the Boltzmann equation: \begin{eqnarray*} \dot{n}_{m}+3n_{m}H &=& \int \frac{d^{3}{\bf p}_{\nu}}{(2\pi)^3 2|{\bf p}_{\nu}|} \frac{d^{3}{\bf p}_{{\overline\nu}}}{(2\pi)^3 2|{\bf p}_{\overline{\nu}}|} \frac{d^{3}{\bf p}_{m}}{(2\pi)^3 2\sqrt{|{\bf p}_{m}|^2+m^2}}\\ &&\hspace{1in} \times (2\pi)^4 \delta^4(p_m - p_{\nu} - p_{{\overline \nu}}) \sum |{\cal M}|^2 e^{-\frac{|{\bf p}_{\nu}|}{T}} e^{-\frac{|{\bf p}_{\overline{\nu}}|}{T}}, \end{eqnarray*} and the integrations can be performed analytically to obtain \begin{equation} s\dot{Y}_m = \dot{n}_{m}+3n_{m}H = \frac{m^5 T}{128\pi^3\overline{M}^{2}_{P}}{\cal K}_1(\frac{m}{T}), \label{eq:prod} \end{equation} where ${\cal K}_1$ is a Bessel function of the second kind. We have applied entropy conservation to express the evolution in terms of the scaled number density $Y_m=n_{m}/s$, where $s$ is the entropy density. We will be interested in KK states that decay to photons in the MeV range, and from (\ref{eq:prod}) we see that essentially all of the graviton production occurs at temperatures near $m$ and thus at times well within the radiation dominated era. The neutrino temperature $T$ is therefore related to the time by \cite{Kolb} \begin{equation} t=1.5 g_{*}^{-1/2} \overline{M}_{P}T^{-2}, \end{equation} where, since we will be considering temperatures of order MeV and lower, $g_{*}=10.75$. Applying $s \propto T^3$ then leads to a present-day graviton density (neglecting decay) of \begin{equation} n_{0}^{(m)}=(2.3 \times 10^{-4}) \frac{m T_{0}^3}{\overline{M}_{P}}\int^{\infty}_{m/T_{*}}\!dx\, x^3 {\cal K}_1(x), \label{eq:gravdens} \end{equation} where the present day neutrino temperature is $T_{0}=1.96 {\rm K}$. A photon produced in the decay of a KK graviton of mass $m$ will have a detected energy that depends on the redshift, or equivalently, the time, at which the decay occured. Thus, the energy spectrum of photons produced in the decays of mass $m$ KK gravitons can be calculated using \begin{equation} \frac{dn_{\gamma}^{(m)}}{dE}=\frac{dn_{\gamma}^{(m)}}{dt}\frac{dt}{dz}\frac{dz}{dE}. \end{equation} The derivatives are evaluated by applying $E=\frac{m}{2}(1+z)^{-1}$, $t=t_{0}(1+z)^{-3/2}$, and $n_{\gamma}^{(m)}=2n_{0}^{(m)}\Gamma_{\gamma}/\Gamma_{T}(1-e^{-\Gamma_{T} t}$), where $\Gamma_{\gamma}$ is the decay width of the graviton into two photons, and $\Gamma_{T}$ is its total decay width. We use the time-redshift relation that holds for the matter-dominated era, because for KK gravitons that are produced near $T_{*}\sim 1{\rm MeV}$, and which decay into photons during the radiation-dominated era, the redshifted photon energies are far below the MeV range that interests us. The spectrum is evaluated to be \begin{equation} \frac{dn_{\gamma}^{(m)}}{dE} = 3n_{0}^{(m)}\Gamma_{\gamma} t_{0}(2/m)^{3/2}E^{1/2}e^{-\Gamma_{T}t_{0}(2E/m)^{3/2}}. \label{eq:spectrum} \end{equation} To calculate the full photon spectrum all that remains is to sum over KK modes. This is accomplished using the measure \begin{equation} dN=2 S_{n-1}\frac{\overline{M}_{P}^2}{M^{2+n}}m^{n-1}dm, \label{eq:measure} \end{equation} where $S_{n-1}=\frac{2\pi^{n/2}}{\Gamma(n/2)}$ is the surface area of a unit-radius sphere in $n$ dimensions. Using equations (\ref{eq:gravdens}) and (\ref{eq:spectrum}), and (\ref{eq:measure}), and the calculated width \begin{equation} \Gamma(G\rightarrow\gamma\gamma)=\frac{m^{3}}{80\pi\overline{M}_{P}^2 }, \label{eq:width} \end{equation} we obtain the spectrum \begin{equation} \frac{dn_{\gamma}}{dE}=(1.6 \times10^{-5}) S_{n-1}\frac{t_{0}T_{0}^{3}}{M^{2+n}\overline{M}_{P}}E^{1/2}f_{n}(E,T_{*}), \label{eq:fullspectrum} \end{equation} where the function $f_{n}(E,T_{*})$ is given by \begin{equation} f_{n}(E,T_{*})=\int^{\infty}_{2E}\!dm\, m^{n+3/2}e^{-\Gamma_{T}t_{0}(2E/m)^{3/2}} \int^{\infty}_{m/T_{*}}\!dx \,x^{3} {\cal K}_{1}(x). \label{eq:f} \end{equation} Numerically one finds $\Gamma (G \rightarrow \gamma \gamma) t_{0} \sim 3\times 10^{-7}(m/{\rm MeV})^{3}$, so that for the KK excitations that interest us, the graviton lifetime will be much longer that the lifetime of the universe. Even after considering other decay channels, we find that $\Gamma_{T}$ is so small that setting the exponential factor in (\ref{eq:f}) to unity does not significantly change the values of $f_{n}(E,T_{*})$. \begin{table} \begin{center} \begin{tabular}{|c|c|c|c|c|}\hline E(MeV)&$\frac{f_{2}(E,T_{*}=1 {\rm Mev)}}{{\rm MeV}^{9/2}}$&$\alpha_{2}(E)$&$\frac{f_{3}(E,T_{*}=1 {\rm Mev})}{({\rm MeV}^{11/2})}$&$\alpha_{3}(E)$\\ \hline 1&1456&$4.2\times 10^{4}$&9570&$.56$\\ 2&1228&$5.0\times 10^{4}$&8835&$.72$\\ 3&778&$4.0\times 10^{4}$&6571&$.66$\\ 4&379&$2.2\times 10^{4}$&3801&$.44$\\ 5&150&$9.8\times 10^{3}$&1773&$.22$\\ 6&51.1&$3.6\times 10^{3}$&698&$1.0 \times 10^{-1}$\\ 7&15.5&$1.2\times 10^{3}$&240&$3.6 \times 10^{-2}$\\ 8&4.27&$3.6\times 10^{2}$&74&$1.2 \times 10^{-2}$\\ 9&1.09&94&21.2&$3.6 \times 10^{-3}$\\ 10&.263&24&5.6&$1.0 \times 10^{-3}$\\ \hline \end{tabular} \caption{Values of the parameters $\alpha_{n}(E)$ and $f_{n}(E,T_{*}=1{\rm MeV})$ defined in equations (\ref{eq:f}) and(\ref{eq:alpha}).} \label{table:table} \end{center} \end{table} Taking $t_{0}=10^{10}$ years, we find that for $T_{*}=1 {\rm MeV}$, the spectrum can be written as \begin{eqnarray} \left.\frac{dn_{\gamma}}{dE}\right|_{T_{*}=1{\rm MeV}}& =& 4.6 \times 10^{-6(n-2)}S_{n-1}\frac{f_{n}(E,T_{*}=1 {\rm MeV})}{{\rm MeV}^{(n+5/2)}}\nonumber \\ &&\hspace{.5in}\times \left(\frac{E}{{\rm MeV}}\right)^{1/2}\left(\frac{M}{{\rm TeV}}\right)^{-(n+2)} {\rm MeV}^{-1}{\rm cm}^{-2}{\rm s}^{-1}{\rm ster}^{-1}\\ &\equiv&\alpha_{n}(E)\left(\frac{M}{{\rm TeV}}\right)^{-(n+2)}{\rm MeV}^{-1}{\rm cm}^{-2}{\rm s}^{-1}{\rm ster}^{-1}. \label{eq:alpha} \end{eqnarray} Values for $\alpha_{n}(E)$ and $f_{n}(E,T_{*}=1{\rm MeV})$ for $n=2,3$ are given in Table \ref{table:table}. The above photon spectrum was derived by calculating the density of KK gravitons produced by annihilation of a single neutrino species. Repeating the same calculation for $\gamma \gamma$ annihilation, we find a spin-summed amplitude squared \begin{equation} \sum |{\cal M}|^2 = 2 \frac{s^2}{\overline{M}^{2}_{P}}, \end{equation} and, taking into account the symmetry factor of $1/2$ due to the initial state photons, a contribution to the spectrum that is larger than that coming from a single neutrino flavor by a factor of 4. When comparing with the observed spectrum, we will take the sum of contributions from photons and three flavors of neutrinos\footnote{We neglect an additional contribution from $e^{+}e^{-}$ annihilation for the sake of a simplified calculation. Including this contribution enhances the bounds we derive only slightly.}: \begin{equation} \left.\frac{dn_{\gamma}}{dE}\right|_{T_{*}=1{\rm MeV}}=7\alpha_{n}(E)\left(\frac{M}{{\rm TeV}}\right)^{-(n+2)}{\rm MeV}^{-1}{\rm cm}^{-2}{\rm s}^{-1}{\rm ster}^{-1}. \label{eq:calc} \end{equation} Before comparing our results with the CDG background data, we can already obtain an independent bound on $M$ by requiring that the KK gravitons do not overclose the universe. Contributions from photon and neutrino annihilation give a graviton energy density \begin{equation} \rho_{G}=14S_{n-1}\frac{\overline{M}_{P}^2}{M^{2+n}}\int^{\infty}_{0}\!dm\, m^{n}n_{0}^{(m)}, \end{equation} where $n_{0}^{(m)}$ is the density defined in (\ref{eq:gravdens}). For $n=2$ we obtain \begin{equation} \rho_{G}=14\times 10^{-44} \left(\frac{M}{{\rm TeV}}\right)^{-4} {\rm GeV}^{4}, \end{equation} which, upon comparison with $\rho_{c}=8.1 h^{2} 10^{-47} {\rm GeV}^{4}$, leads to $M>6.5/\sqrt{h}{\rm TeV}$. Using the relation between the fundamental scale and the radius of compactification, \begin{equation} r_{n}=2\times 10^{31/n - 16}\left(\frac{1 {\rm TeV}}{M}\right)^{1+2/n}{\rm mm}, \label{eq:radius} \end{equation} we obtain \begin{equation} r_{2}<.015h {\rm mm}. \label{eq:critdens} \end{equation} It may be possible, although certainly challenging, to probe distances of this size in near-future sub-mm gravity experiments. If we take a less conservative bound on $T_{*}$ and instead use $T_{*}=2.15 {\rm MeV}$, the decoupling temperature for the muon and tau neutrinos, we get the more stringent bounds $M>13.9/\sqrt{h}{\rm TeV}$ and $r_{2}<3.3h\times 10^{-3} {\rm mm}$. Distances this small are likely not to be accessible to those experiments. \section{Comparison with Data} The CDG background has been measured recently in the 800 keV to 30 MeV energy range using the COMPTEL instrument\cite{kappadath}. The authors of \cite{kappadath} find that the photon spectrum is well described by the power-law function $A(E/E_{0})^{-a}$, with $a=-2.4 \pm .2$, $E_{0}=5{\rm MeV}$, and $A=(1.05 \pm 0.2) \times 10^{-4}{\rm MeV}^{-1}{\rm cm}^{-2}{\rm s}^{-1}{\rm ster}^{-1}$. They find no evidence for the ``MeV bump'' that was inferred from previous data. Using the COMPTEL results and the calculated contribution to the background from graviton decay in equation (\ref{eq:calc}), we can place a lower bound on the gravitational scale $M$: \begin{equation} \left( \frac{M}{{\rm TeV}}\right)^{n+2} > 7\alpha_{n}(E)\left( \frac{\left.\frac{dn_{\gamma}}{dE}\right|_{measured}}{{\rm MeV}^{-1}{\rm cm}^{-2}{\rm s}^{-1}{\rm ster}^{-1}} \right)^{-1}. \end{equation} We find that the most stringent bounds are obtained for $E \simeq 4{\rm MeV}$. Using the very conservative upperbound $\left.\frac{dn_{\gamma}}{dE}\right|_{measured} < 10^{-3}{\rm MeV}^{-1}{\rm cm}^{-2} {\rm s}^{-1}{\rm ster}^{-1}$ gives, for $n=2$, \begin{equation} M > 110 {\rm TeV}. \label{eq:mbound} \end{equation} This corresponds to a bound on the radius of compactification of \begin{equation} r_{2}< 5.1 \times 10^{-5} {\rm mm}, \end{equation} which is far smaller than the distances at which gravity can be probed in planned experiments. If we instead use $T_{*}=2.15 {\rm MeV}$, we obtain $M > 350 {\rm TeV}$: $M$ must be about $10^{3}$ or more larger than the electroweak VEV, reintroducing a mild hierarchy problem, and hence requiring supersymmetry or some other solution\footnote{The string scale may be lower than $M$, in which case the hierarchy is alleviated slightly. At least in the string scenario described in \cite{ADDtwo}, where standard model particles are localized on a 3-brane, the factor one might gain in this way is $\sim 10$ rather than $\sim 10^{3}$\cite{Nima}. If the standard model particles are instead localized on a brane of higher dimension, one can achieve further suppression of the string scale relative to $M$\cite{shiu}.}. Applying the same experimental bound to the $n=3$ case leads to $M > 5.0{\rm TeV}$ or $M > 13.8 {\rm TeV}$, for $T_{*}=1 {\rm MeV}$ and $T_{*}=2.15 {\rm MeV}$, respectively. \section{Cosmological Uncertainties} Are there ways to evade our bounds on $M$? The authors of \cite{Nima} have pointed out that there may be additional branes, besides our own, on which gravitons can decay. Depending on the decay rate on these branes, their existence can greatly reduce the number of gravitons that decay on our brane. If $1/\Gamma'$, the decay lifetime onto the other brane(s), is significantly longer than the age of the universe $t_{0}$, then the number of decays on our brane will not be substantially reduced. If $1/\Gamma' \ll t_{0}$, on the other hand, the number of decays on our brane, and thus the contribution to the photon background, is reduced by a factor $\sim 1/(\Gamma' t_{0})$. Moreover, in this case nearly all of the gravitons decay at large redshift, so that for $T_{*} \sim 1 {\rm MeV}$ the redshifted photon energies fall below the ${\rm MeV}$ range. We know of two scenarios that give the large $\Gamma'$ required to evade the CDG bound. In the first, $\Gamma'$ is large because the extra brane(s) have higher dimension than ours \cite{Nima}. If one of these so called ``fat-branes'' has thickness $W$ in a single extra dimension, the probability that a graviton will decay on it is enhanced over its probability of decaying on our brane by a factor $\sim WT_{*}$. For $WT_{*}\sim 5\times 10^{6}$, we find that the graviton contribution to the CDG is consistent with the COMPTEL result for $M$ as low as $\sim 1{\rm TeV}$. Taking $T_{*}=1{\rm MeV}$, this corresponds to a thickness $W>1 \mu {\rm m}$. Note that introducing a higher-dimensional brane does {\em not} enable us to evade the bound obtained by considering overclosure of the universe, equation (\ref{eq:critdens}). Because the fat-brane is higher-dimensional, the decay products have a momentum component that is perpendicular to our brane, and which therefore does not redshift (recall that the extra spatial dimensions are frozen). Thus the energy density of these decay products will go as $R^{-3}$ rather than $R^{-4}$, regardless of whether or not the particles are relativistic, and we cannot eliminate the graviton contribution to $\Omega$. In the second scenario, $\Gamma'$ is large because there exist a very large number of 4D branes in addition to our own. More precisely, we need at least $\sim 5\times 10^{6}$ additional branes to have a graviton contribution to the CDG background that is consistent with the COMPTEL result when $M \sim 1{\rm TeV}$. An important distinction between this scenario and the one involving higher dimensional branes is that now, provided the foreign branes are parallel to our own, relativistic decay products on them {\em do} redshift, and the bound in equation (\ref{eq:critdens}) can be evaded. \section{Conclusions} We have examined two cosmological constraints on the theories with large extra dimensions proposed in \cite{ADDone, ADDtwo, Nima}. To place limits on $M$, we apply a conservative lower bound on the normalcy temperature, $T_{*}>1{\rm MeV}$, as required by BBN. We find that, ignoring the possibile existence of additional branes, the radius of compactification of the extra dimensions for $n=2$ is bound by the cosmic diffuse gamma ray background, to be $r_{2}< 5.1 \times 10^{-5} {\rm mm}$, well beyond the reach of planned sub-mm gravity experiments. From the constraint that gravitons do not overclose the universe we derive a milder bound, $r_{2}<.015h {\rm mm}$, albeit one that is less dependent on our assumptions regarding foreign branes. If one instead insists on a normalcy temperature above the decoupling temperature for the muon and tau neutrinos, $T_{*}>2.15 {\rm MeV}$, these bounds become $r_{2}< 5.2 \times 10^{-6} {\rm mm}$ and $r_{2}<3.3h \times 10^{-3}{\rm mm}$, respectively. A recent calculation has given the bound $M>50 {\rm TeV}$ for $n=2$, from the requirement that supernovae do not cool too rapidly by graviton emission \cite{astro}. This astrophysical constraint complements the cosmological ones we have studied: it is subject to larger technical calculational uncertainties, while our analysis is subject to uncertainties in the global cosmological picture. In either case, a bound on $M$ can only be translated into a limit on $r_{n}$ if it is assumed that the extra dimensions have the same size. No matter how large $n$ is taken to be, it is always possible that one extra dimension has a size in the mm - $\mu$m range, while the others are much smaller\cite{dvali}. However, in a framework involving vastly different radii, we are unable to argue why gravity would be expected to diverge from $r^{-2}$ behavior specifically at those distance scales accessible to planned experiments. \\ \noindent {\em Acknowledgments}: We are grateful to Nima Arkani-Hamed for useful discussions. D.S. also thanks Michael Graesser for helpful conversations.
\section{Introduction} The use of light and lasers in medicine have increased manifold in the past decade. Enormous efforts have been devoted to the development of new diagnostic techniques such as NIR imaging \cite{tuchin} and flourescence spectroscopy of tissues \cite{kortum}, as well as therapeutic uses such as photothermal coagulation and Photo-dynamic therapy \cite{grossweiner}. Tissues are highly scattering media and an accurate knowledge of light transport parameters of the tissue is indispensible to describe the propagation of light in these media. Several techniques have been developed to measure the light transport parameters in tissues \cite{wilson}. Apart from these, the use of the Coherent Backscattered peak from tissues to estimate the transport parameters has also been suggested \cite{yoo1,yoo2,yoon}.\\ The phenomenon of Coherent Backscattering (CBS) of light by random media has attracted considerable attention since 1985 \cite{albada,wolf1}, when connection was first made between elastic multiple scattering and Anderson localization. This phenomenon also termed weak localization, shows up as a sharp peak in the backscattered direction within a narrow cone of angles. The angular width of the CBS peak is mainly determined by the transport length of light in the medium and the shape of the peak is slightly modified by the absorption. This technique was used to characterize the transport lengths of samples used in experiments on strong localization of light \cite{wiersma1}. Etemad {\it et. al.,}\cite{etemad1} were the first to study the effects of absorption in random media on the CBS peak.\\ There have been some attempts to study the CBS peak from biological tissues and obtain the transport properties \cite{yoo1,yoo2,yoon}. These studies, however, have been inaccurate because of a misinterpretation \cite{error} in applying Akkerman's expression \cite{akkermans} derived for conservative, semi-infinite media to absorptive media. Hence a systematic study of the CBS peaks from biological tissues has not yet been made. Since this is a non-contact method which can be used for online in-vivo measurements also, it is important that a systematic evaluation of this technique for the measurement of transport parameters be carried out. Eddowes {\it et. al.,}\cite{eddowes} have suggested that accurate absorption information may be obtained by using Monte-Carlo simulations to fit experimental measurements and have developed Monte-Carlo based routines to identify the optical coefficients for a given CBS angular profile. However the measurement of the absorption parameter is dependent on both the transport length and the limitations of the experimental setup, as will be explained, and accurate absorption information can be obtained in only limited ranges of the absorption and transport coefficients. In reports that have appeared so far, researchers have concentrated on obtaining the transport and inelastic lengths of tissues from the shape of the CBS angular profile. The possibility of using the intensity information for estimating the anisotropy factor($g$) of the scatterers has not been examined.\\ In this paper, we present the results of our investigations on the suitability of using CBS for measuring transport parameters of biological media. Milk of different concentrations have been used as tissue phantom. It is known that the light transport properties of milk closely resemble those of tissues and milk has been used to model light transport in brain tissue \cite{greenfield}, to study optical imaging \cite{morgan}, photon-density waves \cite{knuttel}, and for investigating the propagation of short laser pulses through a scattering medium \cite{tereshchenko}. To investigate the effects of absorption on CBS, an absorbing dye with known extinction coefficient was added to milk and the CBS from such media were studied. We also examine the possibility of determining the anisotropy factor ($g$) from the the single scattering contribution to the diffuse background.\\ \section{Coherent Backscattering of light} Coherent Backscattering (CBS) of light occurs in all disordered media and is the only major surviving interference effect. When a beam of light is incident on a random medium, there exist partial waves traversing every possible path in the medium. The CBS effect arises from the constructive interference of any partial wave with its time reversed counterpart in the medium. In exactly the backscattered direction, both these two waves have the same phase and constructive interference results. Away from the backscattered direction, the counterpropagating paths develop a phase difference depending on the relative positions of the first and last scattering events in the medium. For the ensemble of all possible light paths, these phases will randomize and the reflection is enhanced within a narrow cone in the backward direction with an angular width of the order of $\lambda / {l_t} $ where $\lambda$ is the wavelength of light and $l_t$ is the transport length in the medium. This peak shows up only after the ensemble averaging over the large scale sample specific fluctuations (speckle) that originate from the random medium \cite{etemad2}.\\ The CBS intensity can be described in terms of three contributions. The total normalized angular intensity is described \cite{wiersma2} by \begin{equation} I(\theta, L) = \frac{\gamma_c(\theta, L) + \gamma_l(\theta) + \gamma_s(\theta)}{\gamma_l(0) +\gamma_s(0)} \end{equation} where $\gamma_c$, $\gamma_l$ and $\gamma_s$ are the bistatic coefficients of the coherent, diffuse and single scattering contributions respectively. $\theta$ is the angle of scattering and $L$ represents the set of transport lengths in the medium. The set of relevant transport parameters are the mean scattering length($l_s = 1/\langle \rho \sigma_{s} \rangle$) defined as the reciprocal of the average of the product of the single particle scattering cross-section ($\sigma_s$) and the density of scatterers ($\rho $)(the probability of a photon remaining unscattered after traversing a distance z in the medium is $\exp (-z/l_s)$), the inelastic length ($l_i$) defined as the reciprocal of the absorption coefficient ($\alpha$) (transmitted intensity $I_{t}(z) = I_{0} \exp (-z/l_{i})$ in the absence of scattering), the anisotropy factor ($g= \langle \cos \theta \rangle$) which is defined as the average of the cosine of the scattering angle for a single scattering event and the mean transport length ($l_{t}$) which is a measure of the distance in which, the direction of the photon's motion becomes uncorrelated with its initial direction and is related as $l_{t} = l_{s}/(1 - g)$. The bistatic coefficients for the diffuse and coherent intensities are traditionally described by summing up the ladder diagrams and the most crossed diagrams respectively in a systematic perturbation of the intensity propagator \cite{mark}. The bistatic coefficient for the single scattering contribution for isotropic scatterers assuming normal incidence is given by \cite{ishimaru} \begin{eqnarray} \gamma_s(\theta) & = & \frac{4 \pi}{A} \int_{0}^{L} \exp\left[\frac{-x n_{0}( \sigma_{s}+\sigma_{a})}{\cos\theta}\right] \left( \frac{n_{0}\sigma_{s}}{4\pi} \right) \exp\left[ -x n_{0}(\sigma_{s}+\sigma_{a}) \right] dx\cdot A \nonumber \\ & = & \frac{a \cos\theta}{ 1 + cos{\theta}} \left\{ 1 - \exp\left[-b \left( 1 + \frac{1}{\cos{\theta}} \right)\right] \right\} \end{eqnarray} where $A$ is the area of the target, $n_{0}$ is the density of scatterers, $\sigma_s$ is the single particle total scattering cross-section, $\sigma_a$ is the absorption cross section, $a$ is the albedo defined as the ratio of the total scattering cross-section to the extinction cross-section[$\sigma_{s}/(\sigma_{s} +\sigma_{a})$] and $b ( = n_{0}(\sigma_{s}+\sigma_{a})L)$ is the optical thickness of the slab. However for the case of anisotropic scatterers, because it is the backscattering coefficient which would contribute to the measurement, $a$ would be modified as \begin{equation} a = \frac {\sigma_b}{\sigma_s + \sigma_a} \end{equation} where $\sigma_b$ is the single particle back-scattering coefficient.\\ When the single scattering contribution is supressed by using an isolator \cite{etemad2} and the helicity preserving channel for circularly polarized light is detected from time invariant random media, the enhancement is exactly 2 in the weak disorder regime \cite{wiersma2}. However if the detection is carried out in the linear polarization preserving channel, the enhancement factor will always be less than 2. This is due to the presence of single scattered events which do not contribute to the enhancement in the back-scattering. In earlier work \cite{yoo1,yoo2,yoon}, the contribution of single scattering was ignored. Though Eddowes {\it et. al.} \cite{eddowes} suggest that the experiment must be performed with circularly polarized light, it has not been examined whether the single scattered contribution can be used to estimate the anisotropy factor. The dependence of $a$ (normalized to unity for isotropic lossless scatterers) with the anisotropy factor $g$ (obtained by varying the scatterer size), is shown in figure-1. The cross-sections and the anisotropy factor were computed using Mie scattering theory for spherical particles and a program based on the BHMIE code given in \cite{bohren}. A value of 1.33 for the refractive index of the medium (water) and 1.41 for the refractive index of the particles (typical of tissues\cite{wilson}) is used. Note the presence of resonances for monodisperse particles in figure-1a. In figure-1b, the cross-sections and the $g$ factors were then averaged over a particle size dispersion of 30\% assuming a Gaussian distribution in order to mimic the experimental situation where the particles are polydisperse . Slightly different values of the refractive indices did not lead to much change, as the averaging process washed all resonances present. As one can notice, the single scattering contribution decreases sharply with the increase in anisotropy and becomes very small for large anisotropy factors. For $g\raisebox{-0.6 ex}{\raisebox{-0.6 ex}{$\stackrel{>}{\sim}$}} 0.6$, the single scattering contribution from lossless scatterers is barely $2 \%$ of the total back-scattered intensity and would hardly be measurable. This unfortunately implies that the anisotropy factor of biological tissues (typically in the range 0.8-0.95 \cite{wilson}) would hardly be measurable by this technique. \\ The CBS peak shape is reasonably well described in the diffusion approximation \cite{mark,akkermans}. In this work, the expressions derived in the diffusion approximation for an incident plane wave on a slab of isotropic scatterers and with a finite thickness, have been used \cite{mark} (See Appendix for the expressions). By replacing the mean scattering length ($l_{s}$) in these expressions by the mean transport length ($l_{t}$), these expressions have been shown to be valid for anisotropic scatterers as well\cite{akkermans2}. \\ It is essential to observe the effects of absorption on the CBS peak. The main effect of absorption is a rounding of the central cusp of the peak in an angular range $\theta_a \sim \lambda / \sqrt{l_{t} l_{i}}$. This corresponds to the extinction of the longer path lengths greater than $\sqrt{l_t l_i}$ due to the absorption. The other effect of absorption is to trivially reduce the enhancement factor because the single scattering events become relatively more important as absorption increases. The CBS peak is also rounded by the convolution with the instrumental response function and unless the angular range $\theta_a$ of the rounding due to absorption is larger than the instrumental resolution, the absorption would not be measurable accurately. For the case of $\theta_a = 0.18 mrad$ our experimental resolution, $\lambda = 0.6328 \mu m$ and $l_t = 670 \mu m$ a minimum value absorption coefficient $\alpha \sim 0.6 cm^{-1}$ can just be discerned. Though much smaller absorption coefficients in principle can be measured by improving the angular resolution, even assuming ideal optics, the resolution eventually is limited by diffraction due to the finite beam size. For our beam size of 5.5 mm, a beam divergence ($\theta_d = \lambda / \pi R$ where R is the beam radius) of about $75 \mu rad$ is present, yielding a minimum measurable $\alpha$ of about $0.1 cm^{-1}$ at a $l_t$ of $700 \mu m$. Trying to increase the beam size further is impractical due to problems of inhomogeneity of tissue samples. Hence accurate information would be possible only for samples with moderate absorption ($\alpha > 1 cm^{-1}$). However this is by no means a fundamental limit of the technique and depends on the transport length of the sample and the resolution of the experimental setup.\\ \section{Experimental setup and procedures} The experimental setup used is schematically depicted in figure-2. A 5mW He-Ne laser at 632.8nm was expanded to a beam diameter of 5.5 mm and collimated to diffraction limit. The collimation was checked by shear interferometry. Following the standard practice, the CBS light was viewed through a non-polarizing 50-50 beam-splitter with a small wedge. A computer controlled CCD array (752 X 244 pixels, model EDC-1000HR, Electrim corporation, U.S.A) was placed at the focal plane of a positive lens with a focal length of 20 cm, to analyze the angular peak. Each pixel on the CCD in this configuration corresponds to 0.06mrad. The laser beam was linearly polarized by placing a polarizer just before the beam-splitter. Another polarizer was placed in front of the CCD to record images in the polarization preserving channel. An aperture slightly larger than the input beam was placed just behind the beam-splitter to avoid ghost images from the AR coated surface of the beam-splitter. The free beam going through the beam-splitter was carefully damped using ND filters. \\ The setup was aligned by placing a mirror in place of the sample and making the reflected beam go back into the laser. The beam transmitted by the beamsplitter and focussed by the lens was scanned by the CCD at the focal plane and the intensity profile so obtained was used as the resolution curve characterizing the system response. This was well described by a gaussian having a FWHM of 0.17 mrad. The tissue phantom used was commercially available boiled and skimmed milk. The concentration of the milk was varied by adding distilled water. To prepare samples with different absorptions, the milk was doped with known concentrations of methylene blue dye. Methylene blue is a water soluble dye having a broad absorption band in the red region with two peaks at 610nm and 664nm The extinction coefficient of the methylene blue solutions which were added to the milk, was measured by a Shimadzu spectrophotometer. The dye-doped colloids thus had a well determined inelastic length. The milk did not change its scattering properties measurably in a time of about 3 hours. This was confirmed by recording the CBS profiles at different times and comparing them. All the experiments were hence carried out on the same milk sample within this time frame. The tissue phantom was taken in a cuvette of 10mm path length. It was placed slightly tilted to the incident beam so that the specular reflection was well away from the backscattered direction. The Brownian motion of the milk particles caused an ensemble average over the speckle and sharp symmetric peaks were observed. CBS intensities for samples of different scatterer concentration and absorption were recorded.\\ The CBS peaks from milk suspensions were very sharp (FWHM $\sim 1mrad$) and consequently affected by the finite instrumental response. Hence in order to fit the parameters, the theoretical profiles were first convolved with the system resolution curve using FFT routines. The resulting curve was then least squares fitted to the experimental points treating $l_t, l_i$ and $\gamma_s$ as parameters to be fitted using standard NAG library routines. Deconvolution of the experimental points was avoided as this was found to be a noisy process. \\ To supplement the CBS measurements, independent measurements of the transport parameters were made by measuring the transmission of light through milk solution as a function of the optical thickness or the concentration of milk. The milk was taken in a cuvette of 5mm path length. A 2mW He-Ne laser was used as the source. The light exiting at the back of the cuvette was imaged onto a photodiode, whose signal was detected by a lockin amplifier controlled by a computer. A separate photodiode was used to simultaneously record the laser intensity fluctuations. We have also undertaken angle-resolved scattering measurements for estimating the value of the anisotropy factor at the same wavelength. The sample was taken in a 1.0 mm cylinderical cuvette and is diluted enough, so as to get into the single scattering regime. The transmission of the sample was coupled to the PMT through an optical fibre kept on a rotating stage such that the distance from the sample to the fibre tip does not change and is read by a digital storage oscilloscope. A reference detector monitored the laser intensity fluctuations. Measurements were avoided in the lower angle region upto 8 degrees. \section{Results} Figure-3 shows the CBS profiles obtained from two different concentrations of milk. The solid lines shown are the best fits to the corresponding experimental points. As expected the CBS profiles became narrower with increasing transport length. The enhancement factor for the sample with longer transport length is considerably reduced due to the effect of convolution with the system response. The fits are quite good with a figure of merit ($\chi^{2}/N$ where N is the number of degrees of freedom for the fit) of about 1.2. The variation of the inverse of the transport length with the concentration is shown in the inset of figure-3. The $l_t^{-1}$ increases linearly with concentration as expected. The transport length is well defined and the numerical routines used quickly converge to the correct value to an accuracy of $\pm 5\%$ in a few iterations. However for the plain milk solutions any absorption coefficient less than 1.5$cm^{-1}$ could be fitted with marginal changes in the transport length. This is because the intensity measurement accuracy is limited to about $3\%$ by the CCD noise and the changes in the profile caused by small amounts of absorption are of the same level. \\ Figure-4 shows the CBS intensity obtained from dye-doped milk with a pre-determined absorption coefficient of 5 $cm^{-1}$. The fitted curve is for an absorption coefficient of $(5.2\pm 0.2)cm^{-1}$ and the elastic transport lengths are within reasonable variation for the pure milk ($l_t = 625 \pm 30 \mu m $) and the absorbing sample ($l_t = 670\pm 30\mu m$). At higher absorption, the entire peak appears broadened because the angular range of the rounded cusp is $\theta_a \sim 0.6 mrad$ while the peak width itself is of the same order. Also the height of the peak seems to be relatively unaffected by the absorption even though the weight of single scattering is enhanced. This counter-intuitive behaviour occurs because of the decreased effect of the instrumental response on the broader peak compared to that of the lower absorption narrow peak. This is clearly brought out in figure-5 where the theoretical shape (assuming same $\gamma_s$ as from experiment) for an infinite resolution is plotted. The high absorption peak clearly shows a lower enhancement. Thus the finite resolution and the relative weight of single scattering act in opposite directions on the enhancement factor with increasing absorption. The inset in Figure-4 shows the CBS profile and the theoretical fit from a sample with higher absorption ($\alpha = 10 cm^{-1} $). The backscattered intensity level was small necessitating large exposure times, increasing the CCD noise. However the theoretical fit again correctly yielded the absorption coefficient within an accuracy of $\pm10\%$. We studied the CBS in dye doped milk having absorption coefficients upto 30$cm^{- 1}$. However, the noise at these absorption levels was high (around 30 $\%$ of the peak intensity) and would appear to limit the use of this technique to measure larger absorption coefficients. Anyhow at higher levels of absorption, the expressions for the CBS peak shape derived within the diffusion approximation will no longer be valid. \\ Using the best fitted values of $\gamma_s$, the enhancement factors of the deconvolved curves if the plain milk-water solutions (no dye) are found to be in the range of 1.82 to 1.86. This reduction in the enhancement factor is lesser than the case of isotropic scatterers in a finite slab of same thickness with the same $l_t$ and $l_i$ for which the theoretical enhancement factors are between 1.76 to 1.80. These values are in agreement with those reported by Wolf {\it et. al.}\cite{wolf2}, but as observed by them the reduction in the enhancement factor appears much too large for such forward scattering media. Other processes such as recurrent multiple scattering \cite{wiersma2} which reduce the enhancement are not important at such large mean free paths present here. The lowering of the enhancement factor could also be caused by the finite beam effects and the gaussian intensity profile of the laser beam. However we do not believe this to be the reason as the beam size is reasonably large (5.5 mm FWHM $\sim 10 l_{t}$). Using the equation(60) of Jakeman\cite{jakeman} for Gaussian beams, we estimate that the reduction in the enhancement factor can only be in the range of 0.01 to 0.04 for the different $l_t$ used. In spite of our best efforts, higher enhancements were not observed. The CBS peak from a piece of white paper with a broad peak (FWHM of $10 mRad$) was observed to have an enhancement of only about 1.85 without deconvolution. The exact reason for this lower enhancement is not clear, but it could be due to some of the laser light being scattered by the non-ideal optical elements, cuvette walls, aggregates and dust which cause an additional component to the diffuse background. This would, however, be sample-independent. Assuming this additional contribution to be uniformly distributed, we have attempted to quantify it by studying the enhancement factor with respect to absorption in the sample. We conclude from this procedure that the ratio $\sigma_{b}/\sigma_{s} \raisebox{-0.6 ex}{\raisebox{-0.6 ex}{$\stackrel{<}{\sim}$}}0.1$ and that the g-factor of milk is greater than 0.55, which is what we expect from our theoretical considerations given earlier. \\ The transport properties can also be obtained from the plot of transmission versus the optical thickness \cite{ishimaru}. The results of the transmission measurements through milk-water suspension is shown in figure-6. At low densities (region A), the transmission falls exponentially depending on the extinction coefficient, the slope of the logarithmic transmittance being given by $\sigma_s +\sigma_a$. In the multiple scattering domain (region B), the transmission falls inversely with concentration or optical thickness. At large optical thickness or concentration (region C), the transmission falls exponentially depending on the absorption coefficient. The slope of the logarithmic transmittance is $\sigma_a$ in this region. These yield a value of the scattering length ($l_s$) of 130$\mu m$ and an absorption coefficient of $1.6cm^{-1}$ for milk at the largest concentration used. As it can be noticed the absorption coefficient lies just in the region where the CBS profiles become insensitive to absorption. Using a value of $l_t = 470 \mu m$ obtained from the CBS measurements for this concentration and $l_s = 130 \mu m$ from the transmission experiment, a value of $g=0.74$ is obtained which appears reasonable for such forward scattering media. The inset in figure-6 shows the angle-resolved scattering data, together with a theoretical fit to the Henyey-Greenstein function (solid line) with $g$ as the free parameter to be fitted. The fit yielded a value of about 0.70 for the anisotropy factor ($g $) with an error of about 5\%.\\ \section{Conclusions} In conclusion, we have investigated the application of Coherent backscattering to measure the light transport parameters in tissues using tissue phantoms (milk and dye doped milk), paying particular attention to the limitations of the technique. This method yields very good estimates of the transport length but it cannot be used to estimate the inelastic length when the absorption of the sample is small ($\alpha < 1.0 cm^{-1}$) for the long transport length typical of tissues. The maximum measurable absorption coefficient ($ \alpha \sim 30cm^{-1} $) with our setup appears to be limited by the sensitivity and noise levels of the detector used. It has been pointed out that in principle this technique can also be used to measure the anisotropy factor($g$) by estimating the single scattering contribution to the angle independent diffuse intensity. The single scattering contribution becomes very small for $g \raisebox{-0.6 ex}{\raisebox{-0.6 ex}{$\stackrel{>}{\sim}$}} 0.6$ ($\sim 2\%$ of the total backscattered intensity) and the technique becomes insensitive for larger values of $g$. Thus it does not seem to be a suitable technique to estimate the $g$-factor of biological media which typically are highly anisotropic ($g \sim 0.8$ to $0.95$). Using the value of the transport length obtained from CBS measurements and scattering length from transmission measurements, a value of 0.74 has been obtained for the anisotropy parameter of milk. This value is confirmed independently by angle-resolved scattering measurement where a value of 0.70 is obtained.\\ \section*{Appendix} For the convenience of the reader, the result of Ref.\cite{mark} for the angular shape of the CBS peak from a slab of finite thickness is reproduced here. \\ Let us define $\beta = \sqrt{ \left( l_{t} l_{i}/3 + q_{\perp}^{2} \right)}$, $q_{\perp} = (2 \pi/\lambda) \sin\theta$, $\delta = (2\pi/\lambda) (1- \cos\theta )$, $ \eta = (1+1/\cos \theta)/ 2l_{t}$ and $z_{0} = 0.71 l_{t}$, where $\theta$ is the backscattering angle, $\lambda $ is the wavelength of the light, $l_{t}$ is the mean transport length and $l_{i}$ is the inelastic length. Now the angular shape of the CBS peak in our notation is given by \begin{eqnarray} \gamma_{c} (\theta, l_{t}, l_{i}) & =& \frac{3 e^{-\eta L}}{2 l_{t}^{3} \beta \sinh [\beta (L+2z_{0})]} \frac{1}{(\eta^{2} + \delta^{2} - \beta^{2} )^{2} + (2 \beta \delta)^{2}} \nonumber\\ &\{ & \frac{2\beta}{\eta} (\beta^{2} + \delta^{2} - \eta^{2} ) \sinh [\beta (L+2z_{0})] \sinh (\eta L) + 2 (\eta^{2} + \delta^{2} + \beta^{2} ) \cos (\delta L) \nonumber\\ &+& 2 (\eta^{2} + \delta^{2} - \beta^{2} ) \cosh [\beta (L+2z_{0})] \cosh (\eta L) + 4 \beta \eta \sinh (\beta L) \sinh (\eta L) \nonumber\\ &-& 2 (\eta^{2} + \delta^{2} + \beta^{2} ) \cosh (\beta L) \cosh (\eta L) - 2 (\eta^{2} + \delta^{2} - \beta^{2} ) \cosh (2\eta z_{0}) \cos (\delta L) \nonumber\\ &-& \beta \eta \sinh (2 \beta z_{0}) \sin (\delta L) ~~\} , \end{eqnarray} where L is the thickness of the slab. It should be noted that the diffuse part of the backscattered light only has a weak kinematic dependence on the angle and the $\gamma_{l}(\theta,L) \simeq \gamma_{c}(0,L)$. In the exact backscattering direction $\gamma_{l}(0,L) = \gamma_{c}(0,L)$.\\ \section*{Acknowledgments} One of us (SAR) would like to sincerely acknowledge Diederick Wiersma (European Laboratory for Non-Linear Spectroscopy, Florence, Italy) for encouragement and for patiently answering all our questions over the e-mail. Both the authors thank P.K.Gupta for support and encouragement, and N.Ghosh for his help with the angle-resolved scattering measurements. \\
\section{Introduction and background} Mechanical description of static granular assemblies is an old but still open problem which recently got revisited from different perspectives \cite{PDM98}. In a classical and well established approach, practitioners of soil mechanics developed a conceptual framework in the spirit of a plasticity theory\cite{Schoefield68,Feda82}. This point of view based on an incremental description of stress-strain relations is suited in practice for numerical implementation. To close the description and introduce real granular effects such as dilatancy, Mohr-Coulomb plasticity etc.. , many rheological parameters have to be introduced on a semi-phenomenological basis\cite{Nedderman92}. Hence, the constitutive laws usually proposed are often complex and rather difficult to handle since they can be non-linear, non-differentiable and anisotropic \cite{Harris92,Norris97}. As a consequence, the outcomes usually based on numerical calculations and applied to complicated flow histories are hard to follow and explain physically. Clearly, from a fundamental point of view (maybe not on a practical point of view!), this approach is unsatisfactory and at the moment, there is a vivid interest in establishing a deeper understanding of the passage from a description of local granular contacts and force distributions, to a macroscopic description of stress-strain relations\cite{Cundall83,Goddard98}. A general feature observed both in experiments \cite{Dantu57,Travers86,Tsoungui98,Delyon90,Liu95} and in simulations \cite{Radjai96,Ouagenouni97,Radjai98,Eloy97} is the very heterogenous and anisotropic character of the force network arising from the intergranular contact geometry and the frictional properties of these forces. The difficulty of the problem is such that no rigorous macroscopic theory of granular mechanics is so far available. Moreover, how this disorder relates to the dispersion and the fluctuations of macroscopic stress measurements is an important issue that remains open \cite{Miller96,Ngadi98}. Recently, based on phenomenology and properties of symmetry, theoretical works have suggested new sets of closure relations for mechanical equations which aim to render the existence of long range anisotropic structures in static assemblies, commonly referred to as vaults. These relations contain a large set of possible histories for the granular material\cite{Bouchaud95,Edwards96}. Besides, the passage from a microscopic to a macroscopic description was undertaken using various models of force networks, the resolution of which is based on cellular automatons \cite{Coppersmith96,Claudin97,Hemmingsson97}, stochastic toy-models\cite{Eloy97,Socolar98} and analytical solutions of stochastic stress transport equations \cite{Claudin98a}. The case of a container filled with grains is a basic situation where the presence of boundaries is a natural initiator of a pressure screening effect that Janssen described\ last century in a pioneering contribution \cite{Janssen}. Quite surprisingly, the simple Janssen analysis is still used as a basis for the design of silos \cite{Schwab94,Thompson96} despite many difficulties in testing any models which come from a lack of reproducible experiments. Large fluctuations of the mean vertical pressure are generically observed in silos \cite{Brown70}. Data shows large differences for tests in identical conditions \cite{Shaxby23} and a strong dependency on the filling procedures \cite{Jotaki77}. The numerous theoretical refinements proposed to improve the Janssen analysis (see for example Ref. \cite{Nedderman92}) fail to take into account the existence of widely scattered pressure measurements and, in our opinion, this is a reason why an attempt to comparatively differentiate all theoretical predictions is most often doomed to failure \cite{Lenczner63}. In this work, we design an experimental set-up with the goal of precisely and reproducibly measuring the mean pressure as well as its fluctuations at the bottom of a granular column. We present two sets of well defined procedures which allow to obtain reproducible measurements. We quantitatively analyze these results in the form of an adaptation of the Janssen's classical model, and we discuss several theoretical models that could be used to compare our results with. \section{Experimental set up} \subsection{Apparatus description} A sketch of the experimental set-up is shown on Fig.1. A mass $M$ of grains is poured into a vertical cylinder of inner radius $R=20mm\ $and lies on the horizontal top surface of a piston. The piston is designed in such a way that it never touches the cylinder walls. Thereby, the average pressure of the granular material on the piston is entirely transmitted to an electronic scale. The scale is a horizontal beam of high effective stiffness $K=2.10^{4}N/m$, the deformation of which is detected with a strain gauge bridge which has the property to eliminate temperature drift effects in strain measurements. The corresponding force is measured in unit of mass with a precision $\Delta m=0.1g$ and it will be referred to as the \textit{apparent mass} $M_{a}$, since its value is different from the total mass $M$ of the granular medium. The system ''beam+piston'' rests on a mechanical elevator whose vertical displacement $\Delta Z$ can be varied and measured with a micrometric screw. \subsection{Filling procedure} We used this apparatus to measure the apparent masses of various species of grains such as monodisperse glass spheres, polydisperse iron shots or rugged quartz grains. All the results presented in this paper concern monodisperse glass spheres of diameter $d=2mm$. In order to vary the initial packing fraction, two filling methods are used (see sketch on Fig.1). In method $1,$ the grains fall from a hopper located on the top of the column and due to the ramming effect of falling grains a rather compact packing is obtained. In method $2,$ the grains fill first an intermediate inner cylinder that initially rests on the piston. Next, the inner cylinder is slowly removed so that the grains gently flow into the outer cylinder and settle rather loosely. \subsection{Volume fraction measurements} Throughout the experiments, the average volume fraction of the grains was estimated by monitoring the total height $H$ of the granular column. The top level of filling is detected using a digital vernier calliper (precision $0.01mm$). We checked that within a precision of one bead radius, the top surface was almost flat. This is probably a consequence of the rather small ratio between the column radius and the bead size : $R/d=10$ which prevents the development of a well defined slope at the free surface$.$ A thin and flat cardboard disc of mass $m=0.3g$ was used to obtain a well-defined top surface and reduce perturbations from the rod when coming in contact with the granular material. The bottom displacement of the column is known from the elevator's micrometric screw. The average granular density is therefore computed as $\rho=M/\pi R^{2}H,$ and the packing fraction is then $\nu$ $=\rho/\rho_{b}$, where $\rho_{b}$ is the grain density. The height measurement is always done after the corresponding apparent mass measurement since the latter is very sensitive to all kind of perturbations as we will see in the following. \section{Obtention of reproducible pressure measurements} \subsection{Measuring stresses in granular assemblies} There is a fundamental and technical difficulty in measuring stresses in granular assemblies. One physical reason is rooted in the hysteretic character of the local friction forces between the grains and with the system boundaries which prevent the system from returning to the initial equilibrium after a perturbation. We stress that it is not necessary to reach plastic deformation on a large scale in order to observe an irreversible change in the equilibrium state. For instance, a simple spring-mass system with friction displays hysteretic properties \cite{Duran98} and the coupled frictional and elastic contacts between grains are expected to display similar features \cite{Mindlin53}. Since every pressure probe is by construction associated with some displacement of a reference surface like a membrane or a piston, the measuring process may modify in return the force distribution in an irreversible way and therefore, the measurement values will depend on the specific perturbation history of both the probe and the material. One remedy could be in a quest for the least disturbing probe like a very stiff spring or a feedback device where position is controlled, but this might create other problems such as pathological coupling with temperature (due to thermal dilation\cite{Clement97}), a decrease in the sensitivity and/or incontrollable instabilities. This is exactly the kind of problem we address in this article. We consider the granular column and the pressure probe as a paradigmatic situation where all the difficult aspects of solid friction, boundaries and pressure measurement are \textit{a priori} coupled in a complex way. The basic idea of the measurement device we propose, and in particular the use of a mechanical elevator, is to produce a statistical distribution of the many metastable states corresponding to a specific average equilibrium with the piston. We want to know, for example, whether the dispersion and non-reproducibility of results currently reported in the literature is due to an intrinsically wide distribution of metastable states or if other ''hidden parameters'' should be considered to rationalize the results. In the following, we report two distinct ways of producing controlled and reproducible measurements, and furthermore, we describe basic observations that could explain why previous experiments have yielded non-reproducible data sets. \subsection{ Measurement procedures} Two different kinds of experimental procedures are investigated. Both lead to a set of reproducible results within a rather small fluctuation scale. They are both based on a slow downward displacement of the system ''beam+piston'' at a slow velocity (typically $20\mu m/s$). Note that importantly, the granular column seems to slide as a whole since we realize that any downward displacement of the bottom piston is also evidenced as a simultaneous downward displacement of the column's top. This procedure has \textit{a-priori} two remarkable advantages: (i) the kinetic energy of the falling grains during the filling stage which is partially stored as elastic energy of the beam, is relaxed, and therefore there is a memory loss of the initial pouring dynamics; (ii) the downward motion of the piston allows the granular column to slide down such that the friction forces at the walls should be fully mobilized and directed upwards. This is crucial for a quantitative analysis since generally, the theoretical models assume that the granular material is at the sliding limit at the walls. This important step was missing in previous experimental reports that we know. Generally speaking, the system would reach an equilibrium position depending specifically on the kinetic energy due to the pouring procedure, the effective distribution of friction forces at the boundaries, and the actual displacement of the pressure probe (related to the spring stiffness). Note that in a preliminary set of experiments\cite{Clement97}, we reported a large coupling of the pressure with small temperature variations. We linked this effect to problems of differential thermal dilation between different parts of the setup. Therefore, a double isothermal container was used to reduce the drastic effect of temperature on the plexiglass column as much as possible (this material has a large dilation coefficient and is a poor temperature conductor). In the new set-up, problems of temperature are dampened using a material with a very low temperature dilation coefficient such as an iron-nickel alloy. It turns out that this precaution is sufficient to obtain measurements independent of the actual temperature values. In fact we observe that after each downward displacement, there is a time gap of typically $30s$ which is large enough to relax our mechanical system and to make a measurement, but which is small enough to be able to neglect the effects of temperature drifts. \subsubsection{Descent experiments} The first kind of experiment probes the effect of accumulated series of downward displacements of the piston. It will be referred to as the \textit{descent experiment.} Note that in a preliminary report, the procedure has been presented as well as some of the effects we discuss in the following \cite{Vanel98}. The vertical displacements are performed with a fixed amplitude of $0.125mm$ ($1/16$\ of bead size) on a total distance of $20mm$ corresponding to one column radius. The displacement amplitude has been chosen for practical convenience, but we checked that changing its magnitude does not modify the reported behavior. Fig. 2a shows a typical evolution of the apparent mass both during the displacement of the piston and after the piston was stopped. We observe that when the piston is slowly moved down, the apparent mass abruptly decreases. The accelerated motion communicated to the system piston+scale during this process is responsible for a slight decompression of the scale spring which leads to this abrupt decrease. As soon as the elevator stops, the apparent mass suddenly increases up to a point where a slow relaxation of the apparent mass occurs and a stable value is finally reached. Since a decrease in the apparent mass implies an upward motion of the scale spring, it could be concluded that after the slow relaxation the friction forces are not fully mobilized at the walls.\ Nevertheless, as a practical statement, we note that the amplitude of the relaxation effect on the apparent mass values is less than $2\%$ so that the friction would be in any cases almost fully mobilized. Moreover, a decrease of the apparent mass means that the overall screening effect becomes stronger. This is in contradiction with a loss of friction mobilization and leads us to think that the small upward displacement of the piston is not really disturbing. One should also keep in mind that this displacement is less than $0.5\mu$m, which is smaller than the typical distance to mobilize friction \cite{Berthoud98}. In fact, the increase of the screening effect during this slow relaxation may correspond to a slow reorganization of the packing structure, but this point deserves a more thorough study. In this report, we only concentrate on the final values of the apparent mass as shown on Fig. 2a. Fig. 3 shows typical plots of the apparent mass as a function of the piston displacement for a filling mass $M=300g$. This filling corresponds to a height : $H\simeq8R$. Whatever filling method is used, there is a global increase of the apparent mass with the number of displacement steps and, eventually, it seems a steady state is reached. Around this global behavior, fluctuations of a smaller scale are evidenced. The average packing fraction changes with the number of displacement steps and also tends towards a steady state. When the initial packing is rather compact ($\nu\simeq0.63$), a global decompaction occurs, while there is a global compaction when the initial packing is rather loose ($\nu\simeq0.58$). Furthermore, whatever the initial packing fraction, the final packing fractions tend to be the same. The reproducibility of the method can be evidenced on Fig. 3a where the results of two independent experiments are superposed. \subsubsection{Tapping experiments} The second kind of experiment probes the effect of changing the granular density with the use of vibrations produced by series of taps on the container walls. It will be referred to as the \textit{tapping experiments}. Starting from a loose piling, previous studies have shown that series of taps induce a slow compaction effect\cite{Knight95}. Note that in this last reference, tapping was a vertical impulse; here, we use series of side impacts. The typical evolution of the apparent mass is shown on Fig. 2b. After a tap, there is a sudden increase of the apparent mass. The main reason for this increase is due to an hysteresis effect of the measurement device. A tap on the wall can either break the frictional contacts between the grains and the walls or make the piston vibrate. In both cases this should result in an additional downward compression of the piston. After a tap, the piston cannot decompress and return to its initial equilibrium because of the granular column frictional resistance to motion. In order to release the dynamical compression effect due to a tap, the piston is then slowly moved down and as a consequence, there is a fast decrease of the load to a level which is quite reproducible. A remarkable fact is that the final apparent mass (after a small downward displacement) is independent of the impact itself but is very well correlated to the average volume fraction, and this for the many independent experiments that we did. On Fig. 2b, we also display the mass variation obtained from the initial pouring procedure.\ The corresponding data points are seen on this figure for a time less than $50s$. We observe that the apparent mass indeed saturates at a value of about $80g$. Nevertheless, we tested that this value is difficult to reproduce from one filling to the other. On the other hand, after the first small and slow displacement $\Delta z$, we obtain a value $50\%$ smaller which is reproducible from one experiment to the other within a $5\%$ error bar! On Fig. 4, we display a typical plot of the apparent mass $M_{a}$ as a function of the packing fraction for a filling mass $M=300g$. Two independent sets of data are shown to illustrate the reproducibility of the variation as well as the level of fluctuations. The tapping procedure leads to a progressive compaction of the granular column while the apparent mass is decreasing. \section{Analysis of the Experimental results} \subsection{A phenomenological description and a qualitative interpretation} The results obtained in both descent and tapping experiments can be summarized on a unique diagram. Fig. 5a shows such a diagram where the apparent mass is plotted as a function of the packing fraction for a filling mass $M=300g.$ The arrows indicate the direction of evolution in the course of each experiment. A similar qualitative diagram is obtained with many other granular materials like polydisperse steel spheres or angular quartz grains, but we only report the results for $d=2mm$ glass spheres here. There is clearly a drastic change in the static equilibrium with the experimental procedure followed. Data for the\textit{\ tapping experiments} shows that a density increase of about $8\%$ induces a pressure decrease of about $20\%$ (Curve \{1\}). On the other hand, the descent experiments show that density variations of less than $5\%$ induce a pressure increase as large as $50\%$ (Curves $\{2\}$ and $\{3\}$). Moreover, this last procedure shows a compaction effect for initially loose packings and a decompaction effect for initially dense packing. Experiments on a 2D Schneebeli medium show than in a descent experiment like ours, friction at the wall produces shearing bands \cite{Pouliquen96}. In these experiments, the assembly was initially rather compact and due to shearing, a global decompaction occurred creating a radial density profile with a looser density on the edges and a larger density at the center. We suppose that in our descent experiments, a similar shearing effect occurs. Moreover, our results are consistent with the classical vision of soil mechanics where many standard tests have shown that shearing a dense medium produces a decompaction\cite{Reynolds} and shearing a loose medium produces a compaction\cite{Schoefield68}; this is what happens in curves $\{2\}$ and $\{3\}$. In the limit of large deformation a so-called ''critical density'', independent of the initial stage, is reached. This is probably what could happen around $\nu o\simeq0.593$ where both curves cross. In our case, it simply indicates that an identical density profile might be finally reached from both sides. On the other hand, what happens in $\left\{ 1\right\} $ is \textit{a priori} less clear. The way we understand these results is that the vibrations produce a compaction effect which ''kills'' the density gradients developed through the shear bands. After a \ tap, the displacement of the piston is small enough to avoid the formation of shear bands, and even if a shear band was initiated, the next vibrational shock is likely to destroy it. As a consequence, we have a more homogeneous granular packing along the radius and likely in the whole column. This interpretation is consistent with the fact that the end of curve $\left\{ 1\right\} $, corresponding to a sequence of tapping and descents, seems to join the beginning of curve $\left\{ 3\right\} $ which represents an early situation when the shear bands are not yet developed. In the tapping experiment, we observe a decrease of the apparent mass when the density is increased, which witnesses a stronger screening effect of the boundaries. Such a dependence between friction and density is reminiscent of recent experimental results \cite{Horwarth96} in which the extraction force of a rod buried in sand is shown to increase drastically with the packing fraction. At this point, to go beyond these many conjectures, more information on the local density spatial distribution would be needed. We are in the process of measuring this experimentally using an invasive technique. \subsection{Quantitative study of the pressure saturation} \subsubsection{Comparison with the classical Janssen model} In the following, we study the apparent mass as a function of the total mass poured in the cylinder. As a guide-line for describing the results, we present at first the Janssen classical analysis\cite{Janssen}. Last century, Janssen proposed a simple heuristic argument to account for pressure distribution $P(z)$ at the edges of a container filled with a granular material. The argument is based on mechanical equilibrium of a horizontal slice of thickness $dz$ and horizontal surface area $S=\pi R^{2}.$ The slice is submitted to the action of its own weight, a pressure gradient from the surrounding granular material and friction forces $dF_{frict}$ from the lateral walls such that:% \begin{equation} -\nabla_{z}P(z).Sdz+\rho gSdz-dF_{frict}=0\tag{1}\label{equ1}% \end{equation} where $\rho$ is the granular density. The core of the model assumes that the vertical pressure is transferred into a horizontal pressure via a constitutive Janssen coefficient $K$ and that the friction forces are \textit{fully mobilized} in the \textit{upward} direction. Thus the friction force acting on a surface element $dS=2\pi Rdz$ is: $dF_{frict}=\mu KP(z)2\pi Rdz,$ where $\mu$ is the solid friction coefficient of the grains with the wall. A rough measurement of the coefficient of friction between glass beads and the wall gives $\mu=0.4(\pm0.02)$. Integration of equation(\ref{equ1}) with the boundary condition $P(0)=0$ yields an exponential saturation of the pressure with depth $z$: $P(z)=\rho g\lambda(1-\exp(-z/\lambda))$. A central parameter in the theory is the characteristic length $\lambda$ which accounts for a pressure screening effect due to the boundaries: $\lambda=R/2\mu K$. In fact, the theory assumes that the stress distribution is uniform across any horizontal section of the material. This assumption turns out to be false and many theoretical refinements have aimed to correct it (see, for example, ref. \cite{Nedderman92}). However, the predicted pressure saturation curves are qualitatively very similar to each other. In the following, we only consider the simple saturation equation predicted by Janssen, except that it is expressed in unit of mass. The saturation mass, $M_{\infty}=\rho\pi R^{2}\lambda$, the mass $M$ of grains poured inside the column and the measured apparent mass $M_{a}$ should agree with the equation :% \begin{equation} M_{a}=M_{\infty}(1-\exp(-M/M_{\infty}))\tag{2}\label{equ2}% \end{equation} We experimentally tested this law by varying the mass of grains in the silo. To change the filling mass, the silo was entirely emptied and filled again with the desired amount of grains. Hence, data obtained for different filling mass corresponds to \textit{completely independent sets of experiments}. Since we have good reasons to believe that the packings obtained are more homogeneous, we now focus on the data obtained with the second measurement procedure, i.e. the \textit{tapping} \textit{experiments}. On Fig. 6, the apparent mass is plotted as a function of the filling mass for an average packing fraction $\nu=0.585\pm0.005$. Note that the straight line on the picture would represent a perfectly hydrostatic behavior. We clearly observe a saturation of the bottom pressure as the most elementary Janssen vision would predict. On the other hand, at a quantitative level, we evidence that the experimental data points are systematically \textit{above} Janssen equation (\ref{equ2}) when the fitting parameter $M_{\infty}$ takes the value of the experimental saturation mass (see dashed curve on Fig.6 obtained with: $M_{\infty}=53g$). This remains true for every saturation curve we obtained with other volume fractions. \subsubsection{Analysis using a modified Janssen model} To account for the experimental fact that the hydrostatic behavior is more pronounced than Janssen prediction, we propose a two-parameter model where is explicitly introduced a finite-size zone the apparent behavior of which is \textit{purely hydrostatic}. At the most elementary level, the hydrostatic zone is viewed as an horizontal slice of mass $M_{0}$ located at the bottom of the column, while the rest of the column is supposed to behave according to Janssen's differential equation (\ref{equ1}). The resulting equations are :% \begin{equation}% \begin{array} [c]{l}% -\text{for }M\leq M_{0},M_{a}=M,\\ -\text{for }M>M_{0},M_{a}=M_{0}+M_{\infty}^{C}\left[ 1-\exp\left( -(M-M_{0})/M_{\infty}^{C}\right) \right] \end{array} \tag{3}\label{equ3}% \end{equation} The thin solid curve on Fig. 6 is a best fit of the experimental data points with the previous equation (\ref{equ3}). The corresponding values of the fitting parameters are $M_{0}=13.4g$ et $M_{\infty}^{C}=39.6g$. Although the two-parameter model is based on a very crude assumption, the fit is quite good and clearly captures the main features of the experimental results. From the data obtained with the many \textit{tapping experiments} we performed, we measured the corresponding sets of parameters: $M_{0}$ and $M_{\infty}^{C}$, for other values of the packing fraction. The dependency of the fitting parameters with the packing fraction is shown on Fig. 7. We find that $M_{0}$ has only a slight tendency to decrease with $\nu$ while the parameter $M_{\infty}^{C}$ is clearly a decreasing function of the packing fraction. From the values of $M_{\infty}^{C}$ and the relationship $M_{\infty}^{C}% =\rho\pi R^{3}/2\mu K_{eff}^{C}$, an effective Janssen coefficient $K_{eff}^{C}$ is determined with the hypothesis that the coefficient of friction is independent of the packing fraction. As a comparison, we also determined the effective Janssen coefficient $K_{eff}$ obtained with the original Janssen model where the saturation mass $M_{\infty}$ is simply the sum of $M_{\infty}^{C}$ and $M_{0}.$ In both cases, the effective Janssen coefficients increase with $\nu$ (see Fig. 8). \section{Discussion} In the previous report we experimentally determined the shapes of the saturation curves for columns in static equilibrium with packings of different densities. We evidenced that the effective form of the saturation curve is strongly dependent on the density and as a rough statement, we find (i) that the effective screening length has a tendency to decrease when the density increases, and (ii) that the saturation curves stay very close to the hydrostatic curve for small fillings. We are aware that the use of a two-parameter model to fit the data is not really satisfactory. Nevertheless we believe that this investigation can be useful since any interpretative theoretical model should be consistent with these experimental facts and moreover, any theoretical saturation curve analyzed with our two-parameter model should provide relations identical to the one obtained on Fig. 8. However, from a classical perspective, we found \textit{a priori} surprising that $K_{eff}$ stays smaller than unity while $K_{eff}^{C}$ is almost always larger than unity. In the engineering literature\cite{Nedderman92}, it would be said that the granular material is in an active state when $K<1$ and in a passive state when $K>1.$ The real physical meaning of this classification is usually related to the mobilization of friction forces \textit{inside the granular material} and, \textit{a priori}, it is not obvious to which case our experiments should correspond to. Recently a theoretical prediction was made on the existence of a pronounced hydrostatic zone for a column filled with grains of finite elasticity \cite{deGennes98}. According to this theory, there is a zone at the top and at the bottom of the column where the grain displacements relative to the walls are much smaller than the minimum distance necessary to mobilize friction. Here we are under the impression that our descent procedure should prevent such a condition to occur. Another physical explanation for the hydrostatic zone could be the following. Let us assume that the stresses actually propagate along specific paths. The qualitative argument is that the weight of grains close to the piston can be carried along stress paths that will never reach the walls but instead, directly hit the piston. Then some of the mass cannot be screened by the walls so that the real saturation curve should be closer to the hydrostatic curve as compared to a simple Janssen law. Several models actually predicts that the stresses propagate along specific directions. This is the case of a model Bouchaud et al \cite{Bouchaud95} have proposed recently, but it is also a consequence of the classical I. F. E. hypothesis (Incipient Failure Everywhere). In addition, Socolar has found using a stochastic toy-model that the average vertical force in a silo is closer to the hydrostatic curve than the Janssen prediction. The curve predicted by Socolar in Fig. 4 of ref. \cite{Socolar98} and our experimental curve in Fig. 6 are very similar. Socolar also observed the emergence of long stress chains which are reminiscent of experimental observations \cite{Dantu57}. Those stress chains represent stress paths along which large values of stress are carried. Thus, the qualitative interpretation we described above seem to be supported by Socolar's results. The question arises whether other theories which don't predict the existence of stress chains can also describe the pronounced hydrostatic behavior we observed experimentally. A precise quantitative comparison between various theoretical models and our experimental data is the aim of a subsequent paper \cite{Claudin98b}. \section{Conclusion} In this paper we investigate the fluctuations and variations of the average pressure at the bottom of a granular column. First we design an apparatus and two experimental procedures which allow to obtain reproducible measurements. For both procedures, we are able to separate a level of fluctuations and a level of systematic variation which are very well correlated with the average packing fraction. One of these operating modes is thought to provide information on rather homogeneous packings and we perform a quantitative analysis of the pressure saturation curves for different packing fractions. We find that the data points are systematically above the simplest prediction one can make using a Janssen model. We make a simple extension of this model and introduce an effective hydrostatic zone at the bottom of the column as a new fitting parameter. The agreement with the experimental results is quite good and we find that the Janssen constant which is the second fitting parameter, shows a systematic increase with the packing fraction. On the other hand, the bottom hydrostatic mass shows a weak variation within the experimental errors. The second experimental procedure is thought to achieve a situation where the downward motion of the grains and the friction with the boundaries produce localized shear bands. Therefore a systematic variation of the density is expected along the column radius, i.e. the packing fraction at the walls is less than at the center. Although this density variation is very small, the effect on the mechanical equilibrium is drastic since the apparent saturation mass shows a systematic increase and reaches values which are twice as large as the one obtained with an homogeneous column. We are aware that this series of experiments and the interpretation we propose, though encouraging, are preliminary. The fact that we find reproducible data sets calls for more experimental work with other materials, other boundary conditions and larger column sizes. Moreover, we need a systematic series of tests in order to discriminate clearly between all different theoretical approaches and possibly reach a well established vision of this elementary but still unraveled problem of the static equilibrium of a granular assembly in a column\cite{Claudin98b}.
\section{Introduction} The history of the study of M~82 is a long and rich one. M~82 is the prototypical starburst galaxy and since it is one of the nearest (D = 3.63 Mpc - \cite{F94}) and brightest members of that class, it has been studied at almost every wavelength ranging from the $\gamma$-rays through to low frequency radio waves (e.g., \cite{B94}; \cite{BST95}; \cite{C90};\cite{LS63}; \cite{SBH98}; \cite{L94}; \cite{P94}; \cite{SKB94}; \cite{H94}; \cite{R94}; \cite{ML97}; \cite{P97}; \cite{SPS97}). It is perhaps impossible for any paper, let alone one this brief, to do justice to this rich history. We refer the reader to excellent exposes on M~82 in, for example, \cite{BT88} and \cite{T88}. M~82 is also the prototype of the ``superwind'' phenomenon (\cite{AT78}; \cite{MHvB87}; \cite{BT88}; \cite{HAM90}; \cite{BST95}; \cite{SPS97}; \cite{SBH98}). These flows are powered by the collective kinetic energy and momentum injected by stellar winds and supernovae in starburst galaxies. While recent studies have established the ubiquity of superwinds in local starburst galaxies (e.g., \cite{LH96}; \cite{DWH98}), there is still only a rough understanding of either their dynamics or the processes responsible for the observed optical and X-ray emission (e.g., \cite{Str98} and references therein). Without such understanding, it is difficult to quantitatively assess the role that superwinds have played in the formation and evolution of galaxies (\cite{KC98}; \cite{SPF98}) and in the chemical enrichment, heating, and possible magnetization of the inter-galactic medium (\cite{GLM97}; \cite{PCN98}; \cite{KLH99}). In this paper we discuss the optical and X-ray properties of the spatially-correlated region of X-ray and H$\alpha$ line emission recently reported by \cite{DB99} (DB hereafter) to be located roughly 11 kpc above the disk of M~82, far beyond the region where the superwind has been previously investigated. We are motivated by the important implications of understanding such a region for the superwind theory. If this region could be shown to be physically related to the superwind that M~82 is driving, then it implies that winds driven even by energetically-modest starbursts like M~82 are able to reach large distances into the halo, making them more likely to be able to escape the gravitational potential of the galaxy and thereby supply mass, metal, and energy to the inter-galactic medium (IGM). Such a finding would have implications for our understanding of the evolution of the IGM, the halos of galaxies, and quasar absorption line systems. Our study of this emission region also provides some unique clues about the physics that underlies the observed X-ray and optical emission from superwinds, and strongly suggests that wind-cloud collisions in starburst galaxy halos are an important emission process. \section{Observations and Data Reduction} \subsection{X-ray Data} The ROSAT PSPC data were obtained during the period from March through October 1991 and had a total integration time of about 25 ksec. M~82 was placed approximately in the center of the field for all the observations, and the source size is small enough so that there is no obscuration of the X-ray emission from M~82 due to the mirror ribs. We extracted a spectrum from a region of diffuse X-ray emission located about 11 arc minutes north of M~82, and coincident with the region of H$\alpha$ emission reported by DB. We will subsequently refer to this region as the ``ridge'' of emission. This area of X-ray and optical emission is oriented roughly parallel to the major axis of M~82, and was denoted as sources 8 and 9 in \cite{DWH98}. We extracted source counts in an elliptical region with a major (minor) axis of 3.3 (1.8) arc minutes oriented in a position angle of 43$^\circ$. This region is just sufficient to encompass all of the obvious X-ray emission, but excludes the soft point source at the ENE end of the ridge of emission that is discussed in \S 3.2 below. We have also extracted source counts for the portion of the ridge that is contaminated by this point source using a circular aperture with a diameter of 1.75 arcmin. The background for these spectra was taken from a large region where there were no apparent background sources to the south and west of M~82. The relative distance of the background region from the center of the PSPC field was approximately the same as that for the ridge of emission. The spectrum was then grouped such that each channel contained at least 25 counts. The extracted background subtracted data were subsequently fitted with various models using the package XSPEC. We have also inspected the archival ROSAT HRI data for M~82. These data have lower sensitivity than the PSPC data, and provide no spectral information. However, they do provide significantly higher spatial resolution ($\sim$6 arcsec FWHM on-axis and $\sim$15 arcsec at the position of the X-ray ridge). The data we have used comprise a 53 ksec exposure taken in April and May 1995. We have used these data only to study the morphology of the X-ray ridge. \subsection{Optical Imaging} Our optical images of M~82 were taken on the night of February 27, 1995 (UT) using the KPNO 0.9m telescope. We observed with a Tek 2048$^2$ CCD and through the optics of 0.9m, this yielded a scale of 0.68 arc seconds pixel$^{-1}$ and a total field of view of over 23 arc minutes. The night was not photometric, with cirrus around all night and intermittent patches of thicker clouds. M~82 was observed through two filters, a narrow-band filter with a full width at half maximum of 29\AA \ and a central wavelength of 6562\AA \ and a wider filter with a FWHM of 85\AA \ and a central wavelength of 6658\AA. The first filter effectively isolated H$\alpha$ and underlying continuum and did not include emission from either [NII] lines at 6548 and 6584\AA. The second filter provided a line-free measure of the continuum and was used for continuum subtraction of the first narrow-band filter to produce a H$\alpha$ line-only image of M~82. The total integration time in each filter was 1200 seconds. The images were reduced in the usual way (bias subtracted, flat-fielded, and then the narrow-band continuum image was aligned and scaled to subtract the continuum from the narrow-band H$\alpha$ image) using the reduction package IRAF\footnote{2}{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation}. Since the data were taken under non-photometric conditions, fluxing the data directly using observations of spectrophotometric standards is impossible. However, obtaining fluxes and hence luminosities is very important to our analysis. Therefore, to estimate the H$\alpha$ flux and continuum flux-density of the off-band filter of the continuum and emission-line nebulae of M~82, we bootstrapped from the flux given in a 5.8 arc second circular aperture centered on knot C from \cite{OM78}. Knot C was chosen because it was easily identifiable in our images and it appears in a relatively uncomplicated region of line emission. This procedure yields a H$\alpha$ flux from M~82 of about 4.6 $\times$ 10$^{-11}$ ergs cm$^{-2}$ s$^{-1}$, which compares very well with the H$\alpha$ flux from M~82 estimated by \cite{MHvB87} of 4.5$\times$10$^{-11}$ ergs cm$^{-2}$ s$^{-1}$ over a similar region 90" $\times$ 90". All subsequent results for the H$\alpha$ flux and luminosity of various features within the H$\alpha$ image of M~82 will rely on this bootstrapping procedure. To estimate the continuum flux-density we used the flux-density of O'Connell \& Mangano taken at 6560\AA. Since the center of our narrow-band continuum emission is 6658\AA, we are thus assuming that the continuum has a constant flux-density between 6560\AA \ and 6658\AA. To see if this extrapolation procedure is accurate, we compared the flux in R of M~82 in a 35 arc second aperture of 1.97$\times$ 10$^{-13}$ ergs cm$^{-2}$ s$^{-1}$\AA$^{-1}$ (from NED based on data originally obtained by \cite{J66}). If we also take the total counts in the 6658\AA \ filter and determine the flux-density by extrapolating the conversion factor from measuring knot C of O'Connell \& Mangano, we find a flux-density of 2.11 $\times$ 10$^{-13}$ ergs cm$^{-2}$ s$^{-1}$ \AA$^{-1}$ in a 35 arc second aperture. Our extrapolated flux is within 10\% (0.1 magnitudes) and thus this extrapolation is reasonably accurate. \section{Results} \subsection{Optical Imaging} The H$\alpha$ and narrow-band continuum images of M~82 are quite spectacular (Figures 1 and 2). We find that the line emission is very extended, with discernible H$\alpha$ line emission extending to about 11 kpc above the plane of M~82. This very faint emission can be seen even more clearly in the deeper H$\alpha$ image obtained by DB. Like previous authors (e.g., \cite{BT88}; \cite{MHvB87}), we find that the H$\alpha$ line emission has a very complex morphology that is very suggestive of outflowing gas, with loops, tendrils, and filaments of line emission pointing out of the plane of the galaxy. An especially interesting feature reported by DB is the faint ridge of emission about 11 arc minutes north of M~82 at a position angle of about $-$20$^\circ$ from the nucleus. This ridge is then about 11.6 kpc (in projection) north of M~82 at the adopted distance of 3.63 Mpc (\cite{F94}). The dimensions of the ridge down to surface brightnesses of about 3.5 $\times$ 10$^{-17}$ ergs s$^{-1}$ cm $^{-2}$ arcsec$^{-2}$ are about 3.5 arc minutes long $\times$ 0.9 arc minutes wide, with the long axis being approximately parallel to the major axis of M~82. At the assumed distance of M~82, this size translates into approximately 3.7 kpc $\times$ 0.9 kpc and the total H$\alpha$ flux from this ridge of emission is 1.5 $\times$ 10$^{-13}$ ergs s$^{-1}$ cm $^{-2}$. This flux is reasonably close to the estimate in DB of 1.2 $\times$ 10$^{-13}$ ergs s$^{-1}$ cm$^{-2}$, and yields an H$\alpha$ luminosity of 2.4$\times$ 10$^{38}$ ergs s$^{-1}$. The total H$\alpha$ flux for the whole of M~82 is approximately 4.6 $\times$ 10$^{-11}$ ergs s$^{-1}$ cm $^{-2}$. Thus, the ridge of emission constitutes approximately 0.3\% of the total H$\alpha$ emission from M~82 (not corrected for internal extinction). While the ridge appears rather structureless in our H$\alpha$ image, the deeper image published by DB shows a collection of bright knots and loops or filaments connected to and immersed in the more diffuse and pervasive emission. The narrow-band line-free {\it continuum} image reveals a complex morphology that is as interesting as the H$\alpha$ image. Besides the obvious complex dust absorption that is seen in the image (with some dark features oriented perpendicular to the plane of the galaxy), we also see filaments of continuum emission extended out of the plane of the galaxy (although the filaments are not entirely obvious in Figure 2). These filaments have twists and turns as seen projected on the sky, but generally point perpendicular to the plane of the galaxy. They extend outwards almost 3 arc minutes to the northwest and almost 2.5 arc minutes to the southeast -- corresponding to a projected distance out of the plane of about 3 kpc. Even though the continuum emission is in many ways as complicated as that seen in the H$\alpha$ image, we do not observe any continuum emission from the ridge of spatially coincident H$\alpha$ -- X-ray emission. Using the counts to flux-density conversion for the narrow-band line-free continuum image and the same aperture as used to measure the H$\alpha$ flux (but with the point sources removed from the area), we find an upper limit for the possible continuum emission from the ridge of fainter than 15.8 R magnitudes. This then implies an absolute R magnitude fainter than $-$12.0. While we find no evidence for any diffuse, spatially-extended optical continuum associated with the H$\alpha$ ridge, we do find a relatively bright point source in the line-free image that is coincident with the X-ray point source seen at the ENE end of the ridge in the ROSAT HRI data. This point source has an R magnitude of 12.2. We will discuss its likely nature in the next section. \subsection{X-ray Imaging and Spectroscopy} The optical images described above were aligned with the X-ray images starting with the coordinates given for the X-ray data from ROSAT and using the positions of stars within the frame from the Guide Star Catalogue and using coordinates given in the header from the digital sky survey. We then slightly adjusted the X-ray pointing by identifying and aligning X-ray sources on the digital sky survey image of the region. The close spatial correspondence between the ridge of optical and X-ray emission about 11 arcmin to the North of M~82 is obvious (Figure 1). We do not display the ROSAT HRI image of the ridge, since these data show only that the ridge is largely diffuse and structureless with the 15 arcsec (260 pc) off-axis spatial resolution of the HRI. The only substructure is a point source at the easternmost end of the ridge. This source appears to coincide with a stellar object visible in Figure 2 (the relatively bright star-like object located about 40 arcsec ENE of knot `I' in Figure 2 of DB). This is most likely a foreground star, as we will briefly discuss below. Examination of the 0.25 keV, 0.75 keV, and 1.5 keV PSPC maps of M~82 presented by \cite{DWH98} -- hereafter DWH -- shows that the X-ray morphology of the ridge is energy-dependent. The ENE portion of the ridge (source 9 in DWH) is relatively much more conspicuous in the 0.25 keV map compared to the rest of the ridge (source 8 in DWH). This reflects the contamination of the eastern-most ridge emission by the (presumably) unrelated point-source visible in the HRI image. As described in \S 2.1 above, we have therefore extracted a PSPC spectrum of the main body of the ridge from a region that excluded the point source. The relatively small number of detected photons precludes fitting complex multi-component models, and we fit a model of a single thermal (``MEKAL'') component to the spectrum. We are motivated to use a thermal model by the detailed X-ray spectroscopy of the brighter portions of the M~82 outflow (DWH; \cite{ML97}; \cite{SPS97}). The fits strongly preferred a low absorbing column, so we simply froze the column at the Galactic value of N$_H$ = 3.7 $\times$ 10$^{20}$ cm$^{-2}$. The data quality were not adequate to fit for the metal abundance, so we froze this at the solar value. Changing the assumed metal abundance did not significantly affect the best-fit temperature, but does directly affect the normalization of the fit. The best-fit temperature was kT=0.76$\pm$0.13 keV (90\% confidence interval with a reduced $\chi^{2}$ $\approx$ 1.3; Figure 3). The relatively high temperature is similar to that measured in the ROSAT PSPC spectroscopy of the innermost region of the M~82 outflow ($\sim$0.7 keV; DWH; \cite{SPS97}). The total unabsorbed flux in the 0.1 -- 2.4 keV band from this part of the ridge of emission is 8.5 $\times$ 10$^{-14}$ ergs cm$^{-2}$ s$^{-1}$. For the ENE part of the ridge - whose emission is contaminated by the point source described above - we have fit a two-component MEKAL model (Figure 3). Both components were frozen at solar abundances, one component (the ridge emission) had an absorbing column frozen at N$_H$ = 3.7 $\times$ 10$^{20}$ cm$^{-2}$, and the other (the stellar point source) had an absorbing column frozen at 0. This fit yields kT=0.85$\pm$0.17 keV for the ridge and kT$\sim$0.1 keV for the point source. The corresponding unabsorbed fluxes are 5.8 $\times$ 10$^{-14}$ ergs cm$^{-2}$ s$^{-1}$ and 3.5 $\times$ 10$^{-15}$ ergs cm$^{-2}$ s$^{-1}$ for the ridge and point source respectively. Combining the fits for the two regions of the ridge, the total X-ray flux is then 1.4 $\times$ 10$^{-13}$ ergs cm$^{-2}$ s$^{-1}$. This comprises about 0.7\% of the total unabsorbed X-ray emission from M~82 in the PSPC band (e.g., \cite{ML97}; DWH). The total X-ray luminosity of the ridge of emission is 2.2$\times$10$^{38}$ ergs s$^{-1}$ (in the 0.1 -- 2.4 keV band). As stated above, the ENE point source is identified with a stellar object having an R magnitude of 12.2 ($\lambda$F$_{\lambda}$ = 1.6 $\times$ 10$^{-10}$ erg cm$^{-2}$ s$^{-1}$). The X-ray/optical flux ratio suggests that the most likely nature of this source is a late-type dwarf star, and this would also be consistent with its soft ROSAT PSPC spectrum (e.g., \cite{RGV85}). \section{Discussion} \subsection{Is the Ridge of Emission Associated with M~82?} Being able to isolate the ridge of emission to the north of M~82 allows us to investigate its nature separately from the rest of M~82. Could this ridge of emission be due to another galaxy near M~82, or is it instead gas excited by the M~82 starburst? We have found that the ridge of X-ray and H$\alpha$ emission has fairly high X-ray and H$\alpha$ luminosities -- both luminosities are in the range observed for star-forming dwarf galaxies (e.g., \cite{HHG93}; \cite{H95}). However, we see no evidence for any optical continuum emission from this ridge. The limit for the level of continuum emission implies that any putative galaxy must be very low luminosity, comparable to that of the local dwarf spheroidal galaxies (\cite{M98}). Moreover, the limit on the continuum emission implies an H$\alpha$ equivalent width of greater than 180\AA. Examining the sample of galaxies with M$_B$ fainter than $-$15 in the the H$\alpha$ survey of nearby dwarf irregular galaxies by \cite{HHG93}, we find that these have typical global H$\alpha$ equivalent widths of a few to a few tens of \AA. The only such galaxy with an H$\alpha$ equivalent width comparable to our lower limit for the M~82 ridge is DDO53. This galaxy has a ratio of HI mass to H$\alpha$ luminosity (in solar units) of 420. The corresponding HI mass for the M~82 ridge would have to be 2.5 $\times 10^7$ M$_{\odot}$. The HI map in \cite{Y93} implies a maximum HI mass in the M~82 ridge that is only about 10$^6$ M$_{\sun}$. Moreover, while the M~82 ridge has an X-ray luminosity similar to that of the dwarf starburst galaxy NGC1569 (\cite{H95}), the M~82 ridge would have a ratio of X-ray to optical continuum luminosity that would be over 40 times higher than NGC1569. Thus, it is clear that the ridge in M~82 has properties that are very different from those of star-forming dwarf galaxies. This argument, together with the strong morphological connection between the M~82 ridge and the M~82 outflow (see DB), leads us to conclude that this ridge is {\it not} a dwarf star-forming galaxy. Other possibilities for this region, like a cloud in the halo of the Milky Way are also highly unlikely. The HI detected in the data of \cite{Y93} and the H$\alpha$ emission from the spectrum of DB have velocities such that it they are both likely to be associated with M~82 and not the galaxy (\cite{BH97}; \cite{hrt98}; \cite{r95}). Moreover, while galactic high velocity HI clouds often show soft X-ray excesses (\cite{Kerp96}; \cite{Kerp99}), the associated X-ray emission is very soft ($\approx$0.1-0.2 keV; see discussion in \cite{Kerp96}). \subsection{Photoionization of the Ridge by M~82} The ridge is evidently related to, and excited by M~82. In the case of the H$\alpha$ emission two possibilities are obvious. Either the superwind in M~82 has worked its way out to $\approx$11 kpc above the plane of the galaxy and is shocking halo material that appears in plentiful supply around M~82 (\cite{Y93}, 1994), or Lyman continuum radiation from the starburst is ionizing gas at very large distances from the disk. Only the first possibility could apply to the X-ray emission from the ridge. Let us suppose that ambient gas is being photoionized by the stellar population of M~82. The rate of ionizing photons from M~82 is approximately $10^{54}$ s$^{-1}$ (\cite{McL93} and references therein). Using the projected size and distance from M~82 estimated for the ridge of emission given in the previous section, and assuming that the ridge is really a ``box'' with a depth equal to its projected length (i.e., 3.7 kpc $\times$ 3.7 kpc) we find that the ridge will intercept about 8 $\times$ 10$^{51}$ ionizing photons s$^{-1}$. This of course is an upper limit since some (most?) of the ionizing photons will be absorbed or dust-scattered before they reach distances as great as 11.6 kpc out of the plane of M~82. We can estimate the rate of photons necessary to give the H$\alpha$ luminosity of the ridge (2.4 $\times$ 10$^{38}$ ergs s$^{-1}$) by assuming that the gas is in photo-ionization equilibrium, i.e., the number of recombinations equals the number of absorptions. Using the numbers for the Case B effective recombination rate of H$\alpha$ from \cite{O89} we find that it is necessary to have a rate of $\sim$ 2$\times$ 10$^{50}$ ionizing photons s$^{-1}$. This is $\sim$40 times smaller than the upper limit for the number of photons that M~82 is able to provide, so only about 2 to 3\% of the ionizing radiation initially emitted in the direction of the ridge needs to make it to this distance. Such an `escape fraction' is consistent with HUT spectra of starbursts below the Lyman edge (\cite{L95}; \cite{Hur97}), and it seems plausible that the optimal path along which ionizing radiation could escape from a starburst is along the cavity carved-out by the superwind. DB have made similar arguments in favor of a starburst-photoionization origin for the H$\alpha$ emission. However, photoionization by stars does not produce the hot X-ray emission that is coincident with the H$\alpha$ emission. Instead, shock-heating by the superwind may be an efficient mechanism for producing coincident X-ray and line-emitting plasma. We now explore the energetic plausibility of this model, and revisit the issue of the role of photoionization in \S 4.5. \subsection{Shock Energetics and the Basic Physical Properties of the Ridge} The ridge of emission has an X-ray luminosity of 2.2 $\times$ 10$^{38}$ ergs s$^{-1}$ and a temperature of 9$\times$ 10$^6$ K. Assuming that the shock is strong and adiabatic, we have that $T_{post-shock}={3 \over 16} {\mu v^2_s \over k}$, where $\mu$ is the mass per particle, $v_s$ is the shock velocity, and $k$ is Boltzmann's constant. Thus to produce gas with T = 9 $\times$ 10$^6$ K requires a shock speed of 820 km s$^{-1}$. Such a high shock speed is quite reasonable considering that M~82 has been estimated to have a deprojected outflow velocity in optical line emitting filaments observed along the minor axis of roughly 660 km s$^{-1}$ (\cite{SBH98}), and the outflowing X-ray gas could {\it in principle} have a velocity as high as 3000 km s$^{-1}$ (\cite{CC85}). It would take material traveling at 820 km s$^{-1}$, $\approx$1.4 $\times$ 10$^7$ yrs to travel the 11.6 kpc from M~82 to the ridge of emission. Since this is approximately the estimated lifetime of the starburst in M~82 (\cite{R93}; \cite{B92}; \cite{Sch98}; \cite{S97}), it is very plausible that the starburst could drive a shock over its lifetime this far out of the galactic plane. If we assume that we are observing a ``box'' of emission with dimensions of 3.7 kpc $\times$ 3.7 kpc $\times$ 0.9 kpc, the total volume is 3.7 $\times$ 10$^{65}$ cm$^{-3}$. Assuming this region is occupied by hot gas filling a fraction f$_X$ of this volume, the measured luminosity in the ROSAT band pass of 0.1 to 2.4 keV implies that the X-ray emitting gas has an electron density, n$_{e,X}$ = 5.4 $\times$ 10$^{-3}$ f$_X$$^{-1/2}$ cm$^{-3}$, pressure, P$_X\approx$ 2n$_{e,X}$kT=1.3 $\times$ 10$^{-11}$ f$_X$$^{-1/2}$ dynes cm$^{-2}$, cooling time, t$_{cool} \approx$ $3kT/n_{e,X}\Lambda$=2.7 $\times$ 10$^8$ f$_X$$^{1/2}$ yrs, a mass, M$_{X}$=2.0 $\times$ 10$^6$ f$_X$$^{1/2}$ M$_{\sun}$, and a total thermal energy of 8$\times$ 10$^{54}$ f$_X$$^{1/2}$ ergs. Similarly, assuming that conditions of Case B recombination and a temperature of 10$^4$ K holds in the region producing the H$\alpha$ emission, and that this region also occupies a volume 3.7 $\times$ 10$^{65}$ f$_{H\alpha}$ cm$^{-3}$, then the implied density and mass of this gas is n$_{e,H\alpha}$ = 4.3 $\times$ 10$^{-2}$ f$_{H\alpha}^{-1/2}$ cm$^{-3}$ and M$_{H\alpha}$ =8.0 $\times$ 10$^6$ f$_{H\alpha}^{1/2}$ M$_{\sun}$ respectively. If we make the physically reasonable assumption that this cooler gas is in rough pressure balance with the hot X-ray gas, then the minimum pressure of 1.3$\times$ 10$^{-11}$ dynes cm$^{-2}$ would require that the H$\alpha$ emitting gas is highly clumped (f$_{H\alpha}$ $\sim$ 8 $\times$ 10$^{-5}$). This in turn implies that n$_{e,H\alpha}$ = 5 cm$^{-3}$ and M$_{H\alpha}$ =1.1 $\times$ 10$^5$ M$_{\sun}$. We note that a small volume filling-factor for this gas is consistent with the complex knotty, filamentary structure seen in the deep H$\alpha$ image in DB. The total bolometric luminosity and ionizing luminosity of M~82 (\cite{McL93} and references therein) can be used together with the starburst models of \cite{LH95} to predict the rate at which the starburst injects mechanical energy and momentum. While these predictions will depend to some degree on the evolutionary state of the starburst (cf. Figures 66 through 69 in Leitherer \& Heckman) we estimate that the mechanical luminosity and momentum flux provided by M~82 are $\sim$ 2.5 $\times$ 10$^{42}$ erg s$^{-1}$ and $\sim$ 2 $\times$ 10$^{34}$ dynes respectively. For this wind mechanical luminosity, the rate at which the X-ray ridge intercepts this energy is 2.5 $\times$ 10$^{41}$ $\Omega_{w}$$^{-1}$ erg s$^{-1}$, where $\Omega_{w}$ is the total solid angle into which the wind flows ($\Omega_{w}$ $\leq$ 4$\pi$). Comparing this heating rate to the X-ray luminosity of the ridge (2.2 $\times$ 10$^{38}$ erg s$^{-1}$) shows that the wind could easily power the emission. It could also supply the ridge's estimated thermal energy in a timescale of 10$^{6}$ f$_X$$^{1/2}$ $\Omega_{w}$ years (much less than the wind/starburst lifetime). Matching the observed high pressure in the X-ray ridge is more challenging. A wind momentum flux of 2.5 $\times$ 10$^{34}$ dynes leads to a wind ram pressure at a radius of 11.6 kpc of 2 $\times$ 10$^{-11}$ $\Omega_{w}$$^{-1}$ dyne cm$^{-2}$. Consistency with the X-ray pressure estimated above would require that the X-ray gas has a large volume filling factor (f$_X$ $\sim$ unity) and that the M~82 wind is rather well collimated ($\Omega_{w}$ $\sim$ 1.5 steradian). This latter constraint is consistent with the morphology of the outflow (e.g., \cite{G90}; \cite{McK95}; \cite{SBH98}). Alternatively, we can use the empirical measurements of the gas pressure in the inner region of the M~82 nebula (\cite{HAM90}) and extrapolate these outward, assuming that the wind ram pressure falls as r$^{-2}$. The pressure at a radius of 1 kpc was measured to be $\sim$ 5 $\times$ 10$^{-10}$ dyne cm$^{-2}$, and so the predicted wind ram pressure at a radius of 11.6 kpc would be $\sim$ 4 $\times$ 10$^{-12}$ dyne cm$^{-2}$, about a factor three smaller than the estimated value (for f$_X$ = 1). In view of the uncertainties in the estimated physical parameters for the X-ray ridge, we regard this level of disagreement as acceptable. If the optical line emission is also shock-excited by the outflowing wind, then the luminosity observed at H$\alpha$ should be consistent with the wind energetics. We can estimate the amount of emission from a shock as follows. The total energy dissipated in a shock is, $L_{shock} = 1/2 \rho A v^3_s= 1.0 \times 10^{34} n_0 A_{kpc} v^3_s$ ergs s$^{-1}$, where $A_{kpc}$ is the shock surface area in kpc$^2$, $n_0$ is the preshock density (cm$^{-3}$), and $v_s$ is in units of km s$^{-1}$. From \cite{BDT85}, we have using the equation above, L$(H\alpha)_{shock} \approx 3 L(H\beta)_{shock} \approx 3 L_{shock}/100 = 1.7 \times 10^{41} n_0 A_{kpc} (v_s/820)^3$ ergs s$^{-1}$, if 90 km s$^{-1}$ $<$ $v_s$ $<$ 1000 km s$^{-1}$. Adopting the geometry above, the total surface area that the ridge would present to the wind would be 3.7 by 3.7 kpc. The observed H$\alpha$ luminosity of the ridge (2.4 $\times$ 10$^{38}$ erg s$^{-1}$) then requires that $n_0 (v_s/820)^3$ = 1.0 $\times$ 10$^{-4}$, while the pressure estimated in the X-ray ridge (1.3$\times$ 10$^{-11}$ dynes cm$^{-2}$) requires $n_0 (v_s/820)^2$ = 8 $\times$ 10$^{-4}$. Thus, provided that the shock speed is greater than 100 km s$^{-2}$, it can both produce the observed H$\alpha$ emission-line luminosity and be consistent with the pressure estimated for the hot shocked gas observed in the X-rays. \subsection{The Relationship Between the H$\alpha$ and X-Ray Emission} What is the physical relation between the H$\alpha$ and X-ray emission in this type of model in which the superwind drives a shock into an ambient cloud in the halo of M~82? One possibility is that the shock driven into the cloud by the wind is fast enough (i.e., $\sim$ 800 km s$^{-1}$) to produce the X-ray emission, and that the shocked gas then cools radiatively down to temperatures of-order 10$^4$ K and produces the observed H$\alpha$ emission via recombination. The difficulty with this simple model is that the estimated radiative cooling time in the X-ray gas ($\sim$ 3 $\times$ 10$^8$ years) is very long compared to the characteristic dynamical time for the X-ray ridge: t$_{dyn}$ $\sim$ (0.9 kpc/820 km s$^{-1}$) $\sim$ 1 $\times$ 10$^6$ years. Simply put, this model would require the X-ray ridge to be $\sim$ 300 kpc thick and the H$\alpha$ emission would be at the back edge of this very thick cooling zone! An additional persuasive argument against this model is provided by the rather quiescent kinematics of the H$\alpha$-emitting material reported by DB. Their optical spectra show that the H$\alpha$ emission in the ridge spans a velocity range of only about 10$^2$ km s$^{-1}$, which is inconsistent with material cooling behind a shock with a velocity nearly an order-of- magnitude larger. While the relatively small difference DB report between the mean velocity of the ridge and M~82 proper (blueshift of 50 to 200 km s$^{-1}$) could perhaps be explained as a projection effect in an outflow seen nearly in the plane of the sky, this effect could not explain the narrowness of the H$\alpha$ emission-line in the ridge ($\sim$ 10$^2$ km s$^{-1}$) without appealing to an unreasonably idealized shock geometry (which is bellied by the morphological complexity of the H$\alpha$ emission in the image published by DB). A much more plausible alternative is that we are actually observing {\it two} physically-related shocks in the M~82 ridge. We propose that as the superwind encounters a large cloud in the halo of M~82, it drives a relatively slow (of-order 10$^2$ km s$^{-1}$) radiative shock into the cloud. At the same time, this encounter also results in a fast (820 km s$^{-1}$) stand-off bow-shock in the superwind fluid upstream from the cloud. The slow cloud shock (aided and abetted by ionizing radiation from M~82) produces the H$\alpha$ emission. This is consistent with the quiescent H$\alpha$ kinematics. The fast bow-shock in the wind produces the X-ray emission and is predicted to only be offset from front edge of the cloud by a fraction of the size of the cloud (\cite{kmc94}; \cite{bw90}). Pressure balance between the two shocks requires that $\rho_{wind} v_{bow}^2 = \rho_{cloud} v_{shock,cloud}^2$, where $\rho$$_{wind}$ and $\rho$$_{cloud}$ are the preshock densities in the wind and cloud respectively. The observed X-ray pressure and temperature imply that v$_{bow}$ $\sim$ 820 km s$^{-1}$ and n$_{wind,ion}$ $\sim$ 1 $\times$ 10$^{-3}$ cm$^{-3}$. Thus - for example - if the shock speed in the cloud is v$_{s,cloud}$ = 100 km s$^{-1}$, then n$_{cloud} \sim$ 0.07 cm$^{-3}$. We note further that the velocity of the bow shock is just the difference between the outflow velocity of the superwind itself and that of the shocked cloud. Assuming that latter velocity is much smaller than the former (as is suggested by the kinematic information provided by DB), the wind velocity implied by the X-ray temperature is $\sim$ 820 km s$^{-1}$. To drive a wind with a terminal velocity of 820 km s$^{-1}$, hot gas would need to be injected into the starburst at a mean mass-weighted temperature of 1 $\times$ 10$^7$ K (\cite{CC85}). This temperature is consistent with X-ray spectroscopy of the hot gas in the core of M~82 (DWH, \cite{ML97}; \cite{SPS97}). We have seen above that the radiative cooling time in the X-ray ridge is $\sim$300 times longer than the characteristic dynamical time for the ridge. This implies that the wind bow shock is adiabatic. This can also be seen by comparing the total rate of energy dissipation in the shock ($L_{shock} = 1/2 \rho_{wind} A_{ridge} v_{bow}^3 = 7 \times 10^{40}$ erg s$^{-1}$ for the parameters estimated above) to the observed X-ray luminosity of the ridge (2.2 $\times$ 10$^{38}$ erg s$^{-1}$). Thus, the thickness of the X-ray ridge in the direction of the outflow is set by the dimensions of the region where adiabatic expansion and cooling causes the X-ray surface-brightness to drop precipitously. This region will be of-order the size of the obstacle in the flow that has caused the bow-shock to develop (the cloud). This is consistent with the observed X-ray morphology and the good spatial correlation seen between the H$\alpha$ and X-ray emission. We have also estimated previously that the superwind can supply the ridge with its estimated thermal energy content in a timescale of-order 10$^6$ years. This is roughly equal to the dynamical time in the ridge (the timescale for adiabatic cooling), as required. We conclude that it is energetically feasible for the cloud to be shock-heated by a superwind driven by the starburst in M~82. A wind/cloud shock interpretation leads to no uncomfortable estimates of the luminosity, covering fraction, or time scales. In addition, as noted previously, there is ample evidence that M~82 lies in an ``HI halo'' (\cite{Y93}, 1994). It is likely that the outflowing gas has intercepted and shock-heated some of this halo material, which is observed to be clumpy over large scales (see \cite{Y93}). We also note that perhaps one may have expected to find stronger evidence {\it a priori} for such interactions on the north side of the plane of M~82 compared to the south side: \cite{Y94} find that the velocity gradient in the HI is much higher on the northern side (12 km s$^{-1}$ kpc$^{-1}$) of the plane of M~82 compared to that on the southern side (3--6 km s$^{-1}$ kpc$^{-1}$). This higher gradient may be related to our observation of co-spatial H$\alpha$ and thermal X-ray emission at large distances from the plane of M~82. \subsection{Where is the Cloud?} We have shown that a simple model of the interaction between a fast ($\sim$ 800 km s$^{-1}$) wind and a massive cloud can quantitatively explain the physical and dynamical properties of the ridge. However, it a fair question to ask why the putative cloud is not observed in the sensitive HI map published by \cite{Y93}. More specifically, we have estimated above that the high pressures in the X-ray gas and the low velocity of the shock driven into the cloud by the wind together mean that the preshock density of the cloud must be of-order 10$^{-1}$ cm$^{-3}$. For a cloud with a line-of-sight dimension of 3.7 kpc, the implied cloud column density is then $\sim$10$^{21}$ cm$^{-2}$. The HI map in \cite{Y93} constrains the HI column density of the ridge to be $<$ 2.7$\times$ 10$^{19}$ cm$^{-2}$. This implies that the cloud must now be fully ionized. This could be accomplished by either the wind-driven shock or by the ionizing radiation from M~82. The time for the shock to traverse the cloud will be $\sim$ 10$^{7}$ years for $v_{shock,cloud}$ = 100 km s$^{-1}$, comparable to the estimated age of the starburst and its superwind. Similarly, the velocity of a photo-ionization front moving though the cloud is just $v_i \approx \Phi_{LyC} n_{cloud}^{-1}$ (e.g., \cite{O89}), where $\Phi_{LyC}$ is the incident flux of ionizing radiation. The observed H$\alpha$ luminosity and the assumed cross-sectional area the cloud presents to the starburst (3.7 by 3.7 kpc) imply that the average value of $\Phi_{LyC}$ is 1.5 $\times 10^{6} f_{abs}^{-1}$ s$^{-1}$ cm$^{-2}$, where $f_{abs}$ is the fraction of the incident ionizing photons absorbed by the cloud (i.e., allowing the cloud to be optically-thin). For a preshock density of 0.07 cm$^{-3}$ in the cloud, the implied velocity is $v_i \approx$ 210 f$_{abs}$$^{-1}$ km s$^{-1}$, and the time for the front to cross the cloud is about 4 $\times$ 10$^6$ f$_{abs}$$^{-1}$ years. Thus it is plausible that an initially neutral cloud like those seen in the HI maps of Yun et al. (1993,1994) has indeed been fully ionized in the time since the starburst and its superwind turned on. Our rough estimates suggest that the timescale for the cloud to be fully photoionized is likely to be somewhat shorter than the timescale for it to be fully shock-ionized. If correct, this would mean that the H$\alpha$ emission produced by the gas cooling and recombining behind the shock would have a high density and pressure, while the rest of the ionized cloud (the as-yet unshocked material) would be at a much lower pressure and density. We speculate that the bright knots and loops of H$\alpha$ visible in the image in DB may correspond to shock-excited emission from gas with a high pressure, while the fainter and more diffuse emission may come from the low-pressure starburst-photoionized gas that has not yet been shocked. \subsection{Implications for the ``Superwind'' Phenomenon} It is often difficult to associate discrete X-rays sources seen in the halos of starburst galaxies with activity in the starburst galaxy itself. Often, it is assumed that most discrete objects observed in the halos of starbursts are more distant objects (usually QSOs) seen in projection. Finding a spatial association between H$\alpha$ emission at the redshift of the starburst galaxy and X-ray emission is a irrefutable way of being able to associate the X-ray emission directly with activity within the starburst galaxy. Having argued that this co-spatial region of soft X-ray and H$\alpha$ emission must be due to the outflowing superwind, it is now evident that at least some starburst galaxies are able to drive material to much larger distances from their nuclei than would be inferred using the relatively high surface brightness X-ray or line emission that is generally observed along their minor axes (see e.g., \cite{LH96}; \cite{HAM90}; DWH). In M~82 for example, the inner ``continuous'' region of bright H$\alpha$ and X-ray emission extends only about 5 arcminutes from the nucleus along the minor axis. The detached ``ridge'' of emission that we have discussed is at a projected angular separation from the nucleus of over twice this distance (about 11 arcminutes). Thus without making the association between the H$\alpha$ and X-ray emission, a variety of interpretations for the source of the X-ray emission from this region would be possible. This obviously has important implications for our understanding of superwinds. It emphasizes that one must be careful in using the observed spatial scales of the extended emission in starbursts to either access whether or not the wind material is able to escape the potential of the host galaxy or to estimate the dynamical age of the outflow. It is interesting in this vein to consider the long-term fate of the hot gas in the X-ray ridge of M~82. To do so, we will compare the observed temperature of the gas to the ``escape temperature'' at that location in the M~82 halo. Following \cite{W95}, the escape temperature for hot gas in a galaxy potential with an escape velocity v$_{es}$ is given by: T$_{es}$ $\simeq$ $1.1\times 10^{5}$ (v$_{es}$/100 km s$^{-1}$)$^{2}$ K. The combined CO plus HI rotation curve of M~82 shows a roughly Keplerian decrease with radius for radii greater than 0.4 kpc, and $v_{rot}$ has dropped to 60 km s$^{-1}$ at a radius of 3.5 kpc (\cite{S98}). If this Keplerian fall-off continues to larger radii, the implied escape velocity at a radius of 11.6 kpc is only 45 km s$^{-1}$ and the corresponding escape temperature is T$_{es}$ $\simeq$ $2\times 10^{4}$ K. This is 450 times cooler than the observed X-ray emission. Even if we assume instead that M~82 has an isothermal dark matter halo with a depth corresponding to a circular velocity of 60 km s$^{-1}$ (the maximum permitted by the directly measured rotation curve) and further assume that this halo extends to a radius of 100 kpc, the escape velocity at r = 11.6 kpc is about 150 km s$^{-1}$ and T$_{es}$ = 2.5 $\times 10^5$ K. This is still a factor of 36 below the observed temperature of the gas in the ridge. This implies that the observed hot gas (and by inference, the superwind itself) will be able to escape the gravitational potential of M~82. The X-ray/H$\alpha$ ridge has particularly interesting implications for the physics of superwind emission. While the existence of superwinds is now well-established, the process(es) by which they produce X-ray emission is (are) not clear. This topic has been nicely reviewed by \cite{Str98}. If the outflowing wind consists purely of the thermalized ejecta from supernovae and stellar winds, the wind fluid is hot ($\sim$ 10$^8$ K prior to adiabatic cooling) and is so tenuous that it will be an insignificant source of X-ray emission (\cite{CC85}; \cite{Suc94}). Various alternatives have been proposed to boost the X-ray luminosity produced by a superwind: 1) Substantial quantities of material could be mixed into the wind in or near the starburst as the stellar ejecta interact with the starburst ISM (the wind could be centrally ``mass-loaded'' - \cite{Suc96}; \cite{H97}). 2) Large quantities of gas could be added {\it in situ} as the wind overtakes and then evaporates or shreds gas clouds in the galactic halo (e.g., \cite{Suc94}). These clouds could either be pre-existing clouds in the halo or material from the disk of the starburst galaxy that has been carried into the halo by the wind. 3) The tenuous wind fluid could drive a shock into a denser volume-filling galactic halo, with the observed X-ray emission arising from the shocked-halo material rather than the wind-fluid itself. In the case of the ridge of emission in the halo of M~82, we have argued that the observable X-ray emission arises as the wind fluid -- which due to adiabatic expansion and cooling would otherwise have an X-ray surface-brightness below the level detectable in the existing ROSAT image -- encounters a few-kpc-size-scale cloud in the halo. We have argued that the observed X-ray emission is produced in the bow-shock created in the superwind just upstream of the cloud. This process might be particularly relevant to the case of M~82, which is immersed in an extensive system of tidally-liberated gas clouds, shared with its interaction partners M81 and NGC3077. Such material would also provide a potential source of optical line emission by either intercepting a small portion of the ionizing radiation produced by the central starburst or as it is shock-heated by the superwind. Given the strong connection between galaxy interactions/mergers and the starburst phenomenon (e.g., \cite{SM96} and references therein), this mechanism may be generally applicable to superwinds. \section{Summary and Conclusions} We report an analysis of a newly-discovered (\cite{DB99}) region of spatially-coincident X-ray and H$\alpha$ emission to the north of the prototypical starburst/superwind galaxy M~82. At the distance of M~82 (3.63 Mpc) the dimensions of this correlated ridge of emission are 3.7 $\times$ 0.9 kpc and it lies at a projected distance of about 11.6 kpc above the plane directly along the minor axis of M~82 (and hence along the axis of the superwind). We measure a total H$\alpha$ flux from the ridge of 1.5 $\times$ 10$^{-13}$ ergs s$^{-1}$ cm $^{-2}$. This flux yields a total H$\alpha$ luminosity of 2.4 $\times$ 10$^{38}$ ergs s$^{-1}$. The total H$\alpha$ flux for the whole of M~82 is approximately 4.6 $\times$ 10$^{-11}$ ergs s$^{-1}$ cm $^{-2}$. Thus the H$\alpha$ ridge constitutes approximately 0.3\% of the total observed H$\alpha$ flux from the galaxy. X-ray emission is seen over the same region as the H$\alpha$ emission by the ROSAT PSPC and HRI. With the 15 arcsec (260 pc) off-axis spatial resolution of the HRI, the ridge is diffuse and featureless except for a (probably unrelated) soft point source at its ENE end. The best fit to the X-ray emission from the ridge is a single thermal component with a temperature of kT=0.80$\pm$0.17 keV, absorbed by a column density N$_H$ = 3.7 $\times$ 10$^{20}$ cm$^{-2}$. The absorption is consistent with the foreground Galactic column. The total unabsorbed flux from the ridge of emission is 1.4 $\times$ 10$^{-13}$ ergs cm$^{-2}$ s$^{-1}$ and its luminosity of 2.2 $\times$ 10$^{38}$ erg s$^{-1}$ comprises about 0.7\% of the total X-ray emission from M~82 in the 0.l -- 2.4 keV ROSAT band. We find that the observed H$\alpha$ emission from the ridge could be produced via photoionization by the dilute radiation from the starburst propagating into the M~82 halo along the path cleared by the superwind. On the other hand, the X-ray emission from the ridge can be most naturally explained as the result of shock-heating associated with an encounter between the starburst-driven galactic ``superwind'' and a large photoionized cloud in the halo of M~82. Shock-heating may also contribute significantly to the excitation of the H$\alpha$ emission, especially the regions of the highest surface brightness. We find that the H$\alpha$ and X-ray luminosities and the X-ray temperature and estimated gas pressure can all be quantitatively understood in this context. The specific model we favor is one in which the encounter between the superwind and the cloud drives a slow radiative shock into the cloud (producing some of the H$\alpha$ emission) and also leads to an adiabatic X-ray-emitting bow-shock in the superwind fluid upstream from the cloud. The cloud is now fully ionized, and so is not detected in the HI maps of Yun et al. (1993, 1994). The wind-cloud interaction process may be especially relevant to M~82, whose halo contains tidally-liberated gas clouds. However, it may well occur generally in powerful starbursts (which are usually triggered by galaxy interactions or mergers). The existence of a region of spatially coincident X-ray and H$\alpha$ emission with strong evidence for wind/cloud interaction is important for several reasons. Firstly, our analysis sheds new light on the physical processes by which galactic winds produce X-ray emission, and implies that wind/cloud collisions are an important emission-producing process. The emission ridge in M~82 highlights the potential pitfalls in using the observed X-ray or H$\alpha$ emission to trace the size and infer the ages or fates of galactic winds: the winds may extend to radii far larger than the region over which they have densities and temperatures high enough to produce observable emission. Wind-cloud encounters then offer us the possibility of `lighting up' the wind at radii that are otherwise inaccessible. The X-ray ridge in M~82 implies that the wind is able to reach large heights above the plane of the galaxy, thereby increasing the likelihood that it will be able to eventually escape the galaxy potential well. Indeed, the observed gas temperature exceeds the local ``escape temperature'' from the M~82 gravitational potential well by about two orders-of-magnitude. This result strengthens the case that starburst-driven galactic winds are important sources of the metal-enrichment and heating of the Intergalactic Medium. \acknowledgements The authors wish to thank Kitt Peak National Observatory for their generous allocation of observing time and their staff for making sure that the observations were carried out efficiently and effectively. We thank Ed Moran, David Strickland, and Crystal Martin for illuminating discussions about the physics of superwinds, and David Devine for communicating his discovery paper in advance of of its publication. We would like to express our sincerest thank you to the referee for carefully reading the manuscript and for providing us with detailed comments that have substantially improved this presentation. This work was supported in part by NASA LTSA grants NAGW-3138 to TH and NAGW-4025 to KW. This research has made use of the IRAF package provided by NOAO and NASA/IPAC extragalactic database (NED), which is operated by the Jet Propulsion Laboratory, Caltech, under contract with the National Aeronautics and Space Administration. \newpage
\chapter{#1}} \newcommand{\rchaptno}[1]{\newpage\label{dummy#1} \ifthenelse{\isodd{\pageref{dummy#1}}}{}{\blankpage} \chapter*{#1}} \newcommand{\hookrightarrow}{\hookrightarrow} \newcommand{{\cal M}} \newcommand{\ts}{\ul{\cal M}}{{\cal M}} \newcommand{\ts}{\underline{\cal M}} \newcommand{h} \newcommand{\tg}{\ul{g}} \newcommand{\ig}{g} \newcommand{\nlg}{g'}{h} \newcommand{\tg}{\underline{g}} \newcommand{\ig}{g} \newcommand{\nlg}{g'} \newcommand{e} \newcommand{\tE}{\ul{E}} \newcommand{\iE}{E} \newcommand{\nE}{E}{e} \newcommand{\tE}{\underline{E}} \newcommand{\iE}{E} \newcommand{\nE}{E} \newcommand{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}} \newcommand{\tw}{\Omega} \def\gow{{\mathfrak w}} \def\scW{{\scr W}} \def\dsW{{\mathbb W}} \def\clW{{\cal W}} \newcommand{\iw}{\Omega} \def\gow{{\mathfrak w}} \def\scW{{\scr W}} \def\dsW{{\mathbb W}} \def\clW{{\cal W}} \newcommand{\nw}{\Omega} \def\gow{{\mathfrak w}} \def\scW{{\scr W}} \def\dsW{{\mathbb W}} \def\clW{{\cal W}}{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}} \newcommand{\tw}{\Omega} \def\gow{{\mathfrak w}} \def\scW{{\scr W}} \def\dsW{{\mathbb W}} \def\clW{{\cal W}} \newcommand{\iw}{\Omega} \def\gow{{\mathfrak w}} \def\scW{{\scr W}} \def\dsW{{\mathbb W}} \def\clW{{\cal W}} \newcommand{\nw}{\Omega} \def\gow{{\mathfrak w}} \def\scW{{\scr W}} \def\dsW{{\mathbb W}} \def\clW{{\cal W}} \newcommand{\wT}{{\cal T}} \newcommand{\tT}{\underline{T}} \newcommand{\iT}{T} \newcommand{\nT}{T} \newcommand{\wsr}{{\cal R}} \newcommand{\tR}{\underline{R}} \newcommand{\ir}{R} \newcommand{\nr}{R} \newcommand{\ul{d}} \newcommand{\id}{d} \newcommand{\nd}{d'}{\underline{d}} \newcommand{\id}{d} \newcommand{\nd}{d'} \newcommand{\theta} \def\gos{{\mathfrak s}} \def\scS{{\scr S}} \def\dsS{{\mathbb S}} \def\clS{{\cal S}} \newcommand{\tco}{\ul{\theta} \def\gos{{\mathfrak s}} \def\scS{{\scr S}} \def\dsS{{\mathbb S}} \def\clS{{\cal S}}} \newcommand{\ico}{\clP} \newcommand{\nco}{\clP'\!}{\theta} \def\gos{{\mathfrak s}} \def\scS{{\scr S}} \def\dsS{{\mathbb S}} \def\clS{{\cal S}} \newcommand{\tco}{\underline{\theta} \def\gos{{\mathfrak s}} \def\scS{{\scr S}} \def\dsS{{\mathbb S}} \def\clS{{\cal S}}} \newcommand{\ico}{\clP} \newcommand{\nco}{\clP'\!} \newcommand{\wdel}{{\cal D}} \newcommand{\tdel}{\underline{\nabla}}\newcommand{\idel}{\nabla} \newcommand{\ndel}{\nabla'} \newcommand{T} \newcommand{\ec}{K} \newcommand{\edel}{\tilde{\nabla}}{T} \newcommand{\ec}{K} \newcommand{\edel}{\tilde{\nabla}} \newcommand{\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}}{\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}} \newcommand{\tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}}{\tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}} \newcommand{\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}}{\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}} \newcommand{\ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}}{\underline{\scC}} \newcommand{\aC}{\Tilde{\underline{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\underline{\scC}}}} \newcommand{\Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}}{\Tilde{\Tilde{\underline{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\underline{V}}}} \newcommand{I}{I} \newcommand{{\cal E}} \newcommand{\te}{{\cal E}}{{\cal E}} \newcommand{\te}{{\cal E}} \newcommand{\tilde{\theta} \def\gos{{\mathfrak s}} \def\scS{{\scr S}} \def\dsS{{\mathbb S}} \def\clS{{\cal S}}}{\tilde{\theta} \def\gos{{\mathfrak s}} \def\scS{{\scr S}} \def\dsS{{\mathbb S}} \def\clS{{\cal S}}} \newcommand{{\cal L}} \newcommand{\cec}{{\cal K}}{{\cal L}} \newcommand{\cec}{{\cal K}} \newcommand{\ul{T}}{\underline{T}} \newcommand{\ul{R}}{\underline{R}} \newcommand{\ul{\nabla}}{\underline{\nabla}} \newcommand{\ul{\cal D}}{\underline{\cal D}} \newcommand{\ul{\hat{\nabla}}}{\underline{\hat{\nabla}}} \newcommand{\ul{\hat{\cal D}}}{\underline{\hat{\cal D}}} \newcommand{(\!m^{\ss\minus 1}\!)}{(\!m^{\ss\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}\!)} \newcommand{(\!\te^{\ss\minus 1}\!)}{(\!\te^{\ss\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}\!)} \newcommand{{\ul{M}}} \newcommand{\tm}{{\ul{m}}} \newcommand{\tmm}{{\ul{\mu} \def\gom{{\mathfrak m}} \def\scM{{\scr M}} \def\dsM{{\mathbb M}} \def\clM{{\cal M}}}}{{\underline{M}}} \newcommand{\tm}{{\underline{m}}} \newcommand{\tmm}{{\underline{\mu} \def\gom{{\mathfrak m}} \def\scM{{\scr M}} \def\dsM{{\mathbb M}} \def\clM{{\cal M}}}} \newcommand{{\ul{N}}} \newcommand{\tn}{{\ul{n}}} \newcommand{\tnn}{{\ul{\nu} \def\gon{{\mathfrak n}} \def\scN{{\scr N}} \def\dsN{{\mathbb N}} \def\clN{{\cal N}}}}{{\underline{N}}} \newcommand{\tn}{{\underline{n}}} \newcommand{\tnn}{{\underline{\nu} \def\gon{{\mathfrak n}} \def\scN{{\scr N}} \def\dsN{{\mathbb N}} \def\clN{{\cal N}}}} \newcommand{{\ul{P}}} \newcommand{\tp}{{\ul{p}}} \newcommand{\tpp}{{\ul{\rho} \def\goo{{\mathfrak o}} \def\scO{{\scr O}} \def\dsO{{\mathbb O}} \def\clO{{\cal O}}}}{{\underline{P}}} \newcommand{\tp}{{\underline{p}}} \newcommand{\tpp}{{\underline{\rho} \def\goo{{\mathfrak o}} \def\scO{{\scr O}} \def\dsO{{\mathbb O}} \def\clO{{\cal O}}}} \newcommand{{\ul{Q}}} \newcommand{\tq}{{\ul{q}}} \newcommand{\tqq}{{\ul{\sigma} \def\goq{{\mathfrak q}} \def\scQ{{\scr Q}} \def\dsQ{{\mathbb Q}} \def\clQ{{\cal Q}}}}{{\underline{Q}}} \newcommand{\tq}{{\underline{q}}} \newcommand{\tqq}{{\underline{\sigma} \def\goq{{\mathfrak q}} \def\scQ{{\scr Q}} \def\dsQ{{\mathbb Q}} \def\clQ{{\cal Q}}}} \newcommand{{\ul{A}}} \newcommand{\ta}{{\ul{a}}} \newcommand{\taa}{{\ul{\alpha} \def\goa{{\mathfrak a}} \def\scA{{\scr A}} \def\dsA{{\mathbb A}} \def\clA{{\cal A}}}}{{\underline{A}}} \newcommand{\ta}{{\underline{a}}} \newcommand{\taa}{{\underline{\alpha} \def\goa{{\mathfrak a}} \def\scA{{\scr A}} \def\dsA{{\mathbb A}} \def\clA{{\cal A}}}} \newcommand{{\ul{B}}} \newcommand{\tb}{{\ul{b}}} \newcommand{\tbb}{{\ul{\beta} \def\gob{{\mathfrak b}} \def\scB{{\scr B}} \def\dsB{{\mathbb B}} \def\clB{{\cal B}}}}{{\underline{B}}} \newcommand{\tb}{{\underline{b}}} \newcommand{\tbb}{{\underline{\beta} \def\gob{{\mathfrak b}} \def\scB{{\scr B}} \def\dsB{{\mathbb B}} \def\clB{{\cal B}}}} \newcommand{{\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}}{{\underline{C}}} \newcommand{\tc}{{\underline{c}}} \newcommand{\tcc}{{\underline{\gamma}}} \newcommand{{\ul{D}}} \newcommand{\td}{{\ul{d}}} \newcommand{\tdd}{{\ul{\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}}}}{{\underline{D}}} \newcommand{\td}{{\underline{d}}} \newcommand{\tdd}{{\underline{\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}}}} \newcommand{{\bar{A}}} \newcommand{\ba}{{\bar{a}}} \newcommand{\baa}{{\bar{\alpha} \def\goa{{\mathfrak a}} \def\scA{{\scr A}} \def\dsA{{\mathbb A}} \def\clA{{\cal A}}}}{{\bar{A}}} \newcommand{\ba}{{\bar{a}}} \newcommand{\baa}{{\bar{\alpha} \def\goa{{\mathfrak a}} \def\scA{{\scr A}} \def\dsA{{\mathbb A}} \def\clA{{\cal A}}}} \newcommand{{\bar{B}}} \newcommand{\bb}{{\bar{b}}} \newcommand{\bbb}{{\bar{\beta} \def\gob{{\mathfrak b}} \def\scB{{\scr B}} \def\dsB{{\mathbb B}} \def\clB{{\cal B}}}}{{\bar{B}}} \newcommand{\bb}{{\bar{b}}} \newcommand{\bbb}{{\bar{\beta} \def\gob{{\mathfrak b}} \def\scB{{\scr B}} \def\dsB{{\mathbb B}} \def\clB{{\cal B}}}} \newcommand{{\bar{C}}} \newcommand{\bc}{{\bar{c}}} \newcommand{\bcc}{{\bar{\gamma}}}{{\bar{C}}} \newcommand{\bc}{{\bar{c}}} \newcommand{\bcc}{{\bar{\gamma}}} \newcommand{{\bar{D}}} \newcommand{\bd}{{\bar{d}}} \newcommand{\bdd}{{\bar{\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}}}}{{\bar{D}}} \newcommand{\bd}{{\bar{d}}} \newcommand{\bdd}{{\bar{\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}}}} \usepackage[all]{xy} \begin{document} \begin{titlepage} \title{{\Huge \bf Curvature relations in almost product manifolds}} \author{ Magnus Holm\thanks{<EMAIL>} $\;$ and Niclas Sandstr\"{o}m\thanks{<EMAIL>} } \date{ Institute for Theoretical Physics\\ G\"oteborg University and Chalmers University of Technology\\ S-412 96 G\"oteborg, Sweden\\ \leavevmode } \maketitle \begin{abstract} New relations involving curvature components for the various connections appearing in the theory of almost product manifolds are given and the conformal behaviour of these connections are studied. New identities for the irreducible parts of the deformation tensor are derived. Some direct physical applications in Kaluza--Klein and gauge theory are discussed. \end{abstract} \end{titlepage} \section{Introduction} In modern day theoretical physics one often deals with additional dimensions besides the ordinary four space-time ones. These extra dimensions manifest themselves in different forms. In gauge theory they appear as the dimension of the gauge group, in M-theory and string theory, they are required for self-consistency. In Kaluza--Klein theory, the gauge theory is obtained by compactification over an internal manifold with an isometry group which equals the gauge group. The almost product structure concept makes possible a geometrical formulation which completely describe these theories, without performing the dimensional reduction. This leads to new insights in their geometrical properties which is unobtainable in the dimensionally reduced theories themselves. For instance in ref. \cite{Holm98} it was shown that the Nijenhuis tensor of a certain almost product structure measures the fieldstrength which in the geometrical language is a measure of the non-integrability of the base manifold of the principal bundle. In almost product manifolds, three different connections appear naturally. As is known the Gauss--Codazzi relations connect curvature components of these connections. In this paper a classification of the relations of all curvature components is given which yields a number of new identities. As a result it becomes manifest that the Vidal connection in a principal bundle, or Kaluza--Klein theory, reduces to the gauge-covariant derivative. Since a lot of recent work \cite{CvYo96,CvDu99} has been made concerning rotating branes which are solutions to various supergravity theories, it will here be stressed that the almost product manifold vievpoint would be the most geometrical approach to these problems. Direct relations for the Ricci tensors in terms of the characteristic deformation tensors of an almost product structure will be given. These will then be the most natural starting point when making ans\"atze for new solutions in the supergravity theories. In the generic case, the Vidal connection will not be metric neither torsion free, and in section 2 we give a review of the theory of general connections. We refer to \cite{Schouten} for a more detailed treatise in this respect. In section 2 the properties of an arbitrary connection under conformal transformations are also reviewed. Section 3 gives a quick introduction to the basic connections and tensors involved with almost product manifolds. The naturally occuring connections, besides the Levi--Civita one, is the Vidal and adapted connections. All tensors formed from these connections are investigated in section 4. In that section several new identities are derived, some of which follows directly from the work in ref. \cite{Gr67}, and the conformal properties are studied. In section 5 this is brought to full fruition when the Vidal connection is shown to be identical to the gauge-covariant derivative in gauge or Kaluza--Klein theory. Possible further developments in this area is discussed in section 6. \section{A review on general connections} This section consists of two parts, the first of which treats a general non-metric connection and its curvature relations together with the Bianchi identities. The second part deals with the induced transformation of an arbitrary connection under conformal transformations. \subsection{Non-metric connections} The most frequently used non-Levi-Civita connections are the ones in which the torsion content is non-zero. In the case of a Vidal connection, the connection will not in general be metric nor symmetric. Also in the case of embeddings one might encounter non-metric connections while studying cases with an auxilliary metric on the world-volume. In this subsection a thorough description of connections in the most general case is given. See also ref. \cite{Schouten}. To this end, the following two important tensor are defined as, \defn{qandt}{Let $\nabla$ be a connection in a manifold ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ with non-degenerate metric $\ig$. Now define the torsion tensor, $T$, and the non-metricity tensor, $Q$, respectively with characteristics, \eqal{1}{ T:&\quad \L^1 \times \L^1 \longmapsto \L^1\\ Q:&\quad \L^1 \times \L^1 \times \L^1 \longmapsto \dsR } by the following equations \eqal{1}{ T(X,Y)&:=\nabla_XY-\nabla_YX-[X,Y]\\ Q(X,Y,Z)&:=(\nabla_X\ig)(Y,Z) } where $X,Y,Z\in\L^1$ are vectorfields on ${\cal M}} \newcommand{\ts}{\ul{\cal M}$. } A general connection on a manifold with non-degenerate metric can be decomposed into the Levi-Civita connection and an arbitrary $(2,1)$-tensor. The dimension of this tensor is therefore $m^3$ where $m$ is the dimension of the manifold ${\cal M}} \newcommand{\ts}{\ul{\cal M}$. Below it is shown that it can be decomposed into one part containing only the torsion tensor, $T$, and one part containing only the non-metricity tensor, $Q$. These two tensors have the dimensions ${1\fr 2}m^2(m-1)$ and ${1\fr 2}m^2(m+1)$ respectively which together give $m^3$. The torsion do not appear directly in the connection but as the contorsion and that is also the case with the non-metricity tensor. The following notation will be used in what follows, \eqnono{ \li T(X,Y,Z):=\ig(T(X,Y),Z) } In the next proposition the contorsion and con-metricity tensors are defined. \defn{contandcoq}{Let $\nabla$ be a connection in a manifold ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ with non-degenerate metric $\ig$. Define the contorsion tensor, $S$, and the con-metricity tensor, $P$, respectively, with same characteristics, \eqal{1}{ S,P:&\quad \L^1 \times \L^1 \times \L^1 \longmapsto \dsR } by following equations \eqal{1}{ \li S(X,Y,Z)&:={1\fr 2}(\li T(X,Y,Z)-\li T(Y,Z,X)+\li T(Z,X,Y))\\ \li P(X,Y,Z)&:={1\fr 2}(-Q(X,Y,Z)-Q(Y,Z,X)+Q(Z,X,Y)) } where $X,Y,Z\in\L^1$ are vectorfields on ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ and $S(X,Y)=g^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}(\li S(X,Y,\cdot)), P(X,Y)=g^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}(\li P(X,Y,\cdot))$. } Now, any connection can be expressed in terms of the Levi Civita connection with respect to a non-degenerate metric, denoted by ${}^g\!\nabla$, plus the contorsion and the con-metricity tensors defined above, i.e, \prop{connection}{Let $\nabla$ be an arbitrary connection on a manifold ${\cal M}} \newcommand{\ts}{\ul{\cal M}$, let further $\ig$ be a non-degenerate metric on ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ and ${}^g\!\nabla$ corresponding Levi-Civita connection. Let $S,P$ be the tensors defined in \ref{contandcoq} .Then \eqnono{ \nabla_XY={}^g\!\nabla_XY+S(X,Y)+P(X,Y) } } The curvature tensor of an arbitrary connection, is defined by, \eq{ R(X,Y)Z:=[\nabla_X,\nabla_Y]Z-\nabla_{[X,Y]}Z,\label{Rcurvature} } will no longer take values in the lie algebra $\goo(m)$ as does the curvature tensor of the Levi-Civita connection, but (will in the generic case take values) in $\gog\gol(m)$. The identities of the curvature tensor will therefor be altered, and its irreducible parts look in the generic case like \eq{ \yng(1,1)\otimes\yng(1)\otimes\yng(1)=\yng(1,1,1,1)\oplus\yng(2,1,1)\oplus\yng(2,2)\oplus\young(\hfil\hfil,:\hfil,:\hfil)\oplus \young(\hfil\hfil\hfil,::\hfil). } \prop{curvatureidentities}{The four identities of the Riemann curvature tensor of an arbitrary connection are \eqal{2}{ (i)&\quad& R_{(ab)c}{}^d=&0\\ (ii)&\quad& R_{[abc]}{}^d=&\nabla_{[a}T_{bc]}{}^d-T_{[ab}{}^eT_{c]e}{}^d\\ (iii)&\quad& R_{ab(cd)}=&-(\nabla_{[a}Q)_{b]cd}-{1\fr 2}T_{ab}{}^eQ_{ecd}\\ (iv)&\quad& R_{abcd}-R_{cdab}=&{3\fr 2}(R_{[abc]d}+R_{[bcd]a}-R_{[cda]b}-R_{[dab]c})+\\ &&&+R_{ab(cd)}-R_{bc(da)}-R_{cd(ab)}+R_{da(bc)}+R_{ac(db)}-R_{db(ac)} } } By the skew-tableaux (the two tableaux on the right in equation (2) above) it is stressed that these two irreducible parts will vanish when the connection is metric, i.e, the right hand side of identity 3 vanishes. In the generic case there are two possible contractions that can be made. \defn{ricci}{Let $\nabla$ be an arbitrary connection on a manifold ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ with curvature tensor, $R$. From the curvature it is possible to construct two types of $(2,0)$ tensors by contraction, namely, \eqal{1}{ R_{ab}:=R_{acb}{}^c,\\ V_{ab}:=R_{abc}{}^c. } The first one is the generalized Ricci tensor and the second one will here be refered to as the Schouten two-form. } The identities of the Ricci and Schouten tensors can be read off directly from the original curvature identities. \prop{Ricci-Schouten}{Let $R_{ab}$ be the Ricci tensor and $V_{ab}$ the Schouten tensor of an arbitrary connection, $\nabla$, then the second and third curvature identities implythe relations, \eqal{1}{ 2R_{[ab]}&=V_{ab}-\nabla_cT_{ab}{}^c-2\nabla_{[a}T_{b]}-T_{ab}{}^cT_c,\\ V_{ab}&=-(\nabla_{[a}Q)_{b]}-{1\fr 2}T_{ab}{}^cQ_c } } The only integrability conditions to the curvature identities are the Bianchi identities. \prop{Bianchi}{Let $\nabla$ be an arbitrary connection with curvature tensor $R$, and torsion tensor $T$. Then the Bianchi identity reads \eqnono{ \nabla_{[a}R_{bc]d}{}^e=T_{[ab}{}^fR_{c]fd}{}^e } from which the identities involving the Ricci tensor $R_{ab}$, and the Schouten two-form $V$, are derived, \eqal{1}{ 2\nabla_{[a}R_{b]c}+T_{ab}{}^dR_{dc}&=\nabla_dR_{abc}{}^d-2T_{d[a}{}^eR_{b]ec}{}^d\\ dV&=0 } } \subsection{Conformal transformations} Below, conformal transformations in the case of an arbitrary connection are studied. There will be some changes compared to the ordinary Levi Civita case when the connection involves torsion and non-metricity. \defn{Conformal tensor}{Let ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ be a manifold with metric $\ig$, let further $\nabla$ be an arbitrary connection on ${\cal M}} \newcommand{\ts}{\ul{\cal M}$. Let $\ct \ig:=e^{2\phi}\ig$ denote a conformal transformation then define the {\bf conformal tensor}, denoted by $\scC$, with characteristics \eqal{1}{ \scC:&\qquad \L^1 \times \L^1 \longmapsto \L^1, } by \eqal{1}{ \scC(X,Y):=\ct \nabla_X Y-\nabla_X Y } where $X,Y\in\L^1$ are vectorfields on ${\cal M}} \newcommand{\ts}{\ul{\cal M}$. } By a straight forward calculation one ends up with the transformations of the characteristic tensors of a connection under conformal transformation. \prop{conformal transformations}{Let ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ be a manifold with metric $\ig$ and $\nabla$ be an arbitrary connection on ${\cal M}} \newcommand{\ts}{\ul{\cal M}$. Let $R,T,Q$ denote the Riemann, torsion and non-metricity tensors respectively. Then their transformations under a conformal transformation can be expressed in terms of the conformal tensor, $\scC$, as \eqal{1}{ \ct R(X,Y)Z-R(X,Y)Z=&(\nabla_X\scC)(Y,Z)-(\nabla_Y\scC)(X,Z)+\\ &+\scC(X,\scC(Y,Z))-\scC(Y,\scC(X,Z))+\scC(T(X,Y),Z),\\ \ct T(X,Y)-T(X,Y)=&\scC(X,Y)-\scC(Y,X),\\ \ct Q(X,Y,Z)-e^{2\phi}Q(X,Y,Z)=&e^{2\phi}[2X[\phi]\ig(Y,Z)-\ig(\scC(X,Y),Z)-\ig(Y,\scC(X,Z))] } } In the case of the Levi-Civita connection the conformal tensor is most easily extracted from the above proposition. \prop{Levi-conf}{Let ${}^g\!\nabla$ be the Levi-Civita connection on a manifold with non-degenerate metric $g$, then its conformal tensor, denoted by ${}^g\!\scC$, reads \eqnono{ {}^g\!\scC(X,Y)=X[\phi]Y+Y[\phi]X-g(X,Y)\ri d\phi } } From these two propositions the conformal tensor in the generic case can be derived. \prop{generic-conf}{Let $\nabla$ be an arbitrary connection on a manifold ${\cal M}} \newcommand{\ts}{\ul{\cal M}$ with non-degenerate metric $g$, let further $S,P$ denote the contorsion and the con-metricity tensor respectively. Then the conformal tensor of $\nabla$ reads \eqnono{ \scC(X,Y)={}^g\!\scC(X,Y)+\ct S(X,Y)-S(X,Y)+\ct P(X,Y)-P(X,Y) } } \section{The connections associated with an almost product structure} Here a quick review on the concepts of almost product structures will be given, for a more thorough treatise see refs. \cite{Holm98,Nav84}. \nota{nynot}{We will denote the objects on our space with an underline, {\it i.e.}, \eqal{2}{ &\ts& &\text{Manifold}\\ &T\ts& &\text{Tangent bundle of $\ts$}\\ &T^*\ts\qquad&&\text{Cotangent bundle of $\ts$}\\ &\tg& &\text{Metric on $\ts$}\\ &\td& &\text{Exterior derivative}\\ &\underline{X}&&\text{Vector field on $\ts$} } to list the primarily used objects. We will use this underlining principle for all objects on $\ts$ whenever there may be risk of confusion. } \defn{i2}{Let I be an almost product structure on a manifold $\ts$ with riemannian metric $\tg$ and let $X,\;Y\in T\ts$ be vector fields. Then the triplet ($\ts$, $\tg$, $I$) is called an {\bf riemannian almost product structure} or simply an {\bf almost product manifold} if \eqnono{ \tg(IX,IY)=\tg(X,Y) } or in other words, $I$ is a automorphism of $\tg$ in the sense that the following diagram commutes: \eqnono{ \xy \xymatrix{T\ts\ar[d]_{\tg}\ar[r]^I&T\ts\ar[d]^{\tg}\\ T^*\ts\ar[r]_{I^t}&T^*\ts} \endxy } {\it i.e.}, \eqnono{ I^t\circ\tg\circI=\tg } } \prop{i1}{Let the triplet $(\ts,\tg,I)$ define a riemannian almost product structure on $\ts$ with $\dim \ts=\tm$, then \rnum{ \item $I^2=1$ \item All eigenvalues are $\pm1$. \item $\hbox{\rm tr} I=2k-\tm$, where $k$ is the number of positive eigenvalues. \item $I\in Gr(k,\tm)\equiv O(\tm)/(O(k)\times O(\tm-k))$. \item There is a preferred base called the oriented base in which $I$ is diagonal and ordered, {\it i.e.}, it takes the form \eqnono{ I=\mx{cccccc}{1&&&&&\\ &\ddots&&&&\\ &&1&&&\\ &&&\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1&&\\ &&&&\ddots&\\ &&&&&\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1\\} } } } The almost product structure will serve as a rigging of the tangentbundle by looking at the spaces of eigenvectors to the almost product structure. \defn{ideffol}{Let I be an almost product structure on $\ts$, then $I$ defines two natural distributions of $T\ts$, denoted $\esD$ and $\esD'$ respectively, in the following way. Let \eqsnono{ \esD_x&:=&\{X\in T_x\ts:IX=X\},\\ \esD'_x&:=&\{X\in T_x\ts:IX=\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! X\},\\ } then \eqnono{ \esD:=\bigcup_{x\in\ts}\esD_x,\quad\esD':=\bigcup_{x\in\ts}\esD'_x. } } Seen as an endomorphism of the tangent bundle two projection operators can be formed from the almost product structure as it squares to one. These will now be projective mappings from the tangent bundle to these two sub-bundles defined above. \defn{projs}{From an almost product structure $I$ on a manifold $\ts$ we can define two projection operators through \eqsnono{ \ico&:=&{1\fr 2}(1+I)\\ \nco&:=&{1\fr 2}(1-I).\\ } These will be mappings in the sense $\ico:T\ts\rightarrow \esD$ and $\nco:T\ts\rightarrow \esD'$ respectively. } The Riemann metric in the triplet of a almost product manifold will now split into two parts which will be the induced metrics on these two sub-bundles of the tangent bundle. \defn{gandg}{Let $\ts$ be a riemannian or pseudo-riemannian manifold with metric, $\tg$, $I$ a reflective structure with $\ico$ and $\nco$ the corresponding projectors, then define the two associated metrics with respect to the reflective structure by \eqnono{ \ig(X,Y):=\tg(\ico X,\ico Y),\quad \nlg(X,Y):=\tg(\nco X,\nco Y) } which implies that $\tg$ splits into these two parts, {\it i.e.}, \eqnono{ \tg=\ig+\nlg. } } \subsection{Tensors associated with an almost product structure} There is one main tensor in the context of almost product manifolds and that is the deformation tensor. This tensor is most suitably decomposed into two irreducible parts namely the Nijenhuis tensor and the Jordan tensor. \defn{nijen2}{Let the triplet $(\ts,\tg,I)$ define a riemannian almost product structure and define the {\bf Nijenhuis tensor} as the measure of how much $\td_I$ fails to be a coboundary operator. The Nijenhuis tensor is thus a (2,1) tensor. Let $X,Y\in \L^1$ be vector fields on $\ts$, then the characteristics of the Nijenhuis tensor are \eqnono{ N_I(X,Y): \L^1 \times \L^1 \longmapsto \L^1 } and we define it through the quadratic action of $d_I$ on functions $f \in C^\infty (\ts)$, \eqnono{ <-N_I(X,Y), d f>:=\td_I \td_I f(X,Y). } It follows that the Nijenhuis tensor measures the failure in closure of the operator $d_I$ and can thus be considered as a kind of torsion. Alternatively, as the equivalent definition below shows, it measures the curvature of the endomorphism, {\it i.e.}, \eqnono{ N_I[X,Y]:=I([X,Y]_I)-[I(X),I(Y)] } Alternatively the Nijenhuis tensor can be seen as measuring how far this endomorphism is from being a Lie algebra homomorphism of the infinite-dimensional Lie algebra of vector fields on ${\cal M}} \newcommand{\ts}{\ul{\cal M}$. } \defn{jordantensor}{Let the triplet $(\ts,\tg,I)$ define a riemannian almost product structure, and let $\{\cdot,\cdot\}$ be the Jordan bracket. The {\bf Jordan tensor} associated to $I$, denoted $M_I$, with the following characteristics: \eqnono{ M_I: \L^1\times\L^1 \longmapsto \L^1 } is defined by, \eqnono{ M_I(X,Y):=I\{X,Y\}_I-\{IX,IY\} } where $X,Y\in\L^1$ are vector fields on $\ts$. The analogy to the Nijenhuis tensor is obvious . } Both the Nijenhuis and the Jordan tenor can be expressed entirely in terms of the covariant derivative of the almost product structure. \defn{Ideformation}{Let the triplet $(\ts,\tg,I)$ define a riemannian almost product structure. Let $\tdel$ be the Levi--Civita connection on $\ts$ and define the {\bf deformation tensor} associated with the endomorphism $I$, denoted $H_I$, with the following characteristics: \eqnono{ H_I:\L^1 \times \L^1 \longmapsto \L^1 } $H_I$ is defined by the expression \eqnono{ H_I(X,Y):=(I\tdel_X I-\tdel_{IX}I)(Y), } where $X,Y\in \L^1$ are two vector fields on $\ts$. An equivalent definition is given by, \eqnono{ H_I(X,Y):=N_I(X,Y)+M_I(X,Y). } } Looking at the characteristic tensors of a distribution the deformation tensor of an almost product structure can now be decomposed into the deformation tensors of the two complementary distributions defined by the almost product structure. \defn{deform}{Let $\esD$ be a $k$-distribution with projection $\ico$ on a riemannian manifold $\ts$ with non-degenerate metric $\tg$. Let $\tdel$ be the Levi--Civita connection with respect to this metric and let $\nco:=1-\ico$ be the coprojection of $\esD$. Now define the following tensors with characteristics \eqal{2}{ H,L,K:&\quad&\L^1_\esD\times\L^1_\esD &\longmapsto \L^1_{\esD'}\\ \kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}:&&\L^1_{\esD'}&\longmapsto \dsR } and \eqal{3}{ (i)&& \quad H(X,Y) &:=\nco\tdel_{\ico X}\ico Y &\quad&\text{\bf deformation tensor},\\ (ii)&& L(X,Y) &:=\frac{1}{2}(H(X,Y)-H(Y,X)) &&\text{\bf twisting tensor},\\ (iii)&& K(X,Y) &:=\frac{1}{2}(H(X,Y)+H(Y,X)) &&\text{\bf extrinsic curvature tensor},\\ (iv)&& \ri\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J} &:=\hbox{\rm tr} H &&\text{\bf mean curvature tensor},\\ (v)&& W(X,Y) &:=K(X,Y)-{1\fr k}\ri\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J} \ig(X,Y) &&\text{\bf conformation tensor}. } This gives us the decomposition of the deformation tensor in its anti-symmetric, symmetric-traceless and trace parts accordingly, \eqnono{ H=L+W+{1\fr k}\ri\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}\ig. } } The extrinsic curvature tensor and the twisting tensor can be written in a more elegant fashion. \prop{defrel}{Let $\esD$ be a distribution on a manifold $\ts$ with metric $\tg$, let further $g(X,Y)=\tg(\ico X,\ico Y)$ be the induced metric on the distribution, then the symmetric part of the deformation tensor can be written like \eqnono{ K(X,Y)(\varphi)=\!{\raise1pt\hbox{$\scriptstyle{-}$}}\!{1\fr 2}\Lie_{\ri\varphi'}\ig(X,Y),\;or\quad\li K(X,Y,Z)=\!{\raise1pt\hbox{$\scriptstyle{-}$}}\!{1\fr 2}\Lie_{Z'}\ig(X,Y), } where the prime denotes projection along the normal directions by $\nco$. The relation for the anti-symmetric part on the other hand is \eqnono{ L(X,Y)={1\fr 2}\nco [\ico X,\ico Y] } } The conformal properties of the irreducible parts of the deformation tensor can be found in next proposition. \prop{defrela}{Let $\ts$ be a riemannian manifold with metric $\tg$, let $I$ be an almost product structure on $\ts$ which split the metric in $\tg=\ig+\nlg$ and let $\lambda} \def\gol{{\mathfrak l}} \def\scL{{\scr L}} \def\dsL{{\mathbb L}} \def\clL{{\cal L}=e^{2\phi}$ be a conformal transformation on $\ig$, {\it i.e.}, $\ct\tg=\lambda} \def\gol{{\mathfrak l}} \def\scL{{\scr L}} \def\dsL{{\mathbb L}} \def\clL{{\cal L}\tg$ then the symmetric parts of the deformation tensor will transform like \eqsnono{ \ct K(\varphi)&=&K(\varphi)-\frac{1}{2}\lambda} \def\gol{{\mathfrak l}} \def\scL{{\scr L}} \def\dsL{{\mathbb L}} \def\clL{{\cal L}^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}\ri\varphi'[\lambda} \def\gol{{\mathfrak l}} \def\scL{{\scr L}} \def\dsL{{\mathbb L}} \def\clL{{\cal L}]g=K(\varphi)-\ri\varphi[\phi]g\\ \ct\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}(X)&=&\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}(X)-\frac{1}{2}k\lambda} \def\gol{{\mathfrak l}} \def\scL{{\scr L}} \def\dsL{{\mathbb L}} \def\clL{{\cal L}^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1} X'[\lambda} \def\gol{{\mathfrak l}} \def\scL{{\scr L}} \def\dsL{{\mathbb L}} \def\clL{{\cal L}]=\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}(X)-kX'[\phi]\\ \ct W&=&W\\ \ct L&=&L } } Now denoting the deformation tensor of the complementary distribution $\esD'$ by $H'$ and its irreducible parts by $L',K',\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}',W'$ respectively we can express the Nijenhuis tensor and the Jordan tensor in terms of these characteristic tensors. \lem{nijenigen}{Let $I$ be an almost product structure on a manifold $\ts$ and let its associated projection operators be $\ico:={1\fr 2}(1+I)$, $\nco:={1\fr 2}(1-I)$, then \eqal{2}{ (i)&&\qquad N_{\ico}=&N_{\nco}\cr (ii)&& N_I=&4N_{\ico}\cr (iii)&& {1\fr 2}[\ico,\nco ]=&N_{\ico}\cr (iv)&&\qquad N_{\ico}(X,Y)=&-\nco [\ico X,\ico Y]-\ico [\nco X, \nco Y] } } \prop{nijol}{Let the triplet $(\ts,\tg,I)$ define an riemannian almost product structure and let $L,\; L'$ be the twisting tensors of the distributions defined by $I$. Then \eqnono{ {1\fr 8}N_I=-L-L' } } \lem{almostjordan}{Let $I$ be an almost product structure on a manifold $\ts$ and let its associated projection operators be $\ico:={1\fr 2}(1+I)$, $\nco:={1\fr 2}(1-I)$. Let $M$ denote the Jordan tensor, then \eqal{2}{ (i)&&\qquad M_{\ico}=&M_{\nco}\cr (ii)&& M_I=&4M_{\ico}\cr (iii)&& {1\fr 2}\{\ico,\nco \}=&M_{\ico}\cr (iv)&&\qquad M_{\ico}(X,Y)=&-\nco \{\ico X,\ico Y\}-\ico \{\nco X, \nco Y\} } } \prop{almostjordan2}{Let the triplet $(\ts,\tg,I)$ define a riemannian almost product structure, $K,\; K'$ be the extrinsic curvature tensors of the distributions defined by $I$, then \eqnono{ {1\fr 8}M_I=-K-K' } } \subsection{Three relevant connections} In a almost product manifold there are three different connections of importance, which will be defined in this subsection. The first is of course the Levi--Civita connection, from which the other two will be defined by simply adding a tensor to it. These will be refered to as the adapted and the Vidal connection. Their basic feature is that they commute with the almost product structure which means that they respect the rigging of the tangent space defined by the almost product structure. The additional feature of the adapted connection is that it is metric which together with the above feature implies that it respects the induced metrics on the two characteristic distributions associated with the almost product structure. The Vidal connection which is metric iff the characteristic distributions are geodesic will play an important role when looking at gauge theories and fiber bundles since they need no metric in the total space and are of the type $(GF,GD)$. By adding the group metric to the fiber we can construct an almost product manifold in which the Vidal connection will reduce to the gauge covariant derivative. This will be explicitly done in section 5. The curvature components of the Vidal connections lying entirely in the tensor algebra of the characteristic distributions, also called the semi-basic parts, will in a more natural way measure the curvature in the respective distributions. This is due to the fact that it does not depend on the connections in its co-parts. What this means explicitly will become clear when the relations are derived. \defn{adap}{Let $\ts$ be a riemannian or pseudo-riemannian manifold with non-degenerate metric $\tg$ and corresponding Levi--Civita connection $\tdel$. Let $I$ define distributions as in definition \ref{ideffol}. Then the following two definitions of the {\bf adapted connection} are equivalent \rnum{ \item $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_XY:=\tdel_XY+A(X,Y), \quad A(X,Y):={1\fr 2}I\tdel_XI(Y)$ \item $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_XY:=\ico\tdel_X\ico Y+\nco\tdel_X\nco Y$ } } \defn{rdel}{Let $\ts$ be a riemannian or pseudo-riemannian manifold with non-degenerate metric $\tg$ and corresponding Levi--Civita connection $\tdel$, let $I$ define a foliation. Then the {\bf Vidal connection} is defined by \eqnono{ \rdel_XY:=\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_XY+B(X,Y),\quad B(X,Y):={1\fr 4}(\tdel_{IY}I+I\tdel_YI)(X). } } \prop{Vidalconnection}{Let $\underline{X},\underline{Y}$ be vectorfields on an almost product manifold $(\ts,\tg,I)$, let $X=\ico \underline{X},X'=\nco\underline{X}$ and similar for $\underline{Y}$. Then the Vidal connection can be written \eqnono{ \rdel_{\underline{X}}\underline{Y}=\mx{cc}{\ico \tdel_X Y& \nco [X,Y']\\\ico [X',Y]&\nco\tdel_{X'}Y'} } } The two recently introduced tensors, $A$ and $B$, are in fact related. \prop{boa}{Let $B$ be the tensor defined in \ref{rdel} and $A$ the tensor defined in \ref{adap}, then it is possible express the tensor $B$ in terms of $A$ and the almost product structure $I$ as \eqnono{ B(X,Y)={1\fr 2}\left(A(Y,X)-A(IY,IX)\right). } } The most important property of the two connections defined above, is that they both commute with the almost product structure. \prop{aroni}{Let $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}$ denote the adapted connection defined in \ref{adap} and $\rdel$ the Vidal connection defined in \ref{rdel}, then their principal feature is that they both commute with the almost product structure $I$, {\it i.e.}, \eqnono{ \tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_XI=\rdel_XI=0 } } Only one of them though will, in the generic case, be metric and that is the adapted connection. \prop{adelmetric}{Let the triplet $(\ts,\tg,I)$ be a riemannian almost product structure on $\ts$ and $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}$ the adapted connection defined in \ref{adap}, then this connection is metric with respect to the splitting of $\tg$ according to \ref{gandg}, {\it i.e.}, \eqal{1}{ \tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}} \ig&=0\\ \tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}} \nlg&=0 } } The connection components takes a most pleasant form in the oriented basis. The notation, $$\tE_{\ba}=(\tE_a,\tE_{a'})$$ will be used, where unprimed(primed) index denotes the basis of the characteristic unprimed(primed) distribution. \prop{del}{Let the triplet $(\ts,\tg,I)$ define a riemannian almost product structure, let $\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}$, $\tilde{\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}}$ and $\Tilde{\Tilde{\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}}}$ denote the connection one-forms of the Levi--Civita connection, the adapted connection and the Vidal connection respectively. Let furthermore $H,H'$ denote the deformation tensors with respect to $I$ and $C,C'$ be coefficients of anholonomy Then \eqsnono{ \underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}&=&\left[\mx{cc}{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}&H\\-H^t&\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}},\mx{cc}{\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}'&H'\\-{H'}^t&\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}'}\right]\\ \tilde{\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}}&=&\left[\mx{cc}{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}&\;\;0\;\;\\\;\;0\;\;&\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}},\mx{cc}{\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}'&\;\;0\;\;\\\;\;0\;\;&\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}'}\right]\\ \tilde{\tilde{\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}}}&=&\left[\mx{cc}{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}&\;\;0\;\;\\\;\;0\;\;&C},\mx{cc}{C'&\;\;0\;\;\\\;\;0\;\;&\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}'}\right]\\ } } The coefficients of anholonomy are defined by $[E_{\ba},E_{\bb}]=:C_{\ba\bb}{}^{\bc}E_{\bc}$ and can be used to express all the connection components and the deformation tensors. It is interesting to note their behaviour under a local $O(k)\times O(k')$ transformation: \eqal{2}{ \tilde{C}_{ab}{}^{c}=&u_a{}^du_b{}^eC_{de}{}^fu^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}_f{}^c+2\tilde{E}_{[a}[u_{b]}{}^f]u^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}_f{}^c,&\quad& \text{Coefficients of anholonomy}\\ \tilde{C}_{ab}{}^{c'}=&u_a{}^du_b{}^eC_{de}{}^{f'}u^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}_{f'}{}^{c'},&\quad&\text{Tensor}\\ \tilde{C}_{a'b}{}^{c}=&u_{a'}{}^{d'}u_b{}^eC_{d'e}{}^fu^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}_f{}^c+\tilde{E}_{a'}[u_{b}{}^f]u^{\!{\raise1pt\hbox{$\scriptstyle{-}$}}\! 1}_f{}^c,&\quad&\text{Connection}\\ } This is what enables the defininition of the Vidal connection, i.e. the observation that the $C_{ab'}{}^{c'},C_{a'b}{}^{c}$ parts of the structure coefficients transform as connections under local $O(k)\times O(k')$ transformations. \prop{sructure-connection}{Let $\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V},\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z},H,L,K$ be the components of the Levi-Civita connection and $C$ the coefficients of anholonomy, then they are related through the following relations, \eqal{1}{ C_{abc}&=2\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{[ab]c}\\ C_{a'bc}&=\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}'_{a'bc}+H_{bca'}\\ C_{abc'}&=2L_{abc'}\\ C_{a'b'c}&=2L'_{a'b'c}\\ C_{ab'c'}&=\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{ab'c'}+H'_{b'c'a}\\ C_{a'b'c'}&=2\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}'_{[a'b']c'} } with inverse relations \eqal{1}{ \omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{abc}&=C_{a[bc]}-{1\fr 2}C_{bca}={1\fr 2}C_{abc}+C_{c(ab)}\\ \Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}'_{a'bc}&=C_{a'[bc]}-{1\fr 2}C_{bca'}\\ K_{abc'}&=C_{c'(ab)}\\ L_{abc'}&={1\fr 2}C_{abc'}\\ H_{abc'}&={1\fr 2}C_{abc'}+C_{c'(ab)}\\ H'_{a'b'c}&={1\fr 2}C_{a'b'c}+C_{c(a'b')}\\ L'_{a'b'c}&={1\fr 2}C_{a'b'c}\\ K'_{a'b'c}&=C_{c(a'b')}\\ \Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{ab'c'}&=C_{a[b'c']}-{1\fr 2}C_{b'c'a}\\ \omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}'_{a'b'c'}&=C_{a'[b'c']}-{1\fr 2}C_{b'c'a'}={1\fr 2}C_{a'b'c'}+C_{c'(a'b')}\\ } } \section{The curvature components and their relations} As fiber bundles and fibrations are examples of almost product manifolds with the additional property of the existense of a surjective submersion of the total space down to a base space, it would be interesting to see what parts of the Vidal and adapted connections, which are defined on the total space, that survive under this submersion. In the total space two new differential operators were defined in $\cite{Holm98}$. \defn{nybound}{Let $I$ be an almost product structure on a manifold $\ts$ with exterior derivative $\td$. Let furthermore $\td_I$ denote the exterior derivative associated with $I$ and define two new differential operators by \eqsnono{ \id&:=&{1\fr 2}(\td+\td_I)\\ \nd&:=&{1\fr 2}(\td-\td_I) } An equivalent definition is by the two projection operators defined by the endomorphism $I$, $\ico:={1\fr 2}(1+I)$ and $\nco:={1\fr 2}(1-I)$, then the operators are simply $\id\equiv\td_\ico$ and $\nd\equiv\td_{\nco}$. } These differential operators become coboundary operators if and only if the Nijenhuis tensor vanishes, which is the same as to say that both the characteristic distributions of the almost product structure are integrable. In a fibration for instance this is not normally true, except for the trivial case of a product manifold, so we will keep track of all components surviving the submersion and those who will not. When it comes to these differential operators it is therefor clear that $d$ defined above will in general differ from the exterior derivative defined on the base space. By projecting out the semi-basic parts of all quantities we can keep track of the parts that survives the submersion. \defn{distributionbrackets}{Let the triplet $(\ts,\tg,I)$ denote a riemannian almost product structure and $\esD$, $\esD'$ be the associated distributions then define the brackets associated with these distributions with following characteristics \eqal{1}{ [\;\cdot\;,\;\cdot\;]^{\esD}:&\quad \L_{\esD}^1\times \L_{\esD}^1 \longmapsto \L^1_{\esD},\\ [\;\cdot\;,\;\cdot\;]^{\esD'}:&\quad \L_{\esD'}^1\times \L_{\esD'}^1 \longmapsto \L^1_{\esD'}, } by \eqal{1}{ [X,Y]^{\esD}&:=\ico [\ico X, \ico Y],\\ [X,Y]^{\esD'}&:=\nco [\nco X, \nco Y], } where $X,Y \in \L^1$ are vectorfields on $\ts$. } Here it is clear that the twisting tensor that measures the amount of non-commutativity is non-semi-basic. The two brackets defined above will therefor not satisfy the Jacobi identity in the total space, but will differ with some terms involving the twisting tensors. The same procedure can be made to define the semi-basic torsions and curvature tensor of a distribution. \defn{vert-horiz}{Let the triplet $(\ts,\tg,I)$ denote a riemannian almost product structure and $\esD$, $\esD'$ be the associated distributions. Let further $T^p_q(T\ts)$ denote the set of $(p,q)$-tensors on $\ts$ and $T^p_q(\esD)$ ($T^p_q(\esD')$) denote the set of tensors lying entirely in $\esD$ ($\esD'$). Now define the associated Levi--Civita connections with following characteristics \eqal{1}{ \tdel^{\esD}_X:&\quad T^p_q(T\ts)\longmapsto T^p_q(\esD),\\ \tdel^{\esD'}_X:&\quad T^p_q(T\ts)\longmapsto T^p_q(\esD'), } through \eqal{1}{ \tdel^{\esD}_X Y:=&\ico \tdel_{\ico X} \ico Y,\\ \tdel^{\esD'}_X:=&\nco \tdel_{\nco X} \nco Y, } where $X,Y \in \L^1$ are vectorfields on $\ts$. Further define the torsion and curvature of the corresponding connections by \eqal{1}{ T^{\esD}(X,Y)&:=\tdel^{\esD}_X Y- \tdel^{\esD}_Y X-[X,Y]^{\esD},\\ T^{\esD'}(X,Y)&:=\tdel^{\esD'}_X Y- \tdel^{\esD'}_Y X-[X,Y]^{\esD'}, } and \eqal{1}{ R^{\esD}(X,Y)Z&:=\tdel^{\esD}_X \tdel^{\esD}_Y Z- \tdel^{\esD}_Y \tdel^{\esD}_X Z-\tdel^{\esD}_{[X,Y]^{\esD}}Z,\\ R^{\esD'}(X,Y)Z&:=\tdel^{\esD'}_X \tdel^{\esD'}_Y Z- \tdel^{\esD'}_Y \tdel^{\esD'}_X Z-\tdel^{\esD'}_{[X,Y]^{\esD'}}Z. } } In this case it is clear that the curvature defined above, will not in general be tensorial in the latter indices. As before these non-tensorial parts will vanish under the submersion. The torsion will still be tensorial though. \subsection{The Vidal connection} In this subsection, all tensors associated with the Vidal connection will be derived, that is the torsion tensor, the non-metricity tensor, the Riemann tensor and its traces. The curvature identities are used to express all components but the two totally semi-basic ones only in terms of the different irreducible parts of the deformation tensor and its derivatives. From these curvature identities there also arise a couple of new relations involving only parts of the deformation tensor. Two of these will evidently become the Bianchi identity of the two twisting tensors but two others will appear in a more unfamiliar fashion. From the definition it is clear that the Vidal connection is neither torsion-free, nor metric in the generic case of an almost product manifold. It is therefor interesting to see what the torsion and non-metricity tensor look like in this case. In ref. \cite{Holm98} the following proposition was derived. \prop{rtor}{Let the triplet $(\ts,\tg,I)$ define an almost product manifold, let $N_I$ denote the Nijenhuis tensor of $I$ and $\rdel$ denote the Vidal connection defined in \ref{rdel}, then, \eqnono{ {1\fr 4}N_I(X,Y)=\rT(X,Y). } } Together with proposition \ref{nijol} the torsion tensor can be written in component form. \prop{vidal torsion}{Let $\rT$ be the torsion tensor of the Vidal connection, $\rdel$, then in component form it reads \eqal{1}{ \rT_{ab}{}^{c}=&0,\\ \rT_{ab}{}^{c'}=&-2L_{ab}{}^{c'},\\ \rT_{a'b}{}^{c}=&0,\\ \rT_{ab'}{}^{c'}=&0,\\ \rT_{a'b'}{}^{c}=&-2L'_{a'b'}{}^{c},\\ \rT_{a'b'}{}^{c'}=&0. } and \eqal{1}{ \rT_a=&0,\\ \rT_{a'}=&0. } } As was seen in \cite{Holm98} the torsion tensor measures the non-integrability of the two complementary distributions defined by an almost product structure. The non-metricity of the Vidal connection is put in the next proposition. \prop{vidal-metric}{Let the triple $(\ts,\tg,I)$ denote a riemannian almost product structure with associated metrical decomposition $\tg=\ig+\nlg$. Let further $\rdel$ denote the Vidal-connection then the following relations hold \eqal{1}{ (\rdel_Z\ig)(X,Y)&=0\\ (\rdel_{Z'}\ig)(X,Y)&=-2K_{Z'}(X,Y)=(\Lie_{Z'}\ig)(X,Y)\\ (\rdel_Z\nlg)(X',Y')&=-2K'_{Z}(X',Y')=(\Lie_{Z}\nlg)(X',Y')\\ (\rdel_{Z'}\nlg)(X',Y')&=0\\ } } In component form the non-metricity tensor can be read off from the next proposition. \prop{vidal non-metricity}{Let $\Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}$ be the non-metricity tensor of the Vidal connection, $\rdel$, then in component form it reads \eqal{1}{ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_{abc}=&0\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_{a'bc}=&-2K_{bca'}\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_{abc'}=&0\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_{a'b'c}=&0\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_{ab'c'}=&-2K'_{b'c'a}\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_{a'b'c'}=&0 } The two traces following from a three tensor symmetric in two indices is found to be \eqal{1}{ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}^1_a=&0\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}^1_{a'}=&0\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_a:=&\Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}^2_a=-2\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_a\\ \Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}_{a'}:=&\Tilde{\Tilde{\ul{Q}}}}\newcommand{\rV}{\Tilde{\Tilde{\ul{V}}}^2_{a'}=-2\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{a'}\\ } } The basic curvature relations are found from definitions \ref{Rcurvature} and \ref{vert-horiz}, they are given in the proposition below, \prop{vidalnoindex}{Let $\rdel$ be the Vidal connection and $\rR$ denote the curvature tensor theorof then the different parts reads \eqal{1}{ \rR(X,Y)Z&=R^{\esD}(X,Y)Z-2\ico [L(X,Y),Z]\\ \rR(X,Y)Z'&=-2(\rdel_{Z'}L)(X,Y)\\ \rR(X,Y')Z&=\ico\nabla_{X}\ico[Y',Z]-\ico[Y',\ico\nabla_{X}Z]-\ico\nabla_{\ico[X,Y']}Z-\ico[\nco[X,Y'],Z]\\ } } In order to deal with the curvature identities it is convenient to put these relations in component form. \prop{vidalcurvatures}{Let $\rR$ be the Riemann-tensor with respect to the Vidal-connection, then $\rR$ has the following components, \eqal{1}{ \rR_{abc}{}^d=&R^{\esD}_{abc}{}^d-2L_{ab}{}^{e'} C_{e'c}{}^d\\ \rR_{abc}{}^{d'}=&0\\ \rR_{abc'}{}^d=&0\\ \rR_{abc'}{}^{d'}=&-2(\rdel_{c'}L)_{ab}{}^{d'}\\ \rR_{a'bc}{}^d=&E_{a'}[{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{bc}{}^d]-E_b[C_{a'c}{}^d]-C_{a'b}{}^e{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{ec}{}^d-C_{a'b}{}^{e'}C_{e'c}{}^{d}- C_{a'c}{}^e{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{be}{}^d+\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{bc}{}^eC_{a'e}{}^d\\ \rR_{a'bc}{}^{d'}=&0\\ \rR_{a'bc'}{}^d=&0\\ \rR_{a'bc'}{}^{d'}=&E_{a'}[{C}_{bc'}{}^{d'}]-E_{b}[\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{a'c'}{}^{d'}]-C_{a'b}{}^{e'}{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{e'c'}{}^{d'}-C_{a'b}{}^{e}C_{ec'}{}^{d'}- \omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{a'c'}{}^{e'}{C}_{be'}{}^{d'}+C_{bc'}{}^{e'}\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{a'e'}{}^{d'}\\ \rR_{a'b'c}{}^d=&-2(\rdel_{c}L)_{a'b'}{}^d\\ \rR_{a'b'c}{}^{d'}=&0\\ \rR_{a'b'c'}{}^d=&0\\ \rR_{a'b'c'}{}^{d'}=&R^{\esD'}_{a'b'c'}{}^{d'}-2L_{a'b'}{}^{e}C_{ec'}{}^{d'} } } This is the ``raw'' expressions for the curvature components of the Vidal connection, but after using the identities of the Riemann tensor seen in section \ref{curvatureidentities} these will simplify remarkably. There will also appear a couple of identities involving only parts of the deformation tensor. Starting with the second identity it can be seen, using the Tic-Tac-Toe notation introduced in appendix A, that the second identity can be split according to the rigging into 8 irreducible parts, corresponding to the following young tableaux. \eqal{1}{ \yng(1,1,1)\otimes\yng(1)=&\left(\young(o,o,o)\otimes\young(o)\right)\oplus\left(\young(o,o,o)\otimes\young(x)\right)\oplus\left(\young(o,o,x)\otimes\young(o)\right)\oplus \left(\young(o,o,x)\otimes\young(x)\right)\oplus\\ &\left(\young(o,x,x)\otimes\young(o)\right)\oplus\left(\young(o,x,x)\otimes\young(x)\right)\oplus \left(\young(x,x,x)\otimes\young(o)\right)\oplus\left(\young(x,x,x)\otimes\young(x)\right) } Here will be listed only the first half of the identities as they of course are symmetric upon changing primes and unprimes. \eqs{ \rR_{[abc]}{}^{d}&=&0\\ \rdel_{[a}L_{bc]}{}^{d'}&=&0\\ \rR_{c'[ab]}{}{d}&=&-L_{ab}{}^{e'}L'_{c'e'}{}^{d}\\ \rR_{abc'}{}^{d'}&=&-2\rdel_{c'}L_{ab}{}^{d'} } So in conclusion the first identity is structurally inherited by the totally semi-basic part of the Vidal curvature, the second leads to a bianchi identity for the twisting tensor, the third relating the anti-symmetric part of $\rR_{c'[ab]}{}^{d}$ in terms of the twisting tensors and the fourth a faster way of deriving the $\rR_{abc'}{}^{d'}$ component. Thus, the following proposition is proved: \prop{twistbianchi}{Let $\rdel$ be the Vidal connection associated with a almost product structure and $L,L'$ be the respective twisting tensors of the associated distributions then the following Bianchi identities hold, \eqal{1}{ \rdel_{[a}L_{bc]}{}^{d'}=&0\\ \rdel_{[a'}L'_{b'c']}{}^{d}=&0 } } The third identity which is decomposed as, \eqal{1}{ \yng(1,1)\otimes\yng(2)=&\left(\young(o,o)\otimes\young(oo)\right)\oplus\left(\young(o,x)\otimes\young(oo)\right)\oplus\left(\young(o,o)\otimes\young(ox)\right)\oplus\\ & \left(\young(o,o)\otimes\young(xx)\right)\oplus\left(\young(o,x)\otimes\young(ox)\right)\oplus \left(\young(x,x)\otimes\young(oo)\right)\oplus\\&\left(\young(x,x)\otimes\young(ox)\right)\oplus\left(\young(o,x)\otimes\young(xx)\right)\oplus \left(\young(x,x)\otimes\young(xx)\right) } In this case there are 9 irreducible parts of which five will be listed and the other follows due to symmetry. \eqs{ \rR_{ab(cd)}&=&-2L_{ab}{}^{e'}K_{cde'}\\ \rR_{a'b(cd)}&=&-(\rdel_{b}K)_{cda'}\\ 0&=&0,\\ \rR_{ab(c'd')}&=&2(\rdel_{[a}K')_{c'd'|b]}\\ 0&=&0 } The first identity here gives no new information, the second gives yet another part of the $\rR_{a'bcd}$ component, the third and the fifth contain nothing while the fourth together with the original expression for the $\rR_{abc'd'}$ component gives a new non-trivial identity which proves the next proposition. \prop{vidal-identity}{Let $\rdel$ be the Vidal-connection, $K,L,K',L'$ be the second fundamental tensors of a almost product structure then we have the following identities \eqal{1}{ \rdel_{[Z}K'_{W]}(X',Y')+\rdel_{(X'}L_{Y')}(Z,W)&=0\\ \rdel_{[Z'}K_{W']}(X,Y)+\rdel_{(X}L'_{Y)}(Z',W')&=0 } In component form the same expressions read \eqal{1}{ (\rdel_{[a}K')_{c'd'|b]}+(\rdel_{(c'}L)_{ab|d')}=&0\\ (\rdel_{[a'}K)_{cd|b']}+(\rdel_{(c}L')_{a'b'|d)}=&0 } Contracted the identities read \eqal{1}{ (\rdel_{[a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_{b]}+(\rdel_{c'}L)_{ab}{}^{c'}=&0\\ (\rdel_{[a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{b']}+(\rdel_{c}L)_{a'b'}{}^{c}=&0 } } Here the notation $$L_{X'}(Y,Z):=\tg(L(Y,Z),X')$$ is used. The last identity of the Riemann curvature is the most non-trivial of them all. It can be decomposed through the box symmetry, i.e. \eqal{1}{ \yng(2,2)=\young(oo,oo)\oplus\young(oo,ox)\oplus\young(oo,xx)\oplus\young(ox,ox)\oplus\young(xx,xo)\oplus\young(xx,xx). } The only identity of these which gives new information is the second why it is the only one listed. This identity though gives the opportunity of writning the entire $\rR_{a'bcd}$ component purely in terms of the different parts of the deformation tensor and its derivatives. \eq{ \rR_{a'bcd}=-(\rdel_bK)_{cda'}-2(\rdel_{[c}K)_{d]ba'}+2L_{cd}{}^{e'}L'_{a'e'b}-4L_{b[c}{}^{e'}L'_{a'e'|d]} } This relation together with the identity in \ref{vidal-identity} prooves the following proposition which is the final form of the Vidal curvature components. \prop{vidal curvature}{Let $\rR$ be the curvature tensor of the Vidal connection then its components can be written \eqal{1}{ \rR_{abcd}=&R^{\esD}_{abcd}-2L_{ab}{}^{e'} C_{e'cd}\\ \rR_{abcd'}=&0\\ \rR_{abc'd}=&0\\ \rR_{abc'd'}=&2[(\rdel_{[a}K')_{c'd'|b]}-(\rdel_{[c'}L)_{ab|d']}]\\ \rR_{a'bcd}=&-(\rdel_bK)_{cda'}-2(\rdel_{[c}K)_{d]ba'}+2L_{cd}{}^{e'}L'_{a'e'b}-4L_{b[c}{}^{e'}L'_{a'e'|d]}\\ \rR_{a'bcd'}=&0\\ \rR_{a'bc'd}=&0\\ \rR_{a'bc'd'}=&(\rdel_{a'}K')_{c'd'b}+2(\rdel_{[c'}K')_{d']a'b}-2L'_{c'd'}{}^{e}L_{bea'}+4L'_{a'[c'}{}^{e}L_{be|d']}\\ \rR_{a'b'cd}=&2[(\rdel_{[a'}K)_{cd|b']}-(\rdel_{[c}L')_{a'b'|d]}],\\ \rR_{a'b'cd'}=&0\\ \rR_{a'b'c'd}=&0\\ \rR_{a'b'c'd'}=&R^{\esD'}_{a'b'c'd'}-2L_{a'b'}{}^{e}C_{ec'd'} } } From the final expressions of the Vidal curvature components the Schouten two-form and the Ricci tensor can be derived. For the Schouten two-form it is easily seen that it ends up as a total exterior derivative of the two mean curvatures by looking at the trace of the curvature two-form in the Cartan formalism, namely, \eq{ V:=R_{\bar{c}}{}^{\bar{c}}=\td\Tilde{\Tilde{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}}_{\bar{c}}{}^{\bar{c}}=\td\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}+\td\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'. } Now of course $\Tilde{\Tilde{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}}$ is only a local object, and therefor it is not sure that $V$ can be written as an exact form globally - this leaves us with the following proposition: \prop{vidal schouten}{Let $\rdel$ be the Vidal connection then the Schouten two-form, $\rV$, of the Vidal connection can locally be written \eqal{1}{ V=d\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_I } where $\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_I=\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}+\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'$. From the integrability condition $dV=0$ it is obvious that \eqnono{ V\in H^2(\ts), } where $H^2(\ts)$ denotes the second cohomology group of the manifold $\ts$. In component form the Schouten two-form looks like \eqal{1}{ \rV_{ab}=&2(\rdel_{[a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_{b]}-2L_{ab}{}^{c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{c'},\\ \rV_{a'b}=&(\rdel_{a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_b-(\rdel_{b}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{a'},\\ \rV_{a'b'}=&2(\rdel_{[a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{b']}-2L_{a'b'}{}^{c}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{c}.\\ } } Now finally the Ricci tensor and the curvature scalar of the Vidal connection can easily be derived. \prop{vidal ricci}{Let $\rdel$ be the Vidal connection then the Ricci tensor reads in component form \eqal{1}{ \rR_{ab}=&R^{\esD}_{ab}-2L_{ac}{}^{e'}C_{e'b}{}^c\\ \rR_{a'b}=&-(\rdel_b\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{a'}+4L_{bc}{}^{e'}L'_{a'e'}{}^c\\ \rR_{ab'}=&-(\rdel_{b'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_{a}+4L_{ac}{}^{e'}L'_{b'e'}{}^c\\ \rR_{a'b'}=&R^{\esD'}_{a'b'}-2L'_{a'c'}{}^{e}C_{eb'}{}^{c'}\\ } and the Riemann curvature scalar is given by \eqnono{ \rR=\Tilde{\Tilde{R}}+\Tilde{\Tilde{R}}'=R^{\esD}+R^{\esD'}-2L_{ab}{}^{c'}C_{c'}{}^{ab}-2L'_{a'b'}{}^{c}C_{c}{}^{a'b'} } } \subsection{The adapted connection} In direct analogy to the previous section, several curvature- and torsion relations are derived with respect to the adapted connection. In contrast to the Vidal-connection the adapted one is metric. In this case the Nijenhuis-tensor is related to the torsion in the way showed by the next proposition. The metricity of the connection has its price though, as is seen below the torsion tensor is more complicated in this case. Some generalized Bianchi identities for the twisting tensor $L$ is also yielded in this case. All tensors, except the totally semi-basic ones, derived from the Riemann tensor, are expressed in terms of the irreducible parts of the deformation tensor. \prop{ator}{Let the triplet $(\ts,\tg,I)$ define an riemannian almost product structure, let $N_I$ denote the Nijenhuis tensor of $I$ and $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}$ the adapted connection defined in \ref{adap}, then we have the following relation, \eqnono{ {1\fr 2}N_I(X,Y)=\tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}(X,Y)+\tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}(IX,IY) } } \prop{adapted torsion}{Let $\tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}$ be the torsion tensor of the adapted connection, $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}$, then in component form it reads, \eqal{1}{ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_{ab}{}^{c}=&0\\ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_{ab}{}^{c'}=&-2L_{ab}{}^{c'}\\ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_{a'b}{}^{c}=&-H_b{}^c{}_{a'}\\ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_{ab'}{}^{c'}=&-H'_{b'}{}^{c'}{}_a\\ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_{a'b'}{}^{c}=&-2L'_{a'b'}{}^{c}\\ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_{a'b'}{}^{c'}=&0. } and \eqal{1}{ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_a=&-\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_a\\ \tilde{\tT}} \newcommand{\rT}{\Tilde{\Tilde{\tT}}_{a'}=&-\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{a'} } } The curvature components can further be simplified as shown in the next proposition. Notice that they are expressed in terms of the Vidal connection. \prop{adaptedcurvatures}{Let $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}$ be the Riemann-tensor with respect to the adapted connection, then $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}$ has the following components, \eqal{1}{ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc}{}^d=&\rR_{abc}{}^d+2L_{ab}{}^{e'}H_{c}{}^d{}_{e'}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc}{}^{d'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'}{}^d=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'}{}^{d'}=&\rR_{abc'}{}^{d'}-2(\rdel_{[a}H')_{c'}{}^{d'}{}_{|b]}-2H'_{c'}{}^{e'}{}_{[a}H'_{e'}{}^{d'}{}_{|b]}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc}{}^d=&\rR_{a'bc}{}^d+(\rdel_bH)_c{}^d{}_{a'}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc}{}^{d'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc'}{}^d=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc'}{}^{d'}=&\rR_{a'bc'}{}^{d'}-(\rdel_{a'}H)_{c'}{}^{d'}{}_{b}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c}{}^d=&\rR_{a'b'c}{}^{d}-2(\rdel_{[a'}H)_{c}{}^{d}{}_{|b']}-2H_{c}{}^{e}{}_{[a'}H'_{e}{}^{d}{}_{|b']}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c}{}^{d'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'}{}^d=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'}{}^{d'}=&\rR_{a'b'c'}{}^{d'}+2L_{a'b'}{}^{e}H_{c'}{}^{d'}{}_e } where, \eqal{1}{ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'}{}^{d'}&:=E_a[\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{bc'}{}^{d'}]-E_b[\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{ac'}{}^{d'}]-C_{ab}{}^e\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{ec'}{}^{d'}-2\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{[a|c'}{}^{e'}\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{b]e'}{}^{d'} -2L_{ab}{}^{e'}\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{e'c'}{}^{d'}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c}{}^d&:=E_{a'}[\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{b'c}{}^{d}]+E_{b'}[\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{a'c}{}^{d}]-C_{a'b'}{}^{e'}\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{e'c}{}^d-2\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{[a'|c}{}^e\Omega} \def\goz{{\mathfrak z}} \def\scZ{{\scr Z}} \def\dsZ{{\mathbb Z}} \def\clZ{{\cal Z}_{b']e}{}^d -2L'_{a'b'}{}^e\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}_{ec}{}^d } } Using proposition \ref{vidal curvature} the curvature tensor of the adapted connection can be written entirely in terms of the semi-basic parts of the Vidal curvature, parts of the deformation tensors and the Vidal covariant derivative thereof. \prop{adapted curvature vidal}{Let $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}$ be the curvature tensor of the adapted connection then its components can be written \eqal{1}{ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abcd}=&\rR_{abcd}+2L_{ab}{}^{e'} H_{cde'}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abcd'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'd}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'd'}=&-2[(\rdel_{[a}L')_{c'd'|b]}+(\rdel_{[c'}L)_{ab|d']}-H'_{c'}{}^{e'}{}_{[a}H'_{d'e'|b]}]\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bcd}=&(\rdel_bL)_{cda'}-2(\rdel_{[c}K)_{d]ba'}+2L_{cd}{}^{e'}L'_{a'e'b}-4L_{b[c}{}^{e'}L'_{a'e'|d]}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bcd'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc'd}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc'd'}=&-(\rdel_{a'}L')_{c'd'b}+2(\rdel_{[c'}K')_{d']a'b}-2L'_{c'd'}{}^{e}L_{bea'}+4L'_{a'[c'}{}^{e}L_{be|d']}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'cd}=&-2[(\rdel_{[a'}L)_{cd|b']}+(\rdel_{[c}L')_{a'b'|d]}-H_{c}{}^{e}{}_{[a'}H_{de|b']}]\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'cd'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'd}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'd'}=&\rR_{a'b'c'd'}+2L_{a'b'}{}^{e}H'_{c'd'e} } } These relations can be expressed purely in terms of the adapted connection instead of the Vidal connection. \prop{adapted curvature}{Let $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}$ be the curvature tensor of the adapted connection then its components can be written \eqal{1}{ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abcd}=&\rR_{abcd}+2L_{ab}{}^{e'} H_{cde'}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abcd'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'd}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'd'}=&-2[(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}L')_{c'd'|b]}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c'}L)_{ab|d']}-W'_{c'}{}^{e'}{}_{[a}W'_{d'e'|b]}+L'_{c'}{}^{e'}{}_{[a}L'_{d'e'|b]}+\\ &+2L_{[a|}{}^{e}{}_{c'}L_{b]ed'}+2W_{[a}{}^{e}{}_{[c'}L_{b]e|d']}+{2\fr k}L_{ab[c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{d']}]\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bcd}=&(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_bL)_{cda'}-2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c}K)_{d]ba'}-L_{cd}{}^{e'}H'_{e'a'b}-2K_{b[c|}{}^{e'}H'_{e'a'|d]}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bcd'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc'd}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc'd'}=&-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{a'}L')_{c'd'b}+2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c'}K')_{d']a'b}+L_{c'd'}{}^{e}H'_{eba'}+2K_{a'[c'|}{}^{e}H'_{eb|d']}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'cd}=&-2[(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a'}L)_{cd|b']}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c}L')_{a'b'|d]}-W_{c}{}^{e}{}_{[a'}W_{de|b']}+L_{c}{}^{e}{}_{[a'}L_{de|b']}+\\ &+2L'_{[a'|}{}^{e'}{}_{c}L'_{b']e'd}+2W'_{[a'}{}^{e'}{}_{[c}L'_{b']e'|d]}+{2\fr k'}L'_{a'b'[c}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{d]}]\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'cd'}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'd}=&0\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'd'}=&\rR_{a'b'c'd'}+2L_{a'b'}{}^{e}H'_{c'd'e} } } From previous proposition the Ricci tensor of the adpted connection can simply be deduced by contraction as the adapted connection is metric. It should be stressed that this Ricci tensor is in general not symmetric. \prop{adapted ricci}{Let $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}$ be the adapted connection then the Ricci tensor reads in component form \eqal{1}{ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{ab}=&\rR_{ab}+2L_{ac}{}^{e'}H_{b}{}^c{}_{e'}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b}=&(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c}H)_{b}{}^{c}{}_{a'}-H_{bc}{}^{e'}H'_{e'a'}{}^{c}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_b\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{a'}+\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^{e'}H'_{e'a'b}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{ab'}=&(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}H')_{b'}{}^{c'}{}_{a}-H'_{b'c'}{}^{e}H_{ea}{}^{c'}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{b'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{a}+\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'{}^{e}H_{eab'}\\ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'}=&\rR_{a'b'}+2L'_{a'c'}{}^{e}H'_{b'}{}^{c'}{}_e\\ } and the Riemann curvature scalar is given by \eqnono{ \tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}=\Tilde{R}+\Tilde{R}'=\Tilde{\Tilde{R}}+\Tilde{\Tilde{R}}'+2L_{ab}{}^{c'}L^{ab}{}_{c'}+2L'_{a'b'}{}^{c}{L'}^{a'b'}{}_{c} } } The generalized Biachi-identities for the twisting tensor $L$ are given in next proposition. They were derived by using the antisymmetry ((i) of proposition \ref{curvatureidentities}) and the components expressed in terms of the adapted connection, see proposition \ref{adapted curvature}. \prop{twistbianchi2}{Let $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}$ be the adapted connection associated with an almost product structure and $L,L',K,K'$ be the respective twisting and extrinsic curvature tensors of the associated distributions then the following identities hold \eqal{1}{ \tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}L_{bc]}{}^{d'}+L_{[ab|}{}^{e'}H_{e'}{}^{d'}{}_{|c]}=&0\\ \tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a'}L_{b'c']}{}^{d}+L_{[a'b'|}{}^{e}H_{e}{}^{d}{}_{|c']}=&0 } } As was seen in previous subsection, where the Vidal connection was studied, new identities between parts of the deformation tensors arose as integrability conditions on these while imposing the identities of the curvature tensor. These were derived in proposition \ref{adapted curvature} above and in the same fashion the corresponding identities for the adapted connection arise. \prop{adepted-identity}{Let $\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}$ be the adapted connection, $K,L,K',L'$ be the second fundamental tensors with respect to a almost product structure, $I$, then the following identities hold \eqal{1}{ (\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}K')_{c'd'|b]}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{(c'}L)_{ab|d')}-2L'_{(c'|}{}^{e'}{}_{[a}K'_{e'|d')|b]}-2K_{[a|}{}^{e}{}_{(c'}L_{e|b]d')}=&0,\\ (\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a'}K)_{cd|b']}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{(c}L')_{a'b'|d)}-2L_{(c|}{}^{e}{}_{[a'}K_{e|d)|b']}-2K'_{[a'|}{}^{e'}{}_{(c}L'_{e'|b']d)}=&0. } These identities look in the contracted case like \eqal{1}{ (\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_{b]}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}L)_{ab}{}^{c'}-2W_{[a|}{}^{e}{}_{c'}L_{e|b]}{}^{c'}-{2\fr k}L_{ab}{}^{c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{c'}=&0,\\ (\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{b']}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c}L)_{a'b'}{}^{c}-2W_{[a'|}{}^{e'}{}_{c}L_{e'|b']}{}^{c}-{2\fr k'}L_{a'b'}{}^{c}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{c}=&0. } } \subsection{The Levi-Civita connection} In this section all curvature relations for the Levi-Civita connection is given. The curvature, Ricci and the curvature scalar are expressed in terms of the irreducible components of the deformation-tensor. Starting from Cartan's structure equations and writing the curvature two-form as, \eqal{1}{ \tR_{\bc}{}^{\bd}:=&\td \underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{\bc}{}^{\bd}-\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{\bc}{}^{\bar{e}}\wedge \underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{\bar{e}}{}^{\bd}=\\ =&\mx{cc}{\td\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{c}{}^d-\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{c}{}^{e}\wedge \underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{e}{}^{d}-\underline{H}_{c}{}^{e'}\wedge \underline{H}_{e'}{}^{d},& \td \underline{H}_{c}{}^{d'}-\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{c}{}^{e}\wedge \underline{H}_{e}{}^{d'}-H_{c}{}^{e'}\wedge \underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{e'}{}^{d'}\\ \td \underline{H}_{c'}{}^{d}-\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{c'}{}^{e'}\wedge \underline{H}_{e'}{}^{d}-H_{c'}{}^{e}\wedge \underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{e}{}^{d},& \td\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{c'}{}^{d'}-\underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{c'}{}^{e'}\wedge \underline{\omega} \def\gov{{\mathfrak v}} \def\scV{{\scr V}} \def\dsV{{\mathbb V}} \def\clV{{\cal V}}_{e'}{}^{d'}-\underline{H}_{c'}{}^{e}\wedge \underline{H}_{e}{}^{d'}} } the components can be given in terms of the adapted connection. \prop{levcivcurvatures}{Let $\tR$ be the Riemann-tensor with respect to the Levi-Civita connection, then $\tR$ has the following components, \eqal{1}{ \tR_{abc}{}^d=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc}{}^d+2H_{[a|c}{}^{e'}H_{b]}{}^{d}{}_{e'}\\ \tR_{abc}{}^{d'}=&2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}H)_{b]c}{}^{d'}+2L_{ab}{}^{e'}H'_{e'}{}^{d'}{}_c\\ \tR_{abc'}{}^d=&-2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}H)_{b]}{}^{d}{}_{c'}-2L_{ab}{}^{e'}H'_{e'c'}{}^{d}\\ \tR_{abc'}{}^{d'}=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abc'}{}^{d'}+2H_{[a|}{}^{e}{}_{c'}H_{b]e}{}^{d'}\\ \tR_{a'bc}{}^d=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc}{}^{d}-H'_{a'}{}^{e'}{}_cH_{b}{}^d{}_{e'}+H_{bc}{}^{e'}H'_{a'e'}{}^d\\ \tR_{a'bc}{}^{d'}=&(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{a'}H)_{bc}{}^{d'}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{b}H')_{a'}{}^{d'}{}_c-H_b{}^e{}_{a'}H_{ec}{}^{d'}-H'_{a'}{}^{e'}{}_bH'_{e'}{}^{d'}{}_c\\ \tR_{a'bc'}{}^d=&-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{a'}H)_{b}{}^{d}{}_{c'}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{b}H')_{a'c'}{}^{d}+H_{b}{}^{e}{}_{a'}H_{e}{}^{d}{}_{c'}+H'_{a'}{}^{e'}{}_{b}H'_{e'c'}{}^{d}\\ \tR_{a'bc'}{}^{d'}=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'bc'}{}^{d'}-H'_{a'c'}{}^{e}H_{be}{}^{d'}+H_{b}{}^{e}{}_{c'}H'_{a'}{}^{d'}{}_e\\ \tR_{a'b'c}{}^d=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c}{}^{d}+2H'_{[a'|}{}^{e'}{}_cH'_{b']e'}{}^{d}\\ \tR_{a'b'c}{}^{d'}=&-2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a'}H')_{b']}{}^{d'}{}_c-2L'_{a'b'}{}^{e}H_{ec}{}^{d'}\\ \tR_{a'b'c'}{}^d=&2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a'}H')_{b']c'}{}^{d}+2L'_{a'b'}{}^{e}H_{e}{}^{d}{}_{c'}\\ \tR_{a'b'c'}{}^{d'}=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'}{}^{d'}+2H'_{[a'|c'}{}^eH'_{b']}{}^{d'}{}_e\\ } } These are also known as the Gauss-Codazzi relations. In the case of the Levi--Civita connection though, it is clear that its curvature tensor possesses the box symmetry, i.e. $$\yng(2,2)$$ From the Tic--Tac--Toe notation it follows that the box symmetry reduces to six irreducible parts under a rigging. From the previous analysis of the Vidal and the adapted curvatures these can again be written entirely in terms of the semi-basic components and parts of the deformation tensors. \prop{leviadapt}{Let $\tR$ be the Riemann tensor with respect to the Levi--Civita connection then its components can be written in terms of just the deformation tensor and adapted covariant derivatives thereof plus the complete longitudal and normal parts of the adapted curvature, \eqal{1}{ \tR_{abcd}=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{abcd}+2H_{[a|c}{}^{e'}H_{b]de'}\\ \tR_{abcd'}=&2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}H)_{b]cd'}+2L_{ab}{}^{e'}H'_{e'd'c}\\ \tR_{abc'd'}=&-2[(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c'}L)_{ab|d']}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a}L')_{c'd'|b]}+L_{[a|}{}^e{}_{c'}L_{b]ed'}+L'_{[c'|}{}^{e'}{}_{a}L'_{d']e'b}-\\ &-W_{[a|}{}^e{}_{c'}W_{b]ed'}-W'_{[c'|}{}^{e'}{}_{a}W'_{d']e'b}]\\ \tR_{a'bc'd}=&\frac{1}{2}\tR_{a'c'bd}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{(a'|}K)_{bd|c')}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{(b}K')_{a'c'|d)}+K_{(b|}{}^e{}_{a'}K_{d)ec'}-L_{(b|}{}^e{}_{a'}L_{d)ec'}+\\ &+K'_{(a'|}{}^{e'}{}_{b}K'_{c)e'd}-L'_{(a'|}{}^{e'}{}_{b}L'_{c)e'd}\\ \tR_{a'b'c'd}=&2(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[a'}H')_{b']c'd}+2L'_{a'b'}{}^{e}H_{edc'}\\ \tR_{a'b'c'd'}=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'c'd'}+2H'_{[a'|c'}{}^{e}H'_{b']d'e}\\ } } In ref. \cite{Gr67} these were deduced in terms of the Levi--Civita connection but as it does not preserve the rigging these are better expressed in terms of the adapted or the Vidal connection. Note that the expressions are only decomposed in terms of the irreducible parts of the deformation tensors, where it is necessary in order to make manifest the symmetries. In all other cases it is a straight forward process to do just by insertion. \prop{levividal}{Let $\tR$ be the Riemann tensor with respect to the Levi--Civita connection then its components can be written in terms of just the deformation tensor and Vidal covariant derivatives thereof plus the complete longitudal and normal parts of the Vidal curvature as \eqal{1}{ \tR_{abcd}=&\rR_{abcd}+2L_{ab}{}^{e'}H_{cde'}+2H_{[a|c}{}^{e'}H_{b]de'}\\ \tR_{abcd'}=&2(\rdel_{[a}H)_{b]cd'}+2H_{[b|c}{}^{e'}H'_{d'e'|a]}+2L_{ab}{}^{e'}H'_{e'd'c}\\ \tR_{abc'd'}=&-2[(\rdel_{[c'}L)_{ab|d']}+(\rdel_{[a}L')_{c'd'|b]}+{2\fr k'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{[a}L'_{c'd'|b]}+{2\fr k}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{[c'}L_{ab|d']}\\ &-W_{[a|}{}^e{}_{c'}W_{b]ed'}-W'_{[c'|}{}^{e'}{}_{a}W'_{d']e'b}-2W_{[a|}{}^e{}_{[c'}L_{b]e|d']}-2W'_{[c'|}{}^{e'}{}_{[a}L'_{d']e'|b]}-\\ &-L_{[a|}{}^e{}_{c'}L_{b]ed'}-L'_{[c'|}{}^{e'}{}_{a}L'_{d']e'b}]\\ \tR_{a'bc'd}=&\frac{1}{2}\tR_{bda'c'}-\rdel_{(a'|}K_{bd|c')}-\rdel_{(b|} K'_{a'c'|d)}-W_{(d|}{}^{e}{}_{(a'}W_{b)ec')}- L_{(d|}{}^{e}{}_{(a'}L_{b)ec')}-\\&-2W_{(d}{}^{e}{}_{(a'|}L_{b)e|c')} -\frac{1}{k^2}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{c'}\eta_{bd}-\frac{1}{k}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{(a'|}W_{bd|c')}-\\ &-W'_{(c'|}{}^{e'}{}_{(b|}W'_{a')e'd)}- L'_{(c'|}{}^{e'}{}_{(b}L'_{d)e'a')}-\\&-2W'_{(c'}{}^{e'}{}_{(b|}L' _{d)e'|a')}-\frac{1}{{k'}^2}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{b}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{d}\eta_{a'c'} -\frac{1}{k'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{(b|}W_{a'c'|d)}\\ \tR_{a'b'c'd}=&2(\rdel_{[a'}H')_{b']c'd}+2H'_{[b'|c'}{}^{e}H_{de|a']}+2L'_{a'b'}{}^{e}H_{edc'}\\ \tR_{a'b'c'd'}=&\rR_{a'b'c'd'}+2L'_{a'b'}{}^{e}H'_{c'd'e}+2H'_{[a'|c'}{}^{e}H'_{b']d'e}\\ } } From propositions \ref{leviadapt} and \ref{levividal} the Ricci tensor and the curvature scalar is most easily deduced. \prop{Ricciadapted}{Let $\tdel$ be the Levi-Civita connection then the Ricci tensor reads in terms of the adapted connection, \eqal{1}{ \tR_{ab}=&R_{ab}+R''_{ab}=\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{(ab)}-\frac{1}{k}(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}})^{c'}{\eta}_{ab}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{(a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}')_{b)}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}W)_{ab}{}^{c'} +\frac{1}{k}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2{\eta}_{ab}+\\ &+W_{ab}{}^{c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{c'}+\frac{1}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}_{a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}_b+W'_{c'e'a}{W'}^{c'e'}{}_b-L'_{c'e'a}{L'}^{c'e'}{}_b\\ \tR_{ab'}=&(\frac{1-k}{k})({\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}}_{a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{b'}+(\frac{1-k'}{k'})({\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}}_{b'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_a +(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_cW)_a{}^c{}_{b'}+\\ &+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}W')_{b'}{}^{c'}{}_a+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_cL)_a{}^c{}_{b'}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}L')_{b'}{}^{c'}{}_a+\\ &+4L_{a}{}^{ec'}L'_{b'c'e}-2L_{a}{}^{ce'}{W'}_{b'e'c}-2L'_{b'}{}^{c'e}{W}_{aec'}-\frac{2}{k'}L_{acb'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^{c} -\frac{2}{k}L_{b'c'a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^{c'},\\ \tR_{a'b'}=&R'_{a'b'}+R''_{a'b'}=\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{(a'b')}-\frac{1}{k'}(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'})^{c}{\eta'}_{a'b'}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{(a'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}})_{b')} -(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c}W')_{a'b'}{}^{c}+\frac{1}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2{\eta'}_{a'b'}+\\ &+W'_{a'b'}{}^{c}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{c}+\frac{1}{k}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}_{a'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}_{b'}+W_{cea'}{W}^{ce}{}_{b'}-L_{cea'}{L}^{ce}{}_{b'} } where the following definitions are used \eqal{1}{ R_{ab}:=&\tR_{acb}{}^c=\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{(ab)}-W_{a}{}^{ec'}W_{bec'}+(\frac{k-2}{k})\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^{c'}W_{abc'}+(\frac{k-1}{k^2}) \eta_{ab}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2+L_{a}{}^{ec'}L_{bec'}\\ R''_{ab}:=&\tR_{ac'b}{}^{c'}=-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{(a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_{b)}-(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}W)_{ab}{}^{c'} -\frac{1}{k}(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})^{c'}\eta_{ab}+{W'}^{c'e'}{}_a{W'}_{c'e'b}+ \frac{1}{k'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_a\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_b-\\&-{L'}^{c'e'}{}_aL'_{c'}{}_{e'b} +W_{a}{}^{ec'}W_{bec'}+\frac{2}{k}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^{c'}W_{abc'}+\frac{1}{k^2} \eta_{ab}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2-L_{a}{}^{ec'}L_{bec'} } and similarly for the $R'_{a'b'}$ component. The Riemann curvature scalar is given by \eqal{1}{ \tR=&R+2R''+R'=\Tilde{R}+\Tilde{R}'+\frac{1-k}{k}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^2+\frac{1-k'}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2-2\tdel\cdot\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_I+ W^2+{W'}^2-L^2-{L'}^2 } where the following definitions are used \eqal{1}{ R:=&\tR_{ab}{}^{ab}=\Tilde{R}+{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2-{K}^2+{L}^2\\ R'':=&\tR_{ab'}{}^{ab'}=-\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^{a'}-\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^a+W^2-L^2+\frac{1}{k}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2+{W'}^2-{L'}^2+\frac{1}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2\\ R':=&\tR_{a'b'}{}^{a'b'}=\Tilde{R}'+{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2-{K'}^2+{L'}^2\\ } } Part of which is found in \cite{Roc84}. In the expressions on the curvature scalar above, we have used the following relation \eqs{ \tdel\cdot\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_I=(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})^{a'}+(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')^a-{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2-{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2\\ } and the notation $L^2:=L_{abc'}L^{abc'}$ etc. In terms of the Vidal connection the Ricci tensor and the curvature scalar is given in next proposition. \prop{Riccividal}{Let $\tdel$ be the Levi-Civita connection then the Ricci tensor reads, in component form, in terms of the Vidal connection \eqal{1}{ \tR_{ab}=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{(ab)}-\frac{1}{k}(\rdel_{c'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}})^{c'}{\eta}_{ab} -(\rdel_{(a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}')_{b)}-(\rdel_{c'}W)_{ab}{}^{c'}-2W_a{}^{ec'}W_{ebc'}+\\ &+(\frac{k-2}{k})W_{ab}{}^{c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{c'} +\frac{1}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}_{a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}_b+\frac{1}{k}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2{\eta}_{ab}+W'_{c'e'a}{W'}^{c'e'}{}_b-\\ &-L'_{c'e'a}{L'}^{c'e'}{}_b+2L_{a}{}^{ec'}L_{bec'}\\ \tR_{ab'}=&(\frac{1-k}{k})({\rdel}_{a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})_{b'}+(\frac{1-k'}{k'})({\rdel}_{b'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_a +(\rdel_cW)_a{}^c{}_{b'}+\\ &+(\rdel_{c'}W')_{b'}{}^{c'}{}_a+(\rdel_cL)_a{}^c{}_{b'}+(\rdel_{c'}L')_{b'}{}^{c'}{}_a+\\ &+6L_{a}{}^{ec'}L'_{b'c'e}+2W_{a}{}^{ec'}W'_{b'c'e}-(\frac{k+k'-2}{kk'})\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_a\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{b'}-\\ &-L_{acb'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^{c} -L_{b'c'a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^{c'}-(\frac{k'-2}{k'})W_{acb'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^{c} -(\frac{k-2}{k})W_{b'c'a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^{c'}\\ \tR_{a'b'}=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{(a'b')}-\frac{1}{k'}(\rdel_{c}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'})^{c}{\eta'}_{a'b'}-(\rdel_{(a'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}})_{b')}-(\rdel_{c}W')_{a'b'}{}^{c} 2W'_{a'}{}^{e'c}W'_{e'b'c}-\\&+(\frac{k'-2}{k'})W'_{a'b'}{}^{c}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{c} +\frac{1}{k}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}_{a'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}_{b'}+\frac{1}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2{\eta'}_{a'b'}+W_{cea'}{W}^{ce}{}_{b'}-\\ &-L_{cea'}{L}^{ce}{}_{b'}+2L'_{a'}{}^{e'c}L'_{e'b'c} } where the following definitions were used \eqal{1}{ R_{ab}:=&\tR_{acb}{}^c=\rR_{(ab)}-W_{a}{}^{ec'}W_{bec'}+(\frac{k-2}{k})\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^{c'}W_{abc'}+(\frac{k-1}{k^2}) \eta_{ab}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2\\ &+3L_{a}{}^{ec'}L_{bec'}+2L_{(a|}{}^{ce'}W_{b)ce'}\\ R''_{ab}:=&\tR_{ac'b}{}^{c'}=-(\rdel_{(a}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}')_{b)}-(\rdel_{c'}W)_{ab}{}^{c'} -\frac{1}{k}(\rdel_{c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J})^{c'}\eta_{ab}+{W'}^{c'e'}{}_a{W'}_{c'e'b}+ \frac{1}{k'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_a\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_b-\\&-{L'}^{c'e'}{}_aL'_{c'}{}_{e'b} -W_{a}{}^{ec'}W_{bec'}+\frac{1}{k^2} \eta_{ab}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2-L_{a}{}^{ec'}L_{bec'}-2L_{(a}{}^{ec'}W_{e|b)c'} } again similarly for the $R'_{a'b'}$ component. The Riemann curvature scalar is given by \eqal{1}{ \tR=&R+2R''+R'=\Tilde{\Tilde{R}}+\Tilde{\Tilde{R'}}+\frac{1-k}{k}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^2+\frac{1-k'}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2-2\tdel\cdot\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_I+W^2+{W'}^2+L^2+{L'}^2 } where the following definitions were used \eqal{1}{ R=&\Tilde{\Tilde{R}}+\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^2-K^2+3L^2\\ R''=&-\rdel_{a'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^{a'}-\rdel_{a}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^a+W^2-L^2+\frac{1}{k}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}}^2+{W'}^2-{L'}^2+\frac{1}{k'}{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2\\ R'=&\Tilde{\Tilde{R'}}+{\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'}^2-{K'}^2+3{L'}^2 } } \subsection{Their conformal properties} In an earlier treatment it was found that conformal transformations does affect a connection including torsion and non-metricity in a non-trivial fashion. Here will be given a complete analysis of the induced transformations of the Vidal, adapted and Levi--Civita connections. \defn{Conform trans of con}{Let the triplet $(\ts,\tg,I)$ denote an almost product manifold, let $\tdel,\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}},\rdel$ denote the Levi--Civita, adapted and Vidal connection respectively then define the associated {\bf conformal tensors} denoted $\ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}},\aC,\rC$ with characteristics \eqal{1}{ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}},\aC,\rC:&\qquad \L^1\times\L^1\longmapsto \L^1 } by \eqal{1}{ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}(X,Y):=&\ct \tdel_XY-\tdel_XY\\ \aC(X,Y):=&\ct \tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_XY-\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_XY\\ \rC(X,Y):=&\ct \rdel_XY-\rdel_XY\\ } where $X,Y\in L^1$ are vectorfields on $\ts$. } The conformal tensor, corresponding to the Vidal and the adapted connection, can most easily be expressed in terms of the conformal tensor of the Levi--Civita connection. \prop{conf trans}{Let $\ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}},\aC,\rC$ be the conformal tensors defined in \ref{Conform trans of con} then following relations hold \eqal{1}{ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}(X,Y)&=X[\phi]Y+Y[\phi]X-\tg(X,Y)\ri\td\phi\\ \aC(X,Y)&=\ico \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}(X,\ico Y)+\nco \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}(X,\nco Y)\\ \rC(X,Y)&=\ico \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}(\ico X,\ico Y)+\nco \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}(\nco X,\nco Y)\\ } where $X,Y\in \L^1$ are vectorfields on $\ts$. } \proof{To be added.} The difference between these conformal tensors manifests itself in a clearer way by studying the expressions in component form. The conformal tensor of the Levi--Civita connection is read of from the above proposition and is noticeably symmetric. \prop{levi conf}{Let $\ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}$ be the conformal tensor of the Levi--Civita connection then in component form it reads, \eqal{1}{ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}_{ab}{}^{c}=&2\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{(a}^cE_{b)}[\phi]-\eta_{ab}\eta^{cd}E_d[\phi]\\ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}_{ab}{}^{c'}=&-\eta_{ab}\eta^{c'd'}E_{d'}[\phi]\\ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}_{a'b}{}^{c}=&E_{a'}[\phi]\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b}^c\\ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}_{ab'}{}^{c'}=&E_{a}[\phi]\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b'}^{c'},\\ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}_{a'b'}{}^{c}=&-\eta_{a'b'}\eta^{cd}E_{d}[\phi]\\ \ul{\scC}} \newcommand{\aC}{\Tilde{\ul{\scC}}} \newcommand{\rC}{\Tilde{\Tilde{\ul{\scC}}}_{a'b'}{}^{c'}=&2\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{(a'}^{c'}E_{b')}[\phi]-\eta_{a'b'}\eta^{c'd'}E_{d'}[\phi]\\ } } The conformal tensor of the adapted connection can now be derived from the above expressions. It should be stressed though that it is not symmetric. \prop{adapt conf}{Let $\aC$ be the conformal tensor of the adapted connection then in component form it reads \eqal{1}{ \aC_{ab}{}^{c}=&2\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{(a}^cE_{b)}[\phi]-\eta_{ab}\eta^{cd}E_d[\phi]\\ \aC_{ab}{}^{c'}=&0\\ \aC_{ab'}{}^{c}=&0\\ \aC_{a'b}{}^{c}=&E_{a'}[\phi]\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b}^c\\ \aC_{ab'}{}^{c'}=&E_{a}[\phi]\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b'}^{c'}\\ \aC_{a'b}{}^{c'}=&0\\ \aC_{a'b'}{}^{c}=&0\\ \aC_{a'b'}{}^{c'}=&2\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{(a'}^{c'}E_{b')}[\phi]-\eta_{a'b'}\eta^{c'd'}E_{d'}[\phi]\\ } } Finally in the case of the Vidal connection the conformal tensor takes a very simple form and will, like in the Levi--Civita case, be symmetric. \prop{vidal conf}{Let $\rC$ be the conformal tensor of the Vidal connection then in component form it reads \eqal{1}{ \rC_{ab}{}^{c}=&2\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{(a}^cE_{b)}[\phi]-\eta_{ab}\eta^{cd}E_d[\phi]\\ \rC_{ab}{}^{c'}=&0\\ \rC_{ab'}{}^{c}=&0\\ \rC_{a'b}{}^{c}=&0\\ \rC_{ab'}{}^{c'}=&0\\ \rC_{a'b}{}^{c'}=&0\\ \rC_{a'b'}{}^{c}=&0\\ \rC_{a'b'}{}^{c'}=&2\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{(a'}^{c'}E_{b')}[\phi]-\eta_{a'b'}\eta^{c'd'}E_{d'}[\phi]\\ } } Well known is the fact that when decomposing the Riemann curvature tensor of the Levi--Civita connection into its irreducible parts with respect to its traces, the appearing Weyl tensors measures whether the riemannian manifold is conformally flat or not. More specifically, the vanishing of the Weyl tensor is the condition for local conformal flatness in the case when the dimension of the manifold exceeds three. The tracefree part of the Ricci tensor is defined by, \eq{ \hat{\tR}_{\ba\bb}:=\tR_{\ba\bb}-{1\fr m}\tR\eta_{\ba\bb} } Next the Ricci one-form and the tracefree Ricci one-form are defined. \defn{riccioneform}{Let $\tR_{ab},\hat{\tR}_{ab}$ denote the Ricci tensor and its tracefree part respectively then define their one forms by \eqal{1}{ \tR^\ba:=&E^\bb\tR_\bb{}^\ba\\ \hat{\tR}^\ba:=&E^\bb\hat{\tR}_\bb{}^\ba=\tR^\ba-{1\fr m}\tR E^\ba\\ } } Denoting the Rimann two form $\tR^{\tc\td}={1\fr 2}E^\ta\wedge E^\tb\tR_{\ta\tb}{}^{\tc\td}$ and the Weyl two form by ${\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}^{\tc\td}={1\fr 2}E^\ta\wedge E^\tb{\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{\ta\tb}{}^{\tc\td}$ the decomposition is most elegantly written \eqnono{ \arr{ccccccc}{ \yng(2,2) &=& \widetilde{\yng(2,2)} &\oplus& \widetilde{\yng(2)} &\oplus& \bigodot\vspace{.3cm}\\ {1\fr 12}m^2(m^2-1) &=& {1\fr 12}(m-3)m(m+1)(m+2) &+& {1\fr 2}(m-1)(m+2) &+& 1\vspace{.3cm}\\ \tR^{\bc\bd} &=& {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}^{\bc\bd} &+&{2\fr m-2}\hat{\tR}^{[\bc}\wedge E^{\bd]}&+& {1\fr {m(m-1)}}\tR E^\bc\wedge E^\bd } } After reinserting the expression for the tracefree part of the Ricci tensor and solving for the Weyl tensor the more familiar form is obtained \eq{ {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{\ba\bb}{}^{\bc\bd}=\tR_{\ba\bb}{}^{\bc\bd}-{4\fr m-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[\ba}^{[\bc}\tR_{\bb]}^{\bd]}+{2\fr (m-1)(m-2)}\tR\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[\ba}^{[\bc}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{\bb]}^{\bd]} } From this equation the different components are read off and put in next proposition. \prop{weyl}{Let ${\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}^{\bc\bd}$ be the weyl tensor in an almost product manifold then its components look like \eqal{1}{ {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{ab}{}^{cd}=&\tR_{ab}{}^{cd}-{4\fr m-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a}^{[c}\tR_{b]}^{d]}+{2\fr (m-1)(m-2)}\tR\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a}^{[c}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b]}^{d]}\\ {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{ab}{}^{cd'}=&\tR_{ab}{}^{cd'}-{2\fr m-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a}^{c}\tR_{b]}^{d'}\\ {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{ab}{}^{c'd'}=&\tR_{ab}{}^{c'd'}\\ {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{a'b}{}^{c'd}=&\tR_{a'b}{}^{c'd}-{1\fr m-2}(\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{a'}^{c'}\tR_{b}^{d}+\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b}^{d}\tR_{a'}^{c'})+{1\fr (m-1)(m-2)}\tR\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{a'}^{c'}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b}^{d}\\ {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{a'b'}{}^{c'd}=&\tR_{a'b'}{}^{c'd}-{2\fr m-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'}^{c'}\tR_{b']}^{d}\\ {\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{a'b'}{}^{c'd'}=&\tR_{a'b'}{}^{c'd'}-{4\fr m-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'}^{[c'}\tR_{b']}^{d']}+{2\fr (m-1)(m-2)}\tR\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'}^{[c'}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b']}^{d']}\\ } } From above proposition it is for instance clear that the $\tR_{abc'}{}^{d'}$ component of the Riemann tensor must be conformally invariant. Taking a look at its final expression in proposition \ref{leviadapt} the only non-manifest conformally invariant terms, are those involving derivatives. These can be proved to be independently conformally invariant. \prop{derlconf}{Let $L,L'$ be the respective twisting tensors of the characteristic distributions defined by an almost product structure on $\ts$ then the following relations hold \eqal{1}{ \ct\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[Z'}L_{W']}(X,Y)=&e^{2\phi}\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[Z'}L_{W']}(X,Y)\\ \ct\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[Z}L'_{W]}(X',Y')=&e^{2\phi}\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[Z}L'_{W]}(X',Y') } or put in component form \eqal{1}{ (\ct\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c'}L)_{ab|d']}=&e^{2\phi}(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c'}L)_{ab|d']}\\ (\ct\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c}L')_{a'b'|d]}=&e^{2\phi}(\tilde{\tdel}} \newcommand{\rdel}{\Tilde{\Tilde{\tdel}}_{[c}L')_{a'b'|d]} } } Proceeding in the same fashion as in \cite{Carter},investigating how the curvature components of the adapted connection can be divided into irreducible parts - in the case of an almost product manifold instead as for just an embedding, some immediate differences is noticed. The generalisation of what Carter calls the outer curvature is the $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{ab}{}^{c'd'}$ component which will be seen not to be conformally invariant in the generic case but only if the unprimed distribution is integrable. The same is of course true for the outer curvature of the complementary (primed) distribution. From proposition \ref{adapted curvature} it is manifest that the non-invariant components are $${2\fr k}L_{ab[c'}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}_{d']}$$ and $${2\fr k'}L'_{a'b'[c}\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}'_{d]}$$ respectively. For the internal curvature which in the language of almost product manifolds is the total semi-basic components of the adapted curvature, one can follow the procedure of dividing the tensor components into its irreducible parts according to the above scheme. In the generic case the semi-basic components does not have the box symmetry $$\yng(2,2)$$. Because of torsion though, it has the other symmetry parts. It is clear that when the distribution is integrable its internal curvature will indeed have the box symmetry. Defining the internal Weyl tensors the generalization of \cite{Carter,Carter97} can be made. \defn{intweyl}{Let $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}$ be the curvature tensor of the adapted connection and $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{ab}{}^{cd}$, $\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'}{}^{c'd'}$ the internal curvatures of the two complementary distributions associated with an almost product manifold then define their respective Weyl tensors by \eqal{1}{ \tilde{C}_{ab}{}^{cd}=&\tilde{R}_{ab}{}^{cd}-{4\fr k-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a}^{[c}\tilde{R}_{b]}^{d]}+{2\fr (k-1)(k-2)}\tilde{R}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a}^{[c}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b]}^{d]}\\ \tilde{C}'_{a'b'}{}^{c'd'}=&{{\tilde{R}}'}_{a'b'}{}^{c'd'}-{4\fr k'-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'}^{[c'}\tilde{R}'{}_{b']}^{{d'}]} +{2\fr (k'-1)(k'-2)}{\tilde{R}'}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'}^{[c'}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b']}^{d']}\\ } where \eqal{3}{ \tilde{R}_{ab}{}^{cd}:=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{ab}{}^{cd},& \tilde{R}_{a}^{b}:=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{ac}{}^{bc},& \tilde{R}:=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{ab}{}^{ab},\\ \tilde{R}'{}_{a'b'}{}^{c'd'}:=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'}{}^{c'd'},&\qquad \tilde{R}'{}_{a'}^{b'}:=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'c'}{}^{b'c'},&\qquad \tilde{R}':=&\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}_{a'b'}{}^{a'b'}. } } Following Carter's procedure it is now easy to generalize his relations to the case of an almost product manifold. \prop{weyl-weyl}{Let $\tilde{C},\tilde{C}'$ be the Weyl tensors of the internal curvatures of the two complemenatry distributions associated with an almost product manifold. Then the following relations hold. \eqal{1}{ \tilde{C}_{ab}{}^{cd}=&C_{ab}{}^{cd}-\frac{4}{k-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}^{[c}_{[a}C^{d]}_{b]}+\frac{2}{(k-1)(k-2)}C-\\ &- 2(L_{[a|}{}^c{}_{e'}L_{b]}{}^{de'}+2L_{[a|}{}^{[c}{}_{e'}W_{b]}{}^{d]e'})-\\ & -\frac{4}{k-2}(\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a|}^{[c}L_{b]}{}^e{}_{e'}L_e{}^{d]e'}+ \delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a|}^{[c}W_{b]}{}^e{}_{e'}W_e{}^{d]e'}+\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a|}^{[c}L_{b]}{}^e{}_{e'}W_e{}^{d]e'}+\\ &+\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a|}^{[c}W_{b]}{}^e{}_{e'}L_e{}^{d]e'})+\\ &+\frac{2}{(k-1)(k-2)}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a}^c\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b]}^d(W^2-L^2),\\ \tilde{C}'_{a'b'}{}^{c'd'}=&C'_{a'b'}{}^{c'd'}-\frac{4}{k'-2}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}^{[c'}_{[a'}{C'}^{d']}_{b']} +\frac{2}{(k'-1)(k'-2)}C'-\\ &- 2(L'_{[a'|}{}^{c'}{}_{e}L'_{b']}{}^{d'e}+2L'_{[a'|}{}^{[c'}{}_{e}W'_{b']}{}^{d']e})-\\ & -\frac{4}{k'-2}(\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'|}^{[c'}L'_{b']}{}^{e'}{}_{e}L'_{e'}{}^{d']e}+ \delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'|}^{[c'}W'_{b']}{}^{e'}{}_{e}W'_{e'}{}^{d']e}+\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'|}^{[c'}L'_{b']}{}^{e'}{}_{e}W'_{e'}{}^{d']e}+ \\ &+\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'|}^{[c'}W'_{b']}{}^{e'}{}_{e}L'_{e'}{}^{d']e})+\\ &+\frac{2}{(k'-1)(k'-2)}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{[a'}^{c'}\delta} \def\god{{\mathfrak d}} \def\scD{{\scr D}} \def\dsD{{\mathbb D}} \def\clD{{\cal D}_{b']}^{d'}({W'}^2-{L'}^2)\\ } where \eqal{3}{ C_{ab}{}^{cd}:=&{\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{ab}{}^{cd},& C_{a}^{b}:=&C_{ac}{}^{bc},& C:=&C_{ab}{}^{ab}\\ C'{}_{a'b'}{}^{c'd'}:=&{\ul{C}}} \newcommand{\tc}{{\ul{c}}} \newcommand{\tcc}{{\ul{\gamma}}_{a'b'}{}^{c'd'},&\qquad C'{}_{a'}^{b'}:=&C'_{a'c'}{}^{b'c'},&\qquad C':=&C'_{a'b'}{}^{a'b'}. } } From this proposition it follows that if the semi-basic part of the conformal tensor is zero and the distribution is integrable then the distribution possess local conformal flatness if and only if its conformation tensor vanishes. Of course this is only true in the case where $k>3$. This generalizes Carter's result to the case of almost product manifold. \section{Physical applications} There are lots of physical applications involving almost product manifolds. Because principal bundles can be regarded as a almost product manifold with the $(GF,GD)$ structure ordinary gauge theory can be found in its utmost geometrical form. In Kaluza--Klein theory the internal space need not be a group manifold but could instead be a homogenous space with the proper gauge group as its isometry group. Here will be given an example of the recovered Kaluza--Klein theory from the almost product structure taken first in the most general case where no restriction of the fiber is made. \exam{Kaluza-Klein}{{\bf Kaluza-Klein theory}\\ Here will be seen how, in the case of a $(GF,GD)$ structure, the Vidal connection will reduce to the gauge covariant derivative. The Einstein--Hilbert action will reduce to the inner curvatures plus the gauge field term as in \cite{DuNiPo85}. First note the spliting of the action in the $(GF,GD)$ case, \eq{ \int d^mx\sqrt{\tg}\tR=\int d^kxd^{k'}\!\!y\sqrt{\ig}\sqrt{\nlg}(\Tilde{\Tilde{R}}+\Tilde{\Tilde{R}}'+L^2). } Note that the primed distribution is chosen to be integrable. Following \cite{DuNiPo85} the vielbeins can locally be parametrized as \eqal{2}{ \tE_a=&E_a+A_a^iK_i,\quad&\tE_{a'}&=E_{a'}\\ \tE^a=&E^a, &\tE^{a'}=&E^{a'}-E^aA_{a}^iK_i^{a'} } where $K_i=K_i{}^{a'}(y)E_{a'}(y)$ are the Killing vectors of the integrable internal manifold. These satisfy an algebra \eq{ [K_i,K_j]=f_{ij}{}^kK_k. } where the structure constants $f_{ij}{}^{k}$ of the isometry group and does not depend on $y$. The gauge fieldstrength $F^i=dA^i+{1\fr 2}f_{jk}{}^iA^j\wedge A^k$ is most easily found \cite{Holm98} \eq{ F(X,Y)=\nco [X,Y]=2L(X,Y) } where $X,Y$ are unprimed vector fields which implies that the $L^2$ term in the action reads ${1\fr 4}F^2$ which is the ordinary action term in gauge theory. Now the Vidal connection of the gauge field can be written \eq{ (\rdel_X F)(Y,Z)=(\tdel^{\esD}_XF)(Y,Z)+F^i(Y,Z)\nco [X,K_i], } written in component form, the relations look like, \eq{ (\rdel_aF)_{bc}^i=(\tdel^{\esD}_aF)_{bc}^i+f_{jk}^iA_a^jF_{bc}^k } which is precisely the gauge covariant derivative. Further the identity from proposition \ref{twistbianchi} $(\rdel_{[a}L)_{bc]}{}^{d'}=0$ reduces to the Bianchi identity of the gauge field \eq{ (\rdel_{[a}F)_{bc]}^i=0 } the $\tR_{ab'}$ term from proposition \ref{Riccividal} reduces to \eq{ \tR_{ab'}={1\fr 2}(\rdel_cF)_{a}^i{}^{c}K_{ib'} } which from the Einstein's equations point of view reduces to the equations of motion for the gauge field, i.e. $(\rdel_cF)_{a}^i{}^{c}=0$. So it is clear that gauge theory and Kaluza-Klein theory is contained in the almost product manifold description. In the general case however the Killing vectors could be exchanged to ordinary vielbeins, the structure coefficients need not be constant and the fiber no longer a group space or homogenous space, the almost product structure procedure would still be valid. In the case of Kaluza-Klein theory containing the dilaton field it is easy to see that it is contained in the mean curvature, $\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}$, see \cite{Holm98}. } Another more traditional example is the decomposition of the four dimensional curvature scalar into three space in hamiltonian formulation of ordinary Einstein gravity. From proposition \ref{Ricciadapted} the decomposition of the curvature scalar is immediately found to be $\tR=\tilde{\tR}} \newcommand{\rR}{\Tilde{\Tilde{\tR}}-\kappa} \def\goj{{\mathfrak j}} \def\scJ{{\scr J}} \def\dsJ{{\mathbb J}} \def\clJ{{\cal J}^2+K^2$. From a foliation point of view it is clear that a space with non-degenerate metric of Minkowskian signature must have vanishing euler number which is also the condition for a space to have a codimension one foliation \cite{To88,Rov98}. This is why the $L^2$ term can be set to zero. \section{Conclusions and outlook} The theory of almost product manifolds is seen to overlap with a lot of physical applications. The main areas is of course geometrical phyics such as gravity, Kaluza-Klein theory and ordinary gauge theory. Here was seen for instance that in gauge- or Kaluza-Klein theory the Vidal connection reduces to the ordinary gauge covariant derivative and the second curvature identity of the Vidal curvature gives the Bianchi identity of the gauge field. The relations found propositions \ref{Ricciadapted} and \ref{Riccividal} could perhaps be used to find new solutions to the equations of motions of various supergravity theories. Here the the Ricci tensors are given in the most general case why all kind of brane solutions must fit in this scheme. From these relations it should also be clear that a black hole solution in ordinary space-time carrying a gauge field charge should correspond to a black hole solution in the total space with rotational parameters corresponding to the gauge charges. The reason for this is of course the identification of the twisting tensor as the gauge fieldstrength. In super gravity theories the black hole solutions carry charges from anti-symmetric tensor fields and will be $p$-brane solutions. These can again have rotational parameters in the transverse directions which correspond to gauge field charges of the isometry group of the transverse space \cite{CvDu99,Du98} which by analysis in \cite{Holm98} would correspond to a non-integrability of the brane itself in this context. Another interesting investigation would be to see how the Clifford algebra splits in a almost product manifold. In a appearing paper \cite{Mig} will be shown how flat super space looks in the almost product structure picture. \vskip1cm \underline{Acknowledgments}: The authors would like to thank Martin Cederwall and Robert Marnelius for discussions.
\section{Introduction} The microwave background radiation is principally a relic of the early Universe, containing information on the fluctuations present at a redshift of about 1000. Detailed knowledge of the fluctuations is extremely valuable as a tool for distinguishing structure formation theories. The main alternative to the microwave background for this purpose is the distribution of galaxies at the present day, or at moderate redshifts. The microwave background has two great advantages over galaxy redshift surveys. Firstly, one needs to make no assumptions about the relationship between galaxy clustering and mass clustering (the issue of bias), and secondly, the microwave background probes the Universe when the fluctuations were extremely small, so linear perturbation theory can describe the predicted fluctuations to high accuracy. Against these advantages must be offset the problems of foreground contamination and the technical difficulties of producing a high signal-to-noise, high-resolution map of large areas of sky. Future satellite experiments planned by ESA and NASA (\pcite{PhaseA}, \pcite{Jungman}) will produce such maps, and these will allow the spectrum of temperature fluctuations to be measured with high accuracy. The primary goal of such experiments will be to determine a number of cosmological parameters, such as the density parameter $\Omega_0$, the Hubble constant, and the mix of species of dark matter, all of which, in inflationary theories, influence the detailed shape of the temperature power spectrum. In such theories, the temperature map is a random gaussian field, whose properties are determined by the power spectrum (or correlation function) alone. It is vitally important to test whether the map is indeed gaussian, as if it is not, the proposed parameter estimation method (e.g. \pcite{Jungman}) is invalid. In addition, a non-gaussian map of the primary temperature anisotropies would indicate a quite different structure formation scenario from the standard inflation model. There are several ways to test the gaussian hypothesis, such as the 3-point function (e.g. \pcite{Hinshaw94}, \pcite{FRS93}, \pcite{LS93}, \pcite{GLMM94}), the genus and Euler-Poincar\'e statistic (\pcite{Coles88}, \pcite{Gott90}, \pcite{Luo94b}, \pcite{Smoot94}), the bispectrum (\pcite{Luo94}, \pcite{Hea98}, \pcite{Ferreira98}), studies of tensor modes in the CMB \cite{CCT94}, and peak statistics (\pcite{BE87}, \pcite{Kogut95}, \pcite{Kogut96}). For all of these it is possible in principle to make predictions for gaussian fields, and each method has its advantages. The bispectrum is attractive because the error analysis can be done without recourse to numerical simulation, and can treat the twin problems of missing sky regions and correlated noise \cite{Hea98}, albeit at large computational expense. In this paper we explore an alternative: the correlation function of peaks in the microwave background. This has the advantage of being readily calculable, and also we find that the predicted correlation functions have oscillatory features which are not obviously present in the temperature autocorrelation function. In much the same way as the oscillations in the power spectrum allow high precision determinations of cosmological parameters, the peak correlation function oscillations should allow a high precision test of the gaussian hypothesis. In the latter case, the correlation function of peaks above any given threshold is completely determined once the power spectrum is known; there are no free parameters. In this paper, we calculate the 2-point correlation function of peaks in a 2D gaussian field, for any given power spectrum. We make no approximations other than to compute the correlation function for small angular separations so we assume the sky is flat and translate the sky power spectrum to an analogous 2D spectrum. We follow the peak statistics methods developed initially by \scite{PH85} and \scite{BBKS}. The layout of the paper is as follows. In section 2 we describe the method, and in section 3 present the results and compare with an approximate correlation function method introduced by \scite{BBKS}. In section 4 we discuss the results and the possibilities for Planck and MAP. \section{Method} In this section, we compute the two-point correlation function of local maxima in 2D gaussian random fields. For small separations, this calculation should be accurately applicable to the clustering of peaks in the microwave sky (see \scite{BE87} for further discussion of this point). The calculation is laid out as follows: after some definitions, we compute the $12 \times 12$ covariance matrix for the 12 variables which are required to specify the two peaks. We then invert this matrix to obtain the multivariate gaussian distribution for the variables, then apply the peak constraint and integrate over the remaining variables to obtain the correlation function. \subsection{Peaks in 2D} We define the temperature fluctuation by $\delta({\bf x}) \equiv T({\bf x})/\bar T -1$, and its 2D Fourier transform by $\delta_{\bf k} \equiv \int d^2{\bf k}\, \delta({\bf x}) \exp(i{\bf k}\cdot{\bf x})$. The fluctuations are specified entirely by the power spectrum, $P(k)$, defined by \begin{equation} \langle \delta_{\bf k} \delta_{{\bf k}'}\rangle = (2\pi)^2 P(k) \delta^D({\bf k}+{\bf k}') \label{Pk} \end{equation} where angle brackets indicate ensemble averages, and $\delta^D$ is a 2D Dirac delta function. We will require moments of the power spectrum, defining \begin{equation} \sigma_j^2 \equiv {1\over (2\pi)^2} \int d^2{\bf k} \,P(k) k^{2j}, \end{equation} and spectral parameters \begin{equation} \gamma \equiv \sigma_1^2/(\sigma_0 \sigma_2) \qquad\theta_* \equiv \sqrt{2}{\sigma_1\over \sigma_2}. \label{Parameters} \end{equation} At peaks, the 2D gradient $\delta_i \equiv {\partial} \delta/{\partial} x_i$ vanishes, and the eigenvalues of the second derivative matrix $\delta_{ij} \equiv {\partial}^2 \delta/{\partial} x_i {\partial} x_j$ are all negative. Since $\delta_{ij}$ is symmetric, we are concerned with 6 independent numbers ${\bf v} = \{\delta,\delta_x,\delta_y,\delta_{xx},\delta_{xy},\delta_{yy}\}$ to specify a local maximum. For a gaussian field, the probability density function for the 12 variables ${\bf v} \equiv \{{\bf v}^{(1)},{\bf v}^{(2)}\}$, evaluated at two points separated by a distance ${\bf r}$, is \begin{equation} p({\bf v}_1,{\bf v}_2) = {1\over (2\pi)^6 ||M||^{1/2}} \exp\left(-{1\over 2} v_i M^{-1}_{ij} v_j\right). \end{equation} The covariance matrix is $M_{ij} \equiv \langle (v_i - \bar v_i)(v_j - \bar v_j)\rangle$, overlines indicate ensemble averages, and the summation convention is assumed. Note that $\bar v_i=0$ in our case, and $M^{-1}_{ij}$ is the $i,j$ component of the inverse of $M$. The number density of peaks is a sum of 2D Dirac delta functions, centred on peaks ${\bf x}_{pk,q}$: \begin{equation} n_{pk} = \sum_q \delta^D({\bf x} - {\bf x}_{pk,q}) \label{n} \end{equation} In the vicinity of a peak, we may expand \begin{equation} \delta({\bf x}) \simeq \delta({\bf x}_{pk}) + {1\over 2}\sum_{i,j} {{\partial}^2 \delta\over {\partial} x_i {\partial} x_j} ({\bf x} - {\bf x}_{pk})_i ({\bf x} - {\bf x}_{pk})_j \end{equation} so \begin{equation} ({\bf x} - {\bf x}_{pk})_i \simeq \left({{\partial}^2 \delta\over {\partial} x_i {\partial} x_j}\right)^{-1} {{\partial} \delta\over {\partial} x_j} \end{equation} and the Dirac delta function in (\ref{n}) is \begin{equation} \delta^D({\bf x} - {\bf x}_{pk}) \simeq |\det \delta_{ij}| \delta^D(\delta_i). \end{equation} The mean number density is \begin{eqnarray} \langle n_{pk}({\bf x})\rangle& =& \langle | \det \delta_{ij}| \delta^D(\delta_i) \rangle\nonumber\\ & = & \int d\delta \,d \delta_{xx}\, d \delta_{xy}\, d \delta_{yy}\, | \det \delta_{ij}|\, p(\delta, \delta_i=0, \delta_{ij}) \end{eqnarray} and the range of integration is set by the requirement that the eigenvalues of $-\delta_{ij}$ are all positive. These eigenvalues are \begin{equation} \lambda = {1\over 2}\left[-(\delta_{xx}+\delta_{yy}) \pm \sqrt{(\delta_{xx}-\delta_{yy})^2 +4 \delta_{xy}^2}\right]. \end{equation} If we introduce the notation \begin{eqnarray} \nu & \equiv & {\delta\over \sigma_0}\nonumber\\ X & \equiv & -{(\delta_{xx}+\delta_{yy})\over \sigma_2}\nonumber\\ Y & \equiv & {(\delta_{xx}-\delta_{yy})\over \sigma_2}\nonumber\\ Z & \equiv & {2\delta_{xy}\over \sigma_2} \label{Vars} \end{eqnarray} then the conditions for a maximum are $X>0$ and $Y^2+Z^2<X^2$. The determinant is \begin{equation} \det \delta_{ij} = {X^2-Y^2-Z^2\over 4\sigma_2^2}. \end{equation} To compute the mean number density, we need the covariance matrix for the 6 variables specifying the field and its first and second derivatives. The non-zero components are $\langle \nu^2 \rangle = \langle X^2\rangle = 1$; $\langle \nu X \rangle = \gamma$; $\langle Y^2 \rangle = \langle Z^2 \rangle = 1/2$; $\langle \delta_x^2 \rangle = \langle \delta_y^2 \rangle = \sigma_1^2/2$. The matrix can readily be inverted, and the integrations over $Y$, $Z$ and $X$ performed. The expression for the differential number density as a function of peak height, $n_{pk}(\nu)$ is given in \scite{BE87}, equation (A1.9): \begin{equation} n_{pk}(\nu)={1\over (2\pi)^{3/2} \theta_*^2}\exp(-\nu^2/2)\,G(\gamma,\gamma\nu) \end{equation} where \begin{eqnarray} G(\gamma,x_*)&\equiv& (x_*^2-\gamma^2)\left\{1-{1\over 2}{\rm erfc}\left[ {x_*\over \sqrt{2(1-\gamma^2)}} \right]\right\} + x_*(1-\gamma^2) {\exp\left[-x_*^2\over 2(1-\gamma^2)\right]\over \sqrt{2\pi(1-\gamma^2)}} + \nonumber\\ &&{\exp\left[-x_*^2\over 3-2\gamma^2\right]\over \sqrt{3-2\gamma^2}}\left\{1-{1\over 2}{\rm erfc}\left[{x_*\over \sqrt{2(1-\gamma^2)(3-2\gamma^2)}}\right]\right\} \end{eqnarray} The correlation function is obtained similarly. We find the covariance matrix for the 12 variables, invert it to find their probability distribution, and integrate subject to the constraints. The cross-correlation function of peaks of height $\nu_1$ and $\nu_2$ is then given by \begin{eqnarray} 1+\xi(r| \nu_1, \nu_2) &= & {1\over 4\theta_*^4 n_{pk}(\nu_1) n_{pk}(\nu_2)} \int_{X_1=0}^\infty\int_{X_2=0}^\infty \int_{Y_1=-X_1}^{X_1}\int_{Y_2=-X_2}^{X_2} \int_{Z_1=-\sqrt{X_1^2-Y_1^2}}^{\sqrt{X_1^2-Y_1^2}}\int_{Z_2=-\sqrt{X_2^2-Y_2^2}}^{\sqrt{X_2^2-Y_2^2}} dX_1 dX_2 dY_1 dY_2 dZ_1 dZ_2 \times \nonumber\\ & & \left(X_1^2-Y_1^2-Z_1^2\right)\left(X_2^2-Y_2^2-Z_2^2\right) p(\nu_1,X_1,Y_1,Z_1,\delta^{(1)}_{x}=0, \delta^{(1)}_{y}=0, \nu_2,X_2,Y_2,Z_2,\delta^{(2)}_{x}=0, \delta^{(2)}_{y}=0). \label{Xsi} \end{eqnarray} The correlation function for peaks above a certain threshold $\nu$ is obtained by adding two further integrations over $\nu_1$ and $\nu_2$, and replacing the differential number densities $n_{pk}(\nu)$ in the denominator of (\ref{Xsi}) by numerically-evaluated integrals $n_{pk}(>\nu)$. The calculation of the covariance matrix and the joint probability distribution is rather technical, and so appears in an appendix. \section{Results} In the flat-sky approximation, the power spectrum $P(k)$ which we require is related simply to the conventional power spectrum of spherical harmonic coefficients $C_\ell \equiv \langle |a_{\ell m}|^2\rangle$, where $\delta T(\theta,\phi) \equiv \sum_{\ell m} a_{\ell m} Y_\ell^m(\theta,\phi)$ \cite{BE87}: \begin{equation} P(k) = C_\ell \quad {\rm with\ }k\simeq \ell \end{equation} We run CMBFAST \cite{SZ96} to generate the power spectrum $C_\ell$, and interpolate to obtain a continuous power spectrum $P(k)$. We model the beam with a gaussian of FWHM $b$, so multiply the power spectrum by a gaussian $\exp\left[-\sigma^2 \ell(\ell+1)\right]$, with $\sigma= b/\sqrt{8\ln2}$. For peaks above a threshold, an 8D numerical integration is performed (although one could be done analytically), and, despite the integrand consisting of about 3000 lines of code, each point in the correlation function can be computed accurately in tens of seconds on a fast workstation. In Fig. \ref{CDMpanels}, we show the peak-peak correlation functions for a variety of models, for a 5 arcminute beam and peaks above various thresholds. The striking feature of these graphs is the wealth of structure in the correlation functions. The first peak is not due to the beam, rather due to the sharp turndown in the power spectrum itself at around $\ell=1500$, due to the thickness of the last scattering surface and/or Silk damping. There may be a correspondence in the peak positions with analogous peaks in the power spectrum, but it could also be ringing due to the abrupt cutoff. Note in particular that the temperature autocorrelation function (the Fourier transform of the power spectrum) is very smooth for these separations, without any oscillatory features. \begin{figure} \centering \begin{picture}(200,200) \special{psfile='CDMpanels.eps' angle=270 voffset=220 hoffset=-70 vscale=40 hscale=40} \end{picture} \caption[]{\label{CDMpanels} The correlation function of peaks in the microwave background, for a cold dark matter (CDM) model (top left; baryon density $\Omega_B=0.05$, $\Omega_{CDM}=0.95$), a mixed dark matter model containing hot dark matter (top right; $\Omega_B=0.05$, $\Omega_{CDM}=0.8$, $\Omega_{HDM}=0.15$), and a flat cosmological constant model (bottom left; $\Omega_B=0.05$, $\Omega_{CDM}=0.15$, $\Omega_\Lambda=0.8$). A Hubble constant $H_0=60$ km s$^{-1}$ Mpc$^{-1}$ is assumed and the beam is 5 arcminutes FWHM. At bottom right, we show the effect on the cosmological constant model of altering the beam size. 5, 10, 15 and 20 arcminute beams are shown.} \end{figure} We also show in Fig. \ref{CDMpanels} the effect of beam-smearing. Correlation functions for beams of 5, 10, 15 and 20 arcminute are shown. The oscillations are more numerous and pronounced for smaller beam sizes, so it is advantageous to work at high resolution. Later in this paper we will want to compute the correlations of estimates of the peak-peak correlation function. This is difficult analytically, so we have generated sky realisations, of $3072 \times 3072$ pixels, each 1 arcminute across, located local maxima, and computed the correlation function (see Fig. \ref{Picture}). Edge effects are irrelevant as we wrap the box round for pair counting. Fig. \ref{MDMNum} shows the average of 10 such realisations, with the error on the mean plotted for each bin. The beam is 5 arcminutes and the spectrum is from a mixed dark matter model, which we have apodized with a cosine bell to remove power at long wavelengths. The apodizing affects $k<30$, and is introduced to avoid discreteness errors arising at low-$k$ from our continuum approximation used to evaluate the moments of the power spectrum and the $\lambda_{\alpha\beta\gamma}$, defined in the Appendix. Note the good agreement with the theoretical curve. This numerical approach allows us to compute the correlation coefficients of the estimators $\hat\xi$ of the peak-peak correlation function. In Fig. \ref{Covariance} we show, from 100 realisations of a mixed dark matter model, $\langle (\hat\xi_i - \bar\xi_i)(\hat\xi_j - \bar\xi_j)\rangle/(\sigma_i\sigma_j)$ where $\bar \xi_i$ is the mean value and $\sigma_i$ the variance of the estimates of the correlation function at radius $r_i$. These correlations would be required in order properly to quantify the significance of any departure from gaussian statistics. Note, however, that the correlation matrix is very close to diagonal, so distinct estimators are almost uncorrelated. \begin{figure} \centering \begin{picture}(200,200) \special{psfile='Picture.eps' angle=270 voffset=220 hoffset=-70 vscale=40 hscale=40} \end{picture} \caption[]{\label{Picture} A $10^\circ \times 10^\circ$ patch of a mixed dark matter simulated sky, at 5 arcminute resolution characteristic of the highest frequency Planck channels. Peaks above $1\sigma$ are marked with circles.} \end{figure} \begin{figure} \centering \begin{picture}(200,200) \special{psfile='MDMNum.eps' angle=270 voffset=220 hoffset=-70 vscale=40 hscale=40} \end{picture} \caption[]{\label{MDMNum} The $\nu=1$ threshold from the mixed dark matter model in Fig. \ref{CDMpanels}, with the correlation function from simulated gaussian fields superimposed. Also shown is the approximate correlation function computed by \scite{BE87} by ignoring derivatives of the temperature autocorrelation function.} \end{figure} \begin{figure} \centering \begin{picture}(200,200) \special{psfile='Covariance.eps' angle=270 voffset=220 hoffset=-70 vscale=40 hscale=40} \end{picture} \caption[]{\label{Covariance} Correlation coefficients for estimators of the peak-peak correlation function of peaks above $1\sigma$ in the MDM simulation shown in Fig. \ref{Picture}.} \end{figure} We have also tested the approximate correlation function considered in 3D by \scite{BBKS} and in 2D by \scite{BE87}. In this approximation, more analytic progress is possible. The approximation ignores all derivatives of the temperature autocorrelation function, and it should be accurate for large separations. In the notation of the appendix, this means setting all the $\lambda_{\alpha\beta\eta}=0$ except for the normalised autocorrelation function itself, $\lambda_{000}$. We show an example of the approximation, with the accurate solution, in Fig. \ref{MDMNum}. The approximation does quite well, especially at large separations, but fails to reproduce the oscillations. \begin{figure} \centering \begin{picture}(200,200) \special{psfile='MAPPlanck.eps' angle=270 voffset=220 hoffset=-70 vscale=40 hscale=40} \end{picture} \caption[]{\label{MAPPlanck} Mixed dark matter model correlation functions with errors expected from MAP (left) and Planck (right). The MAP simulations have a 12.6 arcminute beam, and Planck 5.5 arcminutes. Pixel sizes are taken to be 5 and 2 arcminutes. The simulations were done on square grids of side 3072 pixels, and the error bars (from run-to-run variations) rescaled to mimic 75\% sky coverage. Note that the large pixels in MAP introduce discreteness errors for separations up to about 20 arcminutes, and the theoretical MAP curve is reliable only for separations greater than 15 arcminutes.} \end{figure} In Fig. \ref{MAPPlanck}, we show the expected errors from MAP and Planck for a mixed dark matter model. The data were obtained by repeatedly simulating part of the sky on a grid, and rescaling the errors to simulate 75\% sky coverage. Details are given in the caption. \section{Discussion} In this paper we have shown that the peak-peak correlation function for the microwave sky can be computed to high accuracy for gaussian fields. The only approximation we make is that the sky is flat, but since the features in the correlation function appear on scales less than about 2 degrees, the computations should be very accurate. With upcoming high-resolution CMB experiments MAP and Planck, these results should allow a high-precision test of the hypothesis that structure grows from gaussian initial fluctuations, such as predicted in most inflationary models. We have not computed the expected peak-peak correlation function from defect models, but it is clear from the abundance of structure in the correlation function that agreement with these predictions, for all peak heights, would constitute a powerful argument in favour of gaussian models. We have also investigated numerically the errors expected from MAP and Planck; many of the features in the correlation function should be detectable with high significance. We note the similarity in the number of features in the peak-peak correlation function as in the power spectrum (although there may not be a simple correspondence between the two). The use of the two curves will, however, be rather different. To fit the power spectrum, one has the freedom to adjust fifteen or more parameters, under the gaussian hypothesis. With the peak-peak correlation function, there are no free parameters; for a gaussian process, all the statistics are specified once the power spectrum is fixed. Thus there is no freedom. Note also that the information required for the gaussian test is just the power spectrum; it does not rely on a satisfactory parameter estimation. For practical purposes, there are some issues still to address; detector noise can by easily incorporated into the analysis provided it is gaussian. Practical de-striping algorithms (e.g. \pcite{Delab98a}, \pcite{Revenu98}) to reduce $1/f$ noise, or methods to remove sidelobe contamination \cite{Delab98b} will need to be tested numerically to ensure that residual non-gaussian contributions are small. One interesting possibility for dealing with residual point sources in the maps is to put some constraint on the curvature at the peak (through the parameter $x$) to exclude sharply-peaked sources. The same trick could be used to exclude very flat-topped peaks whose precise locations may be difficult to determine accurately after dust foreground subtraction. Of course, both these techniques can be applied to the data as well as incorporated into the theory. There is an alternative to this procedure for point sources uncorrelated with the CMB; the peak distribution is then a superposition of two point processes, and provided the flux and clustering properties of the point sources are known, their contribution to the power spectrum can be subtracted. The gaussian test made on the joint distribution, as the correlation function will be a suitably-weighted combination of the two components. An aside which is probably not relevant for this application is that the analysis applies equally to any monotonic local functions of 2D gaussian fields, as nothing changes except the interpretation of the threshold, although in general the non-gaussianity would be revealed in the one-point temperature distribution. In addition, one might expect any transformation process to be local in 3D, in which case the mixing of information through the last scattering surface would destroy the local nature.
\section{#1}} \unitlength1cm \begin{document} \title{Analysis of the statistical behavior of genetic cluster-exact approximation} \author{Alexander K. Hartmann\\ {\small <EMAIL>}\\ {\small Institut f\"ur theoretische Physik, Bunsenstr. 9}\\ {\small 37073 G\"ottingen, Germany}\\ {\small Tel. +49-551-399570, Fax. +49-551-399631}} \date{\today} \maketitle \begin{abstract} The genetic cluster-exact approximation algorithm is an efficient method to calculate ground states of EA spin glasses. The method can be used to study ground-state landscapes by calculating many independent ground states for each realization of the disorder. The algorithm is analyzed with respect to the statistics of the ground states and the valleys of the energy landscape. Furthermore, the distribution inside each valley is evaluated. It is shown that the algorithm does not lead to a true $T=0$ thermodynamic distribution, i.e. each ground state has not the same frequency of occurrence when performing many runs. An extension of the technique is outlined, which guarantees that each ground states occurs with the same probability. {\bf Keywords (PACS-codes)}: Spin glasses and other random models (75.10.Nr), Numerical simulation studies (75.40.Mg), General mathematical systems (02.10.Jf). \pacs{75.10.Nr, 75.40.Mg, 02.10.Jf} \end{abstract} \mysection{Introduction} The finite-dimensional Edwards-Anderson spin glass \cite{binder86} is a model for disordered systems which has attracted much attention over the last decades. The opinion on its nature, especially for three dimensional systems, is still controversial \cite{parisi2,mcmillan,bray,fisher,bovier,newman}. Beside trying to address the problem with the help of analytic calculations and simulations at finite temperature, it is possible to investigate the behavior of the model by means of ground-state calculations \cite{rieger98}. Since obtaining spin-glass ground states is computationally hard \cite{barahona82}, the study is restricted to relatively small systems. Recently a new algorithm, the cluster-exact approximation (CEA) \cite{alex2} was presented, which allows in connection with a special genetic algorithm \cite{pal96} the calculation of true \cite{alex-stiff} ground states for moderate system sizes, in three dimensions up to size $14^3$. By applying this method it is possible to study the ground-state landscape of systems exhibiting a $T=0$ degeneracy \cite{alex-sg2}. For a thermodynamical correct evaluation it is necessary that each ground state contributes to the results with the same weight, since all ground states have exactly the same energy. Recently it was shown \cite{alex-false}, that the genetic CEA causes a bias on the quantities describing the $T=0$ landscape. The aim of this paper is to analyze the algorithm with respect to its ground-state statistics. The reasons for the deviation from the correct behavior are given and an extension of the method is outlined, which guarantees thermodynamical correct results. In this work, three-dimensional Edwards-Anderson (EA) $\pm J$ spin glasses are investigated. They consist of $N$ spins $\sigma_i = \pm 1$, described by the Hamiltonian \begin{equation} H \equiv - \sum_{\langle i,j\rangle} J_{ij} \sigma_i \sigma_j \end{equation} The sum runs over all pairs of nearest neighbors. The spins are placed on a three-dimensional (d=3) cubic lattice of linear size $L$ with periodic boundary conditions in all directions. Systems with quenched disorder of the interactions (bonds) are considered. Their possible values are $J_{ij}=\pm 1$ with equal probability. To reduce the fluctuations, a constraint is imposed, so that $\sum_{\langle i,j\rangle} J_{ij}=0$. The article is organized as follows: next a description of the algorithms is presented. Then it is shown for small systems, that the method does not result in a thermodynamical correct distribution of the ground states. In section four, the algorithm and its different variants are analyzed with respect to the ground-state statistics. In the last section a summary is given and an extension of the method is outlined, which should guarantee thermodynamical correct results. \mysection{Algorithms} The algorithm for the calculation bases on a special genetic algorithm \cite{pal96,michal92} and on cluster-exact approximation \cite{alex2}. CEA is an optimization method designed specially for spin glasses. Its basic idea is to transform the spin glass in a way that graph-theoretical methods can be applied, which work only for systems exhibiting no bond-frustrations. Now a short sketch of these algorithms is given, because later the influence of different variants on the results is discussed. Genetic algorithms are biologically motivated. An optimal solution is found by treating many instances of the problem in parallel, keeping only better instances and replacing bad ones by new ones (survival of the fittest). The genetic algorithm starts with an initial population of $M_i$ randomly initialized spin configurations (= {\em individuals}), which are linearly arranged using an array. The last one is also neighbor of the first one. Then $n_o \times M_i$ times two neighbors from the population are taken (called {\em parents}) and two new configurations called {\em offspring} are created. For that purpose the {\em triadic crossover} is used which turned out to be very efficient for spin glasses: a mask is used which is a third randomly chosen (usually distant) member of the population with a fraction of $0.1$ of its spins reversed. In a first step the offspring are created as copies of the parents. Then those spins are selected, where the orientations of the first parent and the mask agree \cite{pal95}. The values of these spins are swapped between the two offspring. Then a {\em mutation} with a rate of $p_m$ is applied to each offspring, i.e. a randomly chosen fraction $p_m$ of the spins is reversed. Next for both offspring the energy is reduced by applying CEA: The method constructs iteratively and randomly a non-frustrated cluster of spins. During the construction of the cluster a local gauge-transformation of the spin variables is applied so that all interactions between cluster spins become ferromagnetic. Fig. \ref{figCEAExample} shows an example of how the construction of the cluster works for a small spin-glass system. To increase the performance, spins adjacent to many unsatisfied bonds are more likely to be added to the cluster. This may introduce a bias on the resulting distribution of the ground states. Later this scheme (``BIAS'') is compared to a variant (``SAME''), where all spins may contribute to the cluster with the same probability. For 3d $\pm J$ spin glasses each cluster contains typically 55 percent of all spins. The non-cluster spins remain fixed during the following calculation, they act like local magnetic fields on the cluster spins. Consequently, the ground state of the gauge-transformed cluster is not trivial, although all interactions inside the cluster are ferromagnetic. Since the cluster exhibits no bond-frustration, an energetic minimum state for its spins can be calculated in polynomial time by using graph-theoretical methods \cite{claibo,knoedel,swamy}: an equivalent network is constructed \cite{picard1}, the maximum flow is calculated \cite{traeff,tarjan} and the spins of the cluster are set to orientations leading to a minimum in energy. Please note, that the ground state of the cluster is degenerate itself, i.e. the spin orientations can be chosen in different ways leading all to the same energy. It is possible to calculate within one single run a special graph, which represents all ground states of the cluster \cite{picard2}, and select one ground state randomly. This procedure is called ``BROAD'' here. On the other hand, one can always choose a certain ground state of the cluster directly\footnote{This ground state has the maximum possible magnetization of the {\em gauge-transformed} spins among all cluster ground states.}. Usually this variant, which is called ``QUICK'' here, is applied, because it avoids the construction of the special graph. But this again introduces a certain bias on the resulting distribution of the ground states. Later the influence of the different methods of choosing ground states is discussed. This CEA minimization step is performed $n_{\min}$ times for each offspring. Afterwards each offspring is compared with one of its parents. The offspring/parent pairs are chosen in the way that the sum of the phenotypic differences between them is minimal. The phenotypic difference is defined here as the number of spins where the two configurations differ. Each parent is replaced if its energy is not lower (i.e. not better) than the corresponding offspring. After this whole step is conducted $n_o \times M_i$ times, the population is halved: From each pair of neighbors the configuration which has the higher energy is eliminated. If more than 4 individuals remain the process is continued otherwise it is stopped and the best individual is taken as result of the calculation. The following representation summarizes the algorithm. \newlength{\mpwidth} \setlength{\mpwidth}{\textwidth} \addtolength{\mpwidth}{-2cm} \begin{center} \begin{minipage}[b]{\mpwidth} \newlength{\tablen} \settowidth{\tablen}{xxx} \newcommand{\hspace*{\tablen}}{\hspace*{\tablen}} \begin{tabbing} \hspace*{\tablen} \= \hspace*{\tablen} \= \hspace*{\tablen} \= \hspace*{\tablen} \= \hspace*{\tablen} \= \hspace*{\tablen} \= \kill {\bf algorithm} genetic CEA($\{J_{ij}\}$, $M_i$, $n_o$, $p_m$, $n_{\min}$)\\ {\bf begin}\\ \> create $M_i$ configurations randomly\\ \> {\bf while} ($M_i > 4$) {\bf do}\\ \> {\bf begin}\\ \> \> {\bf for} $i=1$ {\bf to} $n_o \times M_i$ {\bf do}\\ \>\> {\bf begin}\\ \>\>\> select two neighbors \\ \>\>\> create two offspring using triadic crossover\\ \>\>\> do mutations with rate $p_m$\\ \>\>\> {\bf for} both offspring {\bf do}\\ \>\>\> {\bf begin}\\ \>\>\>\> {\bf for} $j=1$ {\bf to} $n_{\min}$ {\bf do}\\ \>\>\>\> {\bf begin}\\ \>\>\>\>\> construct unfrustrated cluster of spins\\ \>\>\>\>\> construct equivalent network\\ \>\>\>\>\> calculate maximum flow\\ \>\>\>\>\> construct minimum cut\\ \>\>\>\>\> set new orientations of cluster spins\\ \>\>\>\> {\bf end}\\ \>\>\>\> {\bf if} offspring is not worse than related parent \\ \>\>\>\> {\bf then}\\ \>\>\>\>\> replace parent with offspring\\ \>\>\> {\bf end}\\ \>\> {\bf end}\\ \>\> half population; $M_i=M_i/2$\\ \> {\bf end}\\ \> {\bf return} one configuration with lowest energy\\ {\bf end} \end{tabbing} \end{minipage} \end{center} The whole algorithm is performed $n_R$ times and all configurations which exhibit the lowest energy are stored, resulting in $n_G$ statistically independent ground-state configurations ({\em replicas}). A priori nothing about the distribution of ground states raised by the algorithm is known. Thus, it may be possible that for one given realization of the disorder some ground states are more likely to be returned by the procedure than others. Consequently, any quantities which are calculated by averaging over many independent ground states, like the distribution of overlaps, may depend on a bias introduced by the algorithm. For a thermodynamical correct evaluation all ground states have to contribute with the same weight, since they all have exactly the same energy. For the preceding work, the distribution of the ground states determined by the algorithm was taken. The method was utilized to examine the ground state landscape of two-dimensional \cite{alex-2d} and three-dimensional \cite{alex-sg2,alex-ultra} $\pm J$ spin glasses by calculating a small number of ground states per realization. Some of these results depend on the statistics of the ground states, as it will be shown in the next section for the $d=3$ case. On the other hand, the main findings of the following investigations are not affected by the bias introduced by genetic CEA: the existence of a spin-glass phase for nonzero temperature was confirmed for the three-dimensional spin glass \cite{alex-stiff}. The method was applied also to the $\pm J$ random-bond model to investigate its $T=0$ ferromagnetic to spin-glass transition\cite{alex-threshold}. Finally, for small sizes up to $L=8$ all ground-state valleys were obtained by calculating a huge number of ground states per realization and applying a new method called {\em ballistic search} \cite{alex-valleys}. \mysection{Numerical evidence} In this section results describing the ground-state landscape of small three-dimensional $\pm J$ spin glasses are evaluated. It is shown that the data emerging from the use of raw genetic CEA and from a thermodynamically correct treatment differ substantially. Several ground states for small systems of size $N=L^3=3^3, 4^3, 5^3$ were calculated. 1000 realizations of the disorder for $L=3,4$ and 100 realizations for $L=5$ were considered. The parameters ($M_i$, $n_o$, $p_m$, $n_{\min}$), for which true ground states are obtained, are shown in \cite{alex-stiff}. For all calculations the variants BIAS and QUICK were used to obtain maximum performance. The effect of different variants on the results is discussed in the next section. Two schemes of calculation were applied: \begin{itemize} \item[A] For each realization $n_R=40$ runs of genetic CEA were performed and all states exhibiting the ground-state energy stored. Consequently, this scheme reflects the ground-state statistics which is determined solely by the genetic CEA method. Configurations which have a higher probability of occurrence contribute with a larger weight to the results. \item[B] For each realization the algorithm was run up to $10^5$ times. Each particular state was stored only once. For later analysis the number of times each state occurred was recorded. Additionally, a systematic local search was applied to add possibly missing ground states which are related by flips of free spins to states already found. Finally, a $L=3$ realization exhibits 25 different ground states on average. For a $L=4$ realization on average 240 states were found and 6900 states for $L=5$. For the evaluation of physical quantities every ground state is taken with the same probability in this scheme. Thus, the statistics obtained in this way reflect the true $T=0$ thermodynamic behavior. \end{itemize} To analyze the ground-state landscape, the distribution of overlaps is evaluated. For a fixed realization $J=\{J_{ij}\}$ of the exchange interactions and two replicas $\{\sigma^{\alpha}_i\}, \{\sigma^{\beta}_i\}$, the overlap \cite{parisi2} is defined as \begin{equation} q^{\alpha\beta} \equiv \frac{1}{N} \sum_i \sigma^{\alpha}_i \sigma^{\beta}_i \end{equation} The ground state of a given realization is characterized by the probability density $P_J(q)$. Averaging over the realizations $J$, denoted by $[\,\cdot\,]_{J}$, results in ($Z$ = number of realizations) \begin{equation} P(q) \equiv [P_J(q)]_{J} = \frac{1}{Z} \sum_{J} P_J(q) \label{def_P_q} \end{equation} Because no external field is present the densities are symmetric: $P_J(q) = P_J(-q)$ and $P(q) = P(-q)$. So only $P(|q|)$ is relevant. The result of $P(|q|)$ for $L=5$ is shown in Fig. \ref{figPqLfive}. For the true thermodynamic result small overlaps occur less frequent than for the data obtained by the application of pure genetic CEA. Large overlap values occur more often. This deviation has an influence on the way the spin glass behavior is interpreted. The main controversy about finite-dimensional spin glasses mentioned at the beginning is about the question whether for the infinite system $P(|q|)$ shows a long tail down to $q=0$ or not \cite{parisi2,mcmillan,bray,fisher,bovier,newman}. To investigate the finite size behavior of $P(|q|)$ the fraction $X_{0.5}$ of the distribution below $q_0=0.5$ is integrated: \begin{equation} X_{q_0} \equiv \int_0^{q_0} P(|q|) \, dq \end{equation} The development of $X_{0.5}$ as a function of system size $L$ is shown in Fig. \ref{figXL}. The datapoints for the larger sizes $L\ge 6$, obtained using pure genetic CEA, are taken from former calculations \cite{alex-sg2}. These values are more or less independent of the system size, while the correct thermodynamic behavior shows a systematic decrease. Whether for $L\to\infty$ the long tail of $P(|q|)$ persists cannot be concluded from the data, because the systems are too small. Nevertheless, the true $T=0$ behavior differs significantly from the former results. \mysection{Analysis of genetic CEA} To understand, why genetic CEA fails in producing the thermodynamical correct results, in this section the statistics of the ground states, which is determined by the algorithm, is analyzed directly. For the case where all ground states were calculated using a huge number of runs, the frequencies each ground state occurred were recorded. In Fig. \ref{figHistogramm} the result for one sample realization of $N=5^3$ is shown. The system has 56 different ground states. For each state the number of times it was returned by the algorithm in $10^5$ runs is displayed. Obviously the large deviations from state to state cannot be explained by the presence of statistical fluctuations. Thus, genetic CEA samples different ground states from the same realization with different weights. To make this statement more precise, the following analysis was performed: Two ground states are called {\em neighbors}, if they differ only by the orientation of one spin. All ground states which are accessible from each other through this neighbor-relation are defined to be in the same ground-state {\em valley}. That means, two ground states belong to the same valley, if it is possible to move from one state to the other by flipping only free spins, i.e. without changing the energy. For all realizations the valleys were determined using a method presented in \cite{alex-valleys}, which allows to treat systems efficiently exhibiting a huge number of ground states. Then the frequencies $h_V$ for each valley $V$ were computed as the sum of all frequencies of the states belonging to $V$. In Fig. \ref{figHistoSample} the result is shown for a sample $N=5^3$ realization, which has 15 different ground state valleys. Large valleys are returned by the algorithm more frequently, but $h_V$ seems to grow slower than linearly. A strict linear behavior should hold for an algorithm which guarantees the correct $T=0$ behavior. For averaging $h_V$ has to be normalized, because the absolute values of the frequency differ strongly from realization to realization, even if the size $|V|$ of a valley, i.e. the number of ground states belonging to it, is the same. For each realization, the normalized frequency $h^{*}_V$ is measured relatively to the average frequency $\overline{h}_1$ of all valleys of size 1: $h^{*}_V\equiv h_V/\overline{h}_1$ If a realization does not exhibit a valley consisting only of one ground state, the frequency $h_{V_s}$ of the smallest valley $V_s$ is taken. It is assumed, that the normalized frequency exhibits a $h^{*}_V=|V|^{\alpha}$ dependence, which is justified by the results shown later. Consequently, for the case the size $|V_s|$ of the smallest valley is larger than one, $\overline{h}_1\equiv h_{V_s}/|V_s|^{\alpha}$ is chosen. The value of $\alpha$ is determined self-consistently. The result for $L=3$ of $h^{*}_V$ as a function of the valley-size $|V|$ is presented in Fig. \ref{figHistoLthree}. A value of $\alpha=0.854(3)$ was determined. Please note, that the fluctuations for larger valleys are higher, because quite often only one valley was available for a given valley-size. The algebraic form is clearly visible, proving that genetic CEA overestimates systematically the importance of small ground-state valleys. For $L=4$ a value of $\alpha=0.705(3)$ was obtained, while the $L=5$ case resulted in $\alpha=0.642(5)$. Consequently, with increasing system size, the algorithm fails more and more to sample configurations from different ground-state valleys according to the size of the valleys. This explains, why the difference of $X_{0.5}(L)$ between the correct result and the values obtained in \cite{alex-sg2} increases with growing system size. Similar results were obtained for two-dimensional systems. For $L=5$ a self-consistent value of $\alpha=0.650(1)$ was found, while the treatment of $L=7$ systems resulted in $\alpha=0.659(2)$. Here only a slight finite-size dependence occurs. This may explain the fact, that the width of the distribution of overlaps, even calculated only by the application of pure genetic CEA, seems to scale to zero \cite{alex-2d}. In the second section of this paper two variants of the algorithm were presented, which may be able to calculate ground states more equally distributed. To investigate this issue, similar ground-state calculations were conducted for $L=4$ and again $h^{*}_V$ was calculated. For the case, were SAME was used instead of BIAS, a value $\alpha= 0.801(2) $ was determined self-consistently. Using BROAD instead of QUICK resulted in $\alpha=0.749(3)$. Finally, by applying SAME and BROAD together, $\alpha=0.843(3)$ was obtained. Consequently, applying different variants of the method decreases the tendency of overestimating small valleys, but the correct thermodynamic behavior is not obtained as well. Even worse, BROAD and SAME are considerably slower than the combination of QUICK and BIAS. So far it was shown, that genetic CEA fails in sampling ground states from different valleys according the size of the valleys. Now we turn to the question, whether at least states belonging to the same valley are calculated with the correct thermodynamic distribution. By investigating the frequencies of different ground states belonging to the same valley it was found again, that these configurations are not equally distributed. But it is possible to study this issue in a more physical way. For that purpose ground states of 100 $L=10$ realizations were calculated. Then the valley structure was analyzed. The average distribution of overlaps was evaluated, but only contributions of pairs of states belonging to the same valley were considered. For comparison, for the same realizations a long $T=0$ Monte-Carlo (MC) simulation was performed, i.e. randomly spins were selected and flipped if they were free. The ground states were used as starting configurations. Since a MC simulation ensures the correct thermodynamic distribution of the states, all ground states of a valley appear with the same frequency, if the simulation is only long enough. A length of 40 Monte-Carlo steps per spin were found to be sufficient for $L=10$. The result for the distribution of overlaps $P_{\mbox{\small valley}}(|q|)$ restricted to the valleys is displayed in Fig. \ref{figPqValley}. Significant differences between the datapoints from the pure genetic CEA and the correct $T=0$ behavior are visible. Consequently, the algorithm does not sample configurations belonging to the same ground-state valley with the same weight as well. \mysection{Conclusion} In this work the genetic cluster-exact approximation method is analyzed. The algorithm can be used to calculate many independent true ground states of EA spin glasses. The results from the raw application of the method and from calculations of {\em all} ground states for small system sizes were compared. By evaluating the distribution of overlaps is was shown, that genetic CEA imposes some bias on the ground-state statistics. Consequently, the results from the application of the raw method do not represent the true $T=0$ thermodynamics. To elucidate the behavior of the algorithm the statistics of the ground states were evaluated directly. It was shown, that different ground states have dissimilar probabilities of occurrence. To understand this effect better, the ground-state valleys were determined. The genetic CEA method finds configurations from small ground-state valleys relative to the size of the valley more often than configurations from large valleys. Additionally, within a valley the states are not sampled with the same weight as well. It was shown that two variants of the algorithm, which decrease its efficiency, weaken the effect, but it still persists. Summarizing, two effects are responsible for the biased ground-state sampling of genetic CEA: small valleys are sampled too frequently and the distribution within the valleys is not flat. For small system sizes it is possible to calculate all ground states, so one can obtain the true thermodynamic average directly. But already for $L=5$ there are realizations exhibiting more than $10^5$ different ground states. Since the ground-state degeneracy grows exponentially with system size \cite{alex-valleys} larger systems cannot be treated in this way. The following receipt should overcome these problems and should allow to obtain the true thermodynamic $T=0$ behavior for larger systems: \begin{itemize} \item Calculate several ground states of a realization using genetic CEA. \item Identify the ground states which belong to the same valleys. \item Estimate the size of each valley. This can be done using a variant of ballistic search \cite{alex-valleys}, which works by flipping free spins sequentially, each spin at most once. The number of spins flipped is a quite accurate measure for the size of a valley. \item Sample from each valley a number of ground states, which is proportional to the size of the valley. This guarantees, that each valley contributes with its proper weight. Each state is obtained by performing a $T=0$ MC simulation of sufficient length, starting with true ground-state configurations. Since MC simulations achieve a thermodynamical correct distribution, it is guaranteed that the states within each valley are equally distributed. \end{itemize} Please note, that it is not necessary to calculate all ground states to obtain the true thermodynamic behavior, because it is possible to estimate the size of a valley by analyzing only some sample ground states belonging to it. Furthermore, it is even only necessary to have configurations from the largest valleys available, since they dominate the ground-state behavior. This condition is fulfilled by genetic CEA, because large valleys are sampled more often than small valleys, even if small valleys appear too often relatively. From the results presented here it is not possible to deduce the correct $T=0$ behavior of the infinite system, because the system sizes are too small. Using the scheme outlined above, it is possible to treat system sizes up to $L=14$ \cite{alex-equi}. \mysection{Acknowledgements} The author thanks K. Battacharya and A.W. Sandvik for interesting discussions. The work was supported by the Graduiertenkolleg ``Modellierung und Wissenschaftliches Rechnen in Mathematik und Naturwissenschaften'' at the {\em In\-ter\-diszi\-pli\-n\"a\-res Zentrum f\"ur Wissenschaftliches Rechnen} in Heidelberg and the {\em Paderborn Center for Parallel Computing} by the allocation of computer time. The author announces financial support from the DFG ({\em Deutsche Forschungsgemeinschaft}).
\subsection{General remarks about scaling limits} It is often the case that grid-based probabilistic models should be considered as a mere substitute, or simplification, of a continuous process. There are definite advantages for working in the discrete setting, where unpleasant technicalities can frequently be avoided, simulations are possible, and the setup is easier to comprehend. On the other hand, one is often required to pay some price for the simplification. When we adopt the grid-based world, we sacrifice rotational of conformal symmetries which the continuous model may enjoy, and often have to accept some arbitrariness in the formulation of the model. There are also numerous examples where the continuous process is easier to analyze than the discrete process, and in such situations the continuous may be a useful simplification of the discrete. Understanding the connections between grid-based models and continuous processes is a project of fundamental importance, and so far has only limited success. As mathematicians, we should not content ourselves with the vague notion that the discrete and continuous models behave ``essentially the same'', but strive to make the relations concrete and precise. One reasonable way to define a continuous process, is by taking a {\bf scaling limit} of a grid process. This means making sense of the limit of a sequence of grid processes on finer and finer grids. Recently, Aizenman \ref b.Aizenman:web/ has proposed a definition for the scaling limit of percolation, and we shall propose a somewhat different definition \ref b.Schramm:inprep/. Although, in general, the understanding of the connections between grid-based models and continuous models is lacking, there have been some successes. The classical and archetypical example is the relation between simple random walk (SRW) and Brownian motion, which is well studied and quite well understood. See, for example, the discussion of Donsker's Theorem in \ref b.Durrett:probability/. We also mention that recently, T\'oth and Werner \ref b.TW/ have described the scaling limit of a certain self-repelling walk on $\Z$. The present paper deals with the scaling limits of two very closely related processes, the loop-erased random walk (LERW) and the uniform spanning tree (UST). While these processes are interesting also in dimensions $3$ and higher, we restrict attention to two dimensions. In the plane, the scaling limits are conjectured to possess conformal invariance (precise statements appear below), and this can serve as one justification for the special interest in dimension $2$. The recent preprint by Aizenman, Burchard, Newman and Wilson \ref b.ABNW/ discusses scaling limits of random tree processes in two dimensions, including the UST. The present work answers some of the questions left open in \ref b.ABNW/. The most fundamental task in studying a scaling limit process is to set up a conceptual foundation for the scaling limit. This means answering the following two questions: what kind of object is the scaling limit, and what does it mean to be the scaling limit? For the first question, there's often more than one \lq\lq right\rq\rq{} answer. For example, the scaling limit of (two-sided) simple random walk in $\Z$ is usually defined as a probability measure on $C(\R)$, the space of continuous, real-valued functions on $\R$, but it could also be defined as a measure on the space of closed subsets of $\R\times\R$, with an appropriately chosen metric. There's also a \lq\lq wrong\rq\rq{} answer here. It is not a good idea to consider the scaling limit of SRW as a probability measure on $\R^\R$, though this might seem at first as more natural. After the conceptual framework is fixed, the next natural question is the existence of the scaling limit. Unfortunately, we cannot report on any progress here. There is every reason to believe that UST and LERW have scaling limits, but a proof is still lacking. However, in the setup we propose below, the existence of subsequential scaling limits is almost a triviality: for every sequence of positive $\delta_j$ tending to $0$, there is a subsequence $\delta_{j_n}$ such that the UST and the LERW on the grids $\delta_{j_n}\Z^2$ do converge to a limit as $n\to\infty$. Such a limit is called a subsequential scaling limit of the model, and is a probability measure on some space. All the above discussion concerns foundational issues, which are important. But it is not less important to prove properties of the (subsequential) scaling limit% \Ftnote{Sometimes, this can be done without even defining the scaling limit. For example, the Russo-Seymour-Welsh Theorem \ref b.Russo/, \ref b.SW/ in percolation theory implies properties of any reasonably defined percolation scaling limit. Similarly, Benjamini's preprint \ref b.Benjamini:ustsl/ does not explicitly discuss the scaling limit of UST, but has implication to the UST scaling limit.}. In this paper, we prove several almost sure properties of the UST and LERW subsequential scaling limit. We now describe these models and explain the results. \subsection{The LERW model, and its scaling limit} Consider some set of vertices $K\neq\emptyset$, in a recurrent graph $G$, and a vertex $v_0$. The LERW from $v_0$ to $K$ in $G$ is obtained by running simple random walk (SRW) from $v_0$, erasing loops as they are created, and stopping when $K$ is hit. Here's a more precise description. Let ${\ss RW}$ be simple random walk starting at ${\ss RW}(0)=v_0$ and stopped at the first time $\tau$ such that ${\ss RW}(\tau)\in K$. Its loop-erasure, ${\ss LE}={\ss LE}_{{\ss RW}}$, is defined inductively as follows: ${\ss LE}(0):=v_0$, and ${\ss LE}(j+1)={\ss RW}(t+1)$ if $t$ is the last time less than $\tau$ such that ${\ss RW}(t)={\ss LE}(j)$. The walk ${\ss LE}$ stops when it gets to ${\ss RW}(\tau)\in K$. Note that this LERW is a random simple path from $v_0$ to $K$. On a transient graph, it may happen that ${\ss RW}$ does not hit $K$. However, one can discuss the loop-erasure of the walk continued indefinitely, since it a.s.\ visits any vertex only finitely many times. LERW on $\Z^d$ was studied extensively by Greg Lawler (see the survey paper \ref b.Lawler:survey/ and the references therein), who considered LERW as a simpler substitute for the self-avoiding random walk (see the survey \ref b.Slade:survey/), which is harder to analyze. However, we believe that the LERW model is just as interesting mathematically, because of its strong ties with SRW and UST. For compactness's sake, in the following we consider the plane ${\Bbb C}=\R^2$ as a subset of the two-sphere, $\Sp2={\Bbb C}\cup\{\infty\}$, which is the one point compactification of the plane, and work with the spherical metric $\dsp$ on $\Sp2$. Let $D$ be a domain (nonempty open connected set) in the plane ${\Bbb C}=\R^2$. We consider a graph $G=G(D,\delta)$, which is an approximation of the domain $D$ in the square grid $\delta\Z^2$ of mesh $\delta$. The {\bf interior vertices}, ${\verts_I}(G)$, of $G$ are the vertices of $\delta\Z^2$ which are in $D$, and the {\bf boundary vertices}, ${\verts_{\partial}}(G)$, are the intersections of edges of $\delta\Z^2$ with $\partial D$. (The precise definition of $G$ appears in \ref s.back/.) Suppose that each component of $\partial D$ has positive diameter. Let $a\in D$, and let ${\ss LE}={\ss LE}_{a,D,\delta}$ be LERW from a vertex $a'\in\delta\Z^2\cap D$ closest to $a$ to ${\verts_{\partial}}(G)$ in $G$. To make sense of the concept of the {\bf scaling limit} of LERW in $D$, we think of ${\ss LE}$ as a random set in $\overline D$. Recall that the Hausdorff distance $\dHa(X,Y)$ between two closed nonempty sets $X$ and $Y$ in a compact metric space $Z$ is the least $t\geqslant 0$ such that each point $x\in X$ is within distance $t$ from $Y$ and each point $y\in Y$ is within distance $t$ from $X$; that is, $\dHa(X,Y)=\inf\left\{t\geqslant 0:\, X\subset\bigcup_{x\in X} B(x,t),\ Y\subset\bigcup_{y\in Y} B(y,t)\right\}$. On the collection ${\Cal H}(D)$ of closed subsets of $D$, we use the metric $\dHc D(X,Y):=\dHa(X\cup\partial D,Y\cup\partial D)$, and ${\Cal H}(D)$ is compact with this metric. Then ${\ss LE}\cap D$ is a random element in ${\Cal H}(D)$, and its distribution $\mu_\delta=\mu_{\delta,D}$ is a probability measure on ${\Cal H}(D)$. Because the space of Borel probability measures on a compact space is compact in the weak topology% \Ftnote{We review the notion of weak convergence in \ref s.back/.}% , there is a sequence $\delta_j\to 0$ such that the weak limit $\mu_0:=\lim_{j\to\infty}\mu_{\delta_j}$ exists. Such a measure $\mu_0$ will be called a {\bf subsequential scaling limit} measure of LERW from $a$ to $\partial D$. If $\mu_0=\lim_{\delta\to0}\mu_\delta$, then we say that $\mu_0$ is the {\bf scaling limit} measure of LERW from $a$ to $\partial D$. Similarly, we may consider the scaling limit of LERW between two distinct points $a,b\in\Sp2$, as follows. For $\delta>0$, we take ${\ss LE}$ to be the loop-erasure of SRW on $\delta\Z^2$ starting from a vertex of $\delta\Z^2$ within distance $2\delta$ of $a$ and stopped when it first hits a vertex within distance $2\delta$ of $b$. Since ${\ss LE}$ is a.s.\ compact, its distribution is an element of the Hausdorff space ${\Cal H}(\Sp2)$, and there exists a Borel probability measure on ${\Cal H}(\Sp2)$, which is a subsequential scaling limit measure of the law of ${\ss LE}$. \procl t.noloop Let $D$ be a domain in $\Sp2$ such that each connected component of $\partial D$ has positive diameter, and let $a\in D$. Then every subsequential scaling limit measure of LERW from $a$ to $\partial D$ is supported on simple paths. Similarly, if $a,b$ are distinct points in $\Sp2$, then every subsequential scaling limit of the LERW from $a$ to $b$ in $\delta\Z^2$ is supported on simple paths. \endprocl Saying that the measure is supported on simple paths means that there's a collection of simple paths whose complement has zero measure. \subsection{The conformal invariance conjecture for LERW} Consider two domains $D,D'\subset \Sp2$. Every homeomorphism $f:D\to D'$ induces a homeomorphism ${\Cal H}(D)\mapsto{\Cal H}(D')$. Consequently, if $\mu$ is a probability measure on ${\Cal H}(D)$, there is an induced probability measure $f_*\mu$ on ${\Cal H}(D')$. \procl g.confinv Let $D\subsetneqq{\Bbb C}$ be a simply connected domain in ${\Bbb C}$, and let $a\in D$. Then the scaling limit of LERW from $a$ to $\partial D$ exists. Moreover, suppose that $f:D\to D'$ is a conformal homeomorphism onto a domain $D'\subset{\Bbb C}$. Then $f_*\mu_{a,D}=\mu_{f(a),D'}$, where $\mu_{a,D}$ is the scaling limit measure of LERW from $a$ to $\partial D$, and $\mu_{f(a),D'}$ is the scaling limit measure of LERW from $f(a)$ to $\partial D'$. \endprocl Although conformal invariance conjectures have been \lq\lq floating in the air\rq\rq{} in the physics literature for quite some time now, we believe that this precise statement has not yet appeared explicitly. Support for this conjecture comes from simulations which we have performed, and from the work of Rick Kenyon~\ref b.Kenyon:conf/, \ref b.Kenyon:5o4/, \ref b.Kenyon:longrange/. We prove that \ref g.confinv/ implies an explicit description of the LERW scaling limit in terms of solutions of L\"owner{}'s differential equation with a Brownian motion parameter. We now give a brief explanation of this. \@midtrue\@ins \SetLabels \R(0.49*0.5)$0$\\ {\ss T}\R(0.55*0.4)$\beta$\\ \L{\ss B}(0.75*0.36)$q$\\ {\ss T}\R(0.79*0.24)$\beta_q$\\ \endSetLabels \centerline{\AffixLabels{\epsfysize=2.4in\epsfbox{gamma.eps}}} \caption{\figlabel{gamma}\enspace} \endinsert Let ${\Bbb U}:=\{z\in{\Bbb C}:\, |z|<1\}$, the unit disk. If $\gamma$ is a compact simple path in $\overline U-\{0\}$, such that $\gamma\cap\partial {\Bbb U}$ is an endpoint of $\gamma$, then there is a unique conformal homeomorphism $f_\gamma:{\Bbb U}\to{\Bbb U}-\gamma$ such that $f_\gamma(0)=0$ and $f_\gamma'(0)>0$ (that is, $f_\gamma'(0)$ is real and positive). Moreover, if $\widehat \gamma$ is another such path, and $\widehat\gamma\supset\gamma$, then $f_\gamma'(0)>f_{\widehat\gamma}'(0)$. Now suppose that $\beta$ is a compact simple path in $\overline{\Bbb U}$ such that $\partial{\Bbb U}\cap\beta$ is an endpoint of $\beta$ and $0$ is the other endpoint of $\beta$, as in \ref f.gamma/. For each point $q\in\beta-\{0\}$, let $\beta_q$ be the arc of $\beta$ extending from $q$ to $\partial{\Bbb U}$, and let $h(q):=\log f'_{\beta_q}(0)$. (If $q$ is the endpoint of $\beta$ on $\partial{\Bbb U}$, then $f_{\beta_q}(z)=z$.) It turns out that $h$ is a homeomorphism from $\beta-\{0\}$ onto $(-\infty,0]$. We let $q(t)$ denote the inverse map $q:(-\infty,0]\to\beta$, and set $f(z,t)=f_t(z):=f_{\beta_{q(t)}}(z)$. In this setting, L\"owner{}'s Slit Mapping Theorem \ref b.Lowner/ (see also \ref b.Pommerenke:Lowner/) states that $f_t(z)$ is the solution of L\"owner{}'s equation $$ {\partial\over\partial t}f_t(z) = z f_t'(z){\zeta(t)+z\over\zeta(t)-z} \,,\qquad\forall z\in{\Bbb U}\,,\forall t\in(-\infty,0] \,, \label e.loew $$ where $\zeta:(-\infty,0]\to\partial{\Bbb U}$ is some continuous function. In fact, $\zeta(t)$ is defined by the equation $$ f\bigl(\zeta(t),t\bigr)=q(t)\,. $$ (The left hand side makes sense, since there is a unique continuous extension of $f_t$ to $\partial{\Bbb U}$.) Note that $f$ also satisfies $$ f(z,0)=z\,,\qquad\forall z\in{\Bbb U} \,. \label e.bdval $$ \procl t.lesl \procname{The differential equation for the LERW scaling limit} Assume \ref g.confinv/. Let ${\ss B}(t)$, $t\geqslant 0$ be Brownian motion on $\partial{\Bbb U}$ starting from a uniform-random point on $\partial{\Bbb U}$. Let $f(z,t)$ be the solution of \ref e.loew/ and \ref e.bdval/, with $\zeta(t):={\ss B}(-2 t)$. Set $$ \sigma(t)=f(\zeta(t),t)\,,\qquad t\leqslant 0 \,. \label e.sledef $$ Then $\{0\}\cup\sigma\bigl((-\infty,0)\bigr)$ has the same distribution as the scaling limit of LERW from $0$ to $\partial{\Bbb U}$. \endprocl The Brownian motion ${\ss B}(t)$ in the theorem can be defined as ${\ss B}(t):=\exp(i\widehat {\ss B}(t))$, where $\widehat{\ss B}(t)$ is ordinary Brownian motion on $\R$, starting at a uniform-random point in $[0,2\pi)$. \procl r.ode Although \ref e.loew/ may look like a PDE, it can in fact be presented as an ODE. Set $\Phi(z,t,s) := f_s^{-1}(f_t(z))$ when $t\leqslant s\leqslant 0$. Then \begineqalno \Phi(z,t,0)&=f_t(z)\,,\quad \cr \Phi(z,t,t)&=z \,. \label e.phis \endeqalno It is immediate to see that $f_t$ satisfies \ref e.loew/ iff $\Phi=\Phi(z,t,s)$ satisfies $$ {\partial\Phi \over\partial s}= -\Phi {\zeta(s)+\Phi\over\zeta(s)-\Phi} \,. \label e.phiode $$ Therefore, $f_t(z)$ can obtained by solving the ODE \ref e.phiode/ with $t$ fixed and $s\in[t,0]$. Note that $(\zeta+\Phi)/(\zeta-\Phi)$ has positive real part when $\Phi\in{\Bbb U}$ and $\zeta\in\partial{\Bbb U}$. Therefore, \ref e.phiode/ implies that $|\Phi(z,t,s)|$ is monotone decreasing as a function of $s$. {}From this it can be deduced that there is a unique solution to the system \ref e.phis/ and \ref e.phiode/ in the interval $s\in[t,0]$. \endprocl Obviously, \ref t.lesl/ together with \ref g.confinv/ describe the LERW scaling limit in any simply connected domain $D\subsetneqq{\Bbb C}$, since such domains are conformally equivalent to ${\Bbb U}$. It would be interesting to extract properties of the LERW scaling limit from \ref t.lesl/. At the heart of the proof of \ref t.lesl/ lies the following simple combinatorial fact about LERW. Conditioned on a subarc $\beta'$ of the LERW $\beta$ from $0$ to $\partial D$, which extends from some point $q\in\beta$ to $\partial D$, the distribution of $\beta-\beta'$ is the same as that of LERW from $0$ to $\partial (D-\beta')$, conditioned to hit $q$. (See \ref l.dprod/.) When we take the scaling limit of this property, and apply the conformal map from $\partial (D-\beta')$ to ${\Bbb U}$, this translates into the Markov property and stationarity of the associated L\"owner{} parameter $\zeta$. \subsection{The uniform spanning tree and its scaling limit} Shortly, a definition of the uniform spanning tree (UST) on $\Z^2$ will be given. The UST is a statistical-physics model. It lies in the boundary of the two-parameter family of random-cluster measures, which includes Bernoulli percolation and the Ising model \ref b.Hag:ustlim/. The UST is very interesting mathematically, partly because it is closely related to the theory of resistor networks, potential theory, random walks, LERW, and in dimension $2$, also domino tilings. The paper \ref b.BLPS:usf/ gives a comprehensive study of uniform spanning trees (and forests), following earlier pioneering work \ref b.Aldous:ust/, \ref b.Broder:ust/, \ref b.Pemantle:dichot/, \ref b.BP:ust/, \ref b.Hag:ustlim/. A survey of current UST theory can be found in \ref b.Lyons:usfsurvey/. Let $G$ be a connected graph. A {\bf forest} is a subgraph of $G$ that has no cycles. A {\bf tree} is a connected forest. A subgraph of $G$ is {\bf spanning} if it contains ${\ss V}(G)$, the set of vertices of $G$. We will be concerned with spanning trees. Since a spanning tree is determined by its edges, we often don't make a distinction between the spanning tree ${\ss T}$ and its set of edges ${\ss E}(T)$. If $G$ is finite, a uniform spanning tree (UST) in $G$ is a random spanning tree ${\ss T}\subset G$, selected according to the uniform measure. (That is, $\P[{\ss T}=T_1]=\P[{\ss T}=T_2]$, whenever $T_1$ and $T_2$ are spanning trees of $G$.) It turns out that UST's are very closely related to LERW's. If $a,b\in{\ss V}(G)$, then the (unique) path in the UST joining $a$ and $b$ has the same law as the LERW from $a$ to $b$ in $G$. (This, in particular, implies that the LERW from $a$ to $b$ has the same law as the LERW from $b$ to $a$.) Wilson's algorithm \ref b.Wilson:alg/, which will be described in \ref s.back/, is a very useful method to build the UST by running LERW's. R.~Lyons proposed (see \ref b.Pemantle:dichot/) to extend the notion of UST to infinite graphs. Let $G$ be an infinite connected graph. Consider a nested sequence of connected finite sugraphs $G_1\subset G_2\subset\cdots\subset G$ such that $G=\bigcup_j G_j$. For each $j$, the uniform spanning tree measure $\mu_j$ on $G_j$ may be considered as a measure on $2^{{\ss E}(G)}$, the $\sigma$-field of subsets of the edges of $G$, generated by the sets of the form $\bigl\{F\subset {\ss E}(G):\, e\in F\bigr\}$, $e\in{\ss E}(G)$. Using monotonicity properties, it can be shown that the weak limit $\mu:=\lim_{j\to\infty}\mu_j$ exists, and does not depend on the sequence $\{G_j\}$. It is called the free uniform spanning forest measure (FSF) on $G$. (The reason for the word `free', is that there's another natural kind of limit, the wired uniform spanning forest (WSF). On $\Z^d$, these two measures agree.) R.~Pemantle \ref b.Pemantle:dichot/ proved that if $d\leqslant 4$, the FSF measure on $\Z^d$ is supported on spanning trees (that is, if ${\ss T}$ is random and its law is the FSF measure, then ${\ss T}$ is a.s.\ a spanning tree), while if $d\geqslant 5$, the measure is supported on disconnected spanning forests. Let us now restrict attention to the case $G=\Z^2$. Since it is supported on spanning trees, we call the FSF measure on $\Z^2$ the {\bf uniform spanning tree} (UST) on $\Z^2$. I.~Benjamini \ref b.Benjamini:ustsl/ and R.~Kenyon~\ref b.Kenyon:longrange/ studied asymptotic properties of the UST on a rescaled grid $\delta\Z^2$, with $\delta$ small, but did not attempt to define the scaling limit. Aizenman, Burchard, Newman and Wilson \ref b.ABNW/ defined the scaling limit of UST (and other tree processes) in $\Z^2$, and studied some of their properties. We present a different definition for the scaling limit of the UST. Let $\delta>0$. Again, we think of $\delta\Z^2$ as a subset of the sphere $\Sp2=\R^2\cup\{\infty\}$. Let ${\hat \T}_\delta$ be the UST on $\delta\Z^2$, union with the point at infinity. Then ${\hat \T}_\delta$ can be thought of as a random compact subset of $\Sp2$. However, it is fruitless to consider the weak limit as $\delta\to0$ of the law of ${\hat \T}_\delta$ as a measure on the Hausdorff space ${\Cal H}(\Sp2)$, since the limit measure is an atomic measure supported on the single point in ${\Cal H}(\Sp2)$, which is all of $\Sp2$. Given two points $a,b\in{\hat \T}_\delta$, let $\omega_{a,b}$ be the unique path in ${\hat \T}_\delta$ with endpoints $a$ and $b$; allowing for the possibility $\omega_{a,b}=\{a\}$, when $a=b$. (It was proved by R.~Pemantle that a.s.\ the UST ${\ss T}$ in $\Z^2$ has a single end; that is, there is a unique infinite ray in ${\ss T}$ starting at $0$. This implies that indeed $\omega_{a,b}$ exists and is unique, not only for ${\ss T}$, but also for ${\ss T}\cup\{\infty\}$.) Let $\ST_\delta=\ST_\delta({\hat \T}_\delta)$ be the collection of all triplets, $(a,b,\omega_{a,b})$, where $a,b\in{\hat \T}_\delta$. Then $\ST_\delta$ is a closed subset of $\Sp2\times\Sp2\times{\Cal H}(\Sp2)$, and the law $\law_\delta$ of $\ST_\delta$ is a probability measure on the compact space ${\Cal H}\bigl(\Sp2\times\Sp2\times{\Cal H}(\Sp2)\bigr)$. By compactness, there is a subsequential weak limit $\mea$ of $\law_\delta$ as $\delta\to 0$, which is a probability measure on ${\Cal H}\bigl(\Sp2\times\Sp2\times{\Cal H}(\Sp2)\bigr)$. We call $\mea$ a {\bf subsequential UST scaling limit in $\Z^2$}. If $\mea:=\lim_{\delta\to 0}\law_\delta$, as a weak limit, then $\mea$ is the {\bf UST scaling limit}. We prove \procl t.ustlim Let $\mea$ be a subsequential UST scaling limit in $\Z^2$, and let $\ST\in {\Cal H}\bigl(\Sp2\times\Sp2\times{\Cal H}(\Sp2)\bigr)$ be a random variable with law $\mea$. Then the following holds a.s. \begingroup\parindent=25pt \itemrm{(i)} For every $(a,b)\in \Sp2\times\Sp2$, there is some $\omega\in{\Cal H}(\Sp2)$ such that $(a,b,\omega)\in\ST$. For almost every $(a,b)\in\Sp2\times\Sp2$, this $\omega$ is unique. \itemrm{(ii)} For every $(a,b,\omega)\in\ST$, if $a\neq b$, then $\omega$ is a simple path; that is, homeomorphic to $[0,1]$. If $a=b$, then $\omega$ is a single point or homeomorphic to a circle. For almost every $a\in\Sp2$, the only $\omega$ such that $(a,a,\omega)\in\ST$ is $\{a\}$. \itemrm{(iii)} The {\bf trunk}, $$ \Tr:=\bigcup_{(a,b,\omega)\in\ST}\bigl(\omega-\{a,b\}\bigr) \,, $$ is a topological tree (in the sense of \ref d.toptree/), which is dense in $\Sp2$. \itemrm{(iv)} For each $x\in\Tr$, there are at most three connected components of $\Tr-\{x\}$. \vskip0pt\endgroup\noindent \endprocl This theorem basically answers all the topological questions about the UST scaling limit on $\Z^2$. It is sharp, in the sense that all the ``almost every'' clauses cannot be replaced by ``every''. Benjamini~\ref b.Benjamini:ustsl/ proved a result which is closely related to item (iv) of the theorem, and \ref b.ABNW/ proved (in a different language) that (iv) holds with ``three'' replaced by some unspecified constant. The {\bf dual} of a spanning tree $T\subset \Z^2$ is the spanning subgraph of the dual graph $(1/2,1/2)+\Z^2$ containing all edges that do not intersect edges in $T$. It turns out that duality is measure preserving from the UST on $\Z^2$ to the UST on the dual grid. The key to the proof of \ref t.ustlim/ is the statement that the trunk is disjoint from the trunk of the dual UST scaling limit. Let us stress that \ref t.ustlim/ is not contingent on \ref g.confinv/. The only contingent theorems proved in this paper are \ref t.lesl/, and \ref t.confust/, which says that \ref g.confinv/ implies conformal invariance for the scaling limit of the UST on subdomains of ${\Bbb C}$. Recent work of R.~Kenyon \ref b.Kenyon:conf/, \ref b.Kenyon:5o4/ proves some conformal invariance results for domino tilings of domains in the plane. There is an explicit correspondence between the UST in $\Z^2$ and domino tilings of a finer grid. Based on this correspondence, some properties of the UST can be proved using Kenyon's machinery. For example, Kenyon has shown \ref b.Kenyon:5o4/ that the expected number of edges in a LERW joining two boundary vertices in $(\delta\Z^2)\cap [0,1]^2$ (whose distance from each other is bounded from below) grows like $\delta^{-5/4}$, as $\delta\to 0$. He can also show \ref b.Kenyon:longrange/ that the weak limit as $\delta\to0$ of the distribution of the UST meeting point of three boundary vertices of $D\cap(\delta\Z^2)$ is equivariant with respect to conformal maps. That can be viewed as a partial conformal invariance result for the UST scaling limit. Another example for the applications of Kenyon's work to the UST appears in \ref s.windl/. It seems plausible that perhaps soon there would be a proof of \ref g.confinv/. \subsection{SLE with other parameters, critical percolation, and the UST Peano curve} Let $\kappa\geqslant 0$, and take $\zeta(t):= {\ss B}(-\kappa t)$, where ${\ss B}$ is as above, Brownian motion on $\partial{\Bbb U}$, started from a uniform random point. Then there is a solution $f(z,t)$ of \ref e.loew/ and \ref e.bdval/, and for each $t\leqslant 0$ $f_t = f(\cdot,z)$ is a conformal map from ${\Bbb U}$ into some subdomain $D_t\subset{\Bbb U}$. We call the process $\sigma^\kappa_t:={\Bbb U}-D_t$, $t\leqslant 0$, the {\bf stochastic L\"owner\ evolution} (SLE) with parameter $\kappa$. It is not always the case that $\sigma^\kappa_t$ is a simple path. Let $\frak K$ be the set of all $\kappa\geqslant 0$ such that for all $t<0$ the set $\sigma^\kappa_t$ is a.s.\ a simple path. We show in \ref s.crit/ that $\sup\frak K\leq4$, and conjecture that $\frak K=[0,4]$. In the past, there has been some work on the question of which L\"owner{} parameters $\zeta$ produce slitted disk mappings (\ref b.Kufarev/, \ref b.Pommerenke:Lowner/), but only limited progress has been made. Partly motivated by the present work, Marshall and Rohde \ref b.MR:lowner/ have looked into this problem again, and have shown that when $\zeta$ satisfies a H\"older condition with exponent $1/2$ and H\"older(1/2) norm less than some constant, the maps $f_t$ are onto slitted disks, and this may fail when $\zeta$ has finite but large H\"older(1/2) norm. Given some $\kappa>0$, even if $\kappa\notin\frak K$, the process $\sigma^\kappa_t$, $t\leqslant 0$, is quite interesting. It is a celebrated conjecture that critical Bernoulli percolation on lattices in $\R^2$ exhibits conformal invariance in the scaling limit \ref b.Langlands:Bull/. Assuming such a conjecture, we plan to prove in a subsequent work that a process similar to SLE describes the scaling limit of the outer boundary of the union of all critical percolation clusters in a domain $D$ which intersect a fixed arc on the boundary of $D$. We also plan to prove that this implies Cardy's \ref b.Cardy/ conjectured formula for the limiting crossing probabilities of critical percolation, and higher order generalizations of this formula. Let us now briefly explain this. In \ref f.pcurve/, each of the hexagons is colored black with probability $1/2$, independently, except that the hexagons intersecting the positive real ray are all white, and the hexagons intersecting the negative real ray are all black. Then there is a boundary path $\beta$, passing through $0$ and separating the black and the white regions adjacent to $0$. Note that the percolation in the figure is equivalent to Bernoulli$(1/2)$ percolation on the triangular grid, which is critical. (See \ref b.Grimmett:book/ for background on percolation.) The intersection of $\beta$ with the upper half plane, ${\Bbb H}:=\{z\in{\Bbb C}:\, \operatorname{Im} z>0\}$, which is indicated in the picture, is a random path in ${\Bbb H}$ connecting the boundary points $0$ and $\infty$. \@midtrue\@ins \centerline{\epsfysize=2.4in\epsfbox{pcurve.ps}} \caption{\figlabel{pcurve}\enspace The boundary curve for critical percolation with mixed boundary conditions.} \endinsert A subsequential scaling limit of $\beta\cap{\Bbb H}$ exists, by compactness, and naturally, we believe that the weak limit exists. Let $\gamma$ be the scaling limit curve. The physics wisdom (unproven, perhaps not even precisely formulated, but well supported) is that the scaling limit of the ``external boundary'' of macroscopic critical percolation clusters in two dimensions has dimension $7/4$ and is {\bf not} a simple path \ref b.ADA/, and we believe that this is true for $\gamma$. In a subsequent paper, we plan to prove (by adapting the proof of \ref t.lesl/), under the assumption of a conformal invariance conjecture for the scaling limit of critical percolation, that $\gamma$ can be described using a L\"owner-like differential equation in the upper half plane with Brownian motion parameter, as follows. Consider the differential equation $$ {\partial\over\partial t}f_t(z) = {-2f_t'(z)\over\zeta(t)-z} \,,\qquad\forall z\in{\Bbb H}\,,\forall t\in(-\infty,0] \,, \label e.loewh $$ where $\zeta(t)={\ss B}(-\kappa t)$, ${\ss B}$ is Brownian motion on $\R$ starting at ${\ss B}(0)=0$, and $f_0(z)=z$. Then $f_t$ is a conformal mapping from ${\Bbb H}$ onto a subdomain of ${\Bbb H}$, which is normalized by the so-called {\bf hydrodynamic normalization} $$ \lim_{z\to\infty}f_t(z)-z=0\,. \label e.hydro $$ The claim is that for $\kappa=6$, the image of the path $t\mapsto f_t\bigl(\zeta(t)\bigr)$ has the same distribution as $\gamma$. {}From this, one can derive Cardy's \ref b.Cardy/ conjectured formula for the limiting crossing probabilities of critical percolation, as well as some higher order generalizations. This will be done in subsequent work, but basically depends on the ideas appearing in \ref s.crit/ below. A similar representation applies to the scaling limit of the Peano curve which winds around the UST (this curve was discussed in \ref b.Duplantier:Peano/ and mentioned in \ref b.BLPS:usf/), but with $\kappa=8$. Given a domain $D\subset\Sp2$, whose boundary is a simple closed path, and given two distinct points $a,b\in\partial D$, there is a naturally defined (subsequential) scaling limit of the Peano curve of the UST in $D$, with appropriate boundary conditions, and the scaling limit is an (unparameterized) curve from $a$ to $b$, whose image covers $\overline D$. One can show that \ref g.confinv/ implies a conformal invariance property for the scaling limit. Based on this, it should be possible to adapt the proof of \ref t.lesl/ to show that \ref g.confinv/ implies a representation of the form \ref e.loewh/ for this Peano scaling limit when $D={\Bbb H}$, $a=0$, $b=\infty$, and $\kappa=8$. We give a brief overview of this in \ref s.peano/, and hope to give a more thorough treatment in a subsequent paper. The differential equation \ref e.loewh/ is very similar to L\"owner{}'s equation, and the only essential difference is that a normalization at an interior point for the maps $f_t$ is replaced by the hydrodynamic normalization \ref e.hydro/ at a boundary point ($\infty$). The interior point normalization is natural for the LERW scaling limit, because the LERW is a path from an interior point to the boundary of the domain. The Peano curve and the boundary of percolation clusters, as discussed above, are paths joining two boundary points, and hence the hydrodynamic normalization is more appropriate for them. Although L\"owner{}'s Slit Mapping Theorem mentioned above applies to domains of the form ${\Bbb U}-\alpha$, where $\alpha$ is a simple path in ${\Bbb U}-\{0\}$ with one endpoint in $\partial {\Bbb U}$, Pommerenke \ref b.Pommerenke:Lowner/ has a generalization, which is valid for some paths $\alpha$ which are not simple paths. It is this genaralization (or rather, its version in ${\Bbb H}$ with the hydrodynamic normalization) which will substitute L\"owner's Slit Mapping Theorem for the treatment of the percolation boundary or Peano curve scaling limits. The emerging picture is that different values of $\kappa$ in the differential equations \ref e.loew/ or \ref e.loewh/ produce paths which are scaling limits of naturally defined processes, and that these paths can be space-filling, or simple paths, or neither, depending on the parameter $\kappa$. \subsection{Acknowledgement} I wish to express gratitude to Itai Benjamini, Rick Kenyon and David Wilson for inspiring discussions and helpful information. \ref l.tightdiam/ has been obtained jointly with Itai Benjamini. Mladen Bestvina, Brian Bodwitch, Steve Evans, Yakar Kannai, Greg Lawler, Russ Lyons, Yuval Peres, Steffen Rohde, Jeff Steif, Benjamin Weiss and Wendelin Werner have provided very helpful advice. \bsection {Some background and terminology}{s.back} \global\advance\consSer by 1\constnum=0\relax This section will introduce some notations which will be used, and discuss some of the necessary background. We begin with a review of uniform spanning trees and forests. The reader may consult \ref b.BLPS:usf/ for a comprehensive treatment of that subject. \subsection{The domination principle} Suppose that $H$ and $H'$ are two random subsets of some set. We say that $H'$ {\bf stochastically dominates} $H$ if there is a probability measure $\mu$ on pairs $(A,B)$ such that $A$ has the same law as $H$, $B$ has the same law as $H'$, and $\mu\bigl\{(A,B):\, B\supset A\bigr\}=1$. Such a $\mu$ is called a {\bf monotone coupling} of $H$ and $H'$. Let $G$ be a finite connected graph, and let $G_0$ be a connected nonempty subgraph. Let $M$ be the set of vertices of $G_0$ that are incident with some edge in ${\ss E}(G)-{\ss E}(G_0)$. (We let ${\ss E}(G)$ and ${\ss V}(G)$ denote the edges and vertices of $G$, respectively.) Let $G_0^W$ be the graph obtained from $G$ by identifying all the vertices in $M$ to a single vertex, called the {\bf wired vertex}. Then $G_0^W$ is called the {\bf wired graph} associated to the pair $(G_0,G)$. Let ${\ss T}$ be the UST on $G$, let ${\ss T}_0^F$ be the UST on $G_0$, and let ${\ss T}_0^W$ be the UST on $G_0^W$. Then ${\ss T}_0^F$ is called the {\bf free spanning tree} of the pair $(G_0,G)$, and ${\ss T}_0^W$ is the {\bf wired spanning tree} of the pair $(G_0,G)$. Sometimes, we call ${\ss T}_0^F$ [respectively, $T_0^W$] the UST on $G_0$ with free [respectively, wired] boundary conditions. The {\bf domination principle} states that ${\ss T}_0^F\cap G_0$ stochastically dominates ${\ss T}\cap G_0$, and that ${\ss T}\cap G_0$, stochastically dominates ${\ss T}_0^W\cap G_0$. Now let $G$ be an infinite connected graph, and let $G_1\subset G_2\subset \cdots$ be an infinite sequence of finite connected subgraphs satisfying $\bigcup_j G_j=G$. Let $\mu_j^W$ be the law of the wired spanning tree of $(G_j,G)$. Based on the domination principle, it is easy to verify that the weak limit $\mu^W$ of $\mu_j^W$ exists, and is a probability measure on spanning forests of $G$. It is called the {\bf wired spanning forest of $G$} (WSF). The domination principle, when appropriatly interpreted, carries over to infinite and to disconnected graphs as well. If $G$ is a disconnected graph, we take the FSF [respectively, WSF] on $G$ to be the spanning forest of $G$ whose intersection with every component of $G$ is the FSF [respectively, WSF] of that component, and with the restriction to the different components being independent. The more general formulation of the domination principle states that when $G_0$ is a subgraph of $G$, then ${\ss T}_0^F\cap G_0$ stochastically dominates ${\ss T}\cap G_0$ and ${\ss T}\cap G_0$ stochastically dominates ${\ss T}_0^W\cap G_0$, where ${\ss T},{\ss T}_0^F$ and ${\ss T}_0^W$ are the FSF on $G,G_0$ and $G_0^W$, respectively, and the same statement holds when FSF is replaced by WSF. On all recurrent connected graphs, the WSF is equal to the FSF, and both are trees. Therefore, on recurrent graphs we shall refer to this measure as the uniform spanning tree (UST). \subsection{Grid approximations of domains} Let $D\subset{\Bbb C}$ be a domain, that is, an open, connected set. Given $\delta>0$, we define a graph $G=G(D,\delta)$, which is a discrete approximation of the domain $D$ in the grid $\delta\Z^2$, as follows. The {\bf interior vertices} ${\verts_I}(G)={\verts_I}(D,\delta)$ of $G$ are the vertices of $\delta\Z^2$ which are in $D$. The {\bf boundary vertices} ${\verts_{\partial}}(G)={\verts_{\partial}}(D,\delta)$ are the points of intersection of the edges of the grid $\delta\Z^2$ with $\partial D$, the boundary of $D$. The vertices of $G$ are ${\ss V}(G)={\verts_{\partial}}(G)\cup{\verts_I}(G)$. If $a,b\in{\ss V}(G)$ are distinct, then $[a,b]$ is an edge of $G$ iff there is an edge $e\in{\ss E}(\delta\Z^2)$ such that the open segment $\bigl\{ta+(1-t)b:\, t\in(0,1)\bigr\}$ is contained in $D\cap e$. We will often be considering random walks on $G(D,\delta)$ starting at $0$, when $0\in D$. It will be useful to denote by ${\verts_{\partial}^0}$ the set of vertices $v\in{\verts_{\partial}}(D,\delta)$ such that there is a path from $0$ in $G(D,\delta)$ whose intersection with ${\verts_{\partial}}$ is $v$. The wired graph, $G^W(D,\delta)$, associated with $D$ is $G(D,\delta)$ with all the vertices ${\verts_{\partial}}(D,\delta)$ collapsed to a single vertex, which we simply denote $\partial D$. We shall often not distinguish between a graph and its planar embedding, if it has an obvious planar embedding. For example, the UST on $G$ will also be interpreted as a random set in the plane. \subsection{Wilson's algorithm} Let $G$ be a finite graph. Wilson's algorithm \ref b.Wilson:alg/ for generating a UST in $G$ proceeds as follows. Let $v_0\in{\ss V}(G)$ be an arbitrary vertex (which we call the root), and set $T_0:=\{v_0\}$. Inductively, assume that a tree $T_j\subset G$ has been constructed. If ${\ss V}(T_j)\neq{\ss V}(G)$, choose a vertex $v_{j+1}\in{\ss V}(G)-{\ss V}(T_j)$, let $W_{j+1}$ be LERW from $v$ to $T_j$ in $G$, and set $T_{j+1}:=T_j\cup W_{j+1}$. Otherwise ${\ss V}(T_j)={\ss V}(G)$, and the algorithm stops and outputs $T_j$. It is somewhat surprising, but true, that no matter how the choices of the vertices $v_j$ are made, the output of the algorithm is a tree chosen according to the uniform measure. If $G$ is infinite, connected and recurrent, Wilson's algorithm also ``works''. When the subtree $T_j$ generated by the algorithm includes all the vertices in a certain finite set $K\subset{\ss V}(G)$, the subtree of $T_j$ spanned by $K$ (that is, the minimal connected subgraph of $T_j$ that contains $K$) has the same law as the subtree of the UST of $G$ spanned by $K$. (There is also a version of Wilson's algorithm which is useful for generating the WSF of a transient graph, but we shall not need this.) \subsection{Harmonic measure estimates} Because of Wilson's algorithm, many questions about the UST can be reduced to questions about simple random walks (SRW's). It is therefore hardly surprising that we often need to obtain a {\bf harmonic measure estimate}; that is, an estimate on the probability that SRW starting from a vertex $v$ will hit a certain set of vertices $K_1$ before hitting another set $K_0$. As a function of $v$, this probability is harmonic\Ftnote{A function $h$ is {\bf harmonic} at $v$, if $h(v)$ is the average of the value of $h$ on the neighbors of $v$.} away from $K_0\cup K_1$. Almost all the harmonic measure estimates which we will use are entirely elementary, and follow from the following easy fact. Consider an annulus $A=A(p,r,R)$, with center $p$, inner radius $r$ and outer radius $R>r$. Suppose that $\delta$ is sufficiently small so that there is a path in $\delta\Z^2\cap A$ which separates the boundary components of $A$. Let $q\in\delta\Z^2\cap A$ be some vertex such that the distance from $q$ to the boundary $\partial A$ is at least $r/c$, where $c>0$ is some constant. Let $X$ be the image of SRW starting from $q$, which is stopped when it first leaves $A$. Then the probability that $X$ contains a path separating the boundary components of $A$ is bounded below by some positive function of $c$. (This can be proved directly using only the Markov property and the invariance of SRW under the automorphisms of $\delta\Z^2$.) One consequence of this fact and the Markov property, which we will often use, is as follows. \procl l.harmm Suppose that $K$ is a connected subgraph of $\delta\Z^2$ of diameter at least $R$, and $v\in\delta\Z^2$. Then the probability that SRW starting from $v$ will exit the ball $B(v,R)$ before hitting $K$ is at most $\nco a\bigl({\rm dist}(v,K)/R\bigr)^{\nco b}$, where $\co a,\co b>0$ are absolute constants. \endprocl This lemma holds for the spherical as well as Euclidean metric. \subsection{Laplacian random walk} Although this will not be needed in the paper, we have to mention another interpretation of LERW. Let $G$ be a finite connected graph, let $K\subset{\ss V}(G)$ be a set of vertices, and let $a\in{\ss V}(G)-K$. The LERW from $a$ to $K$ can also be inductively constructed, as follows. Suppose that the first $n$ vertices $a={\ss LE}(1),{\ss LE}(2),\dots,{\ss LE}(n)$ have been determined and ${\ss LE}(n)\notin K$. Let $h_n:{\ss V}(G)\to[0,1]$ be the function which is $1$ on $K$, $0$ on $\{{\ss LE}(1),\dots,{\ss LE}(n)\}$ and harmonic on ${\ss V}(G)-\bigl(K\cup \{{\ss LE}(1),\dots,{\ss LE}(n)\}\bigr)$. Then ${\ss LE}(n+1)$ is chosen among the neighbors $w$ of ${\ss LE}(n)$, with probability proportional to $h_n(w)$. This formulation of the LERW may serve as a heuristic for \ref g.confinv/, since discrete harmonic functions are good approximation for continuous harmonic functions, and continuous harmonic functions in 2D have conformal invariance properties. However, this heuristic is quite weak, since near a non-smooth boundary of a domain, the approximation is not good. \subsection{Weak convergence of measures} We now recall several facts and definitions regarding weak convergence. The reader may consult \crfl{EK:markov}{Chap.~3} for proofs and further references. Let $(X,d)$ be a compact metric space, and let $\{\mu_j\}$ be a sequence of Borel probability measures on $X$. $\mu$ be a Borel probability measure on $X$. Saying that the sequence $\mu_j$ converges weakly to $\mu$ means that $\lim_j \int f\, d\mu_j=\int f \,d\mu$ for all continuous $f:X\to\R$. The Prohorov metric on the space of Borel probability measures on $X$ is defined by $$ \dlr(\mu,\mu'):= \inf\bigl\{\epsilon>0:\, \mu(K)\leqslant \mu'(N_\epsilon (K)) + \epsilon \hbox{ for all closed }K\subset X\bigr\} \,, $$ where $N_\epsilon(K):=\bigcup_{x\in K} B(x,\epsilon)$ is the $\epsilon$-neighborhood of $K$. The space of Borel probability measures on $X$ is compact with respect to the Prohorov metric, and weak convergence is equivalent to convergence in the Prohorov metric. Let $\Cal M(\mu,\mu')$ be the collection of all Borel measures $\nu$ on $X\times X$ such that $\nu(A\times X)=\mu(A)$ and $\nu(X\times A)=\mu'(A)$ for all measurable $A\subset X$. Such a $\nu$ is called a {\bf coupling} of $\mu$ and $\mu'$. The Prohorov metric satisfies $$ \dlr(\mu,\mu') = \inf_{\nu\in\Cal M(\mu,\mu')} \inf \Bigl\{\epsilon>0:\, \nu\{(x,y):\, d(x,y)\geqslant\epsilon\}\leqslant\epsilon\Bigr\} \,. \label e.pcoup $$ In other words, $\dlr(\mu,\mu')<\epsilon$ means that one can find a probability space $(\Omega,\P)$ and two $X$ valued random variables $x,y:\Omega\to X$, such that $\P[d(x,y)\geqslant\epsilon]\leqslant\epsilon$ and such that $x$ has law $\mu$ and $y$ has law $\mu'$. This is obtained by taking an appropriate $\P=\nu\in\Cal M(\mu,\mu')$, $\Omega:=X\times X$, and letting $x$ and $y$ be the projections on the first and second factors, respectively. \subsection{Conformal maps} We review some elementary facts about conformal (aka univalent) mappings, as may be found in \ref b.Duren:book/, for example. Let $D\subset{\Bbb C}$ be some domain. A continuous map $f:D\to {\Bbb C}$, which is injective and complex-differentiable is {\bf conformal}. If $f$ is conformal, then $f^{-1}:f(D)\to{\Bbb C}$ is also conformal. Let $D\subsetneqq{\Bbb C}$ be simply connected. Then Riemann's Mapping Theorem states that there is a conformal homeomorphism $f=f_D$ from ${\Bbb U}$ onto $D$. Suppose also that $0\in D$, then $f$ can be chosen to satisfy the normalizations $f(0)=0$ and $f'(0)>0$, which render $f$ unique. In this case, the number $f'(0)$ is called the {\bf conformal radius of $D$} (with respect to $0$). The Schwarz Lemma implies that $$ f'(0)\geqslant \inf\bigl\{|z|:\, z\notin f({\Bbb U})\bigr\} \,, \label e.schw $$ while on the other hand, the Koebe $1/4$ Theorem gives $$ f'(0)\leqslant 4 \inf\bigl\{|z|:\, z\notin f({\Bbb U})\bigr\} \,. \label e.k1q $$ Hence, up to a factor of $4$, the conformal radius can be determined from the in-radius. If $0<a<b<\infty$, then the set of all conformal maps $f:{\Bbb U}\to{\Bbb C}$ satisfying $f(0)=0$ and $a\leqslant |f'(0)|\leqslant b$ is compact, in the topology of uniform convergence on compact subsets of ${\Bbb U}$. If $f_j:{\Bbb U}\to{\Bbb C}$ are conformal, $f_j(0)=0$, and $f_j\to f$ locally uniformly, then the image $f({\Bbb U})$ can be described in terms of the images $D_j:=f_j({\Bbb U})$. Let $D$ be the maximal open connected set containing $0$ and contained in $\bigcup_{n=1}^\infty\bigcap_{j=n}^\infty D_j$. If $D\neq\emptyset$, then $f({\Bbb U})=D$; otherwise, $f({\Bbb U})=\{0\}$. This is called Carath\'eodory's Kernel Convergence Theorem. If $f:{\Bbb U}\to D$ is a conformal homeomorphism onto $D$, and $\partial D$ is a simple closed curve, then $f$ extends continuously to $\partial {\Bbb U}$. The same is true if $D={\Bbb U}-\beta$, where $\beta$ is a simple path. \bsection {No Loops}{s.noloops} \global\advance\consSer by 1\constnum=0\relax In this section, we prove that any LERW subsequential scaling limit is supported on the set of simple paths. That is, we prove \ref t.noloop/. Let ${o_1}$ and ${o_2}$ be distinct points $\R^2$. For each $\delta>0$, let ${o_1^*}$ and ${o_2^*}$ be vertices of $\delta\Z^2$ closest to ${o_1}$ and ${o_2}$, respectively. Note that in $\delta\Z^2$ the combinatorial distance between two vertices $v,v'\in\delta\Z^2$ is $\delta^{-1}\|v-v'\|_1$. However, all metric notions we use will refer to the Euclidean or spherical distance. In this section, we will mainly use the Euclidean metric. Let ${\ss RW}$ be a random walk on $\delta\Z^2$ starting at ${o_2^*}$ and stopped when ${o_1^*}$ is reached for the first time. Let $\omega=\omega_\delta$ denote the loop-erasure of ${\ss RW}$, and let $\P_\delta$ denote the law of $\omega_\delta$. The following lemma will show that the diameter of $\omega_\delta$ is ``tight''. \procl l.tightdiam $$ \P_\delta\bigl[{\operatorname{diam} \omega> s\, {\rm dist}({o_1^*},{o_2^*})}\bigr] \leqslant \nco c s^{-\nco d} \,, $$ where $\co c,\co d>0$ are absolute constants. \endprocl The proof is based on Wilson's algorithm and an elementary harmonic measure estimate. A more precise estimate can be obtained by using the discrete Beurling Projection Theorem (see \ref b.Kesten:hitting/ or \ref b.Lawler:makarov/). \proof Set $r:={\rm dist}({o_1^*},{o_2^*})$, $R:=s\, {\rm dist}({o_1^*},{o_2^*})/4$, and let $z\in \delta\Z^2$ be some vertex such that ${\rm dist}({o_1^*},z)\geq10R$. For $a,b\in\delta\Z^2$, let $\Paa(a,b)$ denote the path in the UST of $\delta\Z^2$ joining $a$ and $b$. Let $w$ be the meeting point of ${o_1^*},{o_2^*},z$ in the UST; that is, the vertex in $\Paa({o_1^*},{o_2^*})\cap\Paa({o_1^*},z)\cap\Paa({o_2^*},z)$. By Wilson's algorithm, the distribution of $\omega$ is identical with the distribution of $\Paa({o_1^*},w)\cup\Paa({o_2^*},w)$. We now estimate $\PBig{\operatorname{diam}\bigl(\Paa(a,w)\bigr)> s\, {\rm dist}({o_1^*},{o_2^*})}$. Using Wilson's algorithm, we may generate $\Paa({o_1^*},w)$ by letting $\Paa({o_2^*},z)$ be LERW from ${o_2^*}$ to $z$, and letting $\Paa({o_1^*},w)$ be LERW from ${o_1^*}$ to $\Paa({o_2^*},z)$. Condition on $\Paa({o_2^*},z)$. By \ref l.harmm/, the probability that SRW starting at ${o_1^*}$ will exist $B({o_1^*},R)$ before hitting $\Paa({o_2^*},z)$ is at most $\nco a (r/R)^{\nco b}$, for some constants $\co a,\co b>0$. Therefore, $\P[\operatorname{diam} \Paa({o_1^*},w)>2 R]\leqslant \co a (r/R)^{\co b}$. Moreover, the same estimate holds for $\operatorname{diam} \Paa({o_2^*},w)$. Since $\Paa({o_1^*},{o_2^*})=\Paa({o_1^*},w)\cup\Paa({o_2^*},w)$, we get $\P[\operatorname{diam} \Paa({o_1^*},{o_2^*})>4 R]\leqslant 2\co a (r/R)^{\co b}$. This completes the proof of the lemma. \Qed \procl r.closetopt The proof of the lemma can be easily adapted to show that if $a,b\in\delta\Z^2$ and $K\subset\delta\Z^2$, then the probability that LERW from $a$ to $K$ will intersect $B(b,r)$ is at most $\nco c \bigl(r/{\rm dist}(a,b)\bigr)^{\nco d}$, provided $r\geqslant \delta$, where $\co c,\co d>0$ are absolute constants. \endprocl \procl d.qlp Let $z_0\in\R^2$, $r,\epsilon>0$. An $(z_0,r,\epsilon)$-quasi-loop in a path $\omega$ is a pair $a,b\in\omega$ with $a,b\in B(z_0,r)$, ${\rm dist}(a,b)\leqslant\epsilon$, such that the subarc of $\omega$ with endpoints $a,b$ is not contained in $B(z_0,2r)$. Let $\ev A(z_0,r,\epsilon)$ denote the set of simple paths in $\R^2$ that have a $(z_0,r,\epsilon)$-quasi-loop. \endprocl \procl l.noloop Let $c$ be the distance from ${o_1^*}$ to ${o_2^*}$, let $r\in(0,c/4)$, $\epsilon>0$ and $z_0\in\R^2$. Then $\lim_{\epsilon\to 0} \P_\delta[{\ev A(z_0,r,\epsilon)}] = 0$, uniformly in $\delta$. \endprocl \proof Let $B_1=B(z_0,r)$ and $B_2=B(z_0,2r)$. The distance from $B_2$ to at least one of the points ${o_1^*},{o_2^*}$ is at least $c/2$. By symmetry, we assume with no loss of generality that ${\rm dist}\bigl({o_2^*},B_2\bigr)\geqslant c/2$. Let $\epsilon_1\in(0,c/4)$, and let $q$ be a vertex in $\delta\Z^2$ such that ${\rm dist}\bigl(q,{o_2^*}\bigr)\in[\epsilon_1-\delta,\epsilon_1]$. Let $\Pa^q$ be a LERW from $q$ to ${o_2^*}$ in $\delta\Z^2$. Let ${\ss RW}$ be an independent simple random walk from ${o_1^*}$. Let ${\ss RW}'$ be the part of the walk ${\ss RW}$ until $\Pa^q$ is first hit. Then, by Wilson's algorithm, $\omega_\delta$ has the same distribution as the arc connecting ${o_1^*}$ to ${o_2^*}$ in ${\ss LE}({\ss RW}')\cup\Pa^q$. Let $\ev A'$ be the event that ${\ss LE}({\ss RW}')$ has a $(z_0,r,\epsilon)$-quasi-loop, and let $\ev C$ be the event that $\Pa^q$ intersects $B_2$. Then $$ \P_\delta\bigl[{\ev A(z_0,r,\epsilon)}\bigr] \leqslant \P[\ev A'] + \P[\ev C] \,. \label e.cas $$ Since ${\rm dist}\bigl(q,{o_2^*}\bigr)\leqslant \epsilon_1$ and ${\rm dist}\bigl(B_2,{o_2^*}\bigr)\geqslant c/2$, from \ref l.tightdiam/ we get an estimate of the form $$ \P[\ev C] \leqslant \nco {aa} (\epsilon_1/c)^{\nco {ab}} \,. \label e.pc $$ We now find an upper bound for $\P[\ev A']$. Let $s_1$ be the first time $s\geqslant 0$ such that ${\ss RW}(s)\in B_1$. Let $t_1$ be the first time $t\geqslant s_1$ such that ${\ss RW}(t_1)\notin B_2$. Inductively, define $s_j$ to be the first time $s\geqslant t_{j-1}$ such that ${\ss RW}(s)\in B_1$ and $t_j$ to be the first time $t\geqslant s_j$ such that ${\ss RW}(t)\notin B_2$. Let $\tau$ be the first time $t\geqslant 0$ such that ${\ss RW}(t)\in\Pa^q$. Finally, for each $s\geqslant 0$ let ${\ss RW}^s$ be the restriction of ${\ss RW}$ to the interval $t\in[0,s]$. For each $j=1,2,\dots$, we consider several events depending on ${\ss RW}^{t_j}$ and $\Pa^q$. Let $\ev Y_j$ be the event that ${\ss LE}({\ss RW}^{t_j})$ has a $(z_0,r,\epsilon)$-quasi-loop. Let $\ev T_j$ be the event that $\tau> t_j$. It is easy to see that $$ \ev A' \subset \bigcup_{j=1}^\infty (\ev Y_j \cap \ev T_j) \,. $$ Since $\ev T_j\supset\ev T_{j+1}$ for each $j$, this implies $$ \ev A' \subset \ev T_{m+1} \cup \bigcup_{j=1}^m \ev Y_j \,. \label e.decomp $$ for every $m$. We first estimate $\P[\ev T_{j+1} \mid {\ss RW}^{t_j},\ \Pa^q]$. Conditioned on any $\Pa^q$, the probability that a SRW starting at any vertex outside of $B_2$ will hit $\Pa^q$ before hitting $B_1$ is at least $$ \nco 1 (\epsilon_1)^{\nco 2} \,, $ where $\co 1>0$ depends only on $c$ and $r$, and $\co 2>0$ is an absolute constant. This is based on the fact that $\Pa^q$ is connected, contains ${o_2^*}$, and has diameter at least $\epsilon_1-\delta$. Applying this to the walk ${\ss RW}$ from time $t_j$ on, we therefore get $$ \P[\ev T_{j+1} \mid {\ss RW}^{t_j},\ \Pa^q] \leqslant 1 - \co 1 (\epsilon_1)^{\co 2} \,. $$ By induction, we therefore find that $$ \P[\ev T_m\mid\Pa^q] \leqslant \Bigl( 1 - \co 1 (\epsilon_1)^{\co 2}\Bigr)^{m-1} \,. \label e.hitp $ We now estimate $\P[\ev Y_{j+1}\mid \neg\ev Y_j, {\ss RW}^{t_j}]$. Let $Q_j$ be the set of components of ${\ss LE}({\ss RW}^{s_{j+1}})\cap B_2$ that do not contain ${\ss RW}(s_{j+1})$. Observe that for $\ev Y_{j+1}$ to occur, there must be a $K\in Q_j$ such that the random walk ${\ss RW}$ comes at some time $t\in[s_{j+1},t_{j+1}]$ within distance $\epsilon$ of $K\cap B_1$ but ${\ss RW}(t)\notin K$ for all $t\in[s_{j+1},t_{j+1}]$. But if ${\ss RW}(t)$ is close to $K$, $t\in[s_{j+1},t_{j+1}]$, then \ref l.harmm/ can be applied, to estimate the probability that ${\ss RW}$ will not hit $K$ before time $t_{j+1}$. That is, conditioned on ${\ss RW}^{s_{j+1}}$, for each given $K\in Q_j$, the probability that ${\ss RW}([s_{j+1},t_{j+1}])$ gets to within distance $\epsilon$ of $K$ but does not hit $K$ is at most $\nco 3 (\epsilon/r)^{\nco 4}$, where $\co 3,\co 4>0$ are absolute constants. Consequently, we get, $$ \P[\ev Y_{j+1}\mid \neg\ev Y_j, {\ss RW}^{t_j}] \leqslant \co 3|Q_j|(\epsilon/r)^{\co 4} \,. $$ Observe that $|Q_j|$, the cardinality of $Q_j$, is at most $j$. Therefore, $$ \P[\ev Y_{j+1}\mid \neg\ev Y_j] \leqslant \co 3 j(\epsilon/r)^{\co 4} \,. $$ This gives \begineqalno \P\Bigl[\bigcup_{j=1}^m\ev Y_j\Bigr] \leqslant \sum_{j=1}^{m-1} \P[\ev Y_{j+1} \cap\neg\ev Y_{j}] & \leqslant \sum_{j=1}^{m-1} \P[\ev Y_{j+1} \mid \neg\ev Y_{j}] \cr & \leqslant \sum_{j=1}^{m-1} j \co 3 (\epsilon/r)^{\co 4} \leqslant \nco 5 m^2 (\epsilon/r)^{\co 4} \label e.yest \endeqalno Combining this with \ref e.cas/, \ref e.pc/, \ref e.decomp/ and \ref e.hitp/, we find that $$ \P_\delta\bigl[{\ev A(z_0,r,\epsilon)}\bigr] \leqslant \co {aa} (\epsilon_1/c)^{\co {ab}} + \co 5 m^2 (\epsilon/r)^{\co 4} + \Bigl( 1 - \co 1 \epsilon_1^{\co 2}\Bigr)^{m-1} \,. $$ The lemma follows by taking $m := \lfloor\epsilon^{-\co 4/3}\rfloor$ and $\epsilon_1 :=-1/\log\epsilon$, say. \Qed \procl t.ql Let $(X,d)$ be a compact metric space, let $\zerp,\onep\in X$, let $f:(0,\infty) \to (0,\infty)$ be monotone increasing and continuous, and let $\Gamma=\Gamma(f)$ be the set of all compact simple paths $\gamma\subset X$ with endpoints $\zerp$ and $\onep$ which satify the following property. Whenever $x,y$ are points in $\gamma$ and $D(x,y)$ is the diameter of the arc of $\gamma$ joining $x$ and $y$ we have $$ d(x,y) \geqslant f(D(x,y))\,. \label e.qlc $$ Then $\Gamma$ is compact in the Hausdorff metric. \endprocl For this we will need the following Janiszewski's~\ref b.Janiszewski/ topological characterization of $[0,1]$ (see \crfl{Newman:planesets}{IV.5}): \procl l.topchar \procname{Topological Characterization of Arcs} Let $K$ be a compact, connected metric space, and let $\zerp,\onep\in K$. Suppose that for every $x\in K-\{\zerp,\onep\}$ the set $K-\{x\}$ is disconnected. Then $K$ is homeomorphic to $[0,1]$. \Qed \endprocl \proofof t.ql Let ${\Cal H}={\Cal H}(X)$ denote the space of compact nonempty subsets of $X$ with the Hausdorff metric $\dHa$. Let $\gamma$ be in the closure of $\Gamma$ in ${\Cal H}$. Then $\gamma$ is connected, compact, and $\gamma\supset\{\zerp,\onep\}$. We now use \ref l.topchar/ to show that $\gamma$ is a simple path. Indeed, suppose that $x\in\gamma-\{\zerp,\onep\}$. \comment{ We first show that $\gamma-\{x\}$ has at most two components. Let $y\in\gamma$. Let $B$ be an open ball about $x$ with radius ${1\over 2}f\bigl(d(x,y)/2\bigr)$. Let $\{\gamma_n\}$ be a sequence in $\Gamma$ such that $d_{{\Cal H}}(\gamma_n,\gamma)<1/n$. Then there are sequences $x_n,y_n\in\gamma_n$ with $x_n\to x$ and $y_n\to y$ as $n\to\infty$. Let $N_\zerp$ be the set of $n$'s such that $y_n$ is on the arc of $\gamma_n$ joining $\zerp$ to $x_n$, and let $N_\onep$ be the set of $n$'s such that $y_n$ is on the arc of $\gamma_n$ joining $x_n$ to $\onep$. Then $N_\zerp\cup N_\onep$ contains the positive integers, so at least one of these sets is infinite. Suppose, for example, that $N_\zerp$ is infinite. Let $\gamma_n'$ be the closed arc of $\gamma_n$ joining $y_n$ and $\zerp$. Suppose that $z_n\in\gamma_n\cap B$ and $x_n\in B$. Then the arc in $\gamma_n$ joining $z$ and $x_n$ has diameter at most $d(x,y)/2$, since $\gamma_n\in \Gamma(f)$. Consequently, for large $n$ it cannot contain $y_n$. Therefore, for large $n$ the arc $\gamma_n'$ does not intersect $B$. Let $\gamma'$ be a Hausdorff limit of a subsequence of $\gamma_n'$. Then $\gamma'$ is connected, disjoint from $B$, and contains $\zerp$ and $y$. So $y$ and $\zerp$ are in the same connected component of $\gamma-\{x\}$. Similarly, if $N_\onep$ is infinite, then $y$ and $\onep$ are in the same connected component of $\gamma-\{x\}$. We conclude that each connected component of $\gamma-\{x\}$ intersects $\{\zerp,\onep\}$, and in particular there are at most two components. } We show that $\zerp$ and $\onep$ are in distinct components of $\gamma-\{x\}$. Let $\{\gamma_n\}$ be a sequence in $\Gamma$ such that $\dHa(\gamma_n,\gamma)<1/n$, and let $\{x_n\}$ be a sequence with $x_n\in\gamma_n$ and $d(x_n,x)<1/n$. For each $n$ let $\gamma^\zerp_n$ be the closed arc of $\gamma_n$ with endpoints $\zerp$ and $x_n$, and let $\gamma^\onep_n$ be the closed arc of $\gamma_n$ with endpoints $x_n$ and $\onep$. By passing to a subsequence, if necessary, assume with no loss of generality that the Hausdorff limits $\gamma^\zerp=\lim \gamma^\zerp_n$ and $\gamma^\onep=\lim\gamma^\onep_n$ exist. If $p_n\in\gamma^\zerp_n$, $q_n\in\gamma^\onep_n$, then $d(p_n,q_n)\geqslant f\bigl(d(p_n,x_n)\bigr)$, since $\gamma_n\in\Gamma(f)$. By taking limits we find that if $p\in \gamma^\zerp$ and $q\in\gamma^\onep$, then $d(p,q)\geqslant f\bigl(d(p,x)\bigr)$. Consequently, $\gamma^\zerp-\{x\}$ and $\gamma^\onep-\{x\}$ are disjoint. Because $\gamma^\zerp\cup\gamma^\onep=\gamma$ and $\gamma^\zerp,\gamma^\onep$ are compact, the set $\gamma-\{x\}$ is not connected. Hence, by \ref l.topchar/, $\gamma$, is a simple path. It remains to prove that $\gamma\in\Gamma(f)$. For any simple path $\beta\subset X$ with endpoints $\zerp,\onep$, let $R(\beta)$ be the set of all $(x,y)\in\beta\times\beta$ such that $x$ belongs to the subarc of $\beta$ with endpoints $\zerp$ and $y$. With arguments as above, it is not hard to show that $\lim R(\gamma_n)=R(\gamma)$, in the Hausdorff metric on $X\times X$. Since $f$ is continuous, it then easily follows that $\gamma\in\Gamma(f)$. The details are left to the reader. (Actually, one can see that this statement is not essential for the proof of \ref t.noloop/. There, we only need the fact that the Hausdorff closure of $\Gamma(f)$ is contained in the set of simple paths.) \Qed Although this will not be needed here, we note the following variation on \ref t.ql/. \procl t.qlg Let $(X,d)$ be a compact metric space, let $f:(0,\infty) \to (0,\infty)$ be monotone increasing and continuous, and let $\Gamma'=\Gamma'(f)$ be the set of all compact subsets of $\gamma\subset X$ that are simple paths satifying the following property. Whenever $x,y$ are points in $\gamma$ and $D(x,y)$ is the diameter of the arc of $\gamma$ joining $x$ and $y$ we have $$ d(x,y) \geqslant f(D(x,y))\,. \label e.qlc $$ Set $\Gamma_0:=\bigl\{\{x\}:\, x\in X\bigr\}$. Then $\Gamma'\cup\Gamma_0$ is compact in the Hausdorff metric. \endprocl The proof does not require much more than the proof of \ref t.ql/. We omit the details. \proofof t.noloop We start with the proof of the second statement, and first assume that $a,b\neq\infty$. Let ${o_1^*}$ and ${o_2^*}$ be vertices of $\delta\Z^2$ closest to $a$ and $b$, respectively. Let $\omega_\delta$ be LERW from ${o_1^*}$ to ${o_2^*}$ in $\delta\Z^2$, and let $\omega'_\delta$ be a path from $a$ to $b$, obtained by taking the line segment joining $a$ to a closest point $a'$ on $\omega_\delta$, taking the line segment joining $b$ to a closest point $b'$ on $\omega_\delta$, and taking the path in $\omega_\delta$ joining $a'$ and $b'$. Then the Hausdorff distance from $\omega_\delta$ to $\omega'_\delta$ is less than $2\delta$. Let $\P_\delta'$ be the law of $\omega'_\delta$. Let $\mu$ be some subsequential weak limit of $\P_\delta$ as $\delta\to 0$. Then it is also a subsequential scaling limit of $\P_\delta'$. Let $m$ be large. By \ref l.tightdiam/, there is an $R_m$ such that with probability at least $1-2^{-m-1}$ we have $\omega_\delta\subset B(\aa,R_m)$. For each $j\in\N$, let $z^j_1,\dots,z^j_{k_j}$ be a finite set of points in $\R^2$ such that the open balls of radius $2^{-j-2}$ about these points cover $\overline B(\aa,R_m)$. For each $j\in\N$, let $\epsilon^m_j\in(0,1)$ be sufficiently small so that $$ \P_\delta\bigl[ \ev A(z_i^j,2^{-j},\epsilon^m_j)\bigr]< 2^{-m-2-j}/k_j \,, \label e.rare $$ for all $i=1,\dots, k_j$, and for all $\delta>0$. Such $\epsilon_j^m$ exist, by \ref l.noloop/. Finally, let $f_m:(0,\infty)\to(0,\infty)$ be a continuous monotone increasing function satisfying $f_m(2^{2-j})\leqslant 2^{-3}\min(\epsilon_j^m,2^{-j})$ for each $j=1,2,\dots$ and $\sup_{s>0} f_m(s)\leqslant 2^{-3}\min(\epsilon_1^m,1/2)$. Let $\ev X_m$ be the space of all compact nonempty subsets of $\overline B(\aa,R_m)$, and set $$ \ev Q_m := \ev X_m - \bigcup_{j=1}^\infty\bigcup_{i=1}^{k_j} \ev A(z_i^j,2^{-j-1},\epsilon^m_j) \,. $$ Note that $$ \P_\delta'(\neg \ev Q_m) \leqslant 2^{-m} \,, \label e.musup $$ for all $\delta$ and $m$. Also note that if we set $X:=\overline B(\aa,R_m)$ and $f:=f_m$ in \ref t.ql/, then $$ \ev Q_m\subset\Gamma_m:=\Gamma(f_m) \,. \label e.contain $ Indeed, suppose that $\gamma\in \ev X_m$ is a path in $X$ joining $a$ and $b$ and $\gamma\notin \Gamma_m$. Then there are $x,y\in\gamma$ and $w$ contained in the arc of $\gamma$ joining $x$ and $y$ such that ${\rm dist}(x,y)< 2f({\rm dist}(x,w))$. Since $\sup f\leqslant 2^{-4}$, we have ${\rm dist}(x,y)< 2^{-3}$. Let $j$ be such that $2^{-j+1}\leqslant {\rm dist}(x,w)\leqslant 2^{-j+2}$. Then $$ {\rm dist}(x,y)<2f({\rm dist}(x,w))\leqslant 2f(2^{-j+2})\leqslant 2^{-j-2},\qquad {\rm dist}(x,y)<2^{-3}\epsilon_j^m \,. $$ Let $i\in\{1,\dots,k_j\}$ be such that ${\rm dist}(x,z^j_i)\leqslant 2^{-j-2}$. Then ${\rm dist}(z^j_i,w)\geqslant 2^{-j}$ and $x,y\in B(z^j_i,2^{-j-1})$. Consequently, since $d(x,y) < \epsilon_j^m$, we have $\gamma\in \ev A(z_i^j,2^{-j-1},\epsilon_j^m)$. This proves \ref e.contain/. By \ref e.musup/ and \ref e.contain/, we get $\P_\delta'(\neg\Gamma_m)\leqslant 2^{-m}$. \ref t.ql/ tells us that $\Gamma_m$ is compact. Therefore, we also have $\mu(\neg\Gamma_m)\leqslant 2^{-m}$, and so $$ \mu\left(\neg\bigcup_m \Gamma_m\right)=0 \,. $ This completes the proof for the case $a,b\neq\infty$, because each element of $\bigcup_m\Gamma_m$ is a simple path. The proof when $a$ or $b$ in $\infty$ is similar. One only needs to note that \ref l.noloop/ is valid when $z_0=\infty$ and the distances are measured in the spherical metric. Indeed, the basic harmonic measure estimate \ref l.harmm/ is also valid in the context of the spherical metric. The proof of the first statement of \ref t.noloop/ is also similar. The details are left to the reader. \Qed Note that the first statement of \ref t.noloop/ implies that for almost every subsequential scaling limit path $\gamma$ from $a$ to $\partial D$, the closure of $\gamma\cap D$ intersects $\partial D$ in a single point. This fact is easy to deduce directly, since it is also true for the image of SRW starting near $a$ and stopped when $\partial D$ is hit. \bsection {First steps in the proof of \ref t.lesl/}{s.markov} \global\advance\consSer by 1\constnum=0\relax Throughout this section we assume \ref g.confinv/. Let $\sigma$ be random, with the law of the scaling limit of LERW from $0$ to $\partial {\Bbb U}$. {}From \ref t.noloop/ we know that $\sigma$ is a.s.\ a simple path. Recall that if $D\subset D'\subsetneqq{\Bbb C}$ are simply connected domains with $0\in D$, then the conformal radius of $D'$ is at least as large as the conformal radius of $D$. This follows from the Schwarz Lemma applied to the map $f_{D'}^{-1}\circ f_D:{\Bbb U}\to{\Bbb U}$. For each $t\in(-\infty,0]$, let $\sigma_t$ be the subarc of $\sigma$ with one endpoint in $\partial {\Bbb U}$ such that the conformal radius of ${\Bbb U}-\sigma_t$ is $\exp t$. It is clear that $\sigma_t$ varies continuously in $t$. Let $f_t:{\Bbb U}\to{\Bbb U}-\sigma_t$ be the conformal map satisfying $f_t(0)=0$ and $f_t'(0)=\exp t$. By L\"owner{}'s slit mapping theorem \ref b.Lowner/, there is a unique continuous $\zeta=\zeta_\sigma:(-\infty,0]\to\partial{\Bbb U}$ such that the differential equation \ref e.loew/ holds. Let $\widehat\zeta=\widehat\zeta_\sigma:(-\infty,0]\to\R$ be the continuous function satisfying $\zeta(t)=\exp\bigl(i\widehat\zeta(t)\bigr)$ and $\widehat\zeta(0)\in[0,2\pi)$. Our goal is to prove \procl p.redu The law of $\widehat\zeta$ is stationary, and $\widehat\zeta$ has independent increments. \endprocl This means that for each $s<0$ the law of the map $t\mapsto\widehat\zeta(s+t)$ restricted to $(-\infty,0]$ is the same as the law of $\widehat\zeta$, and that for every $n\in\N$ and $t_0\leqslant t_1\leqslant\cdots\leqslant t_n\leqslant 0$, the increments $\widehat\zeta(t_1)-\widehat\zeta(t_0),\widehat\zeta(t_2)-\widehat\zeta(t_1),\dots, \widehat\zeta(t_n)-\widehat\zeta(t_{n-1})$ are independent. The proof of this proposition, as well as the next, will be completed in later sections. Note that \ref g.confinv/ implies that the distribution of $\sigma$ is invariant under rotations of ${\Bbb U}$ about $0$. Let $\sigma^1$ be random with the law of $\sigma$ conditioned to hit $\partial U$ at $1$. If $\lambda$ denotes the (random) point in $\sigma\cap\partial {\Bbb U}$, then $\sigma^1$ has the same law as $\lambda^{-1}\sigma$. It turns out that \ref p.redu/ will follow quite easily from \procl p.prod Assume \ref g.confinv/. Fix some $t<0$. Take $\sigma^1$ and $\sigma$ to be independent. As above, let $\sigma_t$ be the compact arc of $\sigma$ that has one endpoint on $\partial {\Bbb U}$ and such that the conformal radius of ${\Bbb U}-\sigma_t$ is $\exp(t)$. Let $q(t)$ be the endpoint of $\sigma_t$ that is in ${\Bbb U}$. Let $\phi$ be the conformal map from ${\Bbb U}$ onto ${\Bbb U}-\sigma_t$ satisfying $\phi(0)=0$ and $\phi(1)=q(t)$. Then $\sigma_t\cup\phi(\sigma^1)$ has the same law as $\sigma$. \endprocl Now comes an easy lemma about LERW, and \ref p.prod/ will be obtained from this lemma by passing to the scaling limit. The passage to the scaling limit is quite delicate. Recall the definition of the graph $G(D,\delta)$ approximating a domain $D$, from \ref s.back/. \procl l.dprod Let $\delta>0$ and $t<0$ be fixed, and let $D\subsetneqq{\Bbb C}$ be a simply connected domain with $0\in D$. Let $\beta$ be LERW from $0$ to $\partial D$ in $G(D,\delta)$. Let $\beta_t$ be the compact arc in $\beta$ such that $\beta_t\cap\partial D$ is an endpoint of $\beta_t$ and such that the conformal radius of $D-\beta_t$ is $\exp(t)$. Let $q_1(t)$ be the endpoint of $\beta_t$ that is not on $\partial{\Bbb U}$. Set $D_t=D-\beta_t$. Then the law of $\beta-\beta_t$ conditioned on $\beta_t$ is equal to the law of LERW from $0$ to $\partial D_t$, conditioned to hit $q_1(t)$. \endprocl \proof There are several different ways to prove this lemma. We prove it using the relation between LERW and the UST. Suppose that $\alpha$ is a path such that $\beta_t=\alpha$ has positive probability. We assume for now that the endpoint $q$ of $\alpha$ which is in $D$ lies in the relative interior of an edge $e$ of $\delta\Z^2$ (this must be true except for at most a countable possible choices of $t$), and set $\tilde\alpha:=\alpha-e$. Let $\tilde q$ be the endpoint of $\tilde \alpha$ in $D$. Let ${\ss T}$ be the UST on $G^W(D,\delta)$, the wired graph of $D$. Then $\beta$ may be taken as the path in ${\ss T}$ from $0$ to $\partial D$. We may generate ${\ss T}$ using Wilson's algorithm with root $\partial D$, and starting with vertices $v_1=\tilde q$ and $v_2=0$. Conditioning on $\beta_t$ being equal to $\alpha$ is the same as conditioning on $\alpha\subset \beta$, which is the same as conditioning on the LERW ${\ss LE}_1$ from $v_1$ to $\partial D$ to be $\tilde\alpha$ and that the LERW from $v_2$ to ${\ss LE}_1$ hits $\tilde q$ through the edge $e$. This completes the proof in the case where $q$ is not a vertex of $\delta\Z^2$. The case where $q\in{\ss V}(\delta\Z^2)$ is treated similarly. \Qed \bsection {Getting uniform convergence}{s.gu} \global\advance\consSer by 1\constnum=0\relax The principle goal of this section is to state and prove \ref c.unifsl/ below. The main point there is that \ref g.confinv/ implies that the weak convergence of loop erased random walk in a domain $D\subset{\Bbb U}$ is uniform in $D$. \procl l.apxx Let $K$ be a compact connected set in $\overline {\Bbb U}$ that contains $\partial {\Bbb U}$ but with $0\notin K$. Let $F$ be a compact subset of ${\Bbb U}-K$, and let $\epsilon>0$. Then there is a $\delta_1>0$ with the following property. Let $K'$ be a compact connected set in $\overline {\Bbb U}$ with $K'\supset\partial{\Bbb U}$ and $d_{{\Cal H}({\Bbb U})}(K,K')<\delta_1/5$. Set $D={\Bbb U}-K'$. Let $\delta\in(0,\delta_1/5)$, and let $Q\subset {\verts_{\partial}^0}(D,\delta)$ be nonempty. Let ${\ss RW}^Q$ be the random walk on $G(D,\delta)$ starting at $0$ that stops when it hits ${\verts_{\partial}}(D,\delta)$, conditioned to hit $Q$. Then the probability that ${\ss RW}^Q$ will reach $F$ after visiting some vertex within distance $\delta_1$ of $K$ is less than $\epsilon$. \endprocl \proof We need to recall some basic facts relating the conditioned random walk ${\ss RW}^Q$ to the unconditioned random walk ${\ss RW}$ (that stops when hitting ${\verts_{\partial}}(D,\delta)$). First recall that ${\ss RW}^Q$ is a Markov chain. (This is easy to prove directly. See also the discussion of Doob's $h$-transform in \crfl{Durrett:BMM}{\SSS 3.1}.) Let $v_0$ be some vertex, $t_0\in\N$, and $W$ a set of vertices. Let $\tau=\tau_W$ be the least $t\geqslant t_0$ such that ${\ss RW}^Q(t)\in W$, if such exists, and otherwise set $\tau=\infty$. For $w\in W$ let $$ a_W(v_0,w):=\P[\tau<\infty, {\ss RW}^Q(\tau)=w\mid{\ss RW}^Q(t_0)=v_0] \,, $$ and let $\tilde a_W(v_0,w)$ be the corresponding quantity for ${\ss RW}$. Then $$ a_W(v_0,w) = {\tilde a_W(v_0,w) h(w)\over h(v_0)} \,, \label e.hest $$ where $h(v)$ is the probability that ${\ss RW}$ hits $Q$ when it starts at $v$. This formula is easy to verify. Let $W_1$ be the set of vertices $v$ of $G(D,\delta)$ such that $h(v)< \epsilon h(0)/5$. By \ref e.hest/, $$ \sum_{w\in W_1} a_{W_1}(0,w)\leqslant (\epsilon/5)\sum_{w\in W_1} \tilde a_{W_1}(0,w) \leqslant \epsilon/5 \,. $$ Consequently, the probability that ${\ss RW}^Q$ visits $W_1$ is at most $\epsilon/5$. Let $\rho$ be the distance from $F$ to $K$, and assume that $100\delta_1<\rho$. Then an easy discrete Harnack inequality shows that $h(v)/h(0)<\nco a$ for all vertices $v$ in $F$, where $\co a$ is some constant which does not depend on $K'$ or $\delta$, but may depend on $F$. There is a first vertex, say $\widehat v$, visited by ${\ss RW}^Q$ such that the distance from $\widehat v$ to $K$ is at most $\delta_1$. Let $W_2$ be the set of vertices of $G(D,\delta)$ in $F$. If $\widehat v\notin W_1$, then $h(\widehat v)\geqslant \epsilon h(0)/5$ and hence $$ \sum_{w\in W_2} a_{W_2}(\widehat v,w) =\sum_{w\in W_2}\tilde a_{W_2}(\widehat v,w) {h(w)\over h(0)}{h(0)\over h(\widehat v)} \leqslant 5\co a\epsilon^{-1} \sum_{w\in W_2} \tilde a_{W_2}(\widehat v, w) \,. \label e.eee $$ But since $K$ is connected, \ref l.harmm/ shows that the probability that a simple random walk starting at $\widehat v$ will get to distance $\rho/2$ from $\widehat v$ without hitting $V_\partial(D,\delta)$ is bounded by $\nco b (\delta_1/\rho)^{\nco c}$, where $\co b,\co c>0$ are absolute constants. By \ref e.eee/, $$ \sum_{w\in W_2} a_{W_2}(\widehat v,w) \leqslant 5 \co a \epsilon^{-1} \co b (\delta_1/\rho)^{\co c} \,. $$ Consequently, if $\delta_1$ is chosen sufficiently small, the probability that ${\ss RW}^Q$ starting at $\widehat v$ will hit $W_2$ is less than $\epsilon/5$, provided $\widehat v\notin W_1$. But $\P[\widehat v\in W_1]\leqslant \epsilon/5$, since the probability that $W_1$ is visited is at most $\epsilon/5$. The lemma follows. \Qed In the following, we let ${\ss RW}^{D,\delta}$ denote SRW on $G(D,\delta)$ that stops when it hits ${\verts_{\partial}}(D,\delta)$, and let ${\ss RW}^{D,\delta,Q}$ denote ${\ss RW}^{D,\delta}$ conditioned to hit $Q$, if $Q\subset{\verts_{\partial}^0}(D,\delta)$. Suppose that $\nu$ is a probability measure on ${\verts_{\partial}^0}(D,\delta)$ and $p$ is random with law $\nu$, then ${\ss RW}^{D,\delta,\nu}$ will denote ${\ss RW}^{D,\delta}$ conditioned to hit $p$ given $p$. In other words, the law of ${\ss RW}^{D,\delta,\nu}$ is the convex combination of the laws of the walks ${\ss RW}^{D,\delta,\{p\}}$, with coefficients $\nu(\{p\})$. \procl l.doub Assume \ref g.confinv/. Let $D\subset{\Bbb U}$ be a Jordan domain with $0\in D$. Let $\phi: D \to{\Bbb U}$ be the conformal homeomorphism from $D$ to ${\Bbb U}$ satisfying $\phi(0)=0$, $\phi'(0)>0$. Then as $\delta\to 0$ the law of the pair $\bigl(\phi({\ss LE}({\ss RW}^{D,\delta})),\partial{\Bbb U}\cap\phi({\ss RW}^{D,\delta})\bigr)$ tends weakly to the law of the pair $(\sigma,\partial{\Bbb U}\cap\sigma)$. \endprocl This lemma is easily proved using arguments as in the proof of \ref l.apxx/, and is therefore left to the reader. If $X$ and $Y$ are random closed subsets of ${\Bbb U}$, we let $\dlru(X,Y)$ denote the Prohorov distance between the law of $X\cup\partial{\Bbb U}$ and the law of $Y\cup\partial{\Bbb U}$ (see \ref s.back/, towards the end), where the metric $\dd_{\overline{\Bbb U}}$ is used on ${\Cal H}(\overline{\Bbb U})$. If $F$ is a subset of $\overline{\Bbb U}$, we set $\dlr_F(X,Y):=\dlru\left(X\cup\overline{{\Bbb U}-F},Y\cup\overline{{\Bbb U}-F}\right)$. This is a measure of how much $X$ and $Y$ differ inside $F$. \procl l.apxxx Let $\epsilon>0$. Then there is a $\delta_0>0$ and a finite collection of smooth Jordan domains $D_1,D_2,\dots,D_n\subset{\Bbb U}$ with $0\in D_j$ for all $j$, and with the following property. Let $D\subset {\Bbb U}$ be a simply connected domain with $B(0,\epsilon)\subset D$, let $\delta\in (0,\delta_0)$ and let $Q\subset {\verts_{\partial}^0}(D,\delta)$ be nonempty. Let $F_\epsilon$ be the component of $0$ in the set of points in $D$ that have distance at least $\epsilon$ to $\partial D$. Then there is a $D'\in \{D_1,\dots,D_n\}$ and a probability measure $\nu$ on ${\verts_{\partial}^0}(D',\delta)$ such that $$ \dlr_{F_\epsilon}({\ss LE}({\ss RW}^{D,\delta,Q}),{\ss LE}({\ss RW}^{D',\delta,\nu})) <\epsilon \,. \label e.apxxx $$ Moreover, we may require that the Hausdorff distance from $\partial D$ to $\partial D'$ is at most $\epsilon$. \endprocl \proof Fix some $D$ and $Q$ as above. By \ref l.apxx/, there is a $\delta_1>0$ such that with probability $\geqslant 1-\epsilon/5$ the walk ${\ss RW}^{D,\delta,Q}$ does not reach $F_\epsilon$ after exiting $F_{\delta_1}$, provided $\delta\in (0,\delta_1/5)$. Also, there is a $\delta_1'<\delta_1$ such that with probability $\geqslant 1-\epsilon/5$ this walk does not reach $F_{\delta_1}$ after exiting $F_{\delta_1'}$. Let $D'\subset D$ be a smooth Jordan domain with $D'\supset F_{\delta_1'}$, and such that the Hausdorff distance from $\partial D'$ to $\partial D$ is less than $\epsilon$. For every $\delta<\delta_1'/5$, let $\nu=\nu_\delta$ be the hitting measure of ${\ss RW}^{D,\delta,Q}$ on ${\verts_{\partial}^0}(D',\delta)$. Observe that we may think of ${\ss RW}^{D',\delta,\nu}$ as equal to ${\ss RW}^{D,\delta,Q}$ stopped when $V_\partial(D',\delta)$ is hit. Let $\ev A_1$ be the event that ${\ss RW}^{D,\delta,Q}$ does not visit $F_\epsilon$ after exiting $F_{\delta_1}$, and let $\ev A_2$ be the event that ${\ss RW}^{D,\delta,Q}$ does not visit $F_{\delta_1}$ after exiting $F_{\delta_1'}$. Note that on the event $\ev A_1\cap\ev A_2$, after exiting $F_{\delta_1'}$ the random walk does not visit any vertex $v$ which was already visited prior to the last visit to $F_\epsilon$. Consequently, the intersection of $F_\epsilon$ with the loop erasure of the walk does not change after the first exit of $F_{\delta_1'}$. Since we may couple ${\ss RW}^{D',\delta,\nu}$ to equal ${\ss RW}^{D,\delta,Q}$ stopped on $V_\partial(D',\delta)$, this means that we may obtain a coupling giving $F_\epsilon\cap {\ss LE}({\ss RW}^{D,\delta,Q})=F_\epsilon\cap{\ss LE}({\ss RW}^{D',\delta,\nu})$ on $\ev A_1\cap A_2$. Since $\P[\ev A_1\cap \ev A_2]\geqslant 1-\epsilon/2$, this proves the lemma for a single $D$. However, the same solution would stand for every $D''$ with $\partial D''$ sufficiently close to $\partial D$ in the Hausdorff metric. Hence, the compactness of the Hausdorff space of compact, connected subsets of $\overline{\Bbb U}-B(0,\epsilon)$ completes the proof. \Qed \procl l.unifhit Let $D\subset{\Bbb U}$ be a smooth Jordan domain with $0\in D$, let $\delta>0$, let $q\in {\verts_{\partial}^0}(D,\delta)$, and let ${\ss RW}^q:= {\ss RW}^{D,\delta,\{q\}}$. Then the law of ${\ss LE}({\ss RW}^q)$ is uniformly continuous in $q$. That is, for every $\epsilon>0$ there is a $\delta_1>0$ such that $$ \dlr_D({\ss LE}({\ss RW}^q),{\ss LE}({\ss RW}^{q'}))<\epsilon $$ provided $\delta\in(0,\delta_1)$, and $|q-q'|<\delta_1$. \endprocl \proof Let $\delta_0>0$ be very small. It is easy to see that when $|q-q'|$ is small we may couple ${\ss RW}^q$ and ${\ss RW}^{q'}$ so that with probability at least $1-\epsilon/5$ they are equal until they both come within distance $\delta_0$ of $\partial D$. Hence, the lemma follows by using an argument similar to the one used in the proof of \ref l.apxxx/. \Qed \procl c.unifsl Assume \ref g.confinv/, and let $\epsilon>0$. Then there is a $\delta_1>0$ with the following property. Let $D\subset{\Bbb U}$ be a simply connected domain with $B(0,\epsilon)\subset D$, and let $F_\epsilon$ be the connected component of $0$ in the set of all points $z$ with $d(z,\partial D)\geqslant\epsilon$. Let $\delta\in (0,\delta_1)$, and let $Q\subset {\verts_{\partial}^0}(D,\delta)$ be nonempty. Let $\phi$ be the conformal homeomorphism from ${\Bbb U}$ to $D$ that satisfies $\phi(0)=0$ and $\phi'(0)>0$. Then there is a random $\lambda\in\partial{\Bbb U}$ independent from $\sigma^1$ such that $$ \dlr_{F_\epsilon}\left(\phi(\lambda\sigma^1),{\ss LE}({\ss RW}^{D,\delta,Q})\right)<\epsilon \,. $$ \endprocl The main point here is that $\delta_1$ does not depend on $D$ or on $Q$. \proof Let $\epsilon_1>0$ be much smaller than $\epsilon$. Suppose that $D'$ is a domain in the list appearing in \ref l.apxxx/ that satisfies the requirements there with $\epsilon_1$ in place of $\epsilon$, and let $\nu$ be as in that lemma. Fix some small $\delta_1$. For each $q\in {\verts_{\partial}^0}(D',\delta)$ let $\nu_q$ be restriction of the hitting measure of ${\ss RW}^{D',\delta}$ to $B(q,\delta_1)\cap V_\partial(D,\delta)$, normalized to be a probability measure. \ref l.unifhit/ implies that we may replace $\nu$ by a probability measure $\nu'$, which is a convex combination of such $\nu_q$, while having $$ \dlr_{F_{\epsilon_1}}({\ss LE}({\ss RW}^{D',\delta,\nu}),{\ss LE}({\ss RW}^{D',\delta,\nu'}))<\epsilon_1 \,, \label e.cua $$ provided $\delta_1$ is sufficiently small. Moreover, by \ref g.confinv/, provided $\delta_1$ is sufficiently small and $\delta\in(0,\delta_1)$, we have $$ \dlr_{F_{\epsilon_1}}({\ss LE}({\ss RW}^{D',\delta,\nu_q}),\psi(\lambda_q\sigma^1))\leqslant \epsilon_1 \,, \label e.cub $$ where $\lambda_q\in\partial{\Bbb U}$ is random and independent from $\sigma^1$, and $\psi$ is the conformal homeomorphism from ${\Bbb U}$ to $D'$ satisfying $\psi(0)=0$ and $\psi'(0)>0$. Since the list $D_1,\dots,D_n$ in \ref l.apxxx/ is finite, we may take $\delta_1$ to be independent of $D$. Consequently, there is a random $\lambda\in\partial{\Bbb U}$ independent from $\sigma^1$ with $$ \dlr_{F_{\epsilon_1}}({\ss LE}({\ss RW}^{D',\delta,\nu'}),\psi(\lambda\sigma^1))\leqslant \epsilon_1 \,. \label e.cuc $$ Provided we have chosen $\epsilon_1$ sufficiently small, we have that $|\psi(\phi^{-1}(z))-z|<\epsilon_1$ for $z\in F_{\epsilon/2}$. \ref c.unifsl/ now follows from \ref e.apxxx/ with $\epsilon_1$ in place of $\epsilon$ and from \ref e.cua/, \ref e.cub/ and \ref e.cuc/. \Qed \bsection {Recognizing the L\"owner\ parameter as Brownian motion}{s.cont} \global\advance\consSer by 1\constnum=0\relax \proofof p.prod Let $\beta:={\ss LE}({\ss RW}^{{\Bbb U},\delta})$ and let $\beta_t$, $D_t$ and $q_1(t)$ be defined as in \ref l.dprod/. Set $\gamma:={\ss LE}( {\ss RW}^{D_t,\delta,q_1(t)})$, where ${\ss RW}^{D_t,\delta,q_1(t)}$ is taken to be independent from $\beta$ conditioned on $\beta_t$. Using \ref g.confinv/, $\dlru(\beta,\sigma)\to 0$ as $\delta\to 0$. By \ref e.pcoup/, this means that we may couple $\beta$ and $\sigma$ (that is, make them defined on the same probability space, where they are not necessarily independent) such that $\beta{\,\mathop{\to}\limits^{\P}\,} \sigma$ in ${\Cal H}({\Bbb U})$, where ${\,\mathop{\to}\limits^{\P}\,}$ denotes convergence in probability as $\delta\to0$. Since $\sigma$ is a.s.\ a simple path, this also implies that $\beta_t{\,\mathop{\to}\limits^{\P}\,}\sigma_t$. Let $\widehat\phi$ be the conformal map from ${\Bbb U}$ onto ${\Bbb U}-\sigma_t$ that satisfies $\widehat\phi(0)=0$ and $\widehat\phi'(0)>0$, and let $\widehat\psi$ be the similarly normalized conformal map from ${\Bbb U}$ onto ${\Bbb U}-\beta_t$. Because $\beta_t{\,\mathop{\to}\limits^{\P}\,}\sigma_t$, it follows that $\widehat\psi{\,\mathop{\to}\limits^{\P}\,}\widehat\phi$, in the topology of uniform convergence on compact subsets of ${\Bbb U}$. Set $\widehat\psi_\lambda(z):=\widehat\psi(\lambda z)$ and $\widehat\phi_\lambda(z):=\widehat\phi(\lambda z)$ for $\lambda\in\partial{\Bbb U}$. Given every $\epsilon>0$ and a closed set $A\subset{\Bbb U}$ let $F_\epsilon(A)$ be the connected component of $0$ in the set of points with distance at least $\epsilon$ from $A$ (or the empty set, if $d(0,A)<\epsilon$), and let $W_\epsilon(A):=\overline{{\Bbb U}-F_\epsilon(A)}$. By \ref c.unifsl/, for every $\epsilon>0$ there is a random $\lambda\in\partial{\Bbb U}$ independent from $\sigma^1$ (but not from $\beta_t$) such that $\dlr_{F_\epsilon(\beta_t)}(\widehat\psi_\lambda(\sigma^1),\gamma)\to 0$. (The law of $\lambda$ may depend on $\delta$ and $\epsilon$.) Observe that $\P[F_{2\epsilon}(\sigma_t)\subset F_{\epsilon}(\beta_t)]\to 1$ as $\delta\to 0$, because $\beta_t{\,\mathop{\to}\limits^{\P}\,}\sigma_t$. Therefore, we may conclude that $\dlr_{F_{2\epsilon}(\sigma_t)}(\widehat\psi_\lambda(\sigma^1),\gamma)\to 0$. Since this is true for every $\epsilon>0$, it follows that we may choose $\lambda=\lambda_\delta$ so that $\dlr_{{\Bbb U}-\sigma_t}(\widehat\psi_\lambda(\sigma^1),\gamma)\to 0$, as $\delta\to 0$. Because $\widehat\psi_\lambda{\,\mathop{\to}\limits^{\P}\,}\widehat\phi_\lambda$, we therefore also have $\dlr_{{\Bbb U}-\sigma_t}(\widehat\phi_\lambda(\sigma^1),\gamma)\to 0$; that is, $\dlr_{{\Bbb U}}(\sigma_t\cup\widehat\phi_\lambda(\sigma^1),\sigma_t\cup\gamma)\to 0$. Since $\beta_t\cup\gamma$ has the same law as $\beta$ (by \ref l.dprod/), and since $\beta_t{\,\mathop{\to}\limits^{\P}\,}\sigma_t$, this gives, $$ \dlr_{\Bbb U}(\sigma_t\cup\widehat\phi_\lambda(\sigma^1),\beta)= \dlr_{\Bbb U}(\sigma_t\cup\widehat\phi_\lambda(\sigma^1),\beta_t\cup\gamma)\to 0 \,. \label e.prodc $$ Let $\lambda^*$ be random in $\partial{\Bbb U}$ with a law that is some weak (subsequential) limit of the law of $\lambda$ as $\delta\to0$. It follows from \ref e.prodc/ that $\sigma_t\cup\widehat\phi_{\lambda^*}(\sigma^1)$ has the same law as $\sigma$. In particular, it is a simple path. The only possibility is therefore that $\widehat\phi_{\lambda^*}=\phi$ a.s., which completes the proof of \ref p.prod/. \Qed \proofof p.redu Recall that $\zeta=\zeta_\sigma:(-\infty,0]\to\partial{\Bbb U}$ is the L\"owner\ parameter associated to the LERW scaling limit $\sigma\subset{\Bbb U}$. Let $\tilde\zeta$ be the L\"owner\ parameter associated with the path $\sigma^1$, and let $\tilde f_t$ be the associated solution of the L\"owner\ system. Note that $\tilde\zeta(0)=1$, since $\sigma^1\cap\partial{\Bbb U}=\{1\}$. Fix some $t_0<0$. Using \ref p.prod/ and its notations (with $t_0$ replacing $t$), we know that the path $\check\sigma:=\sigma_{t_0}\cup\phi(\sigma^1)$ has the same law as $\sigma$. Let $\check f_t$ be the solution of the L\"owner\ system associated with the path $\check\sigma$, and let $\check\zeta$ be the associated L\"owner\ parameter. Then $\check\zeta$ has the same law as $\zeta$, by \ref p.prod/. Let $\phi$ be as in \ref p.prod/, and set $\lambda := |\phi'(0)|/\phi'(0)$ When $t<t_0$, we have $$ \check f_t(z) = \phi\circ \tilde f_{t-t_0}(\lambda z)\,, $$ because the right hand side is a suitably normalized conformal map from ${\Bbb U}$ onto ${\Bbb U}-\check\sigma_{t}$. We differentiate with respect to $t$, and use \ref e.loew/, to get \begineqalno {\partial \over\partial t} \check f_t(z) & = \phi'(\tilde f_{t-t_0}(\lambda z)) {\partial \over\partial t} \tilde f_{t-t_0}(\lambda z) \cr & = \phi'(\tilde f_{t-t_0}(\lambda z)) \lambda z\tilde f_{t-t_0}'(\lambda z){\tilde\zeta(t-t_0)+ \lambda z\over\tilde\zeta(t-t_0)-\lambda z} = z\check f_t'(z){\lambda^{-1}\tilde\zeta(t-t_0)+ z\over\lambda^{-1}\tilde\zeta(t-t_0)-z} \,. \endeqalno Consequently, it follows that $\check \zeta(t) = \lambda^{-1}\tilde\zeta(t-t_0)$ for $t<t_0$. It is clear that $\check\zeta = \zeta$ for $t\in[t_0,0]$. Continuity of $\check \zeta$ gives $\lambda^{-1} =\zeta(t_0)/\tilde\zeta(0)= \zeta(t_0)$. Since $\zeta$ and $\tilde\zeta$ are independent, and $\tilde\zeta$ has the same law as $\zeta$ conditioned on $\zeta(0)=1$, \ref p.redu/ follows. \Qed We shall need the following \procl t.ito Let $a(t)$, $t\geqslant 0$, be a real valued process (that is, a random function $a:[0,\infty)\to\R$). Suppose that $a$ is continuous a.s.\ and for every $n\in\N$ and every $(n+1)$-tuple $0=t_0\leqslant t_1\leqslant t_2\leqslant\cdots\leqslant t_n$, the increments $a(t_{j})-a(t_{j-1})$, $j=1,\dots,n$, are independent. Then for every fixed $s_0\in(0,\infty)$, the random variable $a(s_0)$ is Gaussian. \endprocl This theorem follows from the general theory of L\'evy processes. An entirely elementary proof can be found in Section~4.2 of \ref b.Ito:lectures/. \procl c.brows There is a constant $c>0$ such that the process $\zeta(t)$ has the same law as ${\ss B}(-c t)$, where ${\ss B}(t)$ is Brownian motion on $\partial {\Bbb U}$ started at a uniform random point. \endprocl \proof That $\zeta(t)$ has the same law as ${\ss B}(-ct)$ for some $c\geqslant 0$ follows immediately from \ref p.redu/ and \ref t.ito/. The fact that $c>0$ is clear, since the LERW scaling limit is not equal a.s.\ to a line segment. \Qed \bsection{The winding number of SLE}{s.wind} \global\advance\consSer by 1\constnum=0\relax Let $\kappa\geqslant 0$, let ${\ss B}(t)$ be Brownian motion on $\partial {\Bbb U}$ started at a uniform random point on $\partial{\Bbb U}$, and set $$ \zeta=\zeta_\kappa:={\ss B}(-\kappa t) \,. \label e.zetadef $$ \procl d.lep Let $\frak K$ denote the set of all $\kappa\geqslant 0$ such that the L\"owner{} evolution $f_t$ defined by \ref e.zetadef/, \ref e.loew/ and \ref e.bdval/ is a.s.\ for every $t<0$ a Riemann map to a slitted disk. For $\kappa\in\frak K$, let $\xi_\kappa$ denote the (random) path defined by $\xi_\kappa(t)=f_t(\zeta(t))$. That is, $\xi_\kappa$ is the path in $\overline{\Bbb U}$ such that $f_t$ is the nomalized Riemann map to ${\Bbb U}-\xi_\kappa\bigl([t,0]\bigr)$. The random process $\xi_\kappa\bigl([t,0]\bigr)$, $t\leqslant 0$, will be called {\bf stochastic L\"owner\ evolution} (SLE) with constant $\kappa$. \endprocl As before, we let $\widehat{\ss B}:[0,\infty)\to\R$ be the continuous map satisfying ${\ss B}=\exp i \widehat {\ss B}$ and $\widehat{\ss B}(0)\in[0,2\pi)$. \procl t.wind Let $\kappa\in\frak K$. Let $T\leqslant 0$, and let $\theta_\kappa(T)$ be the winding number of the path $\xi_\kappa\bigl([T,0]\bigr)$ around $0$; that is $\theta_\kappa(T)=\arg(\xi_\kappa(0))-\arg(\xi_\kappa(T))$, with $\arg$ chosen continuous along $\xi_\kappa$. Then for all $s>0$, $$ \P\Bigl[\bigl|T - \log\left|\xi_\kappa(T)\right|\bigr|>s\Bigr] \leqslant \nco f\exp (-\nco e s) \,, \label e.1e $$ and $$ \P\Bigl[\bigl|\theta_\kappa(T)-\widehat{\ss B}(0)+\widehat{\ss B}(-\kappa T)\bigr|>s\Bigr] \leqslant \co f\exp (-\co e s) \,, \label e.2e $$ where $\co f,\co e>0$ are constants, which depend only on $\kappa$. \endprocl Loosely speaking, the theorem says that $t+i \widehat {\ss B}(-\kappa t)$ is a good approximation of the path $\log \xi_\kappa(t)$. A consequence of the theorem is that $\theta_\kappa(t)/\sqrt{\kappa t}$ converges to a gaussian of unit variance as $T\to-\infty$. \proof Let $f_t$ be defined by \ref e.zetadef/, \ref e.loew/ and \ref e.bdval/. Set $\xi:=\xi_\kappa$. Let $w(t,z):= f_t^{-1}\bigl(f_T(z)\bigr)$, and let $y=y(t,z):=\arg w(t,z)$, where $\arg w$ is chosen to be continuous in $t$. By \ref r.ode/, $w$ satisfies the differential equation $$ {\partial_t} w = -w {{\ss B}(-\kappa t)+w\over {\ss B}(-\kappa t)-w} \,, \label e.zode $$ where ${\partial_t}$ denotes differentiation with respect to $t$. Set $x=x(t,z):=\log|w(t,z)|$. Then $w=\exp(x+iy)$, and \ref e.zode/ can be rewritten, $$ {\partial_t} x+i{\partial_t} y = { \sinh x + i \sin\bigl(\widehat{\ss B}(-\kappa t)-y\bigr) \over \cosh x - \cos \bigl(\widehat{\ss B}(-\kappa t)-y\bigr) } \,. \label e.ab $$ Let $z_1$ be a random point on $\partial {\Bbb U}$, chosen uniformly, and independent from the Brownian motion $B$. Then $w(0,z_1)=f_T(z_1)$ is some point on the boundary of $D_T:=f_T({\Bbb U})$. Note that $\partial D_T$ is a connected set that contains $\partial {\Bbb U}$ and intersects the circle $\partial B\bigl(0,\exp(T)\bigr)$, by \ref e.schw/. Set $A_s:=\bigl\{z\in \partial D_T:\, |z| > \exp(T+s)\bigr\}$. It follows from the continuous version of \ref l.harmm/ for Brownian motion that the harmonic measure of $A_s$ in $D_T$ at $0$ is bounded by $O(1)\exp(-\nco 1 s)$, for some constant $\co 1>0$ and every $s\in\R$; that is, at zero, the bounded harmonic function on $D_T$ that has boundary values $1$ on $A_s$ and has boundary values $0$ on $\partial D_T-\overline A_s$ is bounded from above by $O(1)\exp(-\co 1 s)$. Since harmonic measure is invariant under conformal maps, we conclude that the measure of $f_T^{-1}(A_s)$ is at most $O(1)\exp(-\co 1 s)$. This means that $$ \P\bigl[\log |w(0,z_1)|-T>s\bigr] = \P\bigl[\log |f_T(z_1)|-T>s\bigr] \leqslant O(1)\exp(-\co 1 s) \,. \label e.ww $$ Now set $z_0={\ss B}(-\kappa T)$. Then $\xi(T)=w(0,z_0)$, and so we need to relate $|w(0,z_0)|$ and $|w(0,z_1)|$. Let $\tau$ be the least $t\in[T,0]$ such that $w(t,z_1)={\ss B}(-\kappa t)$, if such a $t$ exists, and set $\tau=0$ if not. Note that $|w(t,z_1)|=1$ while $t<\tau$, and $|w(t,z_1)|<1$ for $t\in(\tau,0]$. Also observe that conditioned on $\tau<0$, the law of the process $\bigl(w(t,z_1):\, t\in [\tau,0]\bigr)$ is the same as the law of the process $\bigl(w(t+T-\tau,z_0):\, t\in [\tau,0]\bigr)$. Consequently, the random variable $w(\tau-T,z_1)$ (where ${\ss B}$ is taken as two-sided Brownian motion and \ref e.zode/ is extended to the range $t>0$), conditioned on $\tau<0$, has the same distribution as the random variable $w(0,z_0)$. By \ref e.ab/, ${\partial_t} x\leqslant 0$, and therefore $|w(\tau-T,z_1)|\leqslant |w(0,z_1)|$ on the event $\tau<0$. Thus, for every $s\in\R$ we have $\P\bigl[|w(0,z_1)|>s\mid\tau<0\bigr]\geqslant \P\bigl[|w(0,z_0)|>s\bigr]$. Because $|w(0,z_1)|=1$ when $\tau=0$, we may drop the conditioning on $\tau<0$. Now \ref e.ww/ gives $$ \P\bigl[\log |\xi(T)|-T>s\bigr] = \P\bigl[\log |w(0,z_0)|-T>s\bigr] \leqslant O(1)\exp(-\co 1 s) \,. $$ On the other hand, the Koebe $1/4$ Theorem \ref e.k1q/ gives $$ \exp T=|f'_T(0)|\leqslant 4 \inf\bigl\{|z|:\, z\notin f_T({\Bbb U})\bigr\} \leqslant 4|\xi(T)|\,, $$ and so $\log\left|\xi(T)\right|+\log 4\geqslant T$ always. This completes the proof of \ref e.1e/. Now let $\tau_1$ be the least $t\in[T,0]$ such that $x(t)=w(t,z_0)\leqslant -1$, and set $\tau_1:=0$ if such a $t$ does not exist. Since $x(t)$ is monotone decreasing, we may write $y(t)$ as a function of $x$: $y=g(x)$. By \ref e.ab/, $$ g'\bigl(x(t)\bigr) = { \sin\bigl(\widehat {\ss B}(-\kappa t)-y(t)\bigr) \over \sinh x(t) } $$ and hence $$ \left|g'\bigl(x(t)\bigr)\right|\leqslant \bigl|\sinh x(t)\bigr|^{-1} \,. $$ And so we get $$ |y(0)-y(\tau_1)| = \int_{x(0)}^{x(\tau_1)} |g'(x)|\,dx \leqslant \int_{-\infty}^{-1} \bigl|\sinh x\bigr|^{-1}\,dx <\infty \,. \label e.bdy $$ Let $\phi(s,t):= f_t ^{-1} \bigl(\xi(s)\bigr)$ for $T\leqslant s\leqslant t\leqslant 0$. Then $\phi$ is continuous and its image does not contain $0$. Hence, it may be considered as a homotopy in ${\Bbb C}-\{0\}$ from the path $\phi(s,0)=\xi(s)$, $s\in[T,0]$, to the concatenation of the inverse of the path $\phi(T,t)= f_t ^{-1}\bigl(\xi(T)\bigr)=w(t,z_0)$, $t\in[T,0]$, with the path $\phi(t,t)={\ss B}(-\kappa t)$, $t\in[T,0]$. Therefore, its winding number is the sum of the corresponding winding numbers. This means that $$ \theta_\kappa(T) = \widehat{\ss B}(0)-\widehat{\ss B}(-\kappa T)+y(T)-y(0) \,. $$ By \ref e.bdy/, it therefore suffuces to prove the appropriate bound on the tail of $|y(\tau_1)-y(T)|$. Let $\ppn y:= \min\bigl\{|y-2 \pi n|:\, n\in\Z\bigr\}$. Set $t_0=T$, inductively, let $t_j$ be the first $t\in[t_{j-1},0]$ such that $\pi/2 = \ppn {\widehat{\ss B}(-\kappa t) - \widehat{\ss B}(-\kappa t_{j-1})}$, and set $t_j=0$ if no such $t$ exists. Equation \ref e.ab/ shows that for every $s\in(T,0)$ $$ {\partial_t} \ppn {y(t)-\widehat{\ss B}(-\kappa s)}\leqslant 0,\qquad\hbox{at }t=s \,. $$ Consequently, for every $j\in\N$, if there is an $s\in[t_j,t_{j+1}]$ such that $\ppn{y(s)-\widehat{\ss B}(-\kappa t_j)}<\pi/2$, then this is satisfied also for all $s'\in[s,t_{j+1})$, because $y(t)$ cannot get out of the set $\Bigl\{p\in\R:\, \bigl|{p-\widehat{\ss B}(-\kappa t_j)}\bigr|_{2\pi}<\pi/2\Bigr\}$ while $\widehat{\ss B}(-\kappa t)$ is in it. This implies that $|y(t_{j+1})-y(t_j)|<2\pi$. Hence $|y(\tau_1)-y(T)|\leqslant 2\pi \min\{j\in\N :\, t_j>\tau_1\}$. Therefore, for every $a>0$ and $n\in\N$, $$ \Pbig{|y(\tau_1)-y(T)|\geqslant 2\pi n} \leqslant \Pbig{\tau_1-T\geqslant a} + \Pbig{t_n\leqslant T+a} \,. \label e.casa $ The first summand on the right hand side is bounded by $O(1)\exp(-\nco s\, a)$, for some constant $\co s>0$, by \ref e.1e/. To estimate the second summand, observe that conditioned on $t_n\leqslant T+a$, we have probability at least $2^{-n-1}$ for the event $$ B\bigl(-\kappa\, (T+a)\bigr)-\widehat{\ss B}(-\kappa T)\geqslant n\pi/2 \,, \label e.evf $$ because when $t_n\leqslant T+a$ and $B\bigl(-\kappa\, (T+a)\bigr)\geqslant B\bigl(-\kappa t_n) \geqslant\cdots\geqslant B\bigl(-\kappa t_0)$, we have \ref e.evf/. However, \ref e.evf/ has probability $$ (2\pi a\kappa)^{-1/2} \int_{n\pi/2}^\infty \exp\bigl(-s^2/(2\kappa a)\bigr)\,ds \leqslant O(1) \exp\bigl(-n^2/\nco n a\bigr) \,, $$ and hence $$ \Pbig{t_n\leqslant T+a} \leqslant O(1) 2^n\exp\bigl(-n^2/\co n a\bigr) \,. $$ We choose $a$ to be $n$ times a very small constant. Then our above estimates, together with~\ref e.casa/, give $$ \Pbig{|y(\tau_1)-y(T)|\geqslant 2\pi n} \leqslant O(1)\exp(-\nco m n) \,, $$ with the constants depending only on $\kappa$. This completes the proof of the theorem. \Qed \bsection{The twisting constant of LERW}{s.windl} \global\advance\consSer by 1\constnum=0\relax Consider some scaling limit measure $\P$ of LERW from $0$ to $\partial{\Bbb U}$, and let $\gamma$ be random with law $\P$. Assuming \ref g.confinv/, we have established that SLE with some constant ${\kappa_o}$ has law $\P$. In this section we show that ${\kappa_o}=2$, and thereby complete the proof of \ref t.lesl/. Let $\epsilon\in(0,1)$, let $\gamma_\epsilon$ be the connected component of $\gamma-B(0,\epsilon)$ which has a point in $\partial U$, and let $W(\gamma_\epsilon,0)$ be the winding number of $\gamma_\epsilon$ around $0$, in radians. That is, $W(\gamma_\epsilon,0)$ is the imaginary part of $\int_{\gamma_\epsilon}z^{-1} dz$. By symmetry, it is clear that $\E[W(\gamma_\epsilon,0)]=0$. We shall show that $$ \Ebig{ W(\gamma_\epsilon,0)^2}= 2\log (1/\epsilon) + O(1)\sqrt{\log (1/\epsilon)}\,. \label e.wnv $$ Based on this and the results of \ref s.wind/, it will follow that ${\kappa_o} = 2$. The proof of \ref e.wnv/ will use Kenyon's work \ref b.Kenyon:conf/. The overall idea of the proof is very simple, and based on the relations between UST and domino tilings. We now briefly review the relations between the UST on $\Z^2$ and domino tilings, and the height function for domino tilings. For a more thorough discussion, the reader should consult \ref b.Kenyon:conf/. A {\bf domino tiling} of the grid $\Z^2$ is a tiling of $\R^2$ by tiles of the forms $[k,k+1]\times[j,j+2]$ and $[k,k+2]\times [j,j+1]$, where $k,j\in\Z$. A domino tiling of $\Z^2$ may also be thought of as a {\bf perfect matching} of the dual grid $\bigl((1/2)+\Z\bigr)^2$. (A perfect matching of a graph $G$ is a set of edges $M\subset{\ss E}(G)$ such that every vertex is incident with precisly one edge in $M$.) Let us start with finite graphs. Let $D$ be a simply connected domain in $\R^2$ whose boundary is a simple closed curve in the grid $\Z^2$, and let $G:=\Z^2\cap\overline D$. Let $\rho_0$ be some vertex in $\partial D\cap\Z^2$, which we call the {\bf root}. Let $\check G$ be the graph $\bigl((1/2)\Z^2)\cap \overline D$ with $\rho_0$ and its incident edges removed. Then there is a bijection, discovered by Temperley, between the set of perfect matchings on $\check G$ and spanning trees of $G$. Temperley's bijection (see \ref f.temperley/) works as follows. For every edge $[v,u]$ in the matching $M$ such that $v\in\Z^2$, we put in the tree the edge $e_u$ whose center is $u$. This gives the set of edges in the tree $T$. If $[v,u]\in M$ is as above, we may orient the edge $e_u$ away from $v$, and then the tree $T$ will be oriented towards the root $\rho_0$. \@midtrue\@ins \SetLabels \R\E(.24*-.01)$\rho_0$\\ \endSetLabels \tabskip=1em plus2em minus.5em \halign to \hsize{\hfil # \hfil&\hfil # \hfil&\hfil # \hfil\cr \AffixLabels{\epsfxsize=1.5in\epsfbox{temperley1.eps}}&% \epsfxsize=1.5in\epsfbox{temperley2.eps}&% \epsfxsize=1.5in\epsfbox{temperley3.eps}\cr \noalign{\smallskip} $G$&$\check G$&matching and tree\cr } \medskip \narrower\noindent \figlabel{temperley} Temperley's bijection. On the right, the arrows are edges in the matching that containing vertices of $\Z^2$, the solid segments are other edges in the matching, and the thin lines are edges in the tree. \endinsert Temperley's bijection works also in more general situtations. There is a simple modification to make it work for the wired graph associated to the domain $D$. Also, given a perfect matching on all of $(1/2)\Z^2$, there is an associated (oriented) spanning forest of $\Z^2$. The collection of all domino tilings of $\Z^2$ has a natural probability measure (of maximal entropy), and for a.e.\ domino tiling the corresponding spanning forest is a spanning tree. Temperley's map from perfect matchings on $(1/2)\Z^2$ to spanning forests of $\Z^2$ maps the cannonical probability measure on the set of domino tilings to the law of the UST of $\Z^2$. Let $G$ and $\check G$ be as above, and let $\widehat G$ be the graph of the domino tiling, that is, the union of the squares of edge length $1/2$ with centers at the vertices of $\check G$, thought of as a subgraph of the grid $(1/4,1/4)+(1/2)\Z^2$. Associated to a domino tiling of $\widehat G$ is a {\bf height function} $h$ defined on the vertices of $\widehat G$. Here is the definition of $h$. Pick some vertex $v_0\in{\ss V}(\widehat G)$ and some $a_0\in\R$, and set $h(v_0)=a_0$. Color a square face of the grid $(1/4,1/4)+(1/2)\Z^2$ {\bf white} if its center is a vertex of $\Z^2$ or if it is contained in a face of $\Z^2$, and {\bf black} otherwise. If $[u,v]\in{\ss E}(\widehat G)$ is on the boundary of a domino tile in the tiling, then we require that $h(v)-h(u)=1$ if the square to the right of the directed edge $[u,v]$ is white and $h(v)-h(u)=-1$ if the square to the right of $[u,v]$ is black. These constraints uniquely specify the height function $h$ (except that the choices of $v_0$ and $a_0$ are arbitrary). We will work in the upper half plane ${\Bbb H}:=\{z\in{\Bbb C}:\, \operatorname{Im} z>0\}$. Let $G_\delta(\H):=G^W({\Bbb H},\delta)$, the wired graph of mesh $\delta$ associated with the domain ${\Bbb H}$, and $\widehat G_\delta(\H):={\Bbb H}\cap\bigl((1/4,1/4)+(1/2)\Z^2\bigr)$. The discussion above carries through for the grid $\delta\Z^2$, in place of $\Z^2$. (Although the distance between adjacent vertices in the graph $\widehat G$ is $\delta/2$ when $G\subset\delta\Z^2$, we still work with the height function where the height difference along an edge on the boundary of a tile is $\pm1$.) Temperley's bijection induces a measure preserving transformation between domino tilings of the grid $\widehat G_\delta(\H)$ and the UST of $G_\delta(\H)$. (If we keep the orientation, then the UST is directed towards $\partial{\Bbb H}$.) We normalize the height function associated to a domino tiling of $\widehat G_\delta(\H)$ by requiring that $h\bigl((1/4,1/4)\bigr)=1/2$. Then $h\bigl(((2k+1)/4,1/4)\bigr)=(-1)^k/2$, for $k\in\Z$. If $v$ is some vertex in $G_\delta(\H)$, let $h(v)$ be the average of the value of $h$ on the vertices of $\widehat G_\delta(\H)$ closest to $v$. \procl l.windheight Let $T$ be a spanning tree of $G_\delta(\H)$, and let $h$ be the associated height function. Let $v\in{\ss V}(G_\delta(\H))$ be a vertex different from the wired vertex $\partial {\Bbb H}$, and let $a$ be the real part of $v$. Let $Q_v'$ be the path from $v$ to $\partial{\Bbb H}$ in $T$, considered as a path in the plane, and let $Q_v$ be the union of $Q_v'$ with the line segment joining the intersection $Q_v'\cap\partial{\Bbb H}$ to $a$. Then $-\pi h(v)/2=W(Q_v,v)$, the winding number of $Q_v'$ around $v$. \endprocl This lemma is a special case of a more general observation made by Kenyon. (Since $Q_v$ is a path with $v$ as an endpoint, we define $W(Q_v,v):=\lim_{r\to0}W\bigl(Q_v-B(v,r),v\bigr)$; which is the same as $W\bigl(Q_v-B(v,\delta/2),v\bigr)$.) \proof Use induction on the length of the path $Q'_v$. \Qed Symmetry implies that $\E[h(v)]=0$ for all $v\in{\ss V}(G_\delta(\H))$. Kenyon has shown \ref b.Kenyon:conf/ that $$ \E[h(v)^2] =8\pi^{-2}\log(1/\delta)+O(1) \label e.kenyon $$ (provided that $v$ stays in a compact subset of ${\Bbb H}$). Hence $\E[W(Q_v,v)^2]=2\log(1/\delta)+O(1)$, which seems very close to a proof of \ref e.wnv/. However, to make it into a proof of \ref e.wnv/ requires some effort (it seems). The advantage of the height function over the winding number is that the height difference between two vertices can be computed along any path joining them. On the other hand, to compute $W(Q_v,v)$, one might think that it is necessary to follow $Q_v$, which is a random path. It is immediate that $h(v)$ is $\sum_j \lambda_j \chi_j$, where $\chi_j$ is the indicator of the event that a certain domino tile is present in the tiling, and $\lambda_j$ are some explicit easy to compute (non-random) weights. This means that to calculate $\E[h(v)^2]$ one needs to have a good estimate for the behavior of the correlations $\E[\chi_i\chi_j]$ for small $\delta$. That's how Kenyon proves \ref e.kenyon/. Recall that $A(p,r_1,r_2)$ denotes an annulus with center $p$, inner radius $r_1$, and outer radius $r_2$. The following result is an immediate consequence from \ref b.ABNW/. \procl l.expo Let $D\subset{\Bbb C}$ be a domain, and let $v_0\in\delta\Z^2\cap D$ be some vertex. Consider $T$, the UST on $\delta\Z^2\cap D$, with free or wired boundary. Let $r_2>2r_1>2\delta$, and suppose that $r_2$ is smaller than the distance from $v_0$ to $\partial D$. Let $A$ be the annulus $A:= A(v_0,r_1,r_2)$, and let $k(A)$ be the maximum number of disjoint paths in $T$ each of which intersects both boundary components of $A$. Then for each $k\in\N$ $$ \Pbig{k(A)\geqslant k} \leqslant \nco z (r_2/r_1)^{\nco e (k-1)} \,, $$ where $\co z,\co e>0$ are universal constants. \Qed \endprocl \procl l.windtail Let $D\subset{\Bbb C}$ be some domain, and consider $T$, the UST in $G^W(D,\delta)$ (wired boundary). Let $p_0\in D\cap \delta\Z^2$. Given a set $K\subset D$, let $X(K)$ denote the maximum winding number around $p_0$ of a path in $T$ with endpoints in $K$. Let $r\in\bigl(0,d(p_0,\partial D)/2\bigr)$. Then for each $h>0$ and $s\geqslant 2$, $$ \P\Bigr[X\bigl(A(p_0,r/s,r)\bigr)>h\Bigl]\leqslant \nco a \exp\bigl(-\nco b h/\log s \bigr)\log s \,, $$ where $\co a$ and $\co b$ are absolute constants. \endprocl \proof Set $A:= A(p_0,r/s,r)$, $X:= X(A)$. To begin, assume that $s=2$. Let $B_1,B_2,\dots,B_N$ be a covering of $A$ with balls of radius $r/10$, where $N\leqslant \nco N$, with $\co N$ some universal constant. Let $v_0\in A\cap\delta\Z^2$ be some vertex. For any vertex $u$, let $W(u)$ be the signed winding number around $p_0$ of the path in $T$ from $v_0$ to $u$. Consider neighbors $v,w$ in $A$, and let $\gamma$ be the UST path joining $v$ and $w$ in $T$. If $\gamma\neq [v,w]$, then $\gamma\cup [v,w]$ is a simple closed path, and hence has winding number at most $2\pi$ around $p_0$. This implies that $|W(v)-W(u)|$ is at most $2\pi$ plus the absolute value of the winding number of the edge $[v,u]$ around $p_0$, and therefore $|W(v)-W(u)|\leqslant 3\pi$. From this it follows that we may find vertices $v_1,v_2,\dots,v_n$ in $A$ such that $|W(v_j)-W(v_k)|\geqslant 3\pi$ for $j\neq k$ and $n\geqslant X/6\pi$. Since $N\leqslant \co N$, we may find some ball $B_m$ from the above collection, satisfying $|B_m\cap \{v_1,\dots,v_n\}|\geqslant X/6\pi\co N$. However, if $v,u\in B_m$ and $|W(v)-W(u)|\geqslant 2\pi$, then the path in $T$ joining $v$ and $u$ must go around $p_0$. In particular, it must cross twice the annulus $A_m:=A(c_m,r/10,r/5)$, where $c_m$ is the center of $B_m$. It follows that the number of disjoint crossings of $A_m$ in $T$ is at least $|B_m\cap \{v_1,\dots,v_n\}|$. Hence, by \ref l.expo/, for any fixed $m$ the probability that $|B_m\cap \{v_1,\dots,v_n\}|>b$ is at most $O(1)\exp(-\nco q b)$, for some constant $\co q>0$. Consequently, $\P[X>h]\leqslant O(1)\co N\exp(-\co q h/6\pi\co N)$, which completes the proof in the case $s=2$. If $s>2$, then we may cover the annulus $A$ with at most $2\log s+1$ disjoint concentric annuli with radii ratio $2$. In order that $X(A)$ be at least $h$, there must be one of these smaller annuli $A'$ with $X(A')\geqslant h/(2\log s +1)$. The lemma follows. \Qed \procl l.rick \procname{{\rm\ref b.Kenyon:conf/}} Let $p,q\in{\Bbb H}$ be any two points. Consider a uniform domino tiling of the grid $\widehat G_\delta(\H)\subset{\Bbb H}$ of mesh $\delta$, and let $p',q'\in{\ss V}(\widehat G_\delta(\H))$ be vertices closest to $p$ and $q$, respectively. Then $$ \lim_{\delta\to0}\E [h(p')h(q')] = 8\pi^{-2} \log\left|{\overline p-q\over p-q}\right| \,. \Qed $$ \endprocl It may be noted that the right hand side is invariant under conformal automorphisms of ${\Bbb H}$, since it is the log of the square root of a cross ratio of $p,\overline p,q,\overline q$. Let $\delta>0$, let $v_0$ be a vertex of the grid $\delta\Z^2$ which is closest to $i=\sqrt{-1}$. Fix some $r\in(0,1/2)$. Let $\gamma_r^\delta$ be the connected component of $Q_{v_0}-B(v_0,r)$ that intersects $\partial{\Bbb H}$, and let $W_\delta(r)$ be the winding number of $\gamma_r^\delta$ around $v_0$. \procl p.windw Assuming $r<1/4$, $$ \limsup_{\delta\to0}\Bigl| \E\bigl[W_\delta(r)^2\bigr] - 2\log(1/r)\Bigr| \leqslant \nco e \sqrt{\log(1/r)} \,, \label e.pwindw $$ where $\co e$ is an absolute constant. \endprocl \proof Let $v_1$ be a vertex in $\delta\Z^2$ such that $v_1-v_0-r/3\in[0,\delta)$. Let $V_0$ be the set of vertices of $\delta\Z^2$ whose Euclidean distance to $\{v_1,v_0\}$ is in the range $(r/9,2r)$. Then, assuming that $\delta<r/9$, $V_0$ separates $v_0$ from $v_1$ and $\{v_0,v_1\}$ from $\partial{\Bbb H}$ in the grid $\delta\Z^2$. Given a vertex $v$, let $\ray^0_v$ be the union of $Q_v'$ with the line segment joining the intersection $Q_v'\cap\partial{\Bbb H}$ to $0$. Set $Q:=\bigcup\bigl\{\ray^0_v:\, v\in V_0\}$. Set $X_j:= W(Q_{v_j},v_j)$, $j=0,1$. By \ref l.windheight/, have $X_j=-\pi h(v_j)/2$, and consequently, \ref l.rick/ gives, $$ \lim_{\delta\to 0}\E[X_0X_1] = 2\log(1/r) + O(1) \,. \label e.r $$ Since $V_0$ separates $v_0$ from $v_1$, when conditioning on $Q$, $X_0$ becomes independent from $X_1$. Therefore, $$ \E[X_0X_1]=\E\bigl[\E[X_0X_1\mid Q]\bigr]= \E\Bigl[\E[X_0\mid Q]\cdot\E[X_1\mid Q]\Bigr] \,. \label e.inde $$ We shall show that for small $\delta>0$, $$ \E\Bigl[\bigl(W_\delta(r)-\E[X_j\mid Q]\bigr)^2\Bigr]=O(1), \qquad j=0,1\,. \label e.y $$ Using \ref e.inde/, this implies $$ \E[W_\delta(r)^2]-\E[X_0X_1]=O(1)\sqrt{\E[W_\delta(r)^2]}+O(1) \,. $$ Consequently, by \ref e.r/, $$ \limsup_{\delta\to 0}\Bigl|\E[W_\delta(r)^2]-2\log(1/r)+O(1)\sqrt{\E[W_\delta(r)^2]}\Bigr| \leqslant O(1) \,, $$ which implies \ref e.pwindw/. It therefore suffices to prove \ref e.y/. Let $s:=\min\bigl\{|v-v_0|:\, v\in Q\bigr\}$. Given $Q$, let $T'$ be a UST of $G^W(B(v_0,s/2),\delta)$, and let $T''$ be the UST of the graph obtained from $G_\delta(\H)$ by identifying the vertices of $Q$. Note that (by Wilson's algorithm, say) $T$, the UST on $G_\delta(\H)$, has the same law as $T''\cupQ$ (as a set of edges). By the domination principle, given $Q$, we may couple $T''$ and $T'$ so that ${\ss E}(T'')\supset {\ss E}(T')$. Given $Q$, we couple $T$ and $T'$ so that $T\supset T'$. Let $\alpha$ be the path in $T'$ from $v_0$ to $\partial B(v_0,s)$, and let $a$ be the point where $\alpha$ hits $\partial B(v_0,s)$. Then $\E[W(\alpha,v_0)\midQ]=0$, by symmetry, because given $Q$, $T'$ is just ordinary UST on $G^W(B(v_0,s/2),\delta)$. But $\alpha$ is also a path in $T$. Let $\beta$ be the path in $T$ from $a$ to the endpoint of $\gamma_r^\delta$ near $\partial B(v_0,r)$. Then $X_0=W(\alpha,v_0)+W(\beta,v_0)+W_\delta(r)$, and therefore, $$ W_\delta(r)-\E[X_0\mid Q] = -\E[W(\alpha,v_0)\midQ] -\E[W(\beta,v_0)\midQ] = -\E[W(\beta,v_0)\midQ] \,. \label e.break $$ Let $t>\delta$. Note that if $s<t$, then there are at least two disjoint crossings in $T$ of the annulus $A(v_0,t,r/10)$. Therefore, \ref l.expo/ gives $$ \P[s<t]\leqslant O(1) (t/r)^{\nco d} \,. \label e.ss $$ Fix some $y\geqslant 2$. By \ref l.windtail/, we have $$ \P\Bigl[\bigl|W(\beta,v_0)\bigr|>t,\ s\geqslant r/y\Bigr] \leqslant O(1) \exp\bigl(-\co b t/\log y\bigr)\log y \,. $$ Hence, using \ref e.ss/, \begineqalno \P\Bigl[\bigl|W(\beta,v_0)\bigr|>t\Bigr] & \leqslant \P[ s< r/y\Bigr]+ O(1) \exp\bigl(-\co b t/\log y\bigr)\log y \cr & \leqslant O(1) y^{-\co d} + O(1) \exp\bigl(-\co b t/\log y\bigr)\log y \,. \endeqalno Assuming that $t\geq1$, we may choose $y=\exp\sqrt t$, and then get $$ \P\Bigl[\bigl|W(\beta,v_0)\bigr|>t\Bigr] \leqslant O(1)\sqrt t \exp\bigl(-\nco z \sqrt t\bigr) \,, $$ for some constant $\co z>0$. This gives $\E\bigl[W(\beta,v_0)^2\bigr]=O(1)$. But for every random variable $Y$, we have $\E[Y^2]\geqslant \E\bigl[\E[Y\midQ]^2\bigr]$. Therefore, \ref e.break/ implies \ref e.y/ for $j=0$. The proof of~\ref e.y/ for $j=1$ is entirely the same. This completes the proof of the proposition. \Qed \procl p.theconst Assuming \ref g.confinv/, ${\kappa_o}=2$, where ${\kappa_o}$ is the constant such that SLE with parameter ${\kappa_o}$ is the scaling limit of LERW. \endprocl \proof Recall the definition of $W_\delta(r)$, which appears above \ref p.windw/. Set $r_0=1/2$, and let $r_1>0$ be very small. Let $\psi:{\Bbb H}\to{\Bbb U}$ be the conformal map satisfying $\psi(i)=0$ and $\psi'(i)>0$. Let $C_j$ be the circle of radius $r_j$ about $i$, $j=0,1$, and set $C_j':=\psi(C_j)$. Note that $Z_\delta(r_1)=W_\delta(r_1)-W_\delta(r_0)$ is the winding number around $i$ of some arc on $\gamma^\delta$, the LERW from a vertex near $i$ to $\partial{\Bbb H}$ in $\delta\Z^2\cap{\Bbb H}$, and the arc has one endpoint near the circle $\partial B(i,r_0)$ and the other endpoint near the circle $\partial B(i,r_1)$. It follows that $Z_\delta(r_1)$ converges weakly to a winding number $Z(r_1)$ of an arc $\beta$ of $\psi^{-1}(\xi_{\kappa_o})$ with endpoints on $C_0$ and $C_1$, as $\delta\to0$ along some sequence, where $\xi_{\kappa_o}$ is the SLE curve with parameter ${\kappa_o}$. Moreover, since we have good tail estimates on $Z_\delta(r_1)$ (\ref l.windtail/), from the dominated convergence theorem it follows that $$ \E[Z(r_1)^2]=\lim_{\delta\to 0} \E[Z_\delta(r_1)^2] \,. $$ Hence, \ref p.windw/ gives, $ \E\bigl[Z(r_1)^2\bigr] = 2\log(1/r_1) + O(1) \sqrt{\log(1/r_1)} $. Observe that for any path $\alpha$ in ${\Bbb H}-\{i\}$, the winding number of $\alpha$ around $i$ minus the winding number of $\psi(\alpha)$ around $0$ is bounded by some constant. Consequently, the winding number $W'$ of $\beta':=\psi(\beta)$ around $0$ also satisfies $$ \E\bigl[W'^2\bigr] = 2\log(1/r_1) + O(1) \sqrt{\log(1/r_1)} \,. \label e.www $$ Set $t_j=\log r_j$, $j=0,1$, let $\tilde \beta$ be the arc $\xi_{\kappa_o}(t):[t_1,t_0]\to{\Bbb U}$, and let $\tilde W$ be the winding number of $\tilde \beta$ around $0$. By \ref t.wind/, with high probability, the log of the absolute value of the endpoints of $\tilde \beta$ is not far from the log of the absolute value of the endpoints of $\beta'$. Therefore, it is easy to conclude with the help of \ref l.windtail/, that $$ \Ebig{(\tilde W- W')^2}=O(1) \,. \label e.wwp $$ We know from \ref t.wind/ again that $$ \Ebig{{\tilde W}^2}={\kappa_o}|t_1| + O(1)\sqrt{|t_1|}\,. $$ Combining this with \ref e.www/ and \ref e.wwp/ gives $$ (2-{\kappa_o})\left|t_1\right| = O(1)\sqrt{\left|t_1\right|}\,. $$ Letting $t_1\to-\infty$ now completes the proof. \Qed \proofof t.lesl Immediate from \ref c.brows/ and \ref p.theconst/. \Qed \bsection{The critical value for the SLE}{s.crit} \global\advance\consSer by 1\constnum=0\relax \procl t.kk $\sup\frak K\leqslant 4$, where $\frak K$ is as in \ref d.lep/. \endprocl \proof Fix some $\kappa\in\frak K$, and let $f_t$ be the solution of the L\"owner{} equation with parameter $\zeta(t)={\ss B}(-\kappa t)$, where ${\ss B}:[0,\infty)\to\partial{\Bbb U}$ is Brownian motion starting from a uniform point in $\partial{\Bbb U}$. Note the for every $t<0$ the map $f_t^{-1}$ is well defined and injective on $\partial {\Bbb U}-\{\zeta(t)\}$, since $f_t$ is a Riemann map onto a slit domain, and the slit hits $\partial{\Bbb U}$ at $\zeta(t)$. Set $b(t)=-i\log f_t^{-1}(1)$, with $b(0)-\widehat{\ss B}(0)\in[0,2\pi)$ and $b(t)$ continuous in $t$. Then $b(t)$ is real. As in \ref e.ab/, we have $$ b'(t) = {\sin\bigl(\widehat{\ss B}(-\kappa t)-b(t)\bigr)\over 1-\cos\bigl(\widehat{\ss B}(-\kappa t)-b(t)\bigr)} = \cot{1\over 2}\bigl(\widehat{\ss B}(-\kappa t)-b(t)\bigr) \,. \label e.deq $$ Let $p(s)=b(-s)-\widehat{\ss B}(\kappa s)$, and let $\epsilon>0$. Set $\tau_\epsilon=\inf\{s\geqslant 0:\, p(s)=\epsilon\}$ and $\tau_\pi =\inf\{s\geqslant 0:\, p(s)=\pi\}$. For $x\in[\epsilon,\pi]$, let $g_\epsilon(x)$ be the probability that $\tau_\pi<\tau_\epsilon$, conditioned on $p(0)=x$. Also set $g_\epsilon(x)=0$ for $x<\epsilon$ and $g_\epsilon(x)=1$ for $x>\pi$. We now show that $g_\epsilon$ satisfies $$ {\kappa\over 2} g_\epsilon''(x) + g_\epsilon'(x) \cot (x/2)=0 \label e.gde $$ inside $(\epsilon,\pi)$, using It\^o's formula. (The reader unfamiliar with stochastic calculus can have a look at \ref b.Durrett:BMM/, for example, or try to derive \ref e.gde/ directly. The latter is a bit tricky, but can be done.) Observe that $g_\epsilon\bigl(p(s^*)\bigr)$ is a martingale, where $s^*=\min\{s,\tau_\epsilon,\tau_\pi\}$. By \ref e.deq/, we have $$ dp(s) = -b'(-s) ds -d\widehat{\ss B}(\kappa s) = \cot\bigl(p(s)/2\bigr) ds - d\widehat{\ss B}(\kappa s) \,, $$ and therefore, by It\^o's Formula (assuming, for the moment, that $g_\epsilon$ is $C^2$), \begineqalno d g_\epsilon\bigl(p(s)\bigr) &= g_\epsilon'\bigl(p(s)\bigr)\Bigl(\cot\bigl(p(s)/2\bigr)\,ds-d\widehat{\ss B}(\kappa s)\Bigr) +(1/2)g_\epsilon''\bigl(p(s)\bigr)\,d\left<\widehat{\ss B}(\kappa s)\right> \cr & =\Bigl( \cot\bigl(p(s)/2\bigr)g_\epsilon'\bigl(p(s)\bigr) +(\kappa/2)g_\epsilon''\bigl(p(s)\bigr)\Bigr) \,ds -g_\epsilon'\bigl(p(s)\bigr)\,d\widehat{\ss B}(\kappa s) \endeqalno for $s<\min\{\tau_\epsilon,\tau_\pi\}$. Since $g_\epsilon\bigl(p(s^*)\bigr)$ is a martingale, the $ds$ term must vanish, and so \ref e.gde/ holds inside $(\epsilon,\pi)$. Consequently, in that range, $$ g_\epsilon'(x) = c_\epsilon \bigl(\sin (x/2)\bigr)^{-4/\kappa} \,, $$ where $c_\epsilon$ is some constant depending on $\epsilon$. Since $g_\epsilon(\epsilon)=0$ and $g_\epsilon(\pi)=1$, we have $\int_{\epsilon}^\pi g_\epsilon'(x)\,dx=1$, which gives $$ c_\epsilon^{-1} = \int_{\epsilon}^{\pi} \bigl(\sin (x/2)\bigr)^{-4/\kappa}\,dx \,. $$ We know that a.s.\ $p(s)\neq 0$ for all $s$, which is equivalent to $\lim_{\epsilon\to 0} g_\epsilon(x)= 1$ on $(0,\pi)$. This gives $\lim_{\epsilon\to0} g_\epsilon'(x)= 0$; that is, $\lim_{\epsilon\to 0}c_\epsilon=0$. Therefore, $\kappa\leqslant 4$. This completes the proof, except that we have not shown that $g_\epsilon$ is $C^2$ (there should be a reference implying this, but we have not located one). To deal with this, the above procedure is reversed. {\bf Define} $g_\epsilon$ as the solution of \ref e.gde/ satisfying $g_\epsilon(\epsilon)=0$ and $g_\epsilon(1)=1$. Then the above application of It\^o's Formula shows that $g_\epsilon\bigl(p(s^*)\bigr)$ is a martingale. By the Optional Sampling Theorem, this implies that $g_\epsilon(x)$ is the probability that $\tau_\pi<\tau_\epsilon$, conditioned on $p(0)=x$, and completes the proof. \Qed \procl g.kk $\frak K=[0,4]$. \endprocl \bsection{Properties of UST subsequential scaling limits in two dimensions}{s.ustsl} \global\advance\consSer by 1\constnum=0\relax Before we go into the study of the UST scaling limit, let us remark that the definition we have adopted for the scaling limit is by no means the only reasonable one. There are several other reasonable variations, and choosing one is partly a matter of convenience and taste. We now recall some definitions. Again, we think of $\delta\Z^2$ as a subset of the sphere $\Sp2=\R^2\cup\{\infty\}$. Recall that ${\hat \T}_\delta$ denotes the UST on $\delta\Z^2$, with the point $\infty$ added, to make it compact. Given two points $a,b\in{\hat \T}_\delta$, $a\neq b$, $\omega_{a,b}=\omega_{a,b}^\delta$ denotes the unique path in ${\hat \T}_\delta$ with endpoints $a$ and $b$. For the case $a=b$, we set $\omega_{a,a}=\{a\}$. Let $\ST_\delta$ be the collection of all triplets, $(a,b,\omega_{a,b})$, where $a,b\in{\hat \T}_\delta$. $\ST_\delta$ will be called the {\bf paths ensemble} of ${\hat \T}_\delta$. Let $\ST$ denote a random variable in ${\Cal H}\bigl(\Sp2\times\Sp2\times{\Cal H}(\Sp2)\bigr)$ whose law is a weak subsequential limit of the law of $\ST_\delta$ as $\delta\to0$. The trunk is defined by $$ \Tr=\Tr(\ST):=\bigcup_{(a,b,\omega)\in\ST} ( \omega-\{a,b\}) \,. \label e.trunckdef $$ Let $\epsilon\in(0,1]$, and $a,b\in{\hat \T}_\delta$. We define $\omega_{a,b}(\epsilon)$ as follows. Let $a'$ be the first point along the path $\omega_{a,b}$ (which is oriented from $a$ to $b$) where $d(a,a')=\epsilon$, and let $b'$ be the last point along the path where $d(b,b')=\epsilon$, provided that such points exist. If $a'$ and $b'$ exist, and $a'$ appears on the path before $b'$, then let $\omega_{a,b}(\epsilon)$ be the (closed) subarc of $\omega_{a,b}$ from $a'$ to $b'$; and otherwise set $\omega_{a,b}(\epsilon)=\emptyset$. Let $\ST_\delta(\epsilon)$ denote the set of all triplets $\bigl(a,b,\omega_{a,b}(\epsilon)\bigr)$ such that $a,b\in{\hat \T}_\delta$ and $\omega_{a,b}(\epsilon)\neq\emptyset$. Note that if $d(a,b)>2\epsilon$, then $\omega_{a,b}(\epsilon)\neq\emptyset$. We define $$ \Tr_\delta(\epsilon) :=\bigcup\Bigl\{\omega:\, (a,b,\omega)\in\ST_\delta(\epsilon)\Bigr\}\,. $$ Then $\Tr_\delta(\epsilon)$ is a compact subset of ${\hat \T}_\delta$, which we call the {\bf $\epsilon$-trunk} of ${\hat \T}_\delta$. By compactness, for every $\epsilon\in(0,1]$ there is a subsequential scaling limit of the law of $\Tr_\delta(\epsilon)$. By passing to a subsequence, if necessary, we assume that for all $n\in{\N_+}:=\{1,2,\dots\}$ the weak limit $\Tr_0(1/n)$ of $\Tr_\delta(1/n)$, as $\delta\to 0$, exists. Recall that the {\bf dual} $\d T$ of a spanning tree $T\subset \delta\Z^2$ is the spanning subgraph of the dual graph $\d{(\delta\Z^2)}:=(\delta/2,\delta/2)+\delta\Z^2$ containing all edges that do not intersect edges in $T$. If $T$ is the UST on $\delta\Z^2$, then $\d T$ has the law of the UST on $\d{(\delta\Z^2)}$. (See, e.g., \ref b.BLPS:usf/.) Let $\d{\hat \T}_\delta$ be $\d T\cup\{\infty\}$, where $T$ is the UST on $\delta\Z^2$. $\d\Tr$ and $\d\Tr_0(\epsilon)$ are defined for $\d{\hat \T}_\delta$ as $\Tr$ and $\Tr_0(\epsilon)$ were defined for ${\hat \T}_\delta$. We may think of the random variables $\Tr$, $\d\Tr$, $\Tr_0(1/n)$, and $\d\Tr_0(1/n)$ ($n\in{\N_+}$) as defined on the same probability space, by taking a subsequential limit of the joint distribution of $\ST_\delta$, $\d \ST_\delta$, $\bigl<\Tr_\delta(1/n):\, n\in{\N_+}\bigr>$, and $\bigl<\d\Tr_\delta(1/n):\, n\in{\N_+}\bigr>$. It is immediate to verify that a.s. $$ \Tr=\bigcup_{n\in{\N_+}} \Tr_0(1/n)\,, $$ and $\Tr_0\bigl(1/(n+1)\bigr)\supset \Tr_0(1/n)$ for $n\in{\N_+}$. We shall prove that $\Tr$ is a.s.\ a topological tree, in the sense of the following definition. \procl d.toptree \procname{Trees} An {\bf arc} joining two points $x,y$ in a metric space $X$ is a set $J\subset X$ such that there is a homeomorphism $\phi:[0,1]\to J$ with $\phi(0)=x$ and $\phi(1)=y$. A metric space $X$ will be called a {\bf topological tree} if it is uniquely arcwise connected (that is, given $x\neq y$ in $X$ there is a unique arc in $X$ joining $x$ and $y$) and locally arcwise connected (that is, whenever $x\in U$ and $U$ is an open subset of $X$ there is an open $W\subset U$ with $x\in W$ and $W$ is arcwise connected). A {\bf finite topological tree} is a topological space which is homeomorphic to a finite, connected, simply connected, $1$-dimensional simplicial complex. \endprocl Note that a connected subset of a topological tree is a topological tree \ref b.Bowditch:treelike/. Although we shall not need this fact, it is instructive to note that a metric space which is a topological tree is homeomorphic to an $\R$-tree% \Ftnote{An $\R$-tree is a metric space $(T,d)$ such that for every two distinct points $x,y\in T$ there is a unique isometry $\phi$ from $[0,d(x,y)]$ onto a subset of $T$ satisfying $\phi(0)=x$ and $\phi\bigl(d(x,y)\bigr)=y$.} \ref b.MO:treechar/ (see also \ref b.MMOT:treechar/, for a slightly less general but simpler proof). The next theorem establishes a finiteness property of the $\epsilon$-trunks, which is the first step in the proof of \ref t.ustlim/. \procl t.fin \procname{Finiteness} For every $\epsilon>0$ there is a $\widehat\delta>0$ with the following property. Suppose that $0<\delta<\widehat\delta$. Let $\widehat V$ be a set of vertices of $\delta\Z^2$ such that every point in $\Sp2$ is within distance $\widehat\delta$ of some vertex in $\widehat V$. Let $Q:=Q_\delta(\widehat V)$ be the subtree of ${\hat \T}_\delta$ that is spanned by $\widehat V$; that is, the minimal connected subset of ${\hat \T}_\delta$ containing $\widehat V$. Then with probability at least $1-\epsilon$ we have $Q\supset\Tr_\delta(\epsilon)$. \endprocl \proof Fix some small $\widehat \delta>0$, and suppose that $\delta\in(0,\widehat \delta)$. Let $V_0:=\widehat V$, and for each $j\in{\N_+}$ let $V_j$ be a set of vertices containing $V_{j-1}$ such that every vertex of $\delta\Z^2$ is within spherical distance $\delta_j:=2^{-j}\widehat\delta$ of some vertex in $V_j$, and $V_j$ is a minimal set satisfying these properties. Note that the number of vertices in $V_j-V_{j-1}$ is bounded by $O(1)\delta_j^{-2}$. Let $Q_j$ be the subtree of ${\hat \T}_\delta$ spanned by $V_j$. We now estimate the probability that there is some component of $Q_{j+1}-Q_{j}$ whose diameter is large. Let $v$ be some vertex in $\delta\Z^2$, let $Q(v,j)$ be the arc of ${\hat \T}_\delta$ that connects $v$ to $Q_j$, and let $\ev D(v,j,a)$ be the event the diameter of $Q(v,j)$ is at least $a\delta_j$. By Wilson's algorithm, we may obtain $Q(v,j)$ by conditioning on $Q_j$ and loop-erasing a simple random walk from $v$ that stops when $Q_j$ is hit. Every vertex $w\in \delta\Z^2$ is within distance $\delta_j$ from a vertex in $Q_j$. Since $Q_j$ is connected and has diameter at least $1$, \ref l.harmm/ shows that there is a universal constant $\nco 1>0$ so that the probability that a random walk from $w$ gets to distance $\co 1\delta_j$ from $w$ before hitting $Q_j$ is at most $1/2$. Consequently, $\ev D(v,j,a)$ has probability at most $O(1)\exp({-\nco 3 a})$, where $\co 3>0$ is an absolute constant. We choose $a_j:=j^2 (\log\widehat \delta)^2/\co 3$. Since there are at most $O(1)\delta_j^{-2}$ vertices in $V_{j+1}-V_j$, we find that the probability of $$ \ev D := \bigcup_{j=1}^{\infty}\bigcup_{v\in V_{j+1}}\ev D(v,j,a_j) $$ is bounded by $$ O(1)\widehat\delta^{-2}\sum_{j=1}^\infty 2^{2j}\exp\bigl(-j^2(\log\widehat\delta)^2\bigr) \,, $$ which goes to zero as $\widehat\delta\to 0$. Let $v\in \delta\Z^2$. There is a sequence $v_1,v_2,\dots,v_{n}$ with $v_j\in V_j$ such that $Q(v,1)\subset \bigcup_{j=2}^n Q(v_j,j-1)$, and the latter union is connected. If we are in the complement of $\ev D$, it follows that the diameter of $Q(v,1)$ is at most $s:=\sum_{j=1}^\infty a_j\delta_j$. Since $s\to 0$ as $\delta_0\to 0$, this establishes the theorem. \Qed Several corollaries follow from this theorem. \procl c.fintree For each $n\in{\N_+}$, a.s.\ $\Tr_0(1/n)$ is a finite topological tree. \endprocl \proof Let $W\subset\R^2$ be finite. For each $w\in W$ and $\delta>0$, let $w_\delta\in\delta\Z^2$ be closest to $w$, with ties broken arbitrarily, and set $W_\delta=\{w_\delta:\, w\in W\}$. Let $Q_\delta(W)$ be the subtree of ${\hat \T}_\delta$ spanned by $W_\delta$. The theorem shows that we may choose a finite $W\subset\R^2$ such that $\Tr_\delta(1/n)\subset Q_\delta(W)$ with probability at least $1-\epsilon$, for every sufficiently small $\delta>0$. Consequently, we may couple a subsequential scaling limit $Q(W)$ of $Q_\delta(W)$ as $\delta\to0$ so that $\Tr_0(1/n)\subset Q(W)$ with probability at least $1-\epsilon$. Because $\Tr_0(1/n)$ is connected and $\epsilon$ is an arbitrary positive number, it suffices to prove that $Q(W)$ is a.s.\ a finite tree. The latter is easily proved by induction on $|W|$ using \ref t.noloop/, Wilson's algorithm, and the following easy fact: the tree spanned by a subset of the points in $W$ is unlikely to pass close by to the other points. (See \ref r.closetopt/.) \Qed \procl c.ustdim The Hausdorff dimension of $\Tr$ is in $(1,2)$. Moreover, if $I=(s_0,s_1]$ is an interval such that a.s.\ the Hausdorff dimension of any scaling limit of LERW is in $I$, then the Hausdorff dimension of $\Tr(\ST)$ is in $I$. \endprocl \proof The second statement follows immediately from \ref t.fin/. The first is now a consequence of the result of \ref b.ABNW/, showing that there are $s_0,s_1\in(0,1)$ such that a.s.\ the Hausdorff dimension of LERW scaling limit is in $[s_0,s_1]$.% \Ftnote{From \ref r.closetopt/ follows the weaker result that the area measure of any subsequential scaling limit of LERW is zero, hence that the area of $\Tr$ is zero. It is likely that with a bit more effort the proof of \ref r.closetopt/ is sufficient for the stronger claim that the Hausdorff dimension is smaller than $2$.} \Qed \procl r.dimr The above-mentioned lower bound in \ref b.ABNW/ is based on the ideas of \ref b.BJPP/. Kenyon~\ref b.Kenyon:inprep/ can prove that we may take $s_1=5/4$. In earlier work~\ref b.Kenyon:5o4/ he showed that $n^{5/4}$ times the expected number of edges in a LERW from $(0,0)$ to the boundary of the square $[-n,n]^2$ tends to a finite positive constant as $n\to \infty$. This supports the conjecture that the Hausdorff dimension of the scaling limit of LERW is a.s.\ $5/4$, and the same would apply to $\Tr(\ST)$. \endprocl The {\bf degree} of a point $p$ in a topological tree $T$ is the number of connected components of $T-\{p\}$. The following corollary is a strong form of the statement that the maximum degree of points in $\Tr_0(1/n)$ is $3$. {} From this and the fact that $\Tr$ is a tree (which we prove further below) it immediately follows that the maximum degree in $\Tr$ is $3$, because every finite subset of $\Tr$ is contained in some $\Tr_0(1/n)$. Given a point $p\in\Sp2$ and two numbers $0<r_1<r_2<1$, let ${A_{\rm sp}}(p,r_1,r_2)$ denote the annulus with center $p$, inner radius $r_1$, and outer radius $r_2$, in the spherical metric. \procl c.maxdeg Given every $\epsilon\in(0,1)$, there is an $r\in(0,\epsilon)$ with the following property. For every sufficiently small $\delta>0$, the probability that there is a point $p\in \Sp2$ such that there are $4$ disjoint crossings in $\Tr_\delta(\epsilon)$ of the annulus ${A_{\rm sp}}(p,r,\epsilon)$ is at most $\epsilon$. \endprocl By having $4$ disjoint crossings in $\Tr_\delta(\epsilon)$ of an annulus $A$, we mean that there are $4$ disjoint connected subsets of $\Tr_\delta(\epsilon)$ that intersect both boundary components of $A$. Below, \ref c.stmaxdeg/ gives a strengthening of \ref c.maxdeg/. \proof By \ref t.fin/, it is enough to prove the statement with $Q_\delta(W)$ replacing $\Tr_\delta(\epsilon)$, where $W\subset\R^2$ is a set of bounded size, provided that the value of $r$ does not depend on $\delta$. Again, induction on $|W|$ can be used together with Wilson's algorithm. One needs note the following easy facts. The tree $T_{k-1}$ spanned by $k-1$ points of $W$ is unlikely to pass close to the other points of $W$, and when adding a further point, it is unlikely that the attachment point of the new branch on $T_{k-1}$ will be close to another branch point. Also, once a random walk from the new point gets close to $T_{k-1}$ it will hit $T_{k-1}$ close by, with high likelyhood. The easy details are left to the reader. \Qed We now turn to the central issue in the proof of \ref t.ustlim/, which is, \procl t.noint In any subsequential scaling limit of UST in $\Sp2$, a.s.\ the trunk and dual trunk do not intersect. \endprocl \procl l.3d2 Given $\delta,\epsilon>0$, let $T_\delta^3(\epsilon)$ be the set of points of degree $3$ in $\Tr_\delta(\epsilon)$. Let $\d{\Tr}_\delta(\epsilon)$ be the $\epsilon$-trunk of the dual tree $\d{{\hat \T}_\delta}$. Let $D$ be the spherical distance from $T_\delta^3(\epsilon)$ to $\d{\Tr}_\delta(\epsilon)$; that is, the least spherical distance between a point in $T_\delta^3(\epsilon)$ to a point in $\d{\Tr}_\delta(\epsilon)$. Then $\lim_{t\to0}\P[D<t]\to 0$ uniformly in $\delta$. \endprocl The following simple observation is used in the proof. Suppose that we condition on a set of edges $S$ to appear in the UST tree in a planar graph. For the dual tree, this is the same as deleting the edges dual to the edges in $S$. Consequently, one can perform a variation on Wilson's algorithm for a planar graph, where one switches back and forth from building the tree by adding LERW branches and building the dual tree. When building the tree, the LERW acts with the constructed tree as a wired absorbing boundary and the constructed dual tree as a free boundary, and conversely when building the dual tree. \proof We first choose a large but finite collection of points $Q$ in $\delta\Z^2$ so that with high probability the subtree $T$ of ${\hat \T}_\delta$ spanned by $Q$ contains $\Tr_\delta(\epsilon)$ (and $|Q|$ does not depend on $\delta$). This can be done, by \ref t.fin/. Let $\d Q$ be a set of vertices of the dual graph $\d{(\delta\Z^2)}$, such that with high probability the subtree $\d T$ spanned by $\d Q$ in the dual graph contains $\d{\Tr}_\delta(\epsilon)$. Let $a_1,a_2,a_3\in Q$ and $\d a_1,\d a_2\in\d Q$ be distinct points. It suffices to show that the probability that the arc $\beta$ joining $\d a_1$ and $\d a_2$ in $\d T$ comes within distance $t$ of the meeting point $m$ of $a_1,a_2$ and $a_3$ in $T$ goes to zero as $t\to 0$, uniformly in $\delta$. This is easy. We condition on the subtree $T_0$ of $T$ spanned by $a_1,a_2,a_3$. Let $\d z$ be a dual vertex close to $\d a_1$. Then with high probability the dual tree path $\beta$ from $\d z$ to $\d a_1$ has diameter not much larger than the distance from $\d z$ to $\d a_1$. In particular, it does not go close to $T_0$. By the next lemma, conditioned on $\beta$ and $T_0$, the probability that a simple random walk starting at $\d a_2$, with $T_0$ acting as a reflecting boundary, will get to within distance $t$ of $m$ before hitting $\beta$ is as small as we wish. Consequently, the same is true for the loop-erasure of this walk, which can be taken as the path joining $\beta$ and $\d a_2$ in the dual tree. \Qed \procl l.harm Let $D$ be a domain in $\Sp2$ with two boundary components, $B_1,B_2$, and assume that both are not single points. Consider a sequence $\delta_j$, $j\in\N$, of positive numbers tending to zero. Suppose that to each $j\in\N$ there are two connected subgraphs $B_1^j,B_2^j$ of the grid $\delta_j\Z^2$, and that $B_1^j\to B_1$ and $B_2^j\to B_2$ in the Hausdorff metric on compact subsets of $\Sp2$. Let $m\in B_2$ be some point, and for each $t>0$ and $z\in\delta_j\Z^2-B_2^j$, let $h_j(z,t)$ be the probability that simple random walk on $\delta_j\Z^2-B_2^j$ starting at $z$ (with reflecting boundary conditions on $B_2^j$), will get to within distance $t$ of $m$ before hitting $B_1^j$. Let $K\subset\Sp2-B_2$ be compact. Then $$ \lim_{t\to 0\atop j\to\infty} \sup\bigl\{h_j(z,t):\, z\in K\cap \delta_j\Z^2\bigr\}= 0 \,. $$ \endprocl \proof Set $G_j:=\delta_j\Z^2-B_2^j$, let $S(t)$ be the vertices of $G_j$ that are within distance $t$ from $m$, and let $V_j(t)$ be the set of vertices of $G_j-B_1^j-S(t)$. Then $h_j(z,t)$ is discrete-harmonic in $V_j(t)$. Recall that the Dirichlet energy of $h_j(z,t)$ is $\sum \bigl(h_j(z,t)-h_j(z',t)\bigr)^2$, with the sum extending over all edges $[z,z']$ in $G_j$. Let $\lambda=\lambda_j$ be the minimum of $h_j(z,t)$ on $K$, and let $z$ be where the minimum is achieved. Then there is a path $\beta$ from $z$ to $S(t)$ such that $h_j(z,t)\geqslant\lambda$ on $\beta$, by the maximum principle for discrete harmonic functions. Note that one can find a collection of $1/O(\delta_j)$ disjoint paths in $G_j$ which join $B_1^j$ and $\beta$ and each path in the collection has combinatorial length bounded by $O(1)/\delta_j$. The Dirichlet energy of $h_j(z,t)$ restricted to each such path is at least $\lambda/O(\delta_j)$, and therefore the Dirichlet energy of $h_j(z,t)$ is at least $C\lambda$, where $C>0$ is a constant depending only on $K$ and $D$. Let $d_j$ be the distance from $m$ to $B_1^j$. Since $h_j(z,t)$ is harmonic in $V_j(t)$, it minimizes the Dirichlet energy among functions on $G_j$ that are $1$ on $S(t)$ and $0$ on $B_1^j$. Therefore, the Dirichlet energy of $h_j(z,t)$ is at most the Dirichlet energy of the function $f:G_j\to\R$, which is $1$ on $S(t)$, $0$ outside of $S(d_j)$, and equal to $\log \bigl(d_j/|z-m|\bigr)/\log(d_j/t)$ elsewhere, which is $O(1)/\log(d_j/t)$, as $j\to\infty$. This gives, $\lambda_j=O(1)/\log(d_j/t)$, and the lemma follows. \Qed \proofof t.noint Before we go into the actual details, the overall plan of the proof will be given (in a somewhat imprecise manner). Let $t_0>0$. It is not hard to reduce the theorem to the claim that with probability close to $1$ the path $\gamma\subset {\hat \T}_\delta$ which joins two fixed points $a_1,a_2$ does not have points $p$ close to it such that the path $\alpha_p\subset{\hat \T}_\delta$ joining $p$ to $\gamma$ is not contained in a small neighborhood of $\gamma$. Let $Z$ be the set of points $p$ such that $\alpha_p$ does not stay close to $\gamma$. When we condition on $\gamma$, The probability that $p\in Z$ goes to zero as $p$ tends to a point in $\gamma$, by a simple harmonic measure estimate. However, this is not enough, since there are many different $p$'s close to $\gamma$. We fix some collection $L_1$ of points close to $\gamma$, and take a thick collection of points $L_2$ which are much closer to $\gamma$. What we show is that conditioned on $\gamma$ and on $Z\cap L_2\neq\emptyset$, the expectation of $N:=|Z\cap L_1|$ is much larger than $\E[N\mid\gamma]$. This is established by observing that when $p\in L_1\cap Z$ is appropriately chosen, the expected number of points $p'\in L_2$ such that $\alpha_{p'}$ is contained in $\alpha_p$, except for a small initial segment of $\alpha_{p'}$, is quite large. It follows that $$ \P[Z\cap L_2\neq\emptyset\mid\gamma]\leqslant {\E[N\mid\gamma]\over \E[N\mid\gamma,\ Z\cap L_1\neq\emptyset]} $$ is small, which suffices to prove the theorem. We now give the details. Fix four distinct points $a_1,a_2,b_1,b_2\in\R^2$. Given $\delta>0$, let $a_1'$ and $a_2'$ be points of $\delta\Z^2$ that are closest to $a_1$ and $a_2$, respetively, and let $b_1'$ and $b'_2$ be vertices of the grid dual to $\delta\Z^2$ that are closest to $b_1$ and $b_2$, respectively. Let $\gamma$ be the path in ${\hat \T}_\delta$ that joins $a_1'$ and $a_2'$, and given any $p\in\delta\Z^2$, let $\alpha_p$ denote the path in ${\hat \T}_\delta$ from $p$ to $\gamma$. Let $\beta$ be the path of $\d{{\hat \T}_\delta}$ that joins $b_1'$ and $b_2'$. Since $\Tr=\bigcup_n\Tr_0(1/n)$, and $\d\Tr= \bigcup_n\d\Tr_0(1/n)$, \ref t.fin/ shows that it suffices to prove that the probability that the distance between $\gamma$ and $\beta$ is less than $t$ goes to zero, as $t$ goes down to zero, uniformly in $\delta$. We know that with probability close to one, $\beta$ does not come close to $\{a_1,a_2\}$, and $\gamma$ does not come close to $\{b_1,b_2\}$ \ref r.closetopt/. Therefore, we need only consider the situation where there is a point $q$ on $\gamma$, which is close to $\beta$, but not close to $\{a_1,a_2,b_1,b_2\}$. Since $\beta$ and $\gamma$ cannot cross, and since $\beta$ locally separates the sphere near every point of $\beta-\{b_1',b_2'\}$, such a situation implies that there is a point $q'$ in $\delta\Z^2$, which is near $q$, but in order to get to $\gamma$ from $q'$ one must either cross $\beta$, or go ``around'' it. Consequently, $\operatorname{diam}\left(\alpha_{q'}\right)$ must be bounded away from zero, as $\alpha_{q'}$ cannot cross $\beta$. It therefore suffices to rule out the existence of a point $q'\in\delta\Z^2$ close to $\gamma$ but with $\operatorname{diam}\left(\alpha_{q'}\right)$ bounded away from zero. More precisely, let $K$ be a compact set disjoint from $\{a_1,a_2\}$, and let $\epsilon_1\in(0,1)$. Let $\tilde h$ be the least distance from $\gamma$ to some point $q'\in K\cap \delta\Z^2$ such that $\operatorname{diam}\left (\alpha_{q'}\right)\geqslant\epsilon_1$. It suffices to show that $$ \inf_{h_0>0}\limsup_{\delta\to0}\P[\tilde h<h_0]=0 \,. \label e.targg $$ Given any $p\in\delta\Z^2$, let $h(p)$ be the distance from $p$ to $\gamma$ and let $k(p)$ be the maximal distance from a point on $\alpha_p$ to $\gamma$. By \ref c.maxdeg/, the probability that there is an arc $\alpha$ in ${\hat \T}_\delta$, which is disjoint from $\gamma$, satisfies $\operatorname{diam}(\alpha)\geqslant\epsilon_1$, and every point of $\alpha$ is within distance $t$ of $\gamma$, goes to zero as $t\to0$, uniformly in $\delta$. Hence, to prove \ref e.targg/, it suffices to establish that $$ \forall t>0\qquad \inf_{h_0>0}\limsup_{\delta\to 0} \PBig{\exists p\in K\cap\delta\Z^2\,\,\,h(p)<h_0,\,k(p)\geqslant t}=0 \,. \label e.targa $ Since the proof is somewhat involved, we consider first the simpler situation in which $$ \gamma=[0,1]\times \{0\},\hbox{ and } K=[1/3,2/3]\times[0,1] \label e.simp $$ (notwithstanding that this is an unrealistic situation, of extremely low probability). Obviously, it suffices to prove \ref e.targa/ for small $t>0$. In the following arguments, several small positive quantities appear. Their dependence differs from the natural flow of the proof. In order to make it clear that the proof is logically sound, we state now that the dependence order is as follows: $$ \epsilon_0,t_0,r_0,t_1,h_1,r_1,h_2,\delta\,; $$ that is, each of these quantities may depend only on those appearing before it in the list, and should be thought of as much smaller than it predecessors. Set $h_1':=\max\{k\delta:\, k\in\Z,\ k\delta\leqslant h_1\}$, and let $L_1:=\bigl\{(k\delta,h_1'):\, k\in\Z,\ 1/6<k\delta<5/6\bigr\}$. Given $\gamma$ and $p\in\delta\Z^2$, we may choose $\alpha_p$ by loop-erasing a simple random walk from $p$ to $\gamma$. Consequently, an easy harmonic measure estimate shows that $\P[k(p)>t]=O(h_1/t)$ for all $p\in L_1$. Set $h_2':=\max\{k\delta:\, k\in\Z,\ k\delta\leqslant h_2\}$, and $L_2:=\bigl\{(k\delta,h_2'):\, k\in\Z,\ 1/4<k\delta<3/4\bigr\}$. Again, for all $p\in L_2$, $\P[k(p)\geqslant t_0]=O(h_2/t_0)$, so we may assume that the event $\ev Q$ that the leftmost point in $L_2$ satisfies $k(p)< t_0$ has probability at least $1-\epsilon_0$. Let $\ev K$ be the event that there is some $p\in L_2$ with $k(p)\geqslant t_0$, and let $\ev K':= \ev K\cap\ev Q$. For proving \ref e.targa/ in the simpler situation \ref e.simp/, it suffices to show that $\P[\ev K']=O(\epsilon_0)$ for all sufficiently small $\delta>0$. Consider the following procedure for generating ${\hat \T}_\delta$ given $\gamma$. Perform Wilson's algorithm starting with the vertices in $L_2$, in left-to-right order. If we encounter in this procedure some vertex $p\in L_2$ such that $k(p)\geqslant t_0$, we stop, and let $p_0$ denote that vertex. Let $T_0$ be the tree constructed up to that point (including $\alpha_{p_0}$). On the event $\ev K'$, let $p_1$ be the first point on $\alpha_{p_0}$ whose distance to $\gamma$ is at least $t_0$, and let $\alpha$ be the arc of $\alpha_{p_0}$ from $p_0$ to $p_1$. Let $\ev A_1$ be the event that $\alpha$ is not contained in the rectangle $[1/5,4/5]\times [0,t_0]$. Note that $\ev A_1$ implies that there is an arc in $\alpha\cap ([1/5,4/5]\times [0,t_0])$ with diameter at least $1/15$. By considering this arc and $\gamma$, \ref c.maxdeg/ shows that $\P[\ev A_1\cap\ev K']<\epsilon_0$, assuming that $t_0$ is sufficiently small. On the event $\ev K'-\ev A_1$, let $$ x_1:=\max\bigl\{x\in \R:\, (x,h_1/2)\in\alpha\bigr\} \,, $$ let $U$ be the component of $\R^2-\Bigl(\alpha\cup\bigl([x_1,\infty)\times\{h_1/2\}\bigr)\cup \bigl(\R\times\{t_1\}\bigr)\Bigr)$ that contains $(x_1,\infty)\times\{h_1/2\}$, and let $U'$ be the set of point in $U$ that are within distance $t_1$ from $\alpha$. See \ref f.uprime/. Let $\ev A_2$ be the event that $\ev K'-\ev A_1$ occurs and $U'$ intersects $T_0$. \@midtrue\@ins \SetLabels {\ss B}(.1*.19)$T_0$\\ \L{\ss T}(.26*.09)$p_0$\\ \R(.13*.5)$\alpha$\\ \E\L(1.01*.35)$h_1/2$\\ \E\L(1.01*.02)$0$\\ \E\L(1.01*.89)$t_1$\\ {\ss T}(.47*-.01)$x_1$\\ {\ss T}(.47*.7)$U'$\\ \endSetLabels \centerline{% \AffixLabels{% \epsfysize=2.2in\epsfbox{uprime.eps}}} \medskip \caption{\figlabel{uprime } \endinsert We now prove that $\P[\ev K-\ev A_1-\ev A_2]=O(\epsilon_0)$. Let $N$ be the number of points $p\in L_1$ such that $k(p)\geqslant t_0$. For a given $p\in L_1$, the probability of $k(p)\geqslant t_0$ (given \ref e.simp/, but otherwise unconditioned) is $O(h_1/t_0)$. Therefore, $$ \E[N]\leqslant O(1)h_1/(\delta t_0) ,. \label e.enup $$ On the other hand, condition on the event $\ev K'-\ev A_1-\ev A_2$ and on $T_0$. Let $L_1'$ be the set of $p\in L_1\cap U$ such that the distance from $p$ to $\alpha$ is at most $t_1/2$. Note that conditioned on $p\in L_1'$, the probability that $\alpha_p$ joins with $\alpha_{p_0}$ within distance $2t_1$ from $p$ is at least $$ O(1)^{-1}{h_1\over {\rm dist}(p,\alpha)+h_1} \,, \label e.lowest $$ since after generating $T_0$, we may continue by running Wilson's algorithm starting at $p$, and the probability that the random walk starting at $p$ will hit $\alpha$ before the ray $(x_1,\infty)\times\{h_1/2\}$ is at least \ref e.lowest/. It therefore follows that conditioned on $\ev K-\ev A_1-\ev A_2$, we have \begineqalno \E[N\mid \ev K-\ev A_1-\ev A_2] & \geqslant O(1)^{-1}\sum\bigl\{h_1/(k\delta):\, k\in\Z,\ h_1\leqslant k\delta \leqslant t_1/2\bigr\} \cr & \geqslant O(1)^{-1}\delta^{-1}h_1 \log\bigl(t_1/(2 h_1)\bigr) \,. \endeqalno Combining this with \ref e.enup/ gives $$ \P[ \ev K-\ev A_1-\ev A_2] \leqslant {\E[N]\over \E[N\mid \ev K-\ev A_1-\ev A_2]} \leqslant O(1)\Bigl( t_0 \log\bigl(t_1/(2 h_1)\bigr)\Bigr)^{-1} \leqslant \epsilon_0 \,, $$ provided that $h_1$ is sufficiently small. It remains to establish that $\P[\ev A_2]\leqslant O(\epsilon_0)$. First consider the case that there is some $p\in L_2$, to the left of $p_0$, such that $\alpha_p\cap U\neq\emptyset$. Then this must be the case for $p$ being the left neighbor of $p_0$, namely $p=p_0':=p_0-(\delta,0)$, because when $p\in L_2$ is to the left of $p_0'$, the path $\alpha_p$ cannot cross $\alpha_{p_0}\cup[p_0,p_0']\cup\alpha_{p_0'}$, and $\alpha_p$ does not get to $\R\times\{t_0\}$. If $\alpha_{p_0'}$ intersects $U$, then $\alpha_{p_0'}$ must first get to some point $z$ in $[x_1,\infty)\times\{h_1/2\}$. Near $z$ there must be two points $\d z_1,\d z_2$, which are vertices of the dual grid $\d{(\delta\Z^2)}$, and are locally separated from each other by $\alpha_{p_0}$. The path in the dual tree that joins $\d z_1$ and $\d z_2$ has to contain the edge dual to the edge $[p_0',p_0]$. Now consider another dual vertex $\d z_3$ just left of $\alpha$ near the point $(x_1,h_1/2)$. Let $\d m$ be the meeting point of $\d z_1,\d z_2$ and $\d z_3$ in the dual tree. If $\d m$ is not within distance $r_1$ of $p_0$, we get in the $r_1/10$ trunk of the dual tree at least four disjoint crossings of the annulus $A(p_0,2h_2,r_1/2)$. We may assume that this has probability $\leqslant \epsilon_0$, by \ref c.maxdeg/. Similarly, \ref l.3d2/, with the role of the tree and dual tree reversed, shows that we may take the event that $\d m$ is within distance $r_1$ of $p_0$ to have probability $\leqslant \epsilon_0$, provided that $r_1$ is sufficiently small when compared with $h_1$. To establish that $\P[\ev A_2]\leqslant O(\epsilon_0)$, it now suffices to prove that on the event $\ev K'-\ev A_1$, the probability that $\alpha_{p_0}$ intersects $U'$ is $O(\epsilon_0)$. The argument is similar here, but occurs on a larger scale. Suppose that $w\in\alpha_{p_0}\cap U'$. Then $w$ must be on the segment of $\alpha_{p_0}$ from $p_1$ to the point $p_2$ in $\alpha_{p_0}\cap\gamma$. Because $w\in\alpha_{p_0}-\alpha$ and $w$ is within distance $t_1$ to $\alpha$, there is a point near $w$ that is in the $(t_0/2)$-trunk of the dual tree; namely, some point on the path connecting dual two vertices on opposite sides of $\alpha_{p_0}$ near $p_1$. Consequently, by \ref l.3d2/, we may rule out the possibility that $p_2$ is within distance $r_0$ of $w$ as having small probability, since $p_2$ is a point of degree $3$ of the $(t_0/2)$-trunk. But if the distance between $p_2$ and $w$ is more than $r_0$, then there are in the $(t_0/2)$-trunk at least four disjoint crossings of the annulus $A(w,2t_1,r_0)$: two on $\alpha_{p_0}$ and two on $\gamma$. An appeal to \ref c.maxdeg/ now establishes $\P[\ev A_2]\leqslant O(\epsilon_0)$. This completes the proof in the situation \ref e.simp/. We now explain how to modify the above proof to deal with the general case. First note that the restriction on $K$ is entirely inconsequential; we could in the same way deal with any compact set disjoint from the endpoints of $\gamma$. More significant is the special selection of $\gamma$. Observe that under the assumption of conformal invariance of the LERW scaling limit, the general case can be reduced to the case where $\gamma=[0,1]\times \{0\}$, because after $\gamma$ is generated, the rest of the UST is just unconditioned UST on the complement of $\gamma$ with wired boundary conditions. We may then transform $\gamma$ by a conformal homeomorphism to $[0,1]\times\{0\}$, and refer to the above result. Although we do not assume conformal invariance of the scaling limit, it turns out that the proof above is itself conformally invariant. With some care, one can apply the conformal map to the proof, in a manner of speaking. This is actually not very surprising, because the proof is ultimately based on a simple (discrete) harmonic measure estimate, which is conformally invariant. Let us turn to the details. We may couple the UST for a subsequence of $\delta$ tending to zero so that $\gamma$ tends to some path $\gamma_0$ as $\delta\to 0$ along that subsequence (see the discussion of the Prohorov metric in \ref s.back/). Let $f_\delta:\Sp2-([0,1]\times\{0\})\to\Sp2-\gamma$ be the conformal map normalized to take the endpoints of $[0,1]\times\{0\}$ to the endpoints of $\gamma$ and so that $f_\delta(\infty)$ is on the line which is the set of points at equal distance from both endpoints of $\gamma$, say. (The latter normalization is necessary to make $f_\delta$ unique, but otherwise, it is quite arbitrary.) It follows that $f_\delta$ tends to the similarly normalized conformal map $f:\Sp2-([0,1]\times\{0\})\to\Sp2-\gamma_0$. We may assume that $\delta$ is so small that $f$ and $f_\delta$ are very close on compact subsets disjoint from $[0,1]\times\{0\}$. For each $p\in L_1$, where $L_1$ is as before, we let $\widehat p$ denote a point in $\delta\Z^2$ that is closest to $f(p)$. For the general case, we consider $\widehat N$, the number of $p\in L_1$ such that $k(\widehat p)$ is not small, in place of $N$. Let $\widehat L_2$ denote the set of points in $\delta\Z^2$ that are within distance $3\delta$ of $f\bigl([1/4,3/4]\times \{h_2\}\bigr)$. The proof for the general case uses $\widehat L_2$ in place of $L_2$. For traversing $\widehat L_1$, there is no clear notion of the left-right order. But any ordering that starts near $f\bigl((1/4,h_2)\bigr)$ and later does not visit any vertex before visiting an immediate neighbor, will do. Instead of the left neighbor $p_0'$ of a vertex $p_0\in L_1$, we use for $\widehat p_0\in \widehat L_1$ that neighbor of $\widehat p_0$ in $\widehat L_1$ that is ``most counterclockwise'', in the appropriate sense. The rest of the proof proceeds with essentially no modifications, except that the coordinate system used is transformed by $f$. \Qed \procl r.usfconj In \ref b.BLPS:usf/ it has been asked whether the free USF on every planar proper bounded degree graph is a tree. The proof of \ref t.noint/ seems to be relevant. It is plausible that with a similar argument one can prove that for proper planar graphs with bounded degree and a bounded number of sides per face, the free USF is a tree. \endprocl We may now strengthen \ref c.maxdeg/, as follows \procl c.stmaxdeg Given every $\epsilon\in(0,1)$, there is an $r\in(0,\epsilon)$ with the following property. For every sufficiently small $\delta>0$, the probability that there is a point $p\in \Sp2$ such that there are $4$ disjoint crossings in ${\hat \T}_\delta$ of the annulus ${A_{\rm sp}}(p,r,\epsilon)$ is at most $\epsilon$. \endprocl \proof Consider an annulus $A={A_{\rm sp}}(p,r,\epsilon)$, and suppose that there are four disjoint paths $\alpha_0,\alpha_1,\alpha_2,\alpha_3$ in ${\hat \T}_\delta$ that cross it. Let $B_1$ be the component of $\Sp2-A$ inside the inner boundary component of $A$, and let $B_2$ be the outside component. Without loss of generality, we suppose that $\alpha_0\cup\alpha_2$ separate $\alpha_1$ from $\alpha_3$ inside $A$; that is, the circular order of these paths around $A$ agrees with the order of the indices. Assume first that there are no paths in ${\hat \T}_\delta\cap A$ that join two of the paths $\alpha_j$, $j=0,1,2,3$. Then there must be paths in the dual tree $\beta_0,\beta_1,\beta_2,\beta_3$, such that $\beta_j$ is between $\alpha_{j-1}$ and $\alpha_j$ (indices mod $4$), for each $j=0,1,2,3$. If $\alpha_0,\alpha_1,\alpha_2$ and $\alpha_3$ can all be connected to each other by paths in $A\cup B_1$, it follows that there are four crossings of the annulus ${A_{\rm sp}}\left(p,r,\epsilon/3\right)$ in the $\epsilon/3$ trunk of ${\hat \T}_\delta$, and we know that has small probability to happen anywhere, if $r$ is small, by \ref c.maxdeg/. If neither of the paths $\alpha_j$ connects to another in $A\cup B_1$, the same argument applies to the dual tree, because the paths $\beta_j$ must all connect inside $A\cup B_1$. However, if two of the paths $\alpha_j$ connect in $A\cup B_1$, and one of the others does not connect to them, then also two of the paths $\beta_j$ connect. This implies that the $\epsilon/3$ trunk gets within distance of $r$ from the dual trunk, and again this can be discarded as having small likelyhood. We are left to deal with the situation where there is a simple path $\gamma$ in $A\cap{\hat \T}_\delta$ that connects two of the paths $\alpha_j$. Note that for each pair of paths $\alpha_j$ there can be at most one such $\gamma$ connecting them. Also note that any path connecting $\alpha_j$ and $\alpha_{j+2}$ (indices mod $4$) must cross either $\alpha_{j+1}$ or $\alpha_{j+3}$. Consequently, if we consider any four concentric annuli $A_j:={A_{\rm sp}}(p,r_j,r_{j+1})$, $j=0,1,2,3,4$, with $r_j<r_{j+1}$ for each $j=0,1,2,3,4$, $r_0=r$, and $r_5=\epsilon$, at least one of them will have the property that inside it there is no path joining any two paths among the $\alpha_j$'s. This allows a reduction to the previous case, and completes the proof. \Qed \procl t.trpaths A.s., every simple path $\phi:[0,1)\to\Tr$ has a limit $\lim_{t\to1}\phi(t)$ in $\Sp2$, and for every point $z\in\Sp2$ there is a simple path $\phi:[0,1)\to\Tr$ such that $\lim_{t\to1}\phi(t)=z$. \endprocl \proof Suppose that there are two distinct accumulation points, $x$ and $y$, of $\phi(t)$ as $t\to 1$, and let $m\in{\N_+}$ satisfy $\dsp(x,y)>9/m$. Then for each $t\in (0,1)$ the (spherical) diameter of $\phi\bigl([t,1)\bigr)$ is greater than $9/m$. Let $a\in(0,1)$ be such that the diameter of $\phi\bigl([0,a])$ is at least $5/m$. It easily follows that $\phi\bigl([a,1)\bigr)\subset\Tr_0(1/m)$. But since $\Tr_0(1/m)$ is a compact finite tree (\ref c.fintree/), and the restriction of $\phi$ to $[a,1)$ is in $\Tr_0(1/m)$, it follows that $\lim_{t\to 1}\phi(t)$ exists. Contradiction. Let $z\in\Sp2$ and $z'\in\Tr_0(1)$, $z'\neq z$. We want to produce a simple path $\gamma\subset\Tr$ starting at $z'$ and tending to $z$. If $z\in\Tr$, then $z\in\Tr_0(1/n)$ for some $n\in{\N_+}$, and the existence of $\gamma$ is clear. So suppose that $z\notin\Tr$. For each $n\in{\N_+}$, there is a point $z_n\in\Tr_0(1/n)$ which is within distance $2/n$ from $z$. Let $\beta_n$ be the arc from $z'$ to $z_{n}$ in $\Tr_0(1/n)$. For each $n$ and $m$, the intersection $\gamma_n\cap\Tr_0(1/m)$ is a simple path. Since $\Tr_0(1/m)$ is a compact finite topological tree, there is a subsequence $\gamma_{n_j}$ such that for each $m$ the Hausdorff limit $\lim_j\bigl(\gamma_{n_j}\cap\Tr_0(1/m)\bigr)$ exists, and is a simple path. Because $\gamma_n-B(z,3/m)\subset \Tr_0(1/m)$ when $n\geqslant m$, it now follows that $\gamma:=\lim_j\gamma_{n_j}$ is a simple path. Moreover, it is clear that $z\in\gamma$ and $\gamma-\{z\}\subset\Tr$. \Qed \proofof t.ustlim We first prove that $\Tr$ is a topological tree. Clearly, the trunk is arcwise connected, since $\Tr:=\bigcup_n\Tr_0(1/n)$, and each $\Tr_0(1/n)$ is arcwise connected. It is also clear that the trunk is dense in $\Sp2$. Let $x,y\in\Tr$. Then there is some $n\in\N$ such that $x,y\in\Tr_0(1/n)$, and there is a unique arc $\gamma_1$ joining $x$ and $y$ in $\Tr_0(1/n)$. Let $\gamma_2$ be an arc joining $x$ and $y$ in $\Tr$. Since the dual trunk is dense, it must intersect all connected components of $\Sp2-(\gamma_1\cup\gamma_2)$. Since the dual trunk is disjoint from the trunk, it does not intersect $\gamma_1\cup\gamma_2$. Because the dual trunk is connected, it now follows that $\Sp2-(\gamma_1\cup\gamma_2)$ is connected. Consequently, $\gamma_1=\gamma_2$, and the trunk is uniquely arcwise connected. Let $n\in\N$ and let $t$ be the spherical distance between $\Tr_0(1/n)$ and $\d {\Tr}_0(1/n)$. Since these are compact and disjoint, $t>0$. If $x\in\Tr_0(1/n)$ and there is a $y\in\Tr$ such that there is no path in $\Tr\cap B(x,3/n)$ joining $x$ and $y$, then there must be a path in $\d{\Tr}_0(1/n)$ separating $x$ and $y$ in $B(x,1/n)$. (Indeed, if $\gamma$ is the path joining $x$ and $y$ and $p\in\gamma- B(x,3/n)$, then the path $\beta\subset\d\Tr$ connecting two points $p'$ and $p''$ that are near $p$ and are separated from each other by $\gamma$ near $p$, will have a subarc in $\d{\Tr}_0(1/n)$ separating $x$ and $y$ in $B(x,1/n)$.) Consequently, for every point $x\in\Tr_0(1/n)$ and every $y\in\Tr\cap B(x,t)\cap B(x,1/n)$ there is a path in $\Tr\cap B(x,3/n)$ joining $x$ and $y$. Hence, the union of all arcs that contain $x$ and are contained in $\Tr\cap B(x,3/n)$ is an arcwise connected subset of $\Tr\cap B(x,3/n)$ which contains $\Tr\cap B(x,t)\cap B(x,1/n)$. This implies that $\Tr$ is locally arcwise connected, and so it is a topological tree. It is obviously dense in $\Sp2$, and the proof of part (iii) is complete. It is clear that for every $a,b\in\Sp2$ there is some $\omega$ such that $(a,b,\omega)\in\ST$. Let $\ST(\epsilon)$ be a subsequential scaling limit of $\ST_\delta(\epsilon)$. We prove that a.s.\ every $\omega$ such that $(a,b,\omega)\in\ST(\epsilon)$ for some $a,b\in\Sp2$ is a simple path. Let $\epsilon_1>0$. It suffices to prove that the above statement holds with probability at least $1-\epsilon_1$. Let $V_1\subset\Sp2$ be a finite set of points, and for each $\delta>0$ let $V_1^\delta$ be a collection of vertices of $\delta\Z^2$, each close to one point of $V_1$, and with $|V_1|=|V_1^\delta|$. By \ref t.fin/, $V_1$ may be chosen so that with probability at least $1-\epsilon_1$ the subtree of ${\hat \T}_\delta$ spanned by $V_1^\delta$ contains $\Tr_\delta(\epsilon)$, for all sufficiently small $\delta$. This implies that each $\omega_{a,b}(\epsilon)$ is a subarc of $\omega_{v,u}$ for some $v,u\in V_1^\delta$. Because for every pair of points $v,u\in V_1$ the scaling limit of the LERW from $v$ to $u$ is a simple path, it follows that with probability at least $1-\epsilon_1$ for each $(a,b,\omega)\in\ST(\epsilon)$, $\omega$ is a simple path. We may now conclude that a.s.\ for every $(a,b,\omega)\in\ST$, the set $\omega-\bigl(\overline B(a,\epsilon)\cup\overline B(b,\epsilon)\bigr)$ is a $1$-manifold; that is, a disjoint union of simple paths. Therefore, $\omega':=\omega-\{a,b\}$ is a $1$-manifold. This means that each component of $\omega'$ is an arc with endpoints in $\{a,b\}$. It is clear that $\omega$ may be oriented as a path from $a$ to $b$. Suppose that $\omega$, visits $a$ more than once. If $a\neq b$, it then follows that $a\in\Tr$, and there is a simple closed path in $\Tr$ containing $a$. This is impossible, since $\Tr$ is a topological tree. Hence $a$, and similarly $b$, are each visited only once in $\omega$, which implies that $\omega$ is a simple path if $a\neq b$. If $a=b$, the only possibility is that $\omega=\{a\}$ or that $\omega$ is a simple closed path. This proves the first and second statements in (ii). Observe that if there is simple curve $\alpha\subset\Tr$ such that $\overline\alpha=\alpha\cup\{a\}$, then $a$ must be in the dual trunk, for the dual trunk is connected, intersects both components of $\Sp2-\overline\alpha$, and is disjoint from $\Tr$. This is a rare event, by \ref r.closetopt/ (or \ref c.ustdim/). This proves (ii). \ref c.stmaxdeg/ proves (iv). The first claim in (i) is obvious. Suppose that $a\neq b$ are such that there are two sets $\omega$ and $\omega'$ with $(a,b,\omega),(a,b,\omega')\in\ST$. We know that $\omega$ and $\omega'$ are simple paths. If $\omega\neq\omega'$, then there is a simple closed path, say $\gamma$, contained in $\overline\omega\cup\overline\omega'$. But as above, $\gamma$ must intersect the dual trunk, since the dual trunk is connected and dense. This implies that $a$ or $b$ are in the dual trunk. This completes the proof of (i), and of the theorem. \Qed \procl r.uniqp \procname{Uniqueness of paths} We have seen in the above proof that the path in $\Tr$ from $a$ to $b$ is unique when $\{a,b\}\cap\d\Tr=\emptyset$. The converse is also easily established. \endprocl \procl r.constructdual \procname{Reconstructing $\d\Tr$} It can be shown that the scaling limit dual trunk can be reconstructed from the trunk. This can be seen from \ref r.uniqp/. Another description of the dual trunk from the trunk is as follows. Given distinct $x,y\in\Sp2-\Tr$, let $\gamma(x,y)$ be the (unique) arc in $\Sp2-\Tr$ with endpoints $x,y$. Then $$ \d\Tr = \bigcup\Bigl\{\gamma(x,y)-\{x,y\}:\, x\neq y,\, x,y\in\Sp2-\Tr\Bigr\} \,. $$ To prove this, it suffices to establish that $\gamma(x,y)$ is unique, which follows from the fact that $\Tr$ is connected and dense. \endprocl \procl r.overtree Consider the metric $d^*$ on $\Tr$, where $d^*(x,y)$ is the spherical diameter of the unique (possibly degenerate) arc joining $x$ and $y$ in $\Tr$. Since $\Tr$ is locally arcwise connected, this new metric on $\Tr$ is compatible with the topology of $\Tr$ as a subset of $\Sp2$. Let $\Tr_*$ denote the completion of this metric. Then $\Tr_*$ is a compact topological tree, and is naturally homeomorphic with the ends compactification of $\Tr$. Since $d^*$ majorizes the spherical metric, there is a natural projection $\pi:\Tr_*\to\Sp2$, whose restriction to $\Tr$ is the identity. It is easy to see that every point $p\in\Sp2-\d\Tr$ has a unique preimage under $\pi$, and for points $p\in\d\Tr$, the degree of $p$ in $\d\Tr$ is equal to $|\pi^{-1}(p)|$. Consider some $o\in\Sp2$, and let $\ST^o$ be the appropriate ``slice'' of $\ST$; that is, $\ST^o:=\bigl\{(b,\omega):\, (o,b,\omega)\in\ST\bigr\}$. One can show that if $o\notin\d\Tr$, then $\ST^o$ is homeomorphic with $\Tr_*$, and $\pi:\ST^o\to\Sp2$ is the projection onto the first coordinate, when $\ST^o$ is identified with $\Tr_*$ through this homeomorphism. \endprocl \bsection{Free and wired trunks and conformal invariance}{s.ustci} \global\advance\consSer by 1\constnum=0\relax We now want to give a precise formulation to a conformal invariance conjecture for the UST scaling limit, and prove that it follows from the conjectured conformal invariance of the LERW scaling limit. (Such conformal invariance conjectures seem to be floating in the air these days, with roots in the physics community.) The conformal automorphisms of $\Sp2$ are M\"obius transformations. We conjecture that the different notions of scaling limits of UST in $\Sp2$, which where introduced in the previous section, exist (without a need to pass to a subsequence) and are invariant under M\"obius transformations. Moreover, the scaling limits in subdomains $D\subset \Sp2$ should be invariant under conformal homeomorphisms $f: D\to D'\subset\Sp2$. This is a significantly stronger statement, since the M\"obius transformations of $\Sp2$ form a $6$-dimensional group, while the space of conformal homeomorphisms from the unit disk onto subdomains of $\Sp2$ is infinite dimensional. To formulate more precisely the invariance under conformal homeomorphisms of subdomains $f:D\to D'$, we need first to discuss UST scaling limits in subdomains of $\Sp2$. This will be now explained. For simplicity, we restrict our attention to simply connected domains $D\subsetneqq\R^2$ whose boundary $\partial D$ is a simple closed path. Let $p_0\in D$ be some basepoint. Let ${\ss FT}_\delta^D$ be the uniform spanning tree of $G(D,\delta)$, with free boundary conditions, and let ${{\ss WT}}_\delta^D$ be the uniform spanning tree of $G(D,\delta)$ with wired boundary conditions. Let ${\SS^2_D}$ be the metric space obtained from $\Sp2$ by contracting $\Sp2-D$ to a single point. Then we may think of ${{\ss WT}}_\delta^D$ as a random point in ${\Cal H}({\SS^2_D})$, which is a.s.\ a tree. The tree ${{\ss WT}}_\delta^D$ may be thought of as a random point in ${\Cal H}(\overline D)$, which is a.s.\ a tree. Let $\WST_\delta^D:=\ST({{\ss WT}}_\delta^D)$ and $\FST_\delta^D:=\ST({\ss FT}_\delta^D)$; that is, the {\bf wired paths ensemble} $\WST_\delta$ is defined from the wired tree ${{\ss WT}}_\delta$ in exactly the same way that the ordinary paths ensemble $\ST_\delta$ was defined from ${\hat \T}_\delta$, and similarly for $\FST_\delta$. Note that $\WST_\delta\in{\Cal H}\bigl({\SS^2_D}\times{\SS^2_D}\times{\Cal H}({\SS^2_D})\bigr)$ and $\FST_\delta\in{\Cal H}\bigl(\overline D\times\overline D\times{\Cal H}(\overline D)\bigr)$. Also the definitions of the scaling limits and the trunk are the same as in the previous section. \procl t.dom Let $D\subset\R^2$ be a domain whose boundary is a $C^1$-smooth simple closed curve. \begingroup\parindent=25pt \itemrm{(i)} \ref t.fin/, with $\Sp2$ replaced by $D$, holds for the free and wired spanning trees in $D$. \itemrm{(ii)} The free scaling limit trunk in $D$ is disjoint from $\partial D$, in every (subsequential) scaling limit. \itemrm{(iii)} The free scaling limit trunk in $D$ is disjoint from the scaling limit trunk of the dual tree (which is wired), in every (subsequential) scaling limit. \endprocl There are simply connected domains where $\partial D$ is not a simple closed curve and (ii) fails: the domain $(0,2)\times(0,2) - \bigcup_{n=1}^\infty (0,1]\times\{1/n\}$ is an example. \procl l.freefin There is an absolute constant $C>0$ such that the following holds true. Let $D$ be as in \ref t.dom/. Then there is a $\delta_0=\delta_0(D)>0$ with the following property. Suppose that $\delta$ and $\delta_1$ are numbers satisfying $0<\delta\leqslant\delta_1\leqslant \delta_0$ and $A$ is a connected subgraph of $G(D,\delta)$ with diameter at least $\delta_1$. Further suppose that $p\in{\ss V}\bigl(G(D,\delta)\bigr)$ has distance $\delta_1$ to $A$. Then the probability that a random walk on $G(D,\delta)$ starting at $p$ will get to distance $C\delta_1$ from $p$ before hitting $A$ is less than $1/2$. \endprocl \proof The proof is similar to the proof of \ref l.harm/. Let $Z=Z(t)$ be the set of vertices with distance at least $t$ from $p$, where we take $t>4\delta_1$ Let $A'$ be a component of $A\cap B(p,3\delta_1)$ containing some point at distance $\delta_1$ to $p$ and having diameter at least $\delta_1$. For $v\in{\ss V}\bigl(G(D,\delta)\bigr)$, let $h(v)$ be the probability that a random walk starting from $v$ will reach $Z$ before hitting $A'$. Then $h$ is discrete-harmonic, and minimizes Dirichlet energy among functions that are $1$ on $Z$ and $0$ on $A'$. As in the proof of \ref l.harm/, it follows that the Dirichlet energy of $h$ is at most $O(1)/\log(t/\delta_1)$. Let $B$ be the set of vertices $v$ at distance at most $2\delta_1$ from $p$ such that $h(v)\geqslant h(p)$, and let $B'$ be the component of $B$ containing $p$. Note that the diameter of $B'$ is at least $\delta_1$, as $B'$ must neighbor with some vertex with distance $2\delta_1$ from $p$, by the maximum principle for $h$. As in the proof of \ref l.harm/, it can be shown that when $\delta_1$ is sufficiently small (how small depends on the scale in which $\partial D$ appears smooth), one can find $O(1)/\delta$ disjoint paths in $\delta\Z^2\cap D$ connecting $A'$ to $B'$, each of combinatorial length $O(1)/\delta$. Because $h$ is zero on $A'$ and at least $h(p)$ on $B'$, it follows that the Dirichlet energy of $h$ is at least $O(1)h(p)$. We conclude that $h(p)\leqslant O(1)/\log(t/\delta_1)$, which proves the lemma. \Qed \proofof t.dom The proof for (i) in the wired case is the same as the proof of \ref t.fin/ (and we don't need to assume anything about $D$). The free case is the same, except that one needs to appeal to \ref l.freefin/. Assuming (ii), the proof of (iii) is identical to the proof of \ref t.noint/. The proof of (ii) is also the same as the proof of \briefref t.noint/, except that one needs to find the appropriate substitutes for \ref l.3d2/ and \ref c.maxdeg/; namely, for every $\epsilon>0$ there is an $\epsilon_0>0$ such that for all $\delta\in(0,\epsilon_0)$ with probability at least $1-\epsilon$ all points of degree three in the $\epsilon$-trunk of ${\ss FT}_\delta^D$ have distance at least $\epsilon_0$ from $\partial D$, and the probability that there is a point $p\in\partial D$ such that there are two disjoint crossings of the annulus $A(p,\epsilon_0,\epsilon)$ in the $\epsilon$-trunk of ${{\ss WT}}_\delta^D$ is at most $\epsilon$. The latter statement follows from the proof of \ref c.maxdeg/. It remains to prove the appropriate substitute for \briefref l.3d2/. Consider three distinct points in $D$, $p_1,p_2,p_3$, and for $\delta>0$ let $p_1',p_2',p_3'$ be a triple of points in $\delta\Z^2$ which is close to $p_1,p_2,p_3$, respectively. Let $m$ be the meeting point of $p_1',p_2',p_3'$ in ${\ss FT}_\delta^D$, and let $B$ be any disk whose center is in $\partial D$ and which does not intersect $\{p_1',p_2',p_3'\}$. Let $B'$ and $B''$ be disks concentric with $B$ of $1/2$ and $1/4$ of its size, respectively. By part (i), it suffices to prove that with probability going to $1$ as $\epsilon_0\to 0$ and $\delta\to 0$, $m$ is not within distance $\epsilon_0$ from $\partial D\cap B''$. Suppose that $m\in B$. Let $\gamma_1,\gamma_2,\gamma_3$ be the arcs of ${\ss FT}_\delta^D$ that join $m$ to $p_1',p_2',p_3'$, respectively, and let $\gamma_j'$ be the largest initial segment of $\gamma_j$ that is contained in $B$, $j=1,2,3$. There is a unique $k\in\{1,2,3\}$ such that $\bigcup_{j\neq k}\gamma_j'$ separates $\gamma_k'$ from $\partial D$ in $B$. By symmetry, it suffices to estimate the probability that $m$ is close to $B\cap\partial D$ and $k=3$. We generate $m$ in the following way. Let $\gamma$ be a LERW from $p_1'$ to $p_2'$ in $\delta\Z^2\cap D$. Let $X$ be a random walk on $\delta\Z^2\cap D$ starting at $p_3'$, let $\tau_\gamma$ be the first time $t$ where $X(t)\in\gamma$, and let $\tau_1$ be the first time $t$ when $X(t)$ is incident with an edge intersecting $\partial D\cap B'$. Then we may take $m=X(\tau_\gamma)$. Consider the event $\ev A$ where $m\in B$, $k=3$ and $\tau_1\geqslant\tau_\gamma$. With high probability, the points $X(t)$, $t\leqslant\tau_1$, which are close to $B'\cap\partial D$, are also close to $X(\tau_1)$. This is just a property of simple random walk absorbed at $B'\cap\partial D$. Consequently, on $\ev A$, with high probability, if $m$ is close to $B\cap\partial D$, then $\gamma$ passes close to $X(\tau_1)$. Since the probability that $\gamma$ passes near any point, which is not too close to $p_1'$ and $p_2'$, is small (\ref r.closetopt/ applies here), and $X(\tau_1)$ is independent from $\gamma$, we see that $\ev A$ has arbitrarily small probability. We now need to consider the case $k=3$ and $\tau_1<\tau_\gamma$. Let $\tau_1'$ be the first $t\geqslant\tau_1$ such that $X(t)\notin B$, and let $\tau_2$ be the first $t\geqslant\tau_1'$ such that $X(t)$ is incident with an edge intersecting $\partial D\cap B'$. Inductively, let $\tau_n'$ be the first $t\geqslant \tau_n$ such that $X(t)\notin B$ and let $\tau_{n+1}$ be the first $t\geqslant\tau_n'$ such that $X(t)$ is incident with an edge intersecting $\partial D\cap B'$. Note that if $k=3$ and $\tau_\gamma>\tau_1$, then $\tau_\gamma>\tau_1'$, for the random walk $X$ must go around $\gamma_1'\cup\gamma_2'$ before hitting $\gamma$. Similarly, if $k=3$ and $\tau_\gamma>\tau_1$, then $\tau_\gamma\in(\tau_n',\tau_{n+1})$ for some $n\in\N$. If we fix a finite $k\in\N$, then the same argument as above shows that with high probability, $\gamma$ does not pass close to the set $\bigl\{X(\tau_n):\, n=1,2,\dots,k\bigr\}$. Because $\P[\tau_k'<\gamma]\to 0$ as $k\to\infty$, uniformly in $\delta$, the required result follows. \Qed Suppose that $D$ and $D'$ are two domains in $\R^2$ such that the boundaries $\partial D,\partial D'$ are simple closed paths in $\R^2$. Then there is a conformal homeomorphism $f:D\to D'$. Moreover, $f$ extends continuously to a homeomorphism of $\overline D$ onto $\overline D'$, which we will also denote by $f$. It follows that $f$ induces maps \begineqalno f_W &:{\Cal H}\bigl({\SS^2_D}\times{\SS^2_D}\times{\Cal H}({\SS^2_D})\bigr)\to {\Cal H}\bigl({\Bbb S}^2_{D'}\times{\Bbb S}^2_{D'}\times{\Cal H}({\Bbb S}^2_{D'})\bigr) \,, \cr f_F &:{\Cal H}\bigl(\overline D\times\overline D\times{\Cal H}(\overline D)\bigr)\to {\Cal H}\bigl(\overline D\times\overline D\times{\Cal H}(\overline D)\bigr) \,. \cr \endeqalno \procl t.confust Let $D\subset\R^2$ be a domain whose boundary is a $C^1$-smooth simple closed path. Assuming \ref g.confinv/, the following is true. \begingroup\parindent=25pt \itemrm{(i)} The free and the wired UST scaling limits, $\FST,\WST$, in $D$ exist. (That is, do not depend on the sequence of $\delta$ tending to $0$.) \itemrm{(ii)} If $f:D\to D'$ is a conformal homeomorphism between such domains, then $f_W$ is measure preserving from the law of $\WST$ in $D$ to the law of $\WST$ in $D'$, and similarly for free boundary conditions. \vskip0pt\endgroup\noindent \endprocl \proof The proof for wired boundary conditions follows from Wilson's algorithm and \ref t.dom/. The easy details are left to the reader. For the free boundary conditions, observe that ${{\ss WT}}_\delta^D$ is dual to ${\ss FT}_\delta^D$ (on the dual grid). \ref r.constructdual/ is also valid in the present setting, and shows that the free scaling limit trunk can be reconstructed from the wired scaling limit trunk. It is easy to see that the free scaling limit $\FST$ can be reconstructed from the trunk. Hence, conformal invariance of the wired UST implies conformal invariance of the free. \Qed \bsection{Speculations about the Peano curve scaling limit}{s.peano} This section will discuss the Peano curve winding between the UST and its dual. {}From here on, the discussion will be somewhat speculative, and we omit proofs, not because the proofs are particularly hard, but because the paper is long enough as it is, and it is not clear when another paper on this subject will be produced. This Peano curve was briefly mentioned in \ref b.BLPS:usf/. Consider the set of points $\theta_\delta\subset\R^2$ which have the same Euclidean distance from ${\hat \T}_\delta$ as from its dual $\d{\hat \T}_\delta$. It is easy to verify that $\theta_\delta$ is a simple path in a square grid $G_P(\delta)$, of mesh $\delta/2$, which visits all the vertices in that grid. Set $\widehat \theta_\delta:=\theta_\delta\cup\{\infty\}$. Then $\widehat\theta_\delta$ is a.s.\ a simple closed path in $\Sp2$ passing through $\infty$. To consider the scaling limit of $\widehat\theta_\delta$, it is no use to think of it as a set of points in $\Sp2$, because then the scaling limit will be all of $\Sp2$. Rather, one needs to parameterize $\widehat\theta_\delta$ in some way. One natural parameterization would be by the area of its $\delta/2$-neighborhood, but there are several other plausible parameterizations. Another, more sophisticated approach, would be to think of $\widehat\theta_\delta$ as defining a circular order on the set $\widehat\theta_\delta$. The circular order $R_\delta$ is a closed subset of $\bigl(\Sp2)^4$, and $(a,b,c,d)\in R_\delta$ iff $a,b,c,d\in\widehat\theta_\delta$ and $\{a,c\}$ separates $b$ from $d$ on $\widehat\theta_\delta$. Then the (subsequential) scaling limit of $\widehat\theta_\delta$ may be taken as the weak limit of the law of $R_\delta$ in ${\Cal H}\Bigl(\bigl(\Sp2\bigr)^4\Bigr)$. Let $\theta$ denote the Peano curve scaling limit, defined as a path, or as a circular order, or some other reasonable definition. Here is what we believe to be a description of $\theta$, in terms of the scaling limit of the UST. Recall that in \ref r.overtree/, we have introduced a completion $\Tr_*$ of the trunk, in the metric $d^*$, where the distance between any two points of $\Tr$ is the diameter of the arc connecting them, and that $\pi:\Tr_*\to\Sp2$ is the natural projection. Consider the joint distribution of $\Tr_*$ and the dual $\d\Tr_*$, and let $\d\pi$ denote the projection $\d\pi:\Tr_*\to\Sp2$. Let $\tilde\theta$ be the set of points $(p,q)\in\Tr_*\times\d\Tr_*$ such that $\pi p=\d\pi q$. Then $\tilde\theta$ is a simple closed path, and the map $\tilde\theta\mapsto\Sp2$ defined by $(p,q)\mapsto \pi p$ gives the scaling limit $\theta$. Fix some $a,b\in\R^2$, $a\neq b$. Let $\omega_{a,b}$ be the path such that $(a,b,\omega_{a,b})\in\ST$ (this $\omega_{a,b}$ is a.s.\ unique), and let $\d\omega_{a,b}$ be such that $(a,b,\d\omega_{a,b})\in\d\ST$. Then $\omega_{a,b}$ and $\d\omega_{a,b}$ are a.s.\ simple paths. Let $D_1$ and $D_2$ be the two components of $\Sp2-(\omega_{a,b}\cup\d\omega_{a,b})$. A.s.\ $\infty\in D_1\cup D_2$, and without loss of generality take $\infty\in D_2$. It is then clear that the part of $\theta$ which is between $a,b$ and does not contain $\infty$ is $\overline D_2$, and that the part which does contain $\infty$ is $\overline D_1$. Suppose that we {\bf condition} on $\omega_{a,b}$ and on $\d\omega_{a,b}$, and look at some point $c\in D_2$. We'd like to know the distribution of the part of $\theta$ between $c$ and $a$ which does not include $b$, say. Recall that on a finite planar graph, we may generate the UST and the dual UST by a modification of Wilson's algorithm, where at each step in which we start from a vertex in the graph, the dual tree built up to that point acts as a free boundary component, and the tree built up to that point acts as an absorbing wired boundary component, and at steps in which we start from a dual vertex, the tree built up to that point acts as a free boundary component, and the dual tree built up to that point acts as an absorbing wired boundary component. Consequently, we let $\alpha$ be the scaling limit of LERW on $\delta\Z^2-\d\omega_{a,b}$ starting at $c$ that stops when it hits $\omega_{a,b}$. Then we let $\d\alpha$ be the scaling limit of LERW on $\delta\Z^2-\bigl(\alpha\cup\omega_{a,b}\bigr)$ starting at $c$ that stops when it hits $\d\omega{a,b}$. (Since $c\in\alpha$, to define this requires taking a limit as the starting point tends to $c$.) Then $\alpha\cup\d\alpha$ separates $D_2$ into two regions, say $D_2'$ and $D_2''$, and if $b\notin\overline D_2'$, then $D_2'$ is the part of $\theta$ ``separated'' from $b$ by $\{a,c\}$. In the above construction, the domain $D_2$ was considered with mixed boundary conditions. One arc of $\partial D_2-\{a,b\}$ was taken as wired, while the other was free. The resulting Peano path scaling limit $\theta$ is a path joining $a$ and $b$ in $\overline D_2$. {} From \ref g.confinv/ should follow a conformal invariance result for UST in such domains with mixed boundary conditions. Therefore, having an understanding of the law of the Peano curve for one triplet $(D,a,b)$, where $a$ and $b$ are distinct points in $\partial D$, which is a simple closed curve, suffices for any other such triplet. This suggests that we should take the simplest possible such configuration; that is, $D={\Bbb H}$, the upper half plane, $a=0,b=\infty$. Suppose that we then take a point $c\in{\Bbb H}$ and condition on the part of $\theta$ between $0$ and $c$ and separated from $\infty$. The effect of that on $\theta$ in the remaining subdomain $D_c$ of ${\Bbb H}$ is all in the boundary $\partial D_c$. This is a kind of Markovian property for the Peano curve, similar to the property given by \ref l.dprod/. By taking the conformal map from $D_c$ to ${\Bbb H}$, which fixes $\infty$, takes $c$ to $0$, and is appropriately normalized at $\infty$, we may return to base one. This suggests that, as we have claimed in the introduction, a representation of the Peano curve scaling limit similar to the SLE representation of the LERW from $0$ to $\partial U$ which we have introduced. The analogue of the L\"owner{} differential equation for this situation is \ref e.loewh/. Due to the Markovian nature of the Peano curve, the corresponding parameter $\zeta$ in \ref e.loewh/ should have the form $\zeta={\ss B}(\kappa t)$, where ${\ss B}$ is Brownian motion on $\R$ starting at $0$, and $\kappa$ is some constant. One can, in fact, show that $\kappa=8$, by deriving an appropriate analogue of Cardy's \ref b.Cardy/ conjectured formula, using the representation \ref e.loewh/ and the techniques of \ref s.crit/. The details will appear elsewhere. \bibsty{mybibstyle} \bibfile{\jobname.bib} \startbib{MMOT92} \bibitem[Aiz]{Aizenman:web} {\sc M.~Aizenman}. \newblock Continuum limits for critical percolation and other stochastic geometric models. \newblock Preprint. \newblock \urlref{http://xxx.lanl.gov/abs/math-ph/9806004}. \bibitem[ABNW]{ABNW} {\sc M.~Aizenman, A.~Burchard, C.~M. Newman, and D.~B. Wilson}. \newblock Scaling limits for minimal and random spanning trees in two dimensions. \newblock Preprint. \newblock \urlref{http://xxx.lanl.gov/abs/math/9809145}. \bibitem[ADA]{ADA} {\sc M.~Aizenman, B.~Duplantier, and A.~Aharony}. \newblock Path crossing exponents and the external perimeter in 2{D} percolation. \newblock Preprint. \newblock \urlref{http://xxx.lanl.gov/abs/cond-mat/9901018}. \bibitem[Ald90]{Aldous:ust} {\sc D.~J. Aldous}. \newblock The random walk construction of uniform spanning trees and uniform labelled trees. \newblock {\em SIAM J. Discrete Math. 3}, 4 (1990), pages 450--465. \bibitem[Ben]{Benjamini:ustsl} {\sc I.~Benjamini}. \newblock Large scale degrees and the number of spanning clusters for the uniform spanning tree. \newblock In M.~Bramson and R.~Durrett, editors, {\em Perplexing Probability Problems: Papers in Honor of Harry Kesten}, Boston. Birkh{\"a}user. \newblock To appear. \bibitem[BLPS98]{BLPS:usf} {\sc I.~Benjamini, R.~Lyons, Y.~Peres, and O.~Schramm}. \newblock Uniform spanning forests. \newblock Preprint. \newblock \urlref{http://www.wisdom.weizmann.ac.il/\string~schramm/papers/usf/}. \bibitem[BJPP97]{BJPP} {\sc C.~J. Bishop, P.~W. Jones, R.~Pemantle, and Y.~Peres}. \newblock The dimension of the {B}rownian frontier is greater than $1$. \newblock {\em J. Funct. Anal. 143}, 2 (1997), pages 309--336. \bibitem[Bow]{Bowditch:treelike} {\sc B.~H. Bowditch}. \newblock Treelike structures arising from continua and convergence groups. \newblock {\em Mem. Amer. Math. Soc.\/}. \newblock To appear. \bibitem[Bro89]{Broder:ust} {\sc A.~Broder}. \newblock Generating random spanning trees. \newblock In {\em Foundations of Computer Science}, pages 442--447, 1989. \bibitem[BP93]{BP:ust} {\sc R.~Burton and R.~Pemantle}. \newblock Local characteristics, entropy and limit theorems for spanning trees and domino tilings via transfer-impedances. \newblock {\em Ann. Probab. 21}, 3 (1993), pages 1329--1371. \bibitem[Car92]{Cardy} {\sc J.~L. Cardy}. \newblock Critical percolation in finite geometries. \newblock {\em J. Phys. A 25}, 4 (1992), pages L201--L206. \bibitem[DD88]{Duplantier:Peano} {\sc B.~Duplantier and F.~David}. \newblock Exact partition functions and correlation functions of multiple {H}amiltonian walks on the {M}anhattan lattice. \newblock {\em J. Statist. Phys. 51}, 3-4 (1988), pages 327--434. \bibitem[Dur83]{Duren:book} {\sc P.~L. Duren}. \newblock {\em Univalent functions}. \newblock Springer-Verlag, New York, 1983. \bibitem[Dur84]{Durrett:BMM} {\sc R.~Durrett}. \newblock {\em Brownian motion and martingales in analysis}. \newblock Wadsworth International Group, Belmont, Calif., 1984. \bibitem[Dur91]{Durrett:probability} {\sc R.~Durrett}. \newblock {\em Probability}. \newblock Wadsworth \& Brooks/Cole Advanced Books \& Software, Pacific Grove, CA, 1991. \bibitem[EK86]{EK:markov} {\sc S.~N. Ethier and T.~G. Kurtz}. \newblock {\em Markov processes}. \newblock John Wiley \& Sons Inc., New York, 1986. \bibitem[Gri89]{Grimmett:book} {\sc G.~Grimmett}. \newblock {\em Percolation}. \newblock Springer-Verlag, New York, 1989. \bibitem[H{\"a}g95]{Hag:ustlim} {\sc O.~H{\"a}ggstr{\"o}m}. \newblock Random-cluster measures and uniform spanning trees. \newblock {\em Stochastic Process. Appl. 59}, 2 (1995), pages 267--275. \bibitem[It{\^o}61]{Ito:lectures} {\sc K.~It{\^o}}. \newblock {\em Lectures on stochastic processes}. \newblock Notes by K.~M.~Rao. \newblock Tata Institute of Fundamental Research, Bombay, 1961. \bibitem[Jan12]{Janiszewski} {\sc Janiszewski}. \newblock {\em J. de l'Ecole Polyt. 16\/} (1912), pages 76--170. \bibitem[Ken98a]{Kenyon:conf} {\sc R.~Kenyon}. \newblock Conformal invariance of domino tiling. \newblock Preprint. \newblock \urlref{http://topo.math.u-psud.fr/\string~kenyon/confinv.ps.Z}. \bibitem[Ken98b]{Kenyon:5o4} {\sc R.~Kenyon}. \newblock The asymptotic determinant of the discrete laplacian. \newblock Preprint. \newblock \urlref{http://topo.math.u-psud.fr/\string~kenyon/asymp.ps.Z}. \bibitem[Ken99]{Kenyon:longrange} {\sc R.~Kenyon}. \newblock Long-range properties of spanning trees. \newblock Preprint. \bibitem[Ken]{Kenyon:inprep} {\sc R.~Kenyon}. \newblock In preparation. \bibitem[Kes87]{Kesten:hitting} {\sc H.~Kesten}. \newblock Hitting probabilities of random walks on $\Z \sp d$. \newblock {\em Stochastic Process. Appl. 25}, 2 (1987), pages 165--184. \bibitem[Kuf47]{Kufarev} {\sc P.~P. Kufarev}. \newblock A remark on integrals of {L}\"owner's equation. \newblock {\em Doklady Akad. Nauk SSSR (N.S.) 57\/} (1947), pages 655--656. \bibitem[LPSA94]{Langlands:Bull} {\sc R.~Langlands, P.~Pouliot, and Y.~Saint-Aubin}. \newblock Conformal invariance in two-dimensional percolation. \newblock {\em Bull. Amer. Math. Soc. (N.S.) 30}, 1 (1994), pages 1--61. \bibitem[Law93]{Lawler:makarov} {\sc G.~F. Lawler}. \newblock A discrete analogue of a theorem of {M}akarov. \newblock {\em Combin. Probab. Comput. 2}, 2 (1993), pages 181--199. \bibitem[Law]{Lawler:survey} {\sc G.~F. Lawler}. \newblock Loop-erased random walk. \newblock In M.~Bramson and R.~Durrett, editors, {\em Perplexing Probability Problems: Papers in Honor of Harry Kesten}, Boston. Birkh{\"a}user. \newblock To appear. \bibitem[L{\"o}w23]{Lowner} {\sc K.~L{\"o}wner}. \newblock Untersuchungen {\"u}ber schlichte konforme abbildungen des einheitskreises, {I}. \newblock {\em Math. Ann. 89\/} (1923), pages 103--121. \bibitem[Lyo98]{Lyons:usfsurvey} {\sc R.~Lyons}. \newblock A bird's-eye view of uniform spanning trees and forests. \newblock In {\em Microsurveys in discrete probability (Princeton, NJ, 1997)}, pages 135--162. Amer. Math. Soc., Providence, RI, 1998. \bibitem[MR]{MR:lowner} {\sc D.~E. Marshall and S.~Rohde}. \newblock In preparation. \bibitem[MMOT92]{MMOT:treechar} {\sc J.~C. Mayer, L.~K. Mohler, L.~G. Oversteegen, and E.~D. Tymchatyn}. \newblock Characterization of separable metric ${{\R}}$-trees. \newblock {\em Proc. Amer. Math. Soc. 115}, 1 (1992), pages 257--264. \bibitem[MO90]{MO:treechar} {\sc J.~C. Mayer and L.~G. Oversteegen}. \newblock A topological characterization of ${{\R}}$-trees. \newblock {\em Trans. Amer. Math. Soc. 320}, 1 (1990), pages 395--415. \bibitem[New92]{Newman:planesets} {\sc M.~H.~A. Newman}. \newblock {\em Elements of the topology of plane sets of points}. \newblock Dover Publications Inc., New York, second edition, 1992. \bibitem[Pem91]{Pemantle:dichot} {\sc R.~Pemantle}. \newblock Choosing a spanning tree for the integer lattice uniformly. \newblock {\em Ann. Probab. 19}, 4 (1991), pages 1559--1574. \bibitem[Pom66]{Pommerenke:Lowner} {\sc C.~Pommerenke}. \newblock On the {L}oewner differential equation. \newblock {\em Michigan Math. J. 13\/} (1966), pages 435--443. \bibitem[Rus78]{Russo} {\sc L.~Russo}. \newblock A note on percolation. \newblock {\em Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 43}, 1 (1978), pages 39--48. \bibitem[Sch]{Schramm:inprep} {\sc O.~Schramm}. \newblock In preparation. \bibitem[Sla94]{Slade:survey} {\sc G.~Slade}. \newblock Self-avoiding walks. \newblock {\em Math. Intelligencer 16}, 1 (1994), pages 29--35. \bibitem[SW78]{SW} {\sc P.~D. Seymour and D.~J.~A. Welsh}. \newblock Percolation probabilities on the square lattice. \newblock {\em Ann. Discrete Math. 3\/} (1978), pages 227--245. \newblock Advances in graph theory (Cambridge Combinatorial Conf., Trinity College, Cambridge, 1977). \bibitem[TW98]{TW} {\sc B.~T{\'o}th and W.~Werner}. \newblock The true self-repelling motion. \newblock {\em Probab. Theory Related Fields 111}, 3 (1998), pages 375--452. \bibitem[Wil96]{Wilson:alg} {\sc D.~B. Wilson}. \newblock Generating random spanning trees more quickly than the cover time. \newblock In {\em Proceedings of the Twenty-eighth Annual ACM Symposium on the Theory of Computing (Philadelphia, PA, 1996)}, pages 296--303, New York, 1996. ACM. \endbib \endreferences \filbreak \begingroup \eightpoint\sc \parindent=0pt \htmlref{http://www.wisdom.weizmann.ac.il/} {Mathematics Department}, \htmlref{http://www.weizmann.ac.il/} {The Weizmann Institute of Science}, Rehovot 76100, Israel \medskip \emailwww{<EMAIL>} {http://www.wisdom.weizmann.ac.il/\string~schramm/} \bye
\section{Introduction} High-T$_c$ cuprates show strong electronic correlations but also a non-neglegible electron-phonon interaction. The latter causes, for instance, the observed isotope effect in the transition temperature T$_c$, especially, away from optimal doping \cite{Frank}. It is thus of interest to calculate superconducting instabilities taking both the electron-electron and the electron-phonon interaction into account. There are several calculations dealing with the case of weak electronic correlations \cite{Liechtenstein,Schuettler,Dahm,Bulut,Nunner,Pao}. In these calculations the phonon-mediated part of the effective interaction is unaffected by electronic correlations. The electronic part to the effective interaction is calculated using a Hubbard model and assuming $U$ to be small compared to the band width. It is repulsive and strongly peaked near the $M$-point in the Brillouin zone. As an alternative, Refs.\cite{Schuettler,Nunner} use the spin-fluctuation term of Ref.\cite{Monthoux}. Since the phonon part is attractive throughout the Brillouin zone it is plausible that the d-wave pairing due to the electronic part will be weakened if the electrons couple strongly to phonons near the $M-$ point and only weakly affected if only phonons with small momentum are involved \cite{Liechtenstein,Dahm,Bulut,Nunner,Jepsen}. Refs. \cite{Dahm,Pao}, for instance, find that the $T_c$ for d-wave superconductivity is always lowered by phonons and that the isotope coefficient $\beta$ is negative for a constant electron-phonon coupling function. If only phonons with small momenta play a role $\beta$ becomes positive\cite{Dahm}. Ref. \cite{Nunner} even finds a change of sign in $\beta$ as a function of the phonon frequency. The case of strong electronic correlations has quite different features compared to the weak correlated case. The purely electronic part of the effective interaction is rather a smooth function of momentum without a sharp peak at the $M$-point due to spin fluctuations \cite{Zeyher1}. Furthermore, electronic correlations modify substantially the phonon-mediated part of the effective interaction \cite{Kulic,Zeyher2,Zeyher3}. It is the purpose of this communication to present results for this case of strong electronic correlations. A suitable model for such calculations is a generalized $t-J$ model which also includes phonons within a Holstein model. Extending the spin degrees from 2 to $N$ systematic approximations for the effective interaction in terms of powers in $1/N$ can be carried out \cite{Kulic,Zeyher2,Zeyher3,Ruckenstein}. In a first step we consider only the phonon-mediated effective interaction which, unlike in the case of weak correlations, is strongly modified by vertex corrections transforming the original constant electron-phonon coupling of the Holstein model into a strongly momentum- and frequency-dependent function. Explicit results for $T_c$ and $\beta$ will be given for the leading symmetry channels of the superconducting order parameter as a function of the doping. In a second step these results are extended to the case where also the purely electronic contribution to the effective interaction is taken into account. \section{Linearized equation for the superconductivity gap} Our Hamiltonian for the $t-J$ model plus phonons can be written as \begin{eqnarray} H = \sum_{{ij} \atop {p=1...N}} {{t_{ij}} \over N} X_i^{p0} X_j^{0p} + \sum_{{ij} \atop {p,q=1...N}} {{J_{ij}} \over {4N}} X_i^{pq} X_j^{qp} &-& \nonumber\\ \sum_{{ij} \atop {p,q=1...N}} {{J_{ij}} \over {4N}} X_i^{pp} X_j^{qq} +\sum_i \omega_0 (a^{\dagger}_i a_i+\frac{1}{2})&+& \nonumber\\ \sum_{i \atop {p=1...N}} \frac{g}{\sqrt{N}} [a^{\dagger}_i+a_i]({X_{i}}^{pp}-<{X_{i}}^{pp}>). \end{eqnarray} The first three terms correspond for $N=2$ to the usual $t-J$ model. For $N=2$ $X$ is identical with Hubbard's projection operator $X_i^{pq}= |{p \atop i}><{q \atop i}|$, where $|{p \atop i}>$ denotes for $p=0$ an empty and for $p=1,2$ a singly occupied state with spin up and down, respectively, at the site $i$. $t_{ij}$ and $J_{ij}$ are hopping and Heisenberg interaction matrix elements between the sites $i,j$. The fourth term in Eq.(1) represents one branch of dispersionless, harmonic phonons with frequency $\omega_0$. The fifth term in Eq.(1) describes a local coupling between the phonon and the change in the electronic density at site $i$ with the coupling constant $g$. $<X>$ denotes the expectation value of $X$. The extension from $N=2$ in Eq.(1) to a general $N$ has been discussed in detail in Ref. \cite{Zeyher2}. The label $p$ runs then not only over the two spin directions but also over $N/2$ identical copies of the orbital. The symmetry group of $H$ is the symplectic group $Sp(N/2)$ which allows to perform $1/N$ expansions for physical observables. In Eq.(1) the electron-phonon coupling is scaled as $1/\sqrt{N}$ whereas the free phonon part is independent of $N$. As a result the leading contributions to superconductivity from the $t-J$ model alone and from the phonons are both of order $O(1/N)$ which allows to treat them on an equal footing. Instabilities towards superconductivity in the $t-J$ model have been studied in the above framework in Refs. \cite{Zeyher1,Greco,Zeyher4}. The contributions to the anomalous self-energy from the phonon-mediated effective interaction have been derived in Refs. \cite{Kulic,Zeyher2} for the case $J_{ij}=0$. Generalizing the latter treatment to a finite $J$ similar as in Ref. \cite{Zeyher3} the linearized equation for the superconducting gap $\Sigma_{an}$ for the entire Hamiltonian Eq.(1) can be written as \begin{eqnarray} {\Sigma}_{an}(k) = -{T \over {NN_c}}\sum_{k'}\Theta(k,k') {1 \over {\omega_{n'}^2 +\epsilon^2({\bf k'})}} {\Sigma}_{an}(k'). \end{eqnarray} $N_c$ is the number of cells and $k$ the supervector $k=(\omega_n,{\bf k})$, where $\omega_n$ denotes a fermionic Matsubara frequency and $\bf k$ the wave vector. $\epsilon({\bf k})$ is the one-particle energy which is unchanged by the phonons in O(1) and thus given by Eq.(41) of Ref. \cite{Zeyher1}. The kernel $\Theta$ in Eq.(2) consists of two parts \begin{eqnarray} \Theta(k,k') = \Theta^{t-J}(k,k')+\Theta^{e-p}(k,k'). \end{eqnarray} The first contribution $\Theta^{t-J}$ in Eq.(3) comes from the $t-J$ model. Explicit expressions for it have been given in Eqs.(42)-(52) in Ref. \cite{Zeyher1}. The second term $\Theta^{e-p}$ in Eq.(3) is due to phonon-mediated interactions and is given by \begin{eqnarray} \Theta^{e-p}(k,k') &=& - \frac{2g^2\omega_0}{(\omega_{n}- \omega_{n'})^2+{\omega_0}^2} \nonumber\\ &\cdot&{\gamma_{c}(k',k-k')}{\gamma_{c}(k,k'-k)}. \end{eqnarray} $\gamma_c$ is the charge vertex for which an explicit expression has been derived in Ref. \cite{Zeyher3}. For uncorrelated electrons there are no vertex corrections, i.e., $\gamma_c=-1$. Since $g$ and $\omega_0$ are assumed to be independent of $\bf k$ $\Theta^{e-p}$ is also independent of $\bf k$ and only s-wave superconductivity is possible. When correlations are present the kernel $\Theta^{e-p}$ depends on $\bf k$ and $\omega_n$ through the charge vertex $\gamma_c$ and general symmetries for the order parameter may become possible. \section{Calculation of $T_c$ and $\beta$ from the phonon-mediated part} Taking a square lattice with point group $C_{4v}$ $\Theta(k,k')$ and $\epsilon({\bf k})$ in Eq.(2) are invariant under $C_{4v}$ which means that $\Sigma_{an}(k)$ can be classified according to the five irreducible representations $\Gamma_i$ of $C_{4v}$. s-wave symmetry corresponds to $\Gamma_1$, d-wave symmetry to $\Gamma_3$, etc. In the weak-coupling case $\Theta(k,k')$ can be approximated by its static limit $\Theta({\bf k},{\bf k'})$. Putting all momenta right onto the Fermi line the sum over ${\bf k'}$ in Eq.(2) can be transformed into a line integral along the Fermi line. Assuming a certain irreducible representation $\Gamma_i$ for the order parameter, the line integral can be restricted to the irreducible Brillouin zone (IBZ) introducing a symmetry-projected kernel $\Theta_i({\bf k},{\bf k'})$ with ${\bf k},{\bf k'} \in$ IBZ. Finally, the line integral was discretized by a set of points $[{\bf k}^F_\alpha]$ along the Fermi line in the IBZ with line elements $[s({\bf k}^F_\alpha)]$. Denoting the smallest eigenvalue of the symmetric matrix \begin{equation} {1 \over {4 \pi^2}} \sqrt{ {s({\bf k}_\alpha^F)s({\bf k}^F_\beta)} \over {|\nabla \epsilon({\bf k}^F_\alpha)| \cdot |\nabla \epsilon({\bf k}^F_\beta)|}} \Theta_i({\bf k}^F_\alpha,{\bf k}^F_\beta) \end{equation} by $\lambda_i$ Eq.(2) yields for $\lambda_i <0$, $N=2$, and in the weak-coupling case the BCS-formula \begin{equation} T_{ci} = 1.13 \omega_0 e^{1/ \lambda_i} . \end{equation} As usual we took $\omega_0$ as a suitable cutoff. If $\lambda_i >0$ we have, of course, $T_{ci}=0$. According to Eq.(6) the absolute value of $\lambda_i$ characterizes the strength of the effective interaction in the symmetry channel $i$. The overall strength of the electron-phonon coupling is conventionally expressed in terms of the dimensionless coupling $\lambda$ defined by \begin{eqnarray} \lambda=\frac{g^2}{8 \omega_0}. \end{eqnarray} In Eq.(7) we have introduced a factor $1/2$ to account for the prefactor $1/2$ in Eq.(2) after setting there $N=2$. We also used the average density of states $1/(8t)$ for the density of states factor in $\lambda$ and put $t$ equal to 1. For the range of dopings we will be interested in, i.e., $0.15 < \delta < 0.8$, the density of states varies only little with doping so that the definition Eq.(7) of $\lambda$ is appropriate for all dopings. \vspace{-0.5cm} \begin{figure} \epsfysize=75mm \centerline{\epsffile{fig1.eps}} \caption{ Phonon contribution to the lowest eigenvalues $\lambda_i$ of the static kernel $\Theta$ for the representations $\Gamma_1$ and $\Gamma_3$ of $C_{4v}$ as a function of the doping $\delta$.} \label{fig:fig1} \end{figure} Fig. 1 shows the eigenvalues $\lambda_i$ for the two leading symmetries $i=1,3$ using the parameter value $J=0.3$ and $t$ as the energy unit. $\delta$ is the doping measured from half-filling. For $\lambda$ we have chosen the value $0.75$ which is a typical value obtained in LDA calculations \cite{Andersen,Krakauer}. The lowest eigenvalue occurs always in s-wave symmetry, i.e., for $i=1$. It depends only weakly on doping for $0.3 < \delta < 0.8$. In this region the order parameter also varies only slowly along the Fermi line, describing thus rather isotropic s-wave superconductivity. Below $\delta = 0.3$ $\lambda_1$ rapidly decreases with decreasing doping wich is caused by a soft mode which freezes into an incommensurate bond-order wave at $\delta_{BO} \sim 0.14$ \cite{Zeyher1}. As a result the s-wave order parameter becomes less and less isotropic with decreasing $\delta$ changing, for instance, for $\delta =0.2$ by about a factor 3 along the Fermi line but it does not pass through zero. The d-wave coupling constant $\lambda_3$ would be zero in the uncorrelated case. Decreasing $\delta$ from large values correlation effects increase and the static vertex function $\gamma({\bf k},{\bf k-k'})$ develops more and more a forward scattering peak in the transferred momentum $\bf k-k'$ \cite{Kulic,Zeyher2}. In the extreme case where $\gamma$ is proportional to $\delta({\bf k-k'})$ $\lambda_1$ and $\lambda_3$ would become degenerate. Fig. 1 shows that this degeneracy is nearly reached at low dopings. The strong localization of the effective interaction is caused by the forward scattering peak in $\gamma$ and, in addition, by the incipient instability at $\delta_{BO}$ which is also very localized in ${\bf k}$-space. The neglect of retardation effects which leads to the BCS-formula Eq.(6) is doubtful for several reasons. First, $\lambda = 0.75$ no longer corresponds really to a weak-coupling case. Secondly, $T_c$ is no longer small compared to the frequency of the soft mode near $\delta_{BO}$ which means that the frequency dependence of $\Theta$ cannot be neglected for these dopings. Thirdly, and most severely, the formation of the forward scattering peak in the static vertex function $\gamma$ is accompagnied by a strong frequency dependence which should be taken into account on the same footing as its momentum dependence. We thus have solved the gap equation Eq.(2) assuming only that the momenta can be put right onto the Fermi line but keeping the full frequency and momentum dependence along the Fermi line. \vspace{-0.5cm} \begin{figure} \epsfysize=75mm \centerline{\epsffile{fig2.eps}} \caption{ Transition temperature $T_c$ in units of $t$ as a function of the doping $\delta$ for the uncorrelated (dash-dotted line, $\Gamma_1$ symmetry) and correlated (dashed and solid lines for $\Gamma_1$ and $\Gamma_3$ symmetries, resp.) cases, using only the phonon contribution.} \label{fig:fig2} \end{figure} Fig. 2 shows the obtained results for $T_c$ for a s-wave (dashed line) and a d-wave (solid line) order parameter, together with the $T_c$ curve for the uncorrelated case (dash-dotted line). The first observation is that correlation effects always suppress s-wave superconductivity. The suppression is large for $\delta > 0.25$ and becomes less pronounced at smaller dopings in agreement with previous results based on the static approximation \cite{Zeyher2}. The order parameter associated with the dashed line in Fig. 2 varies only slowly along the Fermi line, i.e., we have a usual s-wave order parameter without nodes. In a Gutzwiller description this order parameter would be exactly zero if retardation effects can be neglected and $N=2$ is taken: Integrating out the phonons the effective interaction becomes in the static limit proportional to the double occupancy operator. Any matrix element formed with Gutzwiller wave functions thus would be zero. In contrast to that the enforcement of our constraint at large $N$'s, namely, that only $N/2$ out of the total $N$ states at a given site can be occupied at the same time, gives rise to another effect, which is absent in the Gutzwiller treatment: the effective interaction becomes more and more long-ranged with decreasing doping. This makes isotropic s-wave superconductivity possible even in the presence of a $N=2$ constraint. Fig. 2 nevertheless shows that the reduction of $T_c$ due to the constraint is substantial except at very small dopings where the effective interaction becomes extremely long-ranged and the suppression by the constraint is small. The kernel $\Theta^{e-p}$ is in the absence of correlations and in the static limit negative for all arguments $\bf k$ and $\bf k'$. This implies that the s-wave symmetry has always the highest $T_c$. According to our numerical studies the same is also true in the correlated case which explains why the solid line in Fig. 2, describing d-wave superconductivity, is always below the dashed line. As indicated previously the very existence of finite values for $T_c$ with d-wave symmetry is in our model due to electronic correlation effects. Fig. 1 showed that the momentum dependence of $\gamma_c$ causes in the static limit a finite coupling strength in the d-wave channel. The solid line in Fig. 2 proves that a finite $T_c$ in the d-wave channel results from this even if the strong frequency dependence of $\gamma_c$ is also taken into account. \vspace{-0.5cm} \begin{figure} \epsfysize=75mm \centerline{\epsffile{fig3.eps}} \caption{ Isotope coefficient $\beta$ as a function of the doping $\delta$ considering only the phonon contribution for $\Gamma_1$ (dashed line) and $\Gamma_3$ (solid line) symmetries. The filled circles are calculated values.} \label{fig:fig3} \end{figure} Fig. 3 shows results for the isotope coefficient $\beta$ in the case of s-wave (dashed line) and d-wave (solid line) superconductivty. We assumed hereby that $\omega_0$ is inversely and $g$ directly proportional to the the square root of the mass rendering $\lambda$ independent of the mass. Neglecting the frequency dependence of $\gamma_c$ there is only one energy scale involved in the gap equation which yields $\beta = 1/2$ independent of doping, $\lambda$, etc. In contrast to that the curves in Fig. 3 show that $\beta$ increases monotonically with decreasing doping and is always larger than 1/2. The deviaton of $\beta$ from the value 1/2 is due to the frequency dependence of $\gamma_c$ which increases with decreasing doping \cite{Zeyher1} and acts as a second energy scale. The non-adiabatic effects produced by electronic correlations thus always increase $\beta$ for our range of parameter values. This should be contrasted to the case where vertex corrections due to the electron-phonon interaction has been studied \cite{Grimaldi,Miller}, and where $\beta$ may be larger or smaller than $1/2$. We can conclude from our results that models based on phonon-induced effective interactions in the presence of correlations seem not to be very attractive models for high-T$_c$ superconductivity because a) the largest transition temperatures are found for s-wave symmetry for all dopings and b) the isotope coefficient is larger than $0.5$, especially at low dopings. Experimentally, high-$T_c$ cuprates exhibit d-wave superconductivity and, at least near optimal dopings, small isotope effects. \section{Calculation of $T_c$ and $\beta$ from the full effective interaction} Superconducting instabilites of the pure $t-J$ model have been studied in \cite{Zeyher1,Greco,Zeyher4}. The smallest eigenvalue of $\Theta^{t-J}$ and, correspondingly, the largest $T_{c}$ always occur in the d-wave symmetry channel. On the other hand, the phonon-induced effective interaction described by $\Theta^{e-p}$ favors in the $t-J$ model mainly s-wave superconductivity. In this section, we will consider the case where both effective interactions are present and study the symmetry of the resulting superconducting state, its transition temperature and isotope coefficient. \vspace{-0.5cm} \begin{figure} \epsfysize=75mm \centerline{\epsffile{fig4.eps}} \caption{ Lowest eigenvalues $\lambda_i$ of the total static kernel $\Theta$ for the representations $\Gamma_1$ and $\Gamma_3$ of $C_{4v}$ as a function of the doping $\delta$.} \label{fig:fig4} \end{figure} Fig. 4 shows the two lowest eigenvalues of the matrix Eq.(5) which occur in the s- and d-wave symmetry channel. For $\Theta$ both terms in Eq.(3) have been included. For $\delta < 0.6$ the lowest eigenvalue has d-wave symmetry and decreases steadily and strongly with decreasing doping diverging finally at $\delta_{BO}$. Comparing the solid lines in Figs. 1 and 4 of this work and Fig. 2 in \cite{Zeyher1} one sees that the lowest eigenvalue is roughly additive in the $t-J$ and the phonon contributions though, of course, $\Theta^{t-J}$ and $\Theta^{e-p}$ do not commute with each other. Note also the considerable contribution of the phonon part to $\lambda_3$ though the overall dimensionless coupling constant has the rather moderate value $\lambda = 0.75$. The eigenvalue $\lambda_1$ (dashed line in Fig. 4) is rather constant for $\delta > 0.3$ and decreases much less towards lower dopings than $\lambda_3$. There is a crossover from d-wave to s-wave symmetry in the lowest eigenvalue at $\delta \sim 0.6$ suggesting that the s-wave order parameter is more stable than the d-wave order parameter at high dopings. Such a crossover is expected for general reasons: For large dopings correlation effects play a minor role. As a result the d-wave part in the phonon-induced effective interaction as well as the whole $t-J$ part becomes small whereas the s-wave part of the Holstein model will dominate. A comparison of the solid lines in Figs. 1 and 4 nevertheless shows that the latter is very effectively suppressed by the $t-J$ part at practically all considered dopings. For instance, $|\lambda_1|$ in Fig. 1 is for $\delta > 0.25$ about a factor 4 larger than in Fig. 4. \vspace{-0.5cm} \begin{figure} \epsfysize=75mm \centerline{\epsffile{fig5.eps}} \caption{ Transition temperature $T_c$ in units of $t$ as a function of the doping $\delta$ for $\Gamma_3$ symmetry for $\lambda=0$ (dash-dotted line) and $\lambda =0.75$ (solid line).} \label{fig:fig5} \end{figure} The solution of the gap equation Eq.(2) is not straightforward because $\Theta$ contains both instantaneous and retarded contributions and because the various retarded contributions have different effective cutoff energies. We developed in \cite{Zeyher1} a method to solve Eq.(2) directly avoiding the use of pseudopotentials. The only simplification is that we put the momenta in retarded (but not in instantaneous) contributions to the Fermi line. This approximation has been checked numerically and found to be very well satisfied in our case. Using the fact that the instantaneous kernel consists only of a few separable contributions Eq.(2) can for a given Matsubara frequency reduced to a linear matrix problem of order 1 and 3 for d- and s-wave symmetry, respectively. In Fig. 5 we have plotted the calculated values for $T_c$ for d-wave symmetry $\Gamma_3$ as a function of the doping $\delta$. The dashed-dotted line corresponds to $\lambda=0$, the solid line to $\lambda=0.75$. The Figure shows that phonons always increase the $T_c$ for d-wave superconductivity. The increase is especially large at low dopings. This is quite in contrast to calculations in the weak-coupling case \cite{Dahm,Nunner} where the $T_c$ of d-wave superconductivity is lowered by phonons. The physical picture in the latter case is that the effective interaction of the electronic part is repulsive and strongly peaked near the $M$-point whereas the phononic d-wave part is attractive and diminuishes especially at the $M$-point the repulsion causing a lowering of $T_c$. In our case the dependence of the static effective interaction of the pure $t-J$ model on the transferred momentum can be inferred from Figs.1a) and 1b) in Ref. \cite{Zeyher1}. For small dopings the phonon part is restricted to small momenta which means that it would contribute only near the $X$-point in Fig. 1b) of Ref. \cite{Zeyher1}. As a result the total effective interaction would become even more attractive around the $X$-point and elsewhere not be changed. This clearly would enhance the d-wave part of the effective interaction which is also in agreement with Fig. 4. From Figs. 1, 4 and 5 it is evident that the large lowering of the eigenvalue $\lambda_3$ due to phonons corresponds only to a rather moderate increase in $T_c$. This means that the use of a BCS-formula with a fixed effective cutoff would grossly overestimate the increase in $T_c$ for d-wave superconductivity due to phonons. What has been overlooked in such an approach is that we deal with at least three energy scales, namely $t$, $J$, and $\omega_0$. For instance, the instantaneous contribution of the $t-J$ part is characterized by the energy scale $t$ whereas the phononic one by $\omega_0$. Since $t >> \omega_0$ it is evident that the phononic part in $\lambda_3$ will contribute to $T_c$ much less than the $t-J$ part. We were unable to find any finite transition temperatures $T_{ci}$ for $i \neq 3$, i.e., for symmetries different from the d-wave symmetry.. Taking the accuracy of our calculation into account this means that $T_{ci} < 0.002$ for $i \neq 3$. As shown in Fig. 2 the phonon-induced effective interactions leads to a considerable $T_c$ for s-wave superconductivity. The $t-J$ part to the effective interaction, however, is very repulsive in the s-wave channel prohibiting s-wave superconductivity. We cannot exclude that the crossover at $\delta \sim 0.6$ in Fig. 4 in the lowest eigenvalue from d-wave to s-wave could stabilize a s-wave order parameter at large dopings. Another possibility is that other symmetries than d- or s-wave become stable at large dopings in view of the approximate degeneracy of the eigenvalues of practically all symmetries in the pure $t-J$ model \cite{Zeyher1}. In any case, the corresponding $T_c$'s would be smaller than $\sim 0.002$ for all dopings and thus be rather irrelevant. The behavior of the isotope coefficient $\beta$ as a function of doping is another interesting test for any theory of high-$T_c$ superconductivity. Fig. 6 shows the calculated $\beta$ for $\lambda = 0.75$, $J=0.3$, $\omega_0=0.06$ in the case of $\Gamma_3$ symmetry. It is always positive, starting from small values at small dopings. With increasing doping it increases monotonously approaching values near $1/2$ at large dopings. Our calculated curve shows the typical behavior of the experimetal data, namely, a small value for $\beta$ at optimal doping and substantial values far away from optimal doping \cite{Frank}. The curve in Fig. 6 is the result of several \vspace{-0.5cm} \begin{figure} \epsfysize=75mm \centerline{\epsffile{fig6.eps}} \caption{ Isotope coefficient $\beta$ as a function of the doping $\delta$ for d-wave-like $\Gamma_3$ symmetry. The filled circles are calculated values.} \label{fig:fig6} \end{figure} competing effects. We saw in Fig. 3 that correlation effects in the phonon-induced part cause a doping dependence of $\beta$ which is opposite to that in Fig. 6, namely, a monotonously increasing $\beta$ with decreasing doping starting at the classical value of $1/2$ at large dopings. Including also the $t-J$ part the large values of $\beta$ at low dopings are nearly completely and at larger dopings partially quenched. Part of this effect may be understood from Fig. 5. It shows that the relative importance of the phonon contribution increases steadily with increasing $\delta$ except at very small dopings. In a simple picture one thus may argue that at small dopings $T_c$ is mainly due to the $t-J$ part. This part decays faster than the phononic part with doping so that $T_c$ at larger dopings is mainly due to the phonons. A more realistic interpretation of Fig. 6 should, however, also take into account the complex competition between the three energy scales $t$, $J$, and $\omega_0$, pair breaking effects, etc. In order to get more insight into the behavior of $\beta$ we have calculated $T_c$ and $\beta$ in the $\Gamma_3$ symmetry channel for the fixed doping value $\delta = 0.20$ as a function of the phonon frequency $\omega_0$. Fig. \ref{fig:fig7} shows the result for $J = 0.3$ and $\lambda = 0.75$. The solid line in Fig. \ref{fig:fig7} represents the value for $T_c$, multiplied with 10. Since we keep $\lambda$ fixed the coupling constant $g$ approaches zero in the adiabatic limit $\omega_0 \rightarrow 0$ which means that $T_c$ in this limit is solely due to the $t-J$ part. With increasing phonon frequency $T_c$ \vspace{-0.5cm} \begin{figure} \epsfysize=75mm \centerline{\epsffile{fig7.eps}} \caption{ Transition temperature $T_c$, multiplied by 10, and isotope coefficient $\beta$ as a function of the phonon frequency $\omega_0$ for d-wave-like $\Gamma_3$ symmetry.} \label{fig:fig7} \end{figure} increases monotonously, passing over from an initial sublinear to a linear behavior. The solid line shows that phonons always increase $T_c$ for d-wave superconductivity. The absolute values in the Figure are quite remarkable: Though the employed coupling constant for $\lambda$ is rather moderate and though the phonon contribution to $T_c$ in the d-wave channel is entirely due to correlation effects $T_c$ increases by nearly a factor 3 for $\omega_0 =0.2$ corresponding to about the largest phonons in the cuprates. The dashed line in Fig. \ref{fig:fig7} represents the dependence of $\beta$ on $\omega_0$. For small $\omega_0$ $\beta$ is practically zero. $\beta$ increases monotonously with increasing $\omega_0$ and tends to the classical value $1/2$ at large phonon frequencies. The dashed line in Fig. \ref{fig:fig7} may be understood in physical terms roughly in the following way. In our model the static effective coupling is independent of the phonon mass, i.e., of $\omega_0$. $\beta$ is thus determined by an effective energy cutoff. For small $\omega_0$'s this cutoff is mainly given by electronic parameters in the $t-J$ part leading to a small value for $\beta$. At large $\omega_0$'s the phonon contribution to $T_c$, which is only due to electronic correlations, determines mainly the total effective energy cutoff causing a value of $\beta$ near $1/2$. In conclusion, we have treated the electron-electron and the electron-phonon interactions in a generalized $t-J$ model by means of a systematic $1/N$ expansion and have solved the resulting linearized equation for the superconducting gap by reliable numerical methods. We found that electronic correlations affect phonon-induced superconductivity in several ways: Instabilities towards d-wave or other symmetries different from s-wave become possible. The corresponding transition temperatures are, however, always smaller than that of s-wave superconductivity. Moreover, the $T_c$ for s-wave superconductivity is at larger dopings heavily and at small dopings somewhat suppressed by correlations. The isotope coefficient $\beta$ is $1/2$ at large dopings but increases with increasing correlations, i.e., decreasing dopings, both in the s- and d-wave channel. Including also the $t-J$ part in the effective interaction we found that within our numerical accurracy only d-wave superconductivity is stable for dopings $0.15 < \delta < 0.8$ and that the phononic part always increases $T_c$ for the above dopings. The $t-J$ part in the effective interaction changes the dependence of $\beta$ on doping: $\beta$ assumes now small values at low dopings and increases monotonously with doping towards the classical value $1/2$ at large dopings. Keeping $\lambda$ fixed and varying the phonon mass, i.e., $\omega_0$ we find that $T_c$ and $\beta$ increase monotonously with $\omega_0$ and that $\beta$ varies between zero at small and roughly $1/2$ at large phonon frequencies within the interval $0 < \omega_0 < 0.2$. Our findings differ in many aspects from the corresponding results based on weak-coupling calculations \cite{Schuettler,Dahm,Bulut,Nunner,Pao}. In these calculations $T_c$ for d-wave superconductivity induced either by a $U$ or a spin-fluctuation term is always suppressed by phonons. For instance, in the fluctuation-exchange approximation $T_c$ drops to zero if the phonon-mediated on-site attraction $U_p$ becomes comparable to the Hubbard term $U$\cite{Pao}. We treated the opposite case where $U >> U_p$ and found a different behavior, namely, that phonons enhance the $T_c$ for d-wave superconductivity. Our different result is mainly caused by corrections to the bare electron-phonon vertex due to the strong electron-electron interaction. This vertex develops for not too large dopings a forward scattering peak so that only phonon with small momenta can couple to the electrons. As a result the electron-phonon and the $t-J$ contributions to the gap decouple in $\bf k$ space. Our calculations show that the two contributions no longer cancel each other to a large extent but, on the contrary, enhance each other. For a momentum-independent bare electron-phonon coupling weak-coupling calculations yield a small, often negative value for the isotope coefficient $\beta$ which is rather independent of doping. In our case the strong momentum-dependence of the effective electron-phonon coupling, induced by electronic correlations, causes a strong dependence of $\beta$ on doping: being always positive, $\beta$ is small at optimal doping and assumes values of roughly $1/2$ at large dopings in agreement with the basic features of the experimental data in the cuprates. The presented results are accurate at large $N$'s because we have taken only the leading terms of a $1/N$ expansion. Phonon renormalizations and vertex corrections due to the electron-phonon interaction are of $O(1/N)$ and thus have been omitted. On the other hand it is known that in the physical case $N=2$ anharmonic effects in the atomic potentials and the formation of polarons occur if $\lambda$ is about 1 or larger \cite{Pao1}. This suggests that keeping only the leading order of the $1/N$ expansion cannot describe adequately the case $N=2$ at large coupling strengths or if the Migdal ratio $\omega_0/t$ is no longer small. Correspondingly, we have shown numerical results for rather moderate values for $\lambda$ and small Migdal ratios. {\bf Acknowledgement}: The authors are grateful to Secyt and the International Bureau of the Federal Ministry for Education, Science, Research and Technology of Germany for financial support (Scientific-technological cooperation between Argentina and Germany, Project No. ARG AD 3P). The first and second author thank the MPI-FKF, Stuttgart, Germany, and the Departamento de F\'{\i}sica, Fac. Cs. Ex. e Ingenier\'{\i}a, U.N. Rosario, Argentina, respectively, for hospitality. The authors also thank P. Horsch for a critical reading of the manuscript.
\section{Introduction} Understanding the non--valence quark content of hadrons remains, despite a history spanning over two decades, a major theoretical challenge. The interest in non--valence physics derives mainly from the unique opportunities which it provides for new insights into quantum aspects of hadron structure beyond the naive and spectroscopically very successful quark model. Important questions in this realm include\ modifications of the QCD vacuum inside hadrons \cite{don86}, the mechanism of flavor mixing and the origin of the OZI--rule \cite{gei91,gei97}, the structure of constituent quarks \cite{kap88}, and the role of gluons in the dynamics of the (isoscalar) strange--quark sea in the nucleon \cite{dos98}. The strangeness distribution inside of the nucleon \cite{dec93} represents the most intensely studied example of a hadronic non--valence quark effect. Currently the vector channel of this distribution, as described by the strange vector form factors, is a focus of experimental \cite{mit,Mai,TJN} and theoretical \cite{dec93,mus94} research. Since systematic and model--independent approaches have still little predictive power for the nucleon's strangeness content (as exemplified by studies in chiral perturbation theory \cite{bor97,ram97,hem98,oh99} and on the lattice \cite {lat}), most previous and current theoretical analyses of the strangeness form factors were model--based. Among the first and most transparent models for the vector form factors were those which implement a kaon--cloud of the nucleon \cite{had} and thus complement pole dominance approaches \cite{pole}. In kaon--cloud models the nucleon's strangeness distribution is generated by fluctuations of the ``bare'' (i.e. nonstrange) nucleon into kaon--hyperon intermediate states which are described by the corresponding one--loop Feynman graphs \cite{had}% . The two crucial assumptions underlying the loop model are 1) that the lightest valence--strangeness carrying intermediate states generate the dominant contribution to the strangeness content and hence give at least a rough estimate of its size, and 2) that rescattering (i.e. multi--loop) contributions are suppressed (despite large couplings). Both of these assumptions have recently been challenged. A dispersive analysis on the basis of analytically continued $K-N$ scattering data demonstrated that rescattering corrections are important even at low momentum transfers, both to restore unitarity and to build up resonance strength in the $\phi $ meson region \cite{ram97c}. Furthermore, a study in an ``unquenched'' quark model found the contributions from higher--lying intermediate states (up to surprisingly large invariant masses) indispensable for the calculation of the strange quark distribution \cite {gei97} and prompted our collaborators and us to investigate these issues in a complementary hadronic one--loop model\ \cite{bar98}. In addition to the original $K-Y$ loops ($Y=\Lambda ,\Sigma $), we included the next higher--lying intermediate states, i.e. the $K^{\ast }-Y$ pairs. In the first part of this study the corresponding loops were evaluated using Bonn--J\"{u}lich $K^{\ast }NY$ form factors (see below),\ as in the original $K-Y$ model. The results were discouraging: the $K^{\ast }$ contributions were found to be larger than those from the kaon loop, and their dispersive analysis indicated strong unitarity violations. The anomalously large and apparently unrealistic $K^{\ast }$ contributions could be traced to the large $K^{\ast }$ tensor couplings and, in particular, to the very large cutoff parameter of the $K^{\ast }N\Lambda $ form factor taken from the Bonn--J\"{u}lich potential model \cite{hol89}. Since both the Nijmegen potential \cite{nag77,rij98} and the forthcoming update of the Bonn--J\"{u}lich $NY$ potential\ \cite{hai98} find substantially smaller values for this cutoff, we feel that a detailed and quantitative analysis of \ the cutoff (and coupling) dependence of the $% K^{\ast }$ contributions (covering the whole range of so far proposed values) \ would be a useful contribution to the ongoing discussion \cite {mel97,mel99} of hadron loop model applications to nucleon strangeness. Such an analysis is one of the main objectives of the present paper. Furthermore, we will calculate and discuss the momentum dependence of the strange vector form factors in the loop model at low momentum transfers ($Q^{2}\leq 1$ GeV) relevant for the present and planned measurements at MIT-Bates, TJNAF and MAMI. \section{The meson-hyperon loop model} \label{mod} We begin by recapitulating the definition of the strange vector form factors, the pertinent features of the hadronic loop model \cite{had}, and its extension to additional intermediate states containing $K^{\ast }$ mesons \cite{bar98}. The focus of our investigations will be on the nucleon matrix element of the strangeness current, which is parametrized by two invariant amplitudes, the Dirac and Pauli strangeness form factors $% F_{1,2}^{(s)}$, \begin{equation} \langle N(p^{\prime })|\bar{s}\gamma _{\mu }s|N(p)\rangle ={\bar{U}}% (p^{\prime })\left[ F_{1}^{(s)}(q^{2})\gamma _{\mu }+i{\frac{\sigma _{\mu \nu }q^{\nu }}{2m_{N}}}F_{2}^{(s)}(q^{2})\right] U(p)\ . \end{equation} Here $U(p)$ denotes the nucleon spinor and $F_{1}^{(s)}(0)=0$, due to the absence of an overall strangeness charge of the nucleon. The leading nonvanishing moments of these form factors are the Dirac and Sachs (square)\ strangeness radii \begin{equation} \langle r_{s}^{2}\rangle _{D}=\left. 6{\frac{d}{dq^{2}}}F_{1}^{(s)}(q^{2})% \right| _{q^{2}=0},\qquad \quad \langle r_{s}^{2}\rangle _{S}=\left. 6{\frac{% d}{dq^{2}}}G_{E}^{(s)}(q^{2})\right| _{q^{2}=0}, \end{equation} as well as the strangeness magnetic moment \begin{equation} \mu _{s}=F_{2}^{(s)}(0)\,. \end{equation} (The Sachs radius is obtained from the electric Sachs form factor $% G_{E}^{(s)}(q^{2})=F_{1}^{(s)}(q^{2})+q^{2}/(4m_{N}^{2})\,F_{2}^{(s)}(q^{2})$ and related to the Dirac radius by $\langle r_{s}^{2}\rangle _{S}=\langle r_{s}^{2}\rangle _{D}+3\mu _{s}/(2m_{N}^{2})$.) As mentioned above, we have chosen a hadronic one--loop model containing $K$ and $K^{\ast }$ mesons as the dynamical framework for the calculation of these moments. This model is based on the meson--baryon effective lagrangians \begin{eqnarray} {\cal L}_{MB} &=&-g_{ps}\bar{B}i\gamma _{5}BK\ \ \ , \label{1aa} \\ {\cal L}_{VB} &=&-g_{v}\left[ \bar{B}\gamma _{\alpha }BV^{\alpha }-\frac{% \kappa }{2m_{N}}\bar{B}\sigma _{\alpha \beta }B\partial ^{\alpha }V^{\beta }% \right] \;, \label{la} \end{eqnarray} where $B$ ($=N,\Lambda ,\Sigma $), $K$, and $V^{\alpha }$ are the baryon, kaon, and $K^{\ast }$ vector-meson fields, respectively, $m_{N}=939$ MeV is the nucleon mass and $\kappa $ is the ratio of tensor to vector coupling, $% \kappa =g_{t}/g_{v}$. In order to account for the finite extent of the above vertices, the model includes form factors from the Bonn--J\"{u}lich $N-Y$ potential \cite{hol89} at the hadronic $KNY$ and $K^{\ast }NY$ ($Y=\Lambda ,\Sigma $) vertices, which have the monopole form \begin{equation} F(k^{2})=\frac{m^{2}-\Lambda ^{2}}{k^{2}-\Lambda ^{2}} \label{ff} \end{equation} with meson momenta $k$ and the physical meson masses $m_{K}=495$ MeV and $% m_{K^{\ast }}=895$ MeV \cite{PDG}. These form factors render all encountered loop integrals finite and reproduce the on--shell values of the mesonic couplings. The range of currently favored values for the couplings $% g_{ps},g_{v},\kappa $ and cutoff parameters $\Lambda _{K}$ and $\Lambda _{K^{\ast }}$, as well as their impact on the strangeness observables, will be discussed below. Since the non-locality of the meson-baryon form factors (\ref{ff}) gives rise to vertex currents, gauge invariance was maintained in \cite{bar98} by introducing the photon field via minimal substitution in the momentum variable $k$ \cite{ohta}. (Consequences of the non--uniqueness of this prescription are discussed in Refs. \cite{ohta,bos92,wan96}.) The resulting nonlocal seagull vertices are given explicitly in \cite{bar98}. The diagonal couplings of $\bar{s}\gamma _{\mu }s$ to the strange mesons and baryons in the intermediate states are straightforwardly determined by current conservation, i.e. they are given by the net strangeness charge of the corresponding hadron. The situation is more complex for the non--diagonal (i.e. spin--flipping) coupling $F_{KK^{\ast }}^{(s)}(0)$ of the strange current to $K$ and $K^{\ast }$, which is defined by the transition matrix element \begin{equation} \langle K_{a}^{\ast }(k_{1},\varepsilon )|{\overline{s}}\gamma _{\mu }s|K_{b}(k_{2})\rangle =\frac{F_{KK^{\ast }}^{(s)}(q^{2})}{m_{K^{\ast }}}% \,\epsilon _{\mu \nu \alpha \beta }\,k_{1}^{\nu }\,k_{2}^{\alpha }\,\varepsilon ^{\ast \beta }\,\delta _{ab}\; \label{spinfl} \end{equation} (where $a$ and $b$ are isospin indices and $\varepsilon ^{\beta }$ is the polarization vector of the $K^{\ast }$). This coupling was estimated in \cite {bar98} on the basis of the vector meson dominance model of Ref. \cite{gm}, with the result $F_{KK^{\ast }}^{(s)}(0)=1.84$. The relevant one--loop Feynman graphs of the model are completely determined by the above vertices and the standard meson and baryon propagators. The diagrams containing strange mesons fall into three categories, corresponding to fluctuations of the nucleon into either $K-Y$ or $K^{\ast }-Y$ pairs, or to the strangeness--current induced spin flip transition from a $K-Y$ to a $% K^{\ast }-Y$ intermediate state. The explicit expressions for the corresponding loop amplitudes are collected in Appendix A. In the following calculations we will concentrate on the $\Lambda $ contributions and omit the comparatively negligible $\Sigma $ contribution \cite{had,bar98}. \section{Internal hadronic vertices} In this section we investigate the dependence of the loop-model results on the couplings and cutoffs which parametrize the internal $K(K^{\ast })\Lambda N$ vertices. In our previous analysis \cite{bar98} we have used the ($SU(3)$ based) couplings $g_{ps}/\sqrt{4\pi }=-3.944$, $g_{v}/\sqrt{% 4\pi }=-1.588$, $\kappa =3.26$ and the cutoff parameter values $\Lambda _{K^{\ast }}=2.2$ (2.1), $\Lambda _{K}=1.2(1.4)$ GeV of the Bonn--J\"{u}lich $NY$ potential \cite{hol89}. The cutoffs were determined from hyperon--nucleon scattering data, with the numbers in parenthesis denoting values obtained in an alternative model for the baryon-baryon interaction. The numerical results of the loop model \cite{bar98} are summarized in Table I. A glance at these numbers shows that the magnitude of the $K^{\ast }$ contributions exceeds those of the $K$ contributions by factors of $5-10$. As already mentioned, the main reason for these unrealistically large contributions can been traced to the unusually large $K^{\ast }N\Lambda $ cutoff parameter $\Lambda _{K^{\ast }}=2.2$ GeV found in \cite{hol89}. It has twice the size of the typical hadronic scale $\sim 1$ GeV around which such cutoff parameters lie normally. A substantially larger value (which in our case also exceeds the largest hadron masses in the loops by a factor of two) must be considered suspect in any model with hadronic degrees of freedom since one expects their quark--gluon substructure to become relevant at such scales. Indeed, the appearance of anomalously large cutoffs in a potential model suggests that those cutoffs are burdened with\ short--distance physics (not directly related to the $K^{\ast }$ sector)\ which would otherwise remain unaccounted for. Literally taking such effects over to the loop--model estimates of the nucleon's strangeness content, by fully associating them with the physical $K^{\ast },$ would therefore very likely be misleading. A hint that the $K^{\ast }$ sector of the original Bonn--J\"{u}lich potential might indeed be overburdened can be obtained from a comparison with the conceptually similar Nijmegen $NY$ potential \cite{nag77,rij98}. The Nijmegen potential contains more degrees of freedom in the scalar meson sector (including a pomeron) and finds a much smaller $K^{\ast }$ cutoff $% \Lambda _{K^{\ast }}\simeq 1.2$ GeV. Morevoer, a smaller $K^{\ast }$ cutoff is also favored in the forthcoming update of the Bonn--J\"{u}lich potential \cite{hai98}. The $K^{\ast }$ cutoff of this model is expected to lie around $1.5$ GeV \cite{mel299}. Hence, nowadays smaller cutoffs seem to be consistently favored by $NY$ potential models. This motivated our reanalysis of the loop--model results of Ref. \cite{bar98} for the nucleon's strangeness observables (as also suggested in \cite{mel99}) which we will discuss below. To get a qualitative idea of the range of coupling and cutoff values to be considered, and to maintain the underlying philosophy of the loop model, we will orient ourselves at the values used in the existing potential models. Of course, these values depend to some extent on the particular dynamics and particle content of a given model, and since the various potential models and also our model differ in this respect, the corresponding parameter sets cannot be directly compared or strictly related to each other. However, we expect the parameter ranges considered below, in particular for the cutoffs, to cover most of the physically reasonable parameter space of our model. We begin our discussion of the parameter dependence by documenting the sensitivity of the results to variations in the cutoff. To this end, we display the cutoff dependence in the range $1{\rm GeV}\leq \Lambda _{K^{\ast }}\leq 2.5\,{\rm GeV}$, which covers all so far proposed values for $\Lambda _{K^{\ast }}$. The first two figures show the behavior of the Dirac strangeness radius (Fig. 1a) and the strangeness magnetic moment (Fig. 1b) as a function of $\Lambda _{K^{\ast }}$, with the $K^{\ast }$ couplings fixed at the old, $SU(3)$--based Bonn--J\"{u}lich values given above. The full curves show the total results, the dashed ones represent the $K^{\ast }K^{\ast }\Lambda $ contributions, the dash--dotted ones the $KK^{\ast }\Lambda $ contributions, and the dotted curve corresponds to the result from the $K$ loop. Clearly, the cutoff dependence is very pronounced, reducing e.g. the value of Ref. \cite{bar98} for the magnetic moment by almost an order of magnitude for $\Lambda _{K^{\ast }}\simeq 1.5$ GeV. The reduction is even considerably stronger for the Nijmegen value $\Lambda _{K^{\ast }}\simeq 1.2$ GeV. Although a strong cutoff dependence was to be expected (especially due to the enhanced degree of divergence of $\ $the $K^{\ast }$ loops with derivative couplings), its actual magnitude is still surprising. If cutoff sizes of the order of that used in the Nijmegen potential were to be the most realistic, some of the conclusions reached in \cite{bar98} would have to be revised. In particular, for cutoffs of the Nijmegen size some sort of ``convergence'' of the intermediate--state sum (to one loop)\ could not anymore be excluded. In this case the kaon cloud contribution might well be sufficient for a first orientation about the overall size of the nucleon's strangeness content in hadronic one--loop models, as advocated in \cite {mel99} and in contrast to the findings of Ref. \cite{gei97} in the quark model. Other problematic aspects of hadron--loop models, such as the expected importance of rescattering corrections and unitarity violations \cite{ram97c}, are of course not affected by these arguments. The strong cutoff dependence of the strangeness observables exposes another, both conceptual and practical problem of hadronic loop models. Although the cutoff is principally a physical parameter (which indicates up to which resolution a purely hadronic description might be adequate), the foundations of the approach are not solid enough to give it a precise and quantitative meaning. Moreover, the available $NY$ scattering data do not allow an accurate determination of the $K^{\ast }NY$ form factors even in the framework of a specific potential model, not to mention other sources of uncertainty as, for example, the largely uncontrolled off--shell ambiguities incurred by transplanting such form factors into another model context. As a consequence, the numerical value of the $K^{\ast }$ cutoff cannot be accurately determined, and the corresponding uncertainty propagates, amplified by a hightened sensitivity, into the strangeness observables. These facts already imply that at most semi--quantitative predictions can be expected from the hadron--loop approach. The existing $NY$ potential models differ not only in the momentum dependence of the $K^{\ast }NY$ form factors, but also in the values of the corresponding $K^{\ast }$ couplings. The $K^{\ast }$ vector (Dirac) coupling of the Nijmegen potential is, for example, considerably smaller than the one of Ref. \cite{hol89} which we used above. In order to illustrate the dependence of the strange form--factor moments on these couplings, we compare in Fig. 2 the $K^{\ast }$ contributions to the strangeness radius (Fig. 2a) and magnetic moment (Fig. 2b)\ for the five pairs ($-1.588,3.26$), ($-0.8,2.0$), ($-0.8,4.0$), ($-2.0,2.0$), ($-2.0,4.0$) of coupling values ($% g_{v}/\sqrt{4\pi },\kappa $). The first of these pairs corresponds to the values of Holzenkamp et al. while the others were chosen to encompass the range of values which appear in other potential models. The Figs. 2 a,b demonstrate that the variations of the strangeness observables due to different choices for the couplings can be quite substantial for large values of the $K^{\ast }$ cutoff. The sensitivity to the couplings remains fairly small, however, for $\Lambda _{K^{\ast }}$ of the order of $1$ GeV, in particular for the strange magnetic moment and for couplings in the range between the Bonn-J\"{u}lich ($-1.588,3.26$) and Nijmegen \cite{rij98} ($% -1.45,2.43$) values\footnote{% These Nijmegen couplings are obtained from Table II of Ref. \cite{rij98} by using the SU(3) relations given in their Eq. (2.14).}. Throughout all of the above calculations we have kept the values of Ref. \cite{hol89} for the cutoff and coupling of the kaon fixed. The Nijmegen potential uses a pseudovector coupling which cannot be related without off--shell ambiguities to the pseudoscalar coupling of the Bonn--J\"{u}lich potential. On--shell, the equivalent pseudoscalar coupling of the Nijmegen potential ($g_{ps}/\sqrt{4\pi }\simeq -4.0$) differs by less than 5\% from the Bonn--J\"{u}lich coupling used here, and the Nijmegen cutoff parameter $% \Lambda _{K}\simeq 1.28$ GeV \cite{rij98} is similarly close to the one we use. In any case, the cutoff and coupling values from potential models should not be taken too literally since, as discussed above, the limited available data and the implicit model assumptions do not allow their precise (and unique) determination. We close this section with a remark on alternative choices for the internal form factors. The Bonn monopole form factors (\ref{ff}) have two well-known limitations: an artificial zero for cutoffs of the size of the meson mass (due to on-shell normalization) on which we will comment in the next section, and an unphysical singularity at time-like momenta $q^{2}=\Lambda ^{2}$. Although the singularity can be tamed by the usual causal boundary condition, it will still affect the values of the loop amplitudes to a certain extent. This problem could be avoided, although at the price of additional model dependence, by choosing alternative, singularity-free extrapolations of the Bonn form factors into the time-like region. One such extrapolation was proposed in Ref. \cite{fri99}. \section{Momentum dependence of the strange form factors} All of the currently running or planned experiments measure the strange vector form factors at $Q^{2}\neq 0$. Therefore, their data need to be extrapolated in order to obtain the leading form-factor moments. This extrapolation requires information on the momentum dependence of the form factors (at low energies) and thus on the spatial strangeness distribution inside the nucleon. In the following section we will evaluate the loop-model predictions for the momentum dependence by extending the above calculations to momentum transfers $Q^{2}\leq $ 1 GeV$^{2}.$ Both strange vector form factors will be measured in about the same $Q^{2}$-range by the planned spectrometer experiment G0 at Jefferson Lab \cite{TJN}. \ In the loop model, the coupling of the strange current to hyperon, meson, $% K(K^{\ast })NY$ vertex or $K/K^{\ast }$ transition vertex inside the loop is described by corresponding vertex functions, which we list in the appendix. These vertex functions contain the full momentum dependence of the form factors and can be evaluated numerically. Fixing the cutoffs for the $% K^{\ast }$ and the kaon at $\Lambda =1.2$ GeV and using the Nijmegen values for the coupling, we show in Fig. 3 (solid line) the resulting strange magnetic form factor \begin{equation} G_{SAMPLE}^{(s)}(Q^{2})=G_{M}^{(s)}(Q^{2})=F_{1}^{(s)}(Q^{2})+F_{2}^{(s)}(Q^{2}). \end{equation} together with the data point measured by the SAMPLE experiment at $Q^{2}=0.1$ GeV$^{2}$ \cite{mit}. (For comparison we have also included the older data point from the first runs only.)\ Figure 4 (solid line) shows the loop-model prediction for the combination \begin{equation} G_{HAPPEX}^{(s)}(Q^{2})=G_{E}^{(s)}(Q^{2})+0.39G_{M}^{(s)}(Q^{2}), \end{equation} measured by the HAPPEX\ collaboration, together with their data point at $% Q^{2}=0.48$ GeV$^{2}$ \cite{TJN}. The corresponding loop-model results for the Dirac strangeness radius and magnetic moment are given in Table I. At first sight, the results are in disagreement with both the SAMPLE\ and the HAPPEX data. The measured form factor combinations are both positive (at their respective momentum transfers) while the loop model leads to negative form factors of smaller slopes and magnitudes. In particular, the loop model predicts - like the majority of the other so far investigated models - a small and negative value for the strangeness magnetic moment. With regard to the SAMPLE result, however, there is a caveat. The derivation of the strangeness form factor from the measured asymmetry requires knowledge of the neutral weak axial form factor $G_{A}^{\left( Z\right) }$ of the nucleon and, in particular, its isovector radiative correction \cite {mus94}. The newest published value $G_{SAMPLE}^{(s)}(Q^{2}=0.1$ GeV$% ^{2})=0.61\pm 0.17\pm 0.21$ is based on the theoretical estimate \cite{mus90} for this radiative correction, which carries a substantial amount of uncertainty. The SAMPLE collaboration is currently measuring with a deuterium target \cite{bat94} in order to pin down the axial weak form factor contribution separately. Only after taking these data or more accurate and reliable theoretical estimates into account (which could substantially change the quoted value for $G_{SAMPLE}^{(s)}(Q^{2}=0.1$ GeV$% ^{2})$ both in sign an magnitude) can the strange magnetic moment be extracted unambigously and compared to the small value of the loop-model prediction. Furthermore, it is interesting to note that the results of the loop model could be made consistent with the HAPPEX data by allowing for smaller cutoff values, as suggested by recent applications of the meson cloud model to deeply inelastic scattering \cite{mst,cdnn}. In order to illustrate this point, we show in Figs. 3 and 4 (dashed lines) the results obtained by using $\Lambda =0.9$ GeV, which also brings the magnetic form factor somewhat closer to the SAMPLE data point. Moreover, this cutoff value is very close to the $K^{\ast }$ mass and therefore effectively switches off the internal $% K^{\ast }$ form factors (cf. Eq. (\ref{ff})). As a consequence, the contributions from the $K^{\ast }$ and the $K/K^{\ast }$ transition are completely negligible relative to the kaon contribution. This shows that the $K^{\ast }$ contributions worsen the agreement of the loop model predictions with the current experimental results. However, the zeros of the Bonn form factors have to be regarded as artefacts of the on-shell normalization and the results for the small cutoffs should therefore be viewed with caution. The latter reservations are enhanced by the difficulties with the physical interpretation of cutoffs of the same size as particle masses in a loop. In Figs. 5 and 6 we show the $K$, $K^{\ast }$ and $K/K^{\ast }$ transition contributions to the Dirac ($F_{1}^{(s)}$) and Pauli ($F_{2}^{(s)}$) form factors separately (again for $\Lambda =1.2$ GeV and Nijmegen couplings). These figures demonstrate that even with the new values for cutoffs and couplings adopted in the present paper, the $K^{\ast }$ and $K/K^{\ast }$ contributions are still of the same order of magnitude as the kaon contributions. This is in contrast to the much smaller size of these contributions to the strangeness magnetic moment in the chiral quark model \cite{han99}. In the latter the tensor coupling of the $K^{\ast }$ to the quarks is small, while the contributions from the vector coupling partially cancel each other and become proportional to the difference between the light and strange quark constituent masses, and therefore also small. Finally, we present the loop-model result for the form-factor combination \begin{equation} G_{MAMI}^{(s)}(Q^{2})=G_{E}^{(s)}(Q^{2})+0.22G_{M}^{(s)}(Q^{2}) \end{equation} which the forthcoming MAMI A4 experiment \cite{Mai} plans to measure at $% Q^{2}=0.23$ GeV$^{2}$. We find \begin{equation} G_{{\scriptsize \mbox{MAMI}}}^{(s)}(Q^{2}=0.23\mbox{GeV}^{2})=\left\{ \begin{tabular}{c} $-0.012\;\mbox{ for }\Lambda =0.9\mbox{ GeV}$ \\ $-0.046\;\mbox{ for }\Lambda =1.2\mbox{ GeV}$% \end{tabular} \right. \end{equation} where the same reservations as above apply to the smaller cutoff value. \section{Summary and Conclusions} To summarize, we have analyzed the cutoff, coupling, and momentum dependence of the $K^{\ast }$ contributions to the nucleon's vector strangeness content in the hadronic one--loop model of Ref. \cite{bar98}. It turns out that the softer $K^{\ast }NY$ form factors now generally favored by $NY$ potential models have some welcome consequences for such models. First, they can reduce the $K^{\ast }$ contributions to the strangeness radius and magnetic moment by over an order of magnitude, thereby indicating that the contributions from the lightest $KY$ intermediate states might be sufficient for rough estimates of the strangeness content in one--loop models. Although the $K^{\ast }$ contributions remain non--negligible (in particular towards larger momentum transfers) and worsen the agreement with the data of the HAPPEX experiment, they cease to be unrealistically large and they do not anymore exclude some sort of (slow) ``convergence'' of the intermediate state sum. The very sensitive dependence of the results on the $K^{\ast }$ cutoff emphasizes, however, the limited and mostly\ qualitative character of hadron--loop model predicitions. H.F. would like to thank the National Institute for Nuclear Theory in Seattle for hospitality during the 1998 Strangeness Program, where the plan for this work originated. He would also like to thank Wally Melnitchouk for useful correspondence on the Bonn--J\"{u}lich potential model, E. Kolomeitsev for pointing out Ref. \cite{fri99}, and the Deutsche Forschungsgemeinschaft for support under habilitation grant Fo 156/2--1. F.S.N. and M.N. would like to thank FAPESP and CNPq, Brazil, for support.
\section{INTRODUCTION} This is the first part of our eight presentations in which we consider applications of methods from wavelet analysis to nonlinear accelerator physics problems. This is a continuation of our results from [1]-[8], which is based on our approach to investigation of nonlinear problems -- general, with additional structures (Hamiltonian, symplectic or quasicomplex), chaotic, quasiclassical, quantum, which are considered in the framework of local (nonlinear) Fourier analysis, or wavelet analysis. Wavelet analysis is a relatively novel set of mathematical methods, which gives us a possibility to work with well-localized bases in functional spaces and with the general type of operators (differential, integral, pseudodifferential) in such bases. In the parts 1-8 we consider applications of wavelet technique to nonlinear dynamical problems with polynomial type of nonlinearities. In this part we consider this very useful approximation in the case of orbital motion in storage rings. Approximation up to octupole terms is only a particular case of our general construction for n-poles. Our solutions are parametrized by solutions of a number of reduced algebraical problems one from which is nonlinear with the same degree of nonlinearity and the rest are the linear problems which correspond to particular method of calculation of scalar products of functions from wavelet bases and their derivatives. \section{Orbital Motion in Storage Rings} We consider as the main example the particle motion in storage rings in standard approach, which is based on consideration in [9]. Starting from Hamiltonian, which described classical dynamics in storage rings $ {\cal H}(\vec{r},\vec{P},t)=c\{\pi^2+m_0^2c^2\}^{1/2}+e\phi $ and using Serret--Frenet parametrization, we have after standard manipulations with truncation of power series expansion of square root the following approximated (up to octupoles) Hamiltonian for orbital motion in machine coordinates: \begin{eqnarray} &&{\cal H}= \frac{1}{2}\cdot\frac{[p_x+H\cdot z]^2 + [p_z-H\cdot x]^2} {[1+f(p_\sigma)]}\nonumber\\ &&+p_\sigma-[1+K_x\cdot x+K_z\cdot z]\cdot f(p_\sigma)\\ &&+\frac{1}{2}\cdot[K_x^2+g]\cdot x^2+\frac{1}{2}\cdot[K_z^2-g]\cdot z^2- N\cdot xz \nonumber\\ &&+\frac{\lambda}{6}\cdot(x^3-3xz^2)+\frac{\mu}{24}\cdot(z^4-6x^2z^2+x^4) \nonumber\\ &&+\frac{1}{\beta_0^2}\cdot\frac{L}{2\pi\cdot h}\cdot\frac{eV(s)}{E_0}\cdot \cos\left[h\cdot\frac{2\pi}{L}\cdot\sigma+\varphi\right]\nonumber \end{eqnarray} Then we use series expansion of function $f(p_\sigma)$ from [9]: $ f(p_\sigma)=f(0)+f^\prime(0)p_\sigma+f^{\prime\prime}(0) p_\sigma^2/2+\ldots =p_\sigma- p_\sigma^2/(2\gamma_0^2)+\ldots$ and the corresponding expansion of RHS of equations corresponding to (1). In the following we take into account only an arbitrary polynomial (in terms of dynamical variables) expressions and neglecting all nonpolynomial types of expressions, i.e. we consider such approximations of RHS, which are not more than polynomial functions in dynamical variables and arbitrary functions of independent variable $s$ ("time" in our case, if we consider our system of equations as dynamical problem). \section{Polynomial Dynamics} The first main part of our consideration is some variational approach to this problem, which reduces initial problem to the problem of solution of functional equations at the first stage and some algebraical problems at the second stage. We have the solution in a compactly supported wavelet basis. Multiresolution expansion is the second main part of our construction. The solution is parameterized by solutions of two reduced algebraical problems, one is nonlinear and the second is some linear problem, which is obtained from the method of Connection Coefficients (CC). \subsection{ Variational Method} Our problems may be formulated as the systems of ordinary differential equations \begin{eqnarray} {\mathrm{d} x_i}/{\mathrm{d} t}=f_i(x_j,t), \quad (i,j=1,...,n) \end{eqnarray} with fixed initial conditions $x_i(0)$, where $f_i$ are not more than polynomial functions of dynamical variables $x_j$ and have arbitrary dependence of time. Because of time dilation we can consider only next time interval: $0\leq t\leq 1$. Let us consider a set of functions $ \Phi_i(t)=x_i{\mathrm{d} y_i}/{\mathrm{d} t}+f_iy_i $ and a set of functionals \begin{eqnarray} F_i(x)=\int_0^1\Phi_i (t)dt-x_iy_i\mid^1_0, \end{eqnarray} where $y_i(t) (y_i(0)=0)$ are dual variables. It is obvious that the initial system and the system \begin{equation} F_i(x)=0 \end{equation} are equivalent. In the following parts we consider an approach, which is based on taking into account underlying symplectic structure and on more useful and flexible analytical approach, related to bilinear structure of initial functional. Now we consider formal expansions for $x_i, y_i$: \begin{eqnarray}\label{eq:pol1} x_i(t)=x_i(0)+\sum_k\lambda_i^k\varphi_k(t)\quad y_j(t)=\sum_r \eta_j^r\varphi_r(t), \end{eqnarray} where because of initial conditions we need only $\varphi_k(0)=0$. Then we have the following reduced algebraical system of equations on the set of unknown coefficients $\lambda_i^k$ of expansions (\ref{eq:pol1}): \begin{eqnarray}\label{eq:pol2} \sum_k\mu_{kr}\lambda^k_i-\gamma_i^r(\lambda_j)=0 \end{eqnarray} Its coefficients are \begin{eqnarray} \mu_{kr}&=&\int_0^1\varphi_k'(t)\varphi_r(t){\rm d}t,\\ \gamma_i^r&=&\int_0^1f_i(x_j,t)\varphi_r(t){\rm d}t.\nonumber \end{eqnarray} Now, when we solve system (\ref{eq:pol2}) and determine unknown coefficients from formal expansion (\ref{eq:pol1}) we therefore obtain the solution of our initial problem. It should be noted if we consider only truncated expansion (\ref{eq:pol1}) with N terms then we have from (\ref{eq:pol2}) the system of $N\times n$ algebraical equations and the degree of this algebraical system coincides with degree of initial differential system. So, we have the solution of the initial nonlinear (polynomial) problem in the form \begin{eqnarray}\label{eq:pol3} x_i(t)=x_i(0)+\sum_{k=1}^N\lambda_i^k X_k(t), \end{eqnarray} where coefficients $\lambda_i^k$ are roots of the corresponding reduced algebraical problem (\ref{eq:pol2}). Consequently, we have a parametrization of solution of initial problem by solution of reduced algebraical problem (\ref{eq:pol2}). The first main problem is a problem of computations of coefficients of reduced algebraical system. As we will see, these problems may be explicitly solved in wavelet approach. The obtained solutions are given in the form (\ref{eq:pol3}), where $X_k(t)$ are basis functions and $\lambda_k^i$ are roots of reduced system of equations. In our case $X_k(t)$ are obtained via multiresolution expansions and represented by compactly supported wavelets and $\lambda_k^i$ are the roots of corresponding general polynomial system (\ref{eq:pol2}) with coefficients, which are given by CC construction. According to the variational method to give the reduction from differential to algebraical system of equations we need compute the objects $\gamma ^j_a$ and $\mu_{ji}$, which are constructed from objects: \begin{eqnarray}\label{eq:pol4} \sigma_i&\equiv&\int^1_0X_i(\tau)\mathrm{d}\tau,\quad \nu_{ij}\equiv\int^1_0X_i(\tau)X_j(\tau)\mathrm{d}\tau,\nonumber\\ \mu_{ji}&\equiv&\int X'_i(\tau)X_j(\tau)\mathrm{d}\tau,\\ \beta_{klj}&\equiv&\int^1_0X_k(\tau)X_l(\tau)X_j(\tau)\mathrm{d}\tau \nonumber \end{eqnarray} for the simplest case of Riccati systems (sextupole approximation), where degree of nonlinearity equals to two. For the general case of arbitrary n we have analogous to (\ref{eq:pol4}) iterated integrals with the degree of monomials in integrand which is one more bigger than degree of initial system. \subsection{ Wavelet Computations} Now we give construction for computations of objects (9) in the wavelet case. We present some details of wavelet machinery in part 2. We use compactly supported wavelet basis (Fig.~1, for example): orthonormal basis for functions in $L^2({\bf R})$. \begin{figure}[ht] \centering \epsfig{file=tup61_f.eps, width=82.5mm, bb=0 200 599 590, clip} \caption{Wavelets at different scales and locations } \end{figure} Let be $ f : {\bf R}\longrightarrow {\bf C}$ and the wavelet expansion is \begin{eqnarray} f(x)=\sum\limits_{\ell\in{\bf Z}}c_\ell\varphi_\ell(x)+ \sum\limits_{j=0}^\infty\sum\limits_{k\in{\bf Z}}c_{jk}\psi_{jk}(x) \end{eqnarray} If in formulae (10) $c_{jk}=0$ for $j\geq J$, then $f(x)$ has an alternative expansion in terms of dilated scaling functions only $ f(x)=\sum_{\ell\in{\bf Z}}c_{J\ell}\varphi_{J\ell}(x) $. This is a finite wavelet expansion, it can be written solely in terms of translated scaling functions. Also we have the shortest possible support: scaling function $DN$ (where $N$ is even integer) will have support $[0,N-1]$ and $N/2$ vanishing moments. There exists $\lambda>0$ such that $DN$ has $\lambda N$ continuous derivatives; for small $N,\lambda\geq 0.55$. To solve our second associated linear problem we need to evaluate derivatives of $f(x)$ in terms of $\varphi(x)$. Let be $ \varphi^n_\ell=\mathrm{d}^n\varphi_\ell(x)/\mathrm{d} x^n $. We consider computation of the wavelet - Galerkin integrals. Let $f^d(x)$ be d-derivative of function $f(x)$, then we have $ f^d(x)=\sum_\ell c_l\varphi_\ell^d(x) $, and values $\varphi_\ell^d(x)$ can be expanded in terms of $\varphi(x)$ \begin{eqnarray} \phi_\ell^d(x)=\sum\limits_m\lambda_m\varphi_m(x), \end{eqnarray} where $\lambda_m=\int\varphi_\ell^d(x)\varphi_m(x)\mathrm{d} x$ are wavelet-Galerkin integrals. The coefficients $\lambda_m$ are 2-term connection coefficients. In general we need to find $(d_i\geq 0)$ \begin{eqnarray} \Lambda^{d_1 d_2 ...d_n}_{\ell_1 \ell_2 ...\ell_n}= \int\limits_{-\infty}^{\infty}\prod\varphi^{d_i}_{\ell_i}(x)dx \end{eqnarray} For Riccati case (sextupole) we need to evaluate two and three connection coefficients \begin{eqnarray} \Lambda_\ell^{d_1 d_2}&=&\int^\infty_{-\infty}\varphi^{d_1}(x)\varphi_\ell^{d_2}(x)dx, \\ \Lambda^{d_1 d_2 d_3}&=&\int\limits_{-\infty}^\infty\varphi^{d_1}(x)\varphi_ \ell^{d_2}(x)\varphi^{d_3}_m(x)dx\nonumber \end{eqnarray} According to CC method [10] we use the next construction. When $N$ in scaling equation is a finite even positive integer the function $\varphi(x)$ has compact support contained in $[0,N-1]$. For a fixed triple $(d_1,d_2,d_3)$ only some $\Lambda_{\ell m}^{d_1 d_2 d_3}$ are nonzero: $2-N\leq \ell\leq N-2,\quad 2-N\leq m\leq N-2,\quad |\ell-m|\leq N-2$. There are $M=3N^2-9N+7$ such pairs $(\ell,m)$. Let $\Lambda^{d_1 d_2 d_3}$ be an M-vector, whose components are numbers $\Lambda^{d_1 d_2 d_3}_{\ell m}$. Then we have the first reduced algebraical system : $\Lambda$ satisfy the system of equations $(d=d_1+d_2+d_3)$ \begin{eqnarray} A\Lambda^{d_1 d_2 d_3}&=&2^{1-d}\Lambda^{d_1 d_2 d_3}, \\ A_{\ell,m;q,r}&=&\sum_p a_p a_{q-2\ell+p}a_{r-2m+p}.\nonumber \end{eqnarray} By moment equations we have created a system of $M+d+1$ equations in $M$ unknowns. It has rank $M$ and we can obtain unique solution by combination of LU decomposition and QR algorithm. The second reduced algebraical system gives us the 2-term connection coefficients. For nonquadratic case we have analogously additional linear problems for objects (12). Solving these linear problems we obtain the coefficients of nonlinear algebraical system (6) and after that we obtain the coefficients of wavelet expansion (8). As a result we obtained the explicit time solution of our problem in the base of compactly supported wavelets with the best possible localization in the phase space, which allows us to control contribution from each scale of underlying multiresolution expansions. In the following parts we consider extension of this approach to the case of (periodic) boundary conditions, the case of presence of arbitrary variable coefficients and more flexible biorthogonal wavelet approach. We are very grateful to M.~Cornacchia (SLAC), W.~Her\-r\-man\-nsfeldt (SLAC), Mrs. J.~Kono (LBL) and M.~Laraneta (UCLA) for their permanent encouragement.
\section{Introduction} There are a large number of works devoted to the Hamiltonian description of the ideal hydrodynamics (see, for instance, the review \cite{ZK} and the references therein). This question was first studied by Clebsch (a citation can be found in Ref. \cite{lamb}), who introduced for nonpotential flows of incompressible fluids a pair of variables $\lambda $ and $\mu $ (which later were called as the Clebsch variables). A fluid dynamics in these variables is such that vortex lines represent themselves intersection of surfaces $% \lambda =\mbox{const}$ and $\mu =\mbox{const}$ and these quantities, being canonical conjugated variables, remain constant by fluid advection. However, these variables, as known (see, i.e.,\cite{KM}) describe only partial type of flows. If $\lambda $ and $\mu $ are single-valued functions of coordinates then the linking degree of vortex lines characterizing by the Hopf invariant \cite{fr} occurs to be equal to zero. For arbitrary flows the Hamiltonian formulation of the equation for incompressible ideal hydrodynamics was given by V.I.Arnold \cite{arnold1,arnold2}. The Euler equations for the velocity curl ${\bf \Omega }=\mbox{curl}\,{\bf v}$ \begin{equation} \frac{\partial {\bf \Omega }}{\partial t}=\mbox{curl}\,[{\bf v\times \Omega }% ],\,\,\,\mbox{div}\,{\bf v}=0 \label{euler} \end{equation} are written in the Hamiltonian form, \begin{equation} \frac{\partial {\bf \Omega }}{\partial t}=\{{\bf \Omega },{\cal {H}\},} \label{euler1} \end{equation} by means of the noncanonical Poisson brackets \cite{KM} \begin{equation} \{F,G\}=\int \left( {\bf \Omega }\left[ \mbox{curl}\,\frac{\delta F}{\delta {\bf \Omega }}\times \mbox{curl}\,\frac{\delta G}{\delta {\bf \Omega }}% \right] \right) {d{\bf r}} \label{bracket} \end{equation} where the Hamiltonian \begin{equation} {\cal H}_h=-\frac 12\int {\bf \Omega }\Delta ^{-1}{\bf \Omega }d^3{\bf r}, \label{Hhydro} \end{equation} coincides with the total fluid energy. In spite of the fact that the bracket (\ref{bracket}) allows to describe flows with arbitrary topology its main lack is a degeneracy. By this reason it is impossible to formulate the variational principle on the whole space $% {\cal S}$ of divergence-free vector fields. The cause of the degeneracy, namely, presence of Casimirs annulling the Poisson bracket, is connected with existence of the special symmetry formed the whole group - the relabeling group of Lagrangian markers (for details see the reviews \cite{salmon,ZK}). All known theorems about the vorticity conservation (the Ertel's, Cauchy's and Kelvin's theorems, the frozenness of vorticity and conservation of the topological Hopf invariant) are a sequence of this symmetry. The main of them is the frozenness of vortex lines into fluid. This is related to the local Lagrangian invariant -- the Cauchy invariant. The physical meaning of this invariant consists in that any fluid particle remains all the time on its own vortex line. The similar situation takes place also for ideal magneto-hydrodynamics (MHD) for barotropic fluids: \begin{equation} \rho _t+\nabla (\rho {\bf v})=0, \label{rho-t} \end{equation} \begin{equation} {\bf v}_t+({\bf v}\nabla ){\bf v}=-\nabla w(\rho )+\frac 1{4\pi \rho }[% \mbox{curl}\,{\bf h}\times {\bf h}], \label{v-t} \end{equation} \begin{equation} {\bf h}_t=\mbox{curl}[{\bf v}\times {\bf h}]. \label{h-t} \end{equation} Here $\rho $ is a plasma density, $w(\rho)$\ plasma entalpy, ${\bf v}$ and ${\bf h}$ are velocity and magnetic fields, respectively. As well known (see, for instance, \cite{landau}-\cite{Kuvshinov}), the MHD equations possesses one important feature -- frozenness of magnetic field into plasma which is destroyed only due to dissipation (by finite conductivity). For ideal MHD combination of the continuity equation (\ref{rho-t}) and the induction equation (\ref{h-t}) gives the analog of the Cauchy invariant for MHD. The MHD equations of motion (\ref{rho-t}-\ref{h-t}) can be also represented in the Hamiltonian form, \begin{equation} \rho _t=\{\rho ,{\cal H}\}\qquad {\bf h}_t=\{{\bf h,}{\cal H}\},\qquad {\bf v% }_t=\{{\bf v,}{\cal H\}}, \label{dp/dt} \end{equation} by means of the noncanonical Poisson brackets \cite{MG}: \begin{equation} \label{compres_bracket} \{F,G\}=\int \left( \frac{{\bf h}}\rho \cdot \left( \left[ \mbox{curl}\frac{% \delta F}{\delta {\bf h}}\times \frac{\delta G}{\delta {\bf v}}\right] -\left[ \mbox{curl}\frac{\delta G}{\delta {\bf h}}\times \frac{\delta F}{% \delta {\bf v}}\right] \right) \right) d^3{\bf r}+ \end{equation} \[ +\int \left( \frac{\mbox{curl}\ {\bf v}}\rho \cdot \left[ \frac{\delta F}{% \delta {\bf v}}\times \frac{\delta G}{\delta {\bf v}}\right] \right) d^3{\bf % r}+\int \left( \frac{\delta G}{\delta \rho }\nabla \left( \frac{\delta F}{% \delta {\bf v}}\right) -\frac{\delta F}{\delta \rho }\nabla \left( \frac{% \delta G}{\delta {\bf v}}\right) \right) d^3{\bf r}. \] This bracket is also degenerated. For instance, the integral $\int {\bf (v,h)% }d{\bf r}$, which characterizes mutual linkage knottiness of vortex and magnetic lines, is one of the Casimirs for this bracket. The analog of the Clebsch representation in MHD serves a change of variables suggested in 1970 by Zakharov and Kuznetsov \cite{ZK70}: \begin{equation} {\bf v}=\nabla \phi +\frac{\lbrack {\bf h}\times \mbox{curl}\ {\bf S}\rbrack }{\rho }. \label{phiHS} \end{equation} New variables $(\phi ,\rho )$ and ${\bf h,S}$ represent two pairs canonically conjugated quantities with the Hamiltonian coinciding with the total energy \[ {\cal H}=\int \left( \rho \frac{{\bf v}^{2}}{2}+\rho \varepsilon (\rho )+% \frac{{\bf h}^{2}}{8\pi }\right) d{\bf r}. \] In the present paper we suggest a new approach of the degeneracy resolution of the noncanonical Poisson brackets by introducing new variables, i.e., Lagrangian markers labeling each vortex lines for ideal hydrodynamics or magnetic lines in the MHD case. The basis of this approach is the integral representation for the corresponding frozen-in field, namely, the velocity curl for the Euler equation and magnetic field for MHD. We introduce new objects, i.e., the vortex lines or magnetic lines and obtain the equations of motion for them. This description is a mixed Lagrangian-Eulerian description, when each vortex (or magnetic) line is enumerated by Lagrangian marker, but motion along the line is described in terms of the Eulerian variables. Such representation removes all degeneracy from the Poisson brackets connected with the frozenness, remaining the equations of motion to be gauge invariant with respect to re-parametrization of each line. Important, that the equations for line motion, as the equations for curve deformation, are transverse to the line tangent. It is interesting that the line representation also solves another problem - the equations of line motion follow from the variational principle, being Hamiltonian. This approach allows also simply enough to consider the limit of narrow vortex (or magnetic) lines. For two-dimensional flows in hydrodynamics this ''new'' description corresponds to the well-known fact, namely, to the canonical conjugation of $x$ and $y$ coordinates of vortices (see, for instance, \cite{lamb}). The Hamiltonian structure introduced makes it possible to integrate the three-di\-men\-sional Euler equation (\ref{euler1}) with Hamiltonian ${\cal H}=\int |{\bf \Omega }|d{\bf r}$. In terms of the vortex lines the given Hamiltonian is decomposed into a set of Hamiltonians of noninteracting vortex lines. The dynamics of each vortex lines is, in turn, described by the equation of a vortex induction which can be reduced by the Hasimoto transformation \cite{Hasimoto} to the integrable one-dimensional nonlinear Schrodinger equation. For ideal MHD a new representation - analog of the Weber transformation - is found. This representation contains the whole vector Lagrangian invariant. In the case of ideal hydrodynamics this invariant provides conservation of the Cauchy invariant and, as a sequence, all known conservation laws for vorticity (for details see the review \cite{ZK}). It is important that all these conservation laws can be expressed in terms of observable variables. Unlike the Euler equation, these vector Lagrangian invariants for the MHD case can not be expressed in terms of density, velocity and magnetic field. It is necessary to tell that the analog of the Weber transformation for MHD includes the change of variables (\ref{phiHS}) as a partial case. The presence of these Lagrangian invariants in the transform provides topologically nontrivial MHD flows. The Weber transform and its analog for MHD play a key role in constructing the vortex line (or magnetic line) representation. This representation is based on the property of frozenness. Just therefore by means of such transform the noncanonical Poisson brackets become non-degenerated in these variables and, as a result, the variational principle may be formulated. Another peculiarity of this representation is its locality, establishing the correspondence between vortex (or magnetic) line and vorticity (or magnetic field). This is a specific mapping, mixed Lagrangian-Eulerian, for which Jacobian of the mapping can not be equal to unity for incompressible fluids as it is for pure Lagrangian description. \setcounter{equation}{0} \section{General remarks} We start our consideration from some well known facts, namely, from the Lagrangian description of the ideal hydrodynamics. In the Eulerian description for barotropic fluids, pressure $p=p(\rho )$, we have coupled equations - discontinuity equation for density $\rho $ and the Euler equation for velocity: \begin{equation} \rho _{t}+\mbox{div}\,\rho {\bf v}=0, \label{cont} \end{equation} \begin{equation} {\bf v}_{t}+{\bf (v\nabla )v}=-\nabla w(\rho ),\qquad dw(\rho )=dp/\rho . \label{velocity} \end{equation} In the Lagrangian description each fluid particle has its own label. This is three-dimen\-sional vector ${\bf a}$, so that particle position at time $t$ is given by the function \begin{equation} {\bf x=x(a,}t). \label{map} \end{equation} Usually initial position of particle serves the Lagrangian marker: ${\bf % a=x(a,}0)$. In the Lagrangian description the Euler equation (\ref{velocity}) is nothing more than the Newton equation: \[ \ddot{{\bf x}}=-\nabla w. \] In this equation the second derivative with respect to time $t$ is taken for fixed ${\bf a}$, but the r.h.s. of the equation is a function of $t$ and $% {\bf x}$. Excluding from the latter the $x$-dependence, the Euler equation takes the form: \begin{equation} \ddot{x}_{i}\frac{\partial x_{i}}{\partial a_{k}}=-\frac{\partial w(\rho )}{% \partial a_{k}}\,, \label{Lag} \end{equation} where now all quantities are functions of $t$ and ${\bf a.}$ In the Lagrangian description the continuity equation (\ref{cont}) is easily integrated and the density is given through the Jacobian of the mapping (\ref {map}) $J=\mbox{det}(\partial x_{i}/\partial a_{k})$: \begin{equation} \rho =\frac{\rho _{0}({\bf a})}{J}. \label{density} \end{equation} Now let us introduce a new vector, \begin{equation} u_k=\frac{\partial x_i}{\partial a_k}v_i\,, \label{u} \end{equation} which has a meaning of velocity in a new curvilinear system of coordinates or it is possible to say that this formula defines the transformation law for velocity components. It is worth noting that (\ref{u}) gives the transform for the velocity ${\bf v}$ as a {\it co-vector}. The straightforward calculation gives that the vector ${\bf u}$ satisfies the equation \begin{equation} \frac{du_k}{dt}=\frac \partial {\partial a_k}\left( \frac{{\bf v}^2}% 2-w\right) . \label{motion} \end{equation} In this equation the right-hand-side represents gradient relative to ${\bf a} $ and therefore the ''transverse'' part of the vector ${\bf u}$ will conserve in time. And this gives the Cauchy invariant: \begin{equation} \frac d{dt}\mbox{curl }_a\,{\bf u}=0, \label{curl} \end{equation} or \begin{equation} \mbox{curl}_a\,{\bf u=I}. \label{Cauchy} \end{equation} If Lagrangian markers ${\bf a}$ are initial positions of fluid particles then the Cauchy invariant coincides with the initial vorticity: ${\bf % I=\Omega _0(a)}$. This invariant is expressed through instantaneous value of $% {\bf \Omega (x},t)$ by the relation \begin{equation} {\bf \Omega _0(a)=}J{\bf (\Omega (x},t)\nabla ){\bf a(x},t) \label{Ca} \end{equation} where ${\bf a=a(x,}t)$ is inverse mapping to (\ref{map}). Following from (% \ref{Ca}) relation for ${\bf B=\Omega }/\rho ,$% \[ B_{0i}(a)=\frac{\partial a_i}{\partial x_k}B_k(x,t), \] shows that, unlike velocity, ${\bf B}$ transforms as a vector. By integrating the equation (\ref{motion}) over time $t$ we arrive at the so-called Weber transformation \begin{equation} {\bf u(a},t)={\bf u}_{0}({\bf a})+\nabla _{a}\Phi , \label{weber} \end{equation} where the potential $\Phi $ obeys the Bernoulli equation: \begin{equation} \frac{d\Phi }{dt}=\frac{{\bf v}^{2}}{2}-w(\rho ) \label{Phi} \end{equation} with the initial condition: $\Phi |_{{t=0}}=0$. For such choice of $\Phi $ a new function ${\bf u_{0}(a)}$ is connected with the ''transverse'' part of $% {\bf u}$ by the evident relation \[ \mbox{curl}_{a}\,{\bf u}_{0}({\bf a)=I}. \] The Cauchy invariant ${\bf I}$ characterizes the vorticity frozenness into fluid. It can be got by standard way considering two equations - the equation for the quantity ${\bf B=\Omega /\rho }$, \begin{equation} \frac{d{\bf B}}{dt}=({\bf B\nabla )v}, \label{froz} \end{equation} and the equation for the vector $\delta {\bf x=x(a+\delta a)-x(a)}$ between two adjacent fluid particles: \begin{equation} \frac{d\delta {\bf x}}{dt}=({\delta {\bf x\nabla )v}}, \label{delta} \end{equation} The comparison of these two equations shows that if initially the vectors ${% \delta {\bf x}}$ are parallel to the vector ${\bf B}$, then they will be parallel to each other all time. This is nothing more than the statement of the vorticity frozenness into fluid. Each fluid particle remains all the time at its own vortex line. The combination of Eqs. (\ref{froz}) and (\ref {delta}) leads to the Cauchy invariant. To establish this fact it is enough to write down the equation for the Jacoby matrix $J_{ij}=\partial x_{i}/\partial a_{j}$ which directly follows from (\ref{delta}): \[ \frac{d}{dt}\ \frac{{\partial a_{i}}}{{\partial x}_{k}}=-\frac{\partial a_{i}% }{\partial x_{j}}\frac{\partial v_{j}}{\partial x_{k}}, \] that in combination with Eq. (\ref{froz}) gives conservation of the Cauchy invariant (\ref{Cauchy}). If now one comes back to the velocity field ${\bf v}$ then by use of Eqs. (\ref{u}) and (\ref{weber}) one can get that \begin{equation} {\bf v}=u_{0k}\nabla a_{k}+\nabla \Phi \label{v} \end{equation} where gradient is taken with respect to ${\bf x}$. Here the equation for potential $\Phi $ has the standard form of the Bernoulli equation: \[ \Phi _{t}+({\bf v}\nabla )\Phi -\frac{{\bf v^{2}}}{2}+w(\rho )=0. \] It is interesting to note that relations (\ref{Cauchy}), as equations for determination of ${\bf x(a,}t)$, unlike Eqs (\ref{motion}), are of the first order with respect to time derivative. This fact also reflects in the expression for velocity (\ref{v}) which can be considered as a result of the partial integration of the equations of motion (\ref{motion}). Of course, the velocity field given by (\ref{v}) contain two unknown functions: one is the whole vector ${\bf a(x},t)$ and another is the potential $\Phi $. For incompressible fluids the latter is determined from the condition $% \mbox{div}\,{\bf v}=0$. In this case the Bernoulli equation serves for determination of the pressure. Another important moment connected with the Cauchy invariant is that it follows from the variational principle (written in terms of Lagrangian variables) as a sequence of relabelling symmetry remaining invariant the action (for details, see the reviews \cite{salmon,ZK}). Passing from Lagrangian to Hamiltonian in this description we have no any problems with the Poisson bracket. It is given by standard way and does not contain any degeneracy that the noncanonical Poisson brackets (\ref{bracket}) and (\ref {compres_bracket}) have. One of the main purposes of this paper is to construct such new description of the Euler equation (as well as the ideal MHD) which, from one side, would allow to retain the Eulerian description, as maximally as possible, but, from another side, would exclude from the very beginning all remains from the gauge invariance of the complete Euler description connected with the relabeling symmetry. As for MHD, this system in one point has some common feature with the Euler equation: it also possesses the frozenness property. The equation for ${\bf h}% /\rho $ coincides with (\ref{froz}) and therefore dynamics of magnetic lines is very familiar to that for vortex lines of the Euler equation. However, this analogy cannot be continued so far because the equation of motion for velocity differs from the Euler equation by the presence of pondermotive force. This difference remains also for incompressible case. \setcounter{equation} {0} \section{Vortex line representation} Consider the Hamiltonian dynamics of the divergence-free vector field ${\bf % \Omega }({\bf r},t)$, given by the Poisson bracket (\ref{bracket}) with some Hamiltonian ${\cal {\ H}}$ \footnote{{\normalsize The Hamiltonian (\ref {Hhydro}) corresponds to ideal incompressible hydrodynamics.}}: \begin{equation} \frac{\partial {\bf \Omega }}{\partial t}=\mbox{curl}\left[ \mbox{curl}\, \frac{\delta {\cal H}}{\delta {\bf \Omega }}\times {\bf \Omega }\right] . \label{dOmega_dt} \end{equation} As we have said, the bracket (\ref{bracket}) is degenerate, as a result of which it is impossible to formulate the variational principle on the entire space ${\cal S}$ of solenoidal vector fields. It is known \cite{ZK} that Casimirs $f$, annulling Poisson brackets, distinguish in ${\cal S}$ invariant manifolds ${\cal M}_{f}$ (symplectic leaves) on each of which it is possible to introduce the standard Hamiltonian mechanics and accordingly to write down a variational principle. We shall show that solution of this problem for the equations (\ref{dOmega_dt}) is possible on the base of the property of frozenness of the field ${\bf \Omega (r},t)$, which allows to resolve all constrains, stipulated by the Casimirs, and gives the necessary formulation of the variational principle. To each Hamiltonian ${\cal {H}}$ - functional of ${\bf \Omega (r},t)$ - we associate the generalized velocity \begin{equation} {\bf v}({\bf r})=\mbox{curl}\,\frac{\delta {\cal H}}{\delta {\bf \Omega }}. \label{Vgeneral} \end{equation} However one should note that the generalized ${\bf v}({\bf r})$ is defined up to addition of the vector parallel to ${\bf \Omega }$: \[ \mbox{curl}\,\frac{\delta {\cal H}}{\delta {\bf \Omega }}\rightarrow % \mbox{curl}\,\frac{\delta {\cal H}}{\delta {\bf \Omega }}+\alpha {\bf \Omega }, \] that in no way does change the equation for ${\bf \Omega }$. Under the condition $({\bf \Omega }\cdot \nabla \alpha )=0$ a new generalized velocity will have zero divergence and the frozenness equation (\ref{dOmega_dt}) can be written already for the new ${\bf v}({\bf r})$. A gauge changing of the generalized velocity corresponds to some addition of a Casimir \ to the Hamiltonian : \[ {\cal H}\rightarrow {\cal H}+f;\qquad \{f,..\}=0. \] Hence becomes clear that the transformation \[ {\bf x=x}({\bf a},t) \] of the initial positions of fluid particles ${\bf x}({\bf a},0)={\bf a}$ by the generalized velocity field ${\bf v}({\bf r})$ through solution of the equation \begin{equation} \label{vel} {\bf {\dot x} = v(x,}t) \end{equation} is defined ambiguously due to the ambiguous definition of ${\bf v}({\bf r})$ by means of (\ref{Vgeneral}). Therefore using full Lagrangian description to the systems (\ref{dOmega_dt}) becomes ineffective. Now we introduce the following general expression for ${\bf \Omega }({\bf r}% ) $, which is gauge invariant and fixes all topological properties of the system that are determined by the initial field ${\bf \Omega }_{0}({\bf a})$% \cite{KR}: \begin{equation} {\bf \Omega }({\bf r},t)=\int \delta({\bf r}-{\bf R}({\bf a},t))({\bf \Omega }_{0}({\bf a})\nabla _{{\bf a}}){\bf R}({\bf a},t)d^{3}{\bf a}. \label{OmegaR} \end{equation} Here now \begin{equation} {\bf r= R}({\bf a},t) \label{particle} \end{equation} does not satisfy any more the equation (\ref{vel}) and, consequently, the mapping Jacobian $J=\mbox{det}||\partial {\bf R}/\partial {\bf a}||$ is not assumed to equal 1, as it was for full Lagrangian description of incompressible fluids. It is easily to check that from condition $(\nabla _{{\bf a}}{\bf \Omega }% _{0}({\bf a}))=0$ it follows that divergence of (\ref{OmegaR}) is identically equal to zero. The gauge transformation \begin{equation} {\bf R}({\bf a})\rightarrow {\bf R}(\tilde{{\bf a}}_{\Omega _{0}}({\bf a})) \label{kalibr} \end{equation} leaves this integral unchanged if $\tilde{{\bf a}}_{\Omega _{0}}$ is arisen from ${\bf a}$ by means of arbitrary nonuniform translations along the field line of ${\bf \Omega }_{0}({\bf a})$. Therefore the invariant manifold $% {\cal M}_{\Omega _{0}}$ of the space ${\cal S}$, on which the variational principle holds, is obtained from the space ${\cal R}:{\bf a}\rightarrow {\bf R}$ of arbitrary continuous one-to-one three-dimensional mappings identifying ${\cal R}$ elements that are obtained from one another with the help of the gauge transformation (\ref{kalibr}) with a fixed solenoidal field $\ {\bf \Omega }_{0}({\bf a})$. The integral representation for ${\bf \Omega }$ (\ref{OmegaR}) is another formulation of the frozenness condition - after integration of the relation (% \ref{OmegaR}) over area $\sigma $, transverse to the lines of ${\bf \Omega }$% , follows that the flux of this vector remains constant in time: \[ \int_{\sigma (t)}({\bf \Omega },d{\bf S_{r}})=\int_{\sigma (0)}({\bf \Omega _{0}},d{\bf S_{a}}). \] Here $\sigma (t)$ is the image of $\sigma (0)$ under the transformation (\ref {particle}). It is important also that ${\bf \Omega _0(a)}$ can be expressed explicitly in terms of the instantaneous value of the vorticity and the mapping ${\bf % a=a(r},t)$, inverse to (\ref{particle}). By integrating over the variables $% {\bf a}$ in the relation (\ref{OmegaR}), \begin{equation} {\bf \Omega }({\bf R})=\frac{({\bf \Omega }_0({\bf a})\nabla _{{\bf a}}){\bf % R}({\bf a})}{\mbox{det}||\partial {\bf R}/\partial {\bf a}||}, \label{O_det} \end{equation} where ${\bf \Omega _0(a)}$ can be represented in the form: \begin{equation} {\bf \Omega _0}({\bf a})=\mbox{det}||\partial {\bf R}/\partial {\bf a}||(% {\bf \Omega }({\bf r})\nabla ){\bf a}. \label{C} \end{equation} This formula is nothing more than the Cauchy invariant (\ref{Cauchy}). We note that according to Eq. (\ref{O_det}) the vector ${\bf b=(\Omega }_0({\bf % a})\nabla _{{\bf a}}){\bf R}({\bf a})$ is tangent to ${\bf \Omega }({\bf R})$% . It is natural to introduce parameter $s$ as an arc length of the initial vortex lines ${\bf \Omega _0}({\bf a})$ so that \[ {\bf b}=\Omega _0(\nu )\frac{\partial {\bf R}}{\partial s}. \] In this expression $\Omega _0$ depends on the transverse parameter $\nu $ labeling each vortex line. In accordance with this, the representation (\ref {OmegaR}) can be written in the form \begin{equation} {\bf \Omega }({\bf r},t)=\int \Omega _0(\nu )d^2\nu \int \delta ({\bf r}-% {\bf R}(s,\nu ,t))\frac{\partial {\bf R}}{\partial s}ds, \label{rho} \end{equation} whence the meaning of the new variables becomes clearer: To each vortex line with index $\nu $ there is associated the closed curve \[ {\bf r}={\bf R}(s,\nu ,t), \] and the integral (\ref{rho}) itself is a sum over vortex lines. We notice that the parametrization by introduction of $s$ and $\nu $ is local. Therefore as global the representation (\ref{rho}) can be used only for distributions with closed vortex lines. To get the equation of motion for ${\bf R(\nu ,}s,t)$ the representation (% \ref{rho}) (in the general case - (\ref{OmegaR})) must be substituted in the Euler equation (\ref{dOmega_dt}) and then a Fourier transform with respect to spatial coordinates performed. As a result of simple integration one can obtain: \[ \left[ {\bf k}\times \int \Omega _{0}(\nu )d^{2}\nu \int dse^{-i{\bf kR}% }\lbrack {\bf R}_{s}\times \{{\bf R}_{t}(\nu,s,t)-{\bf v(R},t)\}\rbrack \right] =0. \] This equation can be resolved by putting integrand equal identically to zero: \begin{equation} \lbrack {\bf R}_{s}\times {\bf R}_{t}(\nu ,s,t)\rbrack =\lbrack {\bf R}_{s}% \times {\bf v(R},t)\rbrack . \label{R} \end{equation} With this choice there remains the freedom in both changing the parameter $s$ and relabelling the transverse coordinates $\nu$. In the general case of arbitrary topology of the field ${\bf \Omega _{0}(a)}$ the vector ${\bf R}_{s}$ in the equation (\ref{R}) must be replaced by the vector ${\bf b}=({\bf % \Omega }_{0}({\bf a})\nabla _{{\bf a}}){\bf R}({\bf a},t)$. Notice that, as it follows from (\ref{R}) and (\ref{O_det}), a motion of a point on the manifold ${\cal M}_{\Omega _{0}}$ is determined only by the transverse to $% {\bf \Omega }({\bf r})$ component of the generalized velocity. The obtained equation (\ref{R}) is the equation of motion for vortex lines. In accordance with (\ref{R}) the evolution of each vector ${\bf R}$ is principally transverse to the vortex line. The longitudinal component of velocity does not effect on the line dynamics. The description of vortex lines with the help of equations (\ref{rho}) and (% \ref{R}) is a mixed Lagrangian-Eulerian one: The parameter $\nu $ has a clear Lagrangian origin whereas the coordinate $s$ remains Eulerian. \setcounter{equation} {0} \section{Variational principle} The key observation for formulation of the variational principle is that the following general equality holds for functionals that depend only on $% {\bf \Omega }$: \begin{equation} \left[ {\bf b}\times \mbox{curl}\left( \frac{\delta F}{\delta {\bf \Omega }(% {\bf R})}\right) \right] =\frac{\delta F}{\delta {\bf R}({\bf a})}\Big|_{% {\bf \Omega }_{0}}. \label{peresch} \end{equation} For this reason, the right-hand-side of (\ref{R}) equals the variational derivative $\delta {\cal H}/\delta{\bf R}$: \begin{equation} \left[ ({\bf \Omega }_{0}({\bf a})\nabla _{{\bf a}}){\bf R}({\bf a})\times {\bf R}_{t}({\bf a})\right] =\frac{\delta {\cal H}\{{\bf \Omega }\{{\bf R}% \}\}}{\delta {\bf R}({\bf a})}\Big|_{{\bf \Omega }_{0}}. \label{main} \end{equation} It is not difficult to check now that the equation (\ref{main}) described dynamics of vortex line is equivalent to the requirement of extremum of the action ($% \delta S=0$) with the Lagrangian \cite{KR} \begin{equation} {\cal L}=\frac{1}{3}\int d^{3}{\bf a}(\lbrack {\bf R}_{t}({\bf a})\times {\bf R}({\bf a})\rbrack \cdot ({\bf \Omega }_{0}({\bf a})\nabla _{{\bf a}})% {\bf R}({\bf a}))-{\cal H}(\{{\bf \Omega }\{{\bf R}\}\}). \label{LAGRANGIAN} \end{equation} Thus, we have introduced the variational principle for the Hamiltonian dynamics of the divergence-free vector field topologically equivalent to $% {\bf \Omega }_{0}({\bf a})$. Let us discuss some properties of the equations of motion (\ref{main}), which are associated with excess parametrization of elements of ${\cal M}% _{\Omega _{0}}$ by objects from ${\cal R}$. We want to pay attention to the fact that From Eq. (\ref{peresch}) follows the property that the vector $% {\bf b}$ and ${\delta F}/{\delta {\bf R}({\bf a})}$ are orthogonal for all functionals defined on ${\cal M}_{\Omega _{0}}$. In other words the variational derivative of the gauge-invariant functionals should be understood (specifically, in (\ref{peresch})) as \[ \hat{P}\frac{\delta F}{\delta {\bf R}({\bf a})}, \] where $\hat{P}_{ij}=\delta _{ij}-\tau _{i}\tau _{j}$ is a projector and $% {\bf \tau }={\bf b/|b|}$ a unit tangent (to vortex line) vector. Using this property as well as the transformation formula (\ref{peresch}) it is possible, by a direct calculation of the bracket (\ref{bracket}), to obtain the Poisson bracket (between two gauge-invariant functionals) expressed in terms of vortex lines: \begin{equation} \{F,G\}=\int \frac{d^{3}{\bf a}}{|{\bf b}|^{2}}\left( {\bf b}\cdot \left[ \hat{P}\frac{\delta F}{\delta {\bf R}({\bf a})}\times \hat{P}\frac{\delta G}{% \delta {\bf R}({\bf a})}\right] \right) . \label{NonCanR} \end{equation} The new bracket (\ref{NonCanR}) does not contain variational derivatives with respect to ${\bf \Omega }_{0}({\bf a})$. Therefore, with respect to the initial bracket the Cauchy invariant ${\bf \Omega }_{0}({\bf a})$ is a Casimir fixing the invariant manifolds ${\cal M}_{\Omega _{0}}$ on which it is possible to introduce the variational principle (\ref{LAGRANGIAN}). In the case of the hydrodynamics of a superfluid liquid a Lagrangian of the form (\ref{LAGRANGIAN}) was apparently first used by Rasetti and Regge \cite {RR} to derive an equation of motion, identical to Eq. (\ref{R}), but for a separate vortex filament. Later, on the base of the results \cite{RR}, Volovik and Dotsenko Jr. \cite{VD} obtained the Poisson bracket between the coordinates of the vortices and the velocity components for a continuous distribution of vortices. The expression for these brackets can be extracted without difficulty from the general form for the Poisson brackets (\ref{NonCanR}) . However, the noncanonical Poisson brackets obtained in \cite{RR,VD} must be used with care. Their direct application gives for the equation of motion of the coordinate of a vortex filament an answer that is not gauge-invariant. For a general variation, which depends on time, additional terms describing flow along a vortex appear in the equation of motion. For this reason, the dynamics of curves (including vortex lines) is in principle "transverse" with respect to the curve itself. We note that for two-dimensional (in the $x-y$ plane) flows the variational principle for action with the Lagrangian (\ref{LAGRANGIAN}) leads to the well-known fact that $X(\nu ,t)$- and $Y(\nu ,t)$- coordinates of each vortex are canonically conjugated quantities (see \cite{lamb}). \setcounter{equation} {0} \section{Integrable hydrodynamics} Now we present an example of the equations of the hydrodynamic type (\ref {dOmega_dt}), for which transition to the representation of vortex lines permits to establish of the fact of their integrability \cite{KR}. Consider the Hamiltonian \begin{equation} {\cal H}\{{\bf \Omega (r)}\}=\int |{\bf \Omega }|d{\bf r} \label{hamilt} \end{equation} and the corresponding equation of frozenness (\ref{dOmega_dt}) with the generalized velocity \[ {\bf v}=\mbox{curl}\left( {\bf \Omega }/{\Omega }\right) . \] We assume that vortex lines are closed and apply the representation (\ref {rho}). Then due to (\ref{O_det}) the Hamiltonian in terms of vortex lines is decomposed as a sum of Hamiltonians of vortex lines: \begin{equation} {\cal H}\{{\bf R}\}=\int |\Omega _{0}(\nu )|d^{2}\nu \int \left| \frac{% \partial {\bf R}}{\partial s}\right| ds. \label{ham} \end{equation} The standing here integral over $s$ is the total length of a vortex line with index $\nu $. According to (\ref{main}), with respect to these variables the equation of motion for the vector ${\bf R}(\nu ,s)$ is local, it does not contain terms describing interaction with other vortices: \begin{equation} \eta \lbrack {\bf \tau }\times {\bf R}_{t}(\nu ,s,t)\rbrack ={\bf \lbrack \tau \times \lbrack \tau \times \tau} _{s}\rbrack \rbrack . \label{motion1} \end{equation} Here $\eta =\mbox{sign}(\Omega _{0})$, $ {\bf \tau =R}_{s}/|{\bf R}_{s}|$ is the unit vector tangent to the vortex line. This equation is invariant against changes $s\rightarrow \tilde{s}(s,t)$. Therefore the equation (\ref{motion1}) can be resolved relative to ${\bf R}_t$ up to a shift along the vortex line -- the transformation unchanged the vorticity ${\bf \Omega }$. This means that to find ${\bf \Omega }$ it is enough to have one solution of the equation \begin{equation} \eta |{\bf R}_{s}|{\bf R}_{t}={\bf \lbrack \tau \times \tau}_s \rbrack +\beta{\bf R}_{s}, \label{motion2} \end{equation} which follows from (\ref{motion1}) for some value of $\beta $. Arisen from here equation for ${\bf \tau }$ as a function of filament length $l$ ($dl=|% {\bf R}_{s}|ds$) and time $t$ (by choosing a new value $\beta =0$) reduces to the integrable one-dimensional Landau-Lifshits equation for a Heisenberg ferromagnet: \[ \eta \frac{\partial {\bf \tau }}{\partial t}=\left[ {\bf \tau }\times \frac{% \partial ^{2}{\bf \tau }}{\partial l^{2}}\right] . \] This equation is gauge-equivalent to the 1D nonlinear Schr\"{o}dinger equation \cite{ZT} and, for instance, can be reduced to the NLSE by means of the Hasimoto transformation \cite{Hasimoto}: \[ \psi (l,t)=\kappa (l,t)\cdot \exp (i\int^{l}\chi (\tilde{l},t)d\tilde{l}), \] where $\kappa (l,t)$ is a curvature and $\chi (l,t)$ the line torsion. The considered system with the Hamiltonian (\ref{hamilt}) has direct relation to hydrodynamics. As known (see the paper \cite{Hasimoto} and references therein), the local approximation for thin vortex filament (under assumption of smallness of the filament width to the characteristic longitudinal scale) leads to the Hamiltonian (\ref{ham}) but only for one separate line. Respectively, the equation (\ref{dOmega_dt}) with the Hamiltonian (\ref{hamilt}) can be used for description of motion of a few number of vortex filaments, thickness of which is small compared with a distance between them. In this case (nonlinear) dynamics of each filament is independent upon neighbor behavior. In the framework of this model singularity appearance ( intersection of vortices) is of an inertial character very similar to the wave breaking in gas-dynamics. Of course, this approximation does not work on distances between filaments comparable with filament thickness. It should be noted also that for the given approximation the Hamiltonian of vortex line is proportional to the filament line whence its conservation follows that, however, in no cases is adequate to behavior of vortex filaments in turbulent flows where usually process of vortex filament stretching takes place. It is desirable to have the better model free from this lack. A new model must necessarily describe nonlocal effects. In addition we would like to say that the list of equations (\ref{dOmega_dt}% ) which can be integrated with the help of representation (\ref{rho}) is not exhausted by (\ref{hamilt}). So, the system with the Hamiltonian \begin{equation} {\cal H}_\chi \{{\bf \Omega (r)}\}=\int |{\bf \Omega }|\chi d{\bf r} \label{hamilt_chi} \end{equation} is gauge equivalent to the modified KdV equation \[ \psi _t+\psi _{lll}+\frac 32|\psi |^2\psi _l=0\,\,\,- \] the second one after NLSE in the hierarchy generated by Zakharov-Shabat operator. As against previous model (\ref{hamilt}) some physical application of (\ref{hamilt_chi}) has not yet been found. \setcounter{equation}{0} \section{Lagrangian description of MHD} Consider now how the relabelling symmetry works in the ideal MHD. First, rewrite equations of motion (\ref{rho-t}-\ref{h-t}) in the Lagrangian representation by introducing markers ${\bf a}$ for fluid particles \[ {\bf x=x(a},t) \] with $$ {\bf v(x},t)={\dot {{\bf x}}}{\bf (a},t). $$ In this case the continuity equation (\ref{rho-t}) and the equation for magnetic field (\ref{h-t}) can be integrated. The density and the magnetic field are expressed in terms of the Jacoby matrix by means of Eq. (\ref{density}) and by the equation \begin{equation} B_i(x,t)=\frac{\partial x_i}{\partial a_k}B_{0k}(a)\,, \label{htr} \end{equation} where ${\bf B=h}/\rho .$ In the latter transformation the Jacoby matrix serves the evolution operator for vector ${\bf B.}$ The vector ${\bf B}$, in turn, transforms as a vector. In terms of Lagrangian variables the equation of motion (\ref{v-t}) is written as follows \begin{equation} \frac{\partial x_i}{\partial a_k}\ddot {x_i}=-\frac{\partial w(\rho )}{% \partial a_k}+\frac J{4\pi \rho _0({\bf a})} [\mbox{curl}\,{\bf h}\times {\bf h}]_i% \frac{\partial x_i}{\partial a_k} \label{motion3} \end{equation} With the help of relation (\ref{htr}) and Eq. (\ref{motion}) the vector $% {\bf u}$ given by (\ref{u}) will satisfy the equation \begin{equation} \frac{d{\bf u}}{dt}=\nabla \left( \frac{{\bf v}^2}2-w\right) -\frac 1{4\pi }\left[ {\bf B}_0({\bf a})\times \mbox{curl}_a{\bf H}\right] . \label{umhd} \end{equation} Here vector ${\bf B}_0({\bf a})={\bf h}_0({\bf a})/\rho_0({\bf a})$ is a Lagrangian invariant and ${\bf H}$ represents the co-adjoint transformation of the magnetic field, analogous to (\ref{u}): \[ H_i(a,t)=\frac{\partial x_m}{\partial a_i}h_m(x,t). \] Now by analogy with (\ref{motion}) and (\ref{weber}), integration of Eq.(\ref{umhd}) over time leads to the Weber type transformation: \begin{equation} {\bf u(a,}t)={\bf u}_0{\bf (a})+\nabla _a\Phi +\left[ {\bf B}_0({\bf a}% )\times \mbox{curl}_a\tilde {{\bf S}}\right] . \label{webermhd} \end{equation} Here ${\bf u}_0{\bf (a})$ is a new Lagrangian invariant which can be chosen as pure transverse, namely, with $\mbox{div}_a$ ${\bf u}_0=0.$ This new Lagrangian invariant cannot be expressed through the observed physical quantities such as magnetic field, velocity and density. In spite of this fact, as it will be shown in the next section, the vector Lagrangian invariant ${\bf u}_0{\bf (a})$ has a clear physical meaning. As far as new variables $\Phi $ and $\tilde {{\bf S}}$, they obey the equations: \begin{eqnarray*} \frac{d\Phi }{dt} &=&\frac{{\bf v}^2}2-w, \\ \frac{d\tilde {{\bf S}}}{dt} &=&-\frac{{\bf H}}{4\pi }+\nabla _a\psi . \end{eqnarray*} The transformation (\ref{webermhd}) for velocity ${\bf v(x},t)$ takes the form: \begin{equation} {\bf v=}u_{0k}{\bf (a})\nabla a_k+\nabla \Phi +\left[ \frac{{\bf h}}\rho \times \mbox{curl}\,{\bf S}\right] \label{ZK} \end{equation} where ${\bf S}$ is the vector $\tilde {{\bf S}}$ transformed by means of the rule (\ref{u}): \[ S_i(x,t)=\frac{\partial a_k}{\partial x_i}\,\tilde S_k(a,t). \] In Eulerian description $\Phi $ satisfies the Bernoulli equation \begin{equation} \frac{\partial \Phi }{\partial t}+({\bf v\nabla })\Phi -\frac{{\bf v}^2}2+w=0 \label{bern} \end{equation} and equation of motion for ${\bf S}$ is of the form: \begin{equation} \frac{\partial {\bf S}}{\partial t}+\frac{{\bf h}}{4\pi }{\bf -}\left[ {\bf v% }\times \mbox{curl}{\bf S}\right] +\nabla \psi _1=0. \label{S} \end{equation} For ${\bf u}_0=0$ the transformation (\ref{ZK}) was introduced for ideal MHD by Zakharov and Kuznetsov in 1970 \cite{ZK70}. In this case magnetic field $% {\bf h}$ and vector ${\bf S}$ as well as $\Phi $ and $\rho $ are two pairs of canonically conjugated variables. It is interesting to note that in the canonical case the equations of motion for ${\bf S}$ and $\Phi $ obtained in \cite{ZK70} coincide with (\ref{bern}) and (\ref{S}). However, the canonical parametrization describes partial type of flows, in particular, it does not describe topological nontrivial flows for which mutual knottiness between magnetic and vortex lines is not equal to zero. This topological characteristics is given by the integral $\int ({\bf v,h})d{\bf x.}$ Only when ${\bf u}_0\neq 0$ this integral takes non-zero values. \setcounter{equation} {0} \section{Frozen-in MHD fields} To clarify meaning of new Lagrangian invariant ${\bf u_0(a)}$ we remind that the MHD equations (\ref{rho-t}-\ref{h-t}) can be obtained from two-fluid system where electrons and ions are considered as two separate fluids interacting each other by means of self-consistent electromagnetic field. The MHD equations follow from two-fluid equations in the low-frequency limit when characteristic frequencies are less than ion gyro-frequency. The latter assumes i) neglecting by electron inertia, ii) smallness of electric field with respect to magnetic field, and iii) charge quasi-neutrality. We write down at first some intermediate system called often as MHD with dispersion \cite{karpman}: \begin{equation} \mbox{curl}\ \mbox{curl}{\bf A}=\frac{4\pi e}c(n_1{\bf v}_1-n_2{\bf v}_2), \label{varA} \end{equation} \begin{equation} (\partial _t+{\bf v}_1\nabla )m{\bf v}_1=\frac ec\left( -{\bf A}_t+[{\bf v}% _1\times \mbox{curl}\ {\bf A}]\right) -\nabla \frac{\partial \varepsilon }{% \partial n_1}, \label{var1} \end{equation} \begin{equation} 0=-\frac ec\left( -{\bf A}_t+[{\bf v}_2\times \mbox{curl}\ {\bf A}]\right) -\nabla \frac{\partial \varepsilon }{\partial n_2}. \label{var2} \end{equation} In these equations ${\bf A}$ is the vector potential so that the magnetic field ${\bf h}=\mbox{curl}{\bf A}$ and electric field ${\bf E=-\frac 1cA_t}$. This system is closed by two continuity equations for ion density $n_1$ and electron density $n_2$: \begin{equation} n_{1,t}+\nabla (n_1{\bf v}_1)=0,\qquad n_{2,t}+\nabla (n_2{\bf v}_2)=0. \label{nepr2} \end{equation} In this system ${\bf v}_{1,2}$ are velocities of ion and electron fluids, respectively. The first equation of this system is a Maxwell equation for magnetic field in static limit. The second equation is equation of motion for ions. The next one is equation of motion for electrons in which we neglect by electron inertia. By means of the latter equation one can obtain the equation of frozenness of magnetic field into electron fluid (this is another Maxwell equation): \[ {\bf h}_t=\mbox{curl}[{\bf v}_2\times {\bf h}]. \] Applying the operator $\mbox{div}$ to (\ref{varA}) gives with account of continuity equations the quasi-neutrality condition: $n_1=n_2=n$. Next, by excluding $n_2$ and ${\bf v}_2$ we have finally the MHD equations with dispersion in its standard form \cite{karpman}: \[ (\partial _t+{\bf v}\nabla )m{\bf v}=-\nabla w(n)+\frac 1{4\pi n}[\mbox{curl}% \ {\bf h}\times {\bf h}],\qquad w(n)=\frac \partial {\partial n}\varepsilon (n,n), \] \begin{equation} n_t+\nabla (n{\bf v})=0,\qquad {\bf h}_t=\mbox{curl}\left[ \left( {\bf v}% -\frac c{4\pi en}\mbox{curl}\ {\bf h}\right) \times {\bf h}\right] , \label{MHD2} \end{equation} where ${\bf v}_1={\bf v}$, and $\varepsilon(n,n)$ is internal energy density so that $w(n)$ is entalpy per one pair ion-electron. The classical MHD follows from this system in the limit when the last term $% c/(4\pi en)\mbox{curl}\,{\bf h}$ in equation (\ref{MHD2}) should be neglected with respect to ${\bf v}$. At the same time, the vector potential ${\bf A}$ must be larger characteristic values of $(mc/e){\bf v}$ in order to provide inertia and magnetic terms in Eq. (\ref{var1}) being of the same order of magnitude. Both requirements are satisfied if $\epsilon =c/(\omega_{pi}L)<<1$ where $L$ is a characteristic scale of magnetic field variation and $\omega _{pi}=\sqrt{4\pi ne^2/m}$ is ion plasma frequency. Unlike MHD equations (\ref{rho-t}-\ref{h-t}), the given system has two frozen-in fields. These are the field ${\bf \Omega }_2=-\frac e{mc}{\bf h}$ frozen into electron fluid and the field $$ {\bf \Omega }_1=\mbox{curl}({\bf v}% +\frac e{mc}{\bf A})={\bf \Omega }-{\bf \Omega }_2 $$ frozen into ion component: \[ {\bf \Omega }_{1t}=\mbox{curl}\left[ {\bf v\times \Omega }_1\right] , \] \[ {\bf \Omega }_{2t}=\mbox{curl}\left[ {\bf v}_2{\bf \times \Omega }_2\right] \] where $$ {\bf v}_2={\bf v}-\frac c{4\pi en}\mbox{curl}\ {\bf h.} $$ Hence for both fields one can construct two Cauchy invariants by the same rule (\ref{Cauchy}) as for ideal hydrodynamics: \begin{equation} {\bf \Omega _{10}(}{\bf a)}=J_1({\bf \Omega _1(x},t)\nabla ){\bf a(x,}t) \label{C1} \end{equation} where ${\bf a(x,}t)$ is inverse mapping to ${\bf x=x}_1{\bf (a},t)$ which is solution of the equation ${\bf \dot x}{={\bf v(x,}}t);$ \begin{equation} {\bf \Omega _{20}(}{\bf a}_2)=J_2({\bf \Omega _2(x},t)\nabla ){\bf a}_2{\bf (x,% }t) \label{C2} \end{equation} with ${\bf a}_2{\bf (x,}t)$ inverse to the mapping ${\bf x=x}_2{\bf (a}_2,t)$ and ${\bf \dot x=v_2(x,}t).$ In order to get the corresponding Weber transformation for MHD as a limit of the system it is necessary to introduce two momenta for ion and electron fluids: \begin{eqnarray} {\bf p}_1 &=&m{\bf {v}}+\frac ec{\bf {A}} \\ {\bf p}_2 &=&-\frac ec{\bf A}. \end{eqnarray} In these expressions the terms containing the vector potential are greater sum of ${\bf p}_1$ and ${\bf p}_2$ in parameter $\epsilon $. For each momentum in Lagrangian representation one can get equations, analogous to (\ref{Lag}), (\ref{motion}): \begin{eqnarray} \frac{\partial x_k}{\partial a_{1i}}\frac{dp_{1k}}{dt}&=& -p_{1k}\frac{\partial v_k}{\partial a_{1i}}+ \frac \partial {\partial a_{1i}} \left(-\frac{\partial\varepsilon}{\partial{n_1}} +\frac{e}{c}({\bf v\cdot A})+m\frac{v^2}2\right)\\ \frac{\partial x_k}{\partial a_{2i}}\frac{dp_{2k}}{dt}&=& -p_{2k}\frac{\partial v_{2k}}{\partial a_{2i}}+ \frac \partial {\partial a_{2i}} \left(-\frac{\partial\varepsilon}{\partial{n_2}} -\frac{e}{c}({\bf v}_2\cdot{\bf A})\right). \end{eqnarray} By introducing the vector $\tilde{{\bf p}}$ for each type of fluids, by the same rule as (\ref{u}), \[ \tilde{p}_i=\frac{\partial x_k}{\partial a_i}p_k, \] after integration over time of equations of motion for $\tilde{{\bf p}}$ one can arrive at two Weber transformations for each momentum: \begin{eqnarray} {\bf p}_1 &=&\tilde p_{1i}(a_1)\nabla a_{1i}+\nabla \Phi _1, \label{p1} \\ {\bf p}_2 &=&\tilde p_{2i}(a_2)\nabla a_{2i}+\nabla \Phi _2. \label{p2} \end{eqnarray} In the limit $\epsilon \rightarrow 0$ the markers ${\bf a_1}$ and ${\bf a}_2$ can be put approximately equal. This means that their difference will be small: \[ {\bf a}_2-{\bf a_1}={\bf d \, \sim }\epsilon . \] Besides, due to charge quasi-neutrality, Jacobians with respect to $a_1$ and $a_2$ must be equal each other (here we put $n_{10}({\bf a_1})=n_{20}({\bf a}_2)=1$ without loss of generality): \[ \mbox{det}||{\partial {\bf x}}/{\partial {\bf a_1}}||= \mbox{det}||{\partial {\bf x}} /{\partial {\bf a_2}}||. \] As a result, the infinitesimal vector ${\bf d(a,}t)$ relative to the argument ${\bf a}$ occurs divergence free: $\partial d_i/\partial a_i=0.$ Then, summing (\ref{p1}) and (\ref{p2}) and considering the limit $\epsilon \rightarrow 0$, we obtain the Weber-type transformation coinciding with (\ref{webermhd}): \begin{equation} {\bf u(a,}t)={\bf u}_0{\bf (a})+\nabla _a\Phi +\left[ {\bf B}_0({\bf a}% )\times \mbox{curl}_a\,\tilde {{\bf S}}\right] , \label{sum} \end{equation} where vectors ${\bf u}_0{\bf (a})$ and $\tilde {{\bf S}}$ are expressed through the Lagrangian invariants $\tilde{\bf p}_1{\bf (a})$ and $\tilde{\bf p}_2{\bf (a})$ and displacement ${\bf d}$ between electron and ion by means of relations \cite{ruban}: \begin{eqnarray*} {\bf u(a,}t)&=&\frac 1m(\tilde{\bf p}_1{\bf (a})+\tilde{\bf p}_2{\bf (a})), \\ {\bf d}&=&-\frac{mc}e\mbox{curl}_a\tilde {{\bf S}}. \end{eqnarray*} Important that in (\ref{sum}) all terms are of the same order of magnitude (zero order relative to $% \epsilon) .$ Curl of vectors $\tilde{\bf p}_1{\bf (a})$ and $\tilde{\bf p}_2{\bf (a}_2)$ yield the corresponding Cauchy invariants (\ref{C1}) and (\ref{C2}). \setcounter{equation} {0} \section{Relabeling symmetry in MHD} Now let us show how existence of new Lagrangian invariants corresponds to the relabeling symmetry. Consider the MHD Lagrangian \cite{ZK}, \[ {\cal L}_*=\int\left(\rho\frac{{\bf v}^2}{2}-\rho\tilde\varepsilon(\rho) -% \frac{{\bf h}^2}{8\pi}\right)d{\bf r}, \] where we neglect by contribution from electric field in comparison with that from magnetic field. Here $\tilde\varepsilon(\rho)$ is specific internal energy. In terms of mapping ${\bf x(a},t)$ the Lagrangian $% {\cal L}_*$ is rewritten as follows \cite{IP}: \begin{equation} \label{L*} {\cal L}_*=\int\frac{{\bf \dot x}^2}{2}d^3{\bf a}- \int \tilde\varepsilon(J_{% {\bf x}}^{-1}({\bf a}))d^3{\bf a} -\frac{1}{8\pi} \int\left(\frac{({\bf % h_0(a)\nabla_a)x}}{J_{{\bf x}}({\bf a})}\right)^2 J_{{\bf x}}({\bf a})d^3% {\bf a}. \end{equation} Here density and magnetic field are expressed by means of relations $$ \rho={1}/{J_{\bf x}},\qquad {\bf h}=({\bf h}_0({\bf a})\nabla_a){\bf x}/J_{\bf x}, $$ and $$ J_x({\bf a},t)= \mbox{det}||\partial {\bf x}/\partial {\bf a}|| $$ is the Jacobian of mapping ${\bf x=x(a,}t)$ and initial density is put to equal 1. Notice, that variation of the action with by the Lagrangian (\ref{L*}) relative to ${\bf x(a)}$ gives the equation of motion (\ref{motion3}) (or the equivalent equation for vector ${\bf u}$ (\ref{umhd})). Due to the presence of magnetic field in the Lagrangian (\ref{L*}), the relabeling symmetry, in comparison with ideal hydrodynamics, reduces. If the first two terms in (\ref{L*}) are invariant with respect to all incompressible changes ${\bf a\rightarrow a(b)}$ with $J|_{b}=1$, invariance of the last term, however, restricts the class of possible deformations up to the following class $$ ({\bf h_0(a)\nabla_a)b=h_0(b)}. $$ For infinitesimal transformations $$ {\bf a\rightarrow\ a+\tau g(a)} $$ where $\tau$ is a (small) group parameter the vector ${\bf g}$ must satisfy two conditions: \begin{equation} \label{condition} \mbox{div}_a{\bf g}=0, \,\,\, \mbox{curl}_a[{\bf g\times h_0}]=0. \end{equation} The first condition is the same as for ideal hydrodynamics, the second one provides conservation of magnetic field frozenness. The conservation laws generating by this symmetry, in accordance with Noether theorem, can be obtained by standard scheme from the Lagrangian (\ref{L*}). They are written through the infinitesimal deformation ${\bf g(a)}$ as integral over ${\bf a}$: \begin{equation} \label{conserv} I=\int({\bf u,g(a)}) d{\bf a} \end{equation} where the vector ${\bf u}$ is given by (\ref{u}). Putting ${\bf g=h_0}$ from this (infinite) family of integrals one gets the simplest one $$ I_{ch}=\int {\bf (v,h)}d{\bf r} $$ which represents a cross-helicity characterizing degree of mutual knottiness of vortex and magnetic lines. The conservation laws (\ref{conserv}) are compatible with the Weber-type transformation. Really, substituting (\ref{webermhd}) into (\ref{conserv}) and using (\ref{condition}) one leads to the relation $$ \int({\bf u_0(a),g(a)}) d{\bf a}. $$ Hence conservation of (\ref{conserv}) also follows. Note that if one would not suppose an independence of ${\bf u_0}$ on $t$ then, due to arbitrariness of ${\bf g(a)}$, this could be considered as independent verification of conservation of solenoidal field ${\bf u_0}$: $$ \frac d{dt}{\bf u_0} =0. $$ The MHD equations expressed in terms of Lagrangian variables become Hamiltonian ones, as in usual mechanics, for momentum ${\bf p=\dot{{\bf x}}}$ and coordinate ${\bf x}$. These variables assign the canonical Poisson structure. In the Eulerian representation the MHD equations can be written also in the Hamiltonian form \cite{MG}: $$ \rho_t=\{\rho, H \},\,\,\, {\bf v}_t=\{{\bf v}, H \},\,\,\, {\bf h}_t=\{{\bf h}, H \}, $$ where noncanonical Poisson bracket $\{F,G\}$ is given by the expression (\ref{compres_bracket}). As for ideal hydrodynamics, this Poisson bracket occurs to be degenerated. For example, the cross helicity $I_{ch}$ serves a Casimir for the bracket (\ref{compres_bracket}). The reason of the Poisson bracket degeneracy is the same as for one-fluid hydrodynamics - it is connected with a relabeling symmetry of Lagrangian markers. For incompressible case the Poisson bracket (\ref{compres_bracket}) reduces so that it can be expressed only through magnetic field ${\bf h}$ and vorticity ${\bf \Omega}$: \begin{equation} \label{brack} \{F,G\}= \int\left( \frac{{\bf h}}{\rho}\cdot \left(\left[ \mbox{curl}\frac{% \delta F}{\delta{\bf h}}\times \mbox{curl}\frac{\delta G}{\delta{\bf \Omega}} \right] - \left[ \mbox{curl}\frac{\delta G}{\delta{\bf h}}\times \mbox{curl} \frac{\delta F}{\delta% {\bf \Omega}}\right] \right)\right)d^3{\bf r} \end{equation} \[ +\int\left( {\bf\Omega }\left[ \mbox{curl}\frac{\delta F}{\delta {\bf\Omega}}\times \mbox{curl}\frac{\delta G}{\delta{\bf \Omega}}\right] \right)d^3 {\bf r}. \] This bracket remains also degenerated. \setcounter{equation}{0} \section{Variational principle for incompressible MHD} By analogy with incompressible hydrodynamics, one can introduce magnetic line representation: \begin{equation} \label{hR} {\bf h}({\bf r},t)=\int\delta({\bf r}-{\bf R}({\bf a},t)) ({\bf h}_0({\bf a}% )\nabla_{{\bf a}}){\bf R}({\bf a},t)d^3{\bf a}. \end{equation} For vorticity the analog of vortex line parametrization (\ref{OmegaR}) can be obtained, for instance, as a limit $\epsilon\rightarrow 0$ of the corresponding representations for the two-fluid system. Calculations give \cite{ruban}: \begin{equation} \label{OmegaR1} {\bf \Omega}({\bf r},t)=\int\delta({\bf r}-{\bf R}({\bf a},t)) (({\bf \Omega}% _0({\bf a})+ \mbox{curl}_{{\bf a}} [{\bf h}_0({\bf a})\times{\bf U}({\bf a}% ,t)]) \nabla_{{\bf a}}){\bf R}({\bf a},t)d^3{\bf a}, \end{equation} Here the field ${\bf {U}}({\bf {a}},t)$ is not assumed solenoidal, as well as the Jacobian of mapping ${\bf r=R(a,}t)$ is not equal to unity. From the corresponding limit of the two-fluid system to incompressible MHD it is possible also to get the expression for Lagrangian \begin{equation} \label{LAGRANGIAN1} L=\int d^3{\bf a} ([({\bf h}_0\nabla_{{\bf a}}){\bf R}\times ({\bf U}\nabla_{% {\bf a}}){\bf R}]\cdot{\bf R}_t)+ \end{equation} \[ +{1}/{3} \int d^3{\bf a} ([{\bf R}_t \times {\bf R}]\cdot ({\bf \Omega}% _0\nabla_{{\bf a}}){\bf R}) -{\cal H}\{{\bf \Omega}\{{\bf R,U}\},{\bf h}\{% {\bf R}\}\}. \] The Hamiltonian of the incompressible MHD ${\cal H}_{MHD}$ in terms of ${\bf {U}}({\bf {a}},t)$ and ${\bf R(a,}t)$ takes the form \[ {\cal H}_{MHD}= \frac{1}{8\pi}\int\frac {(({\bf h}_0({\bf {a}})\nabla_{{\bf {% a}}}){\bf R}({\bf {a}}))^2} {\mbox{det}||\partial{\bf {R}}/\partial{\bf {a}}% ||}d^3{\bf {a}}+ \] \begin{equation} \label{H_MHD} +\frac{1}{8\pi}\int\int \frac{ (({\bf \Omega}({\bf {a}}_1)\nabla_1){\bf R}(% {\bf {a}}_1)\cdot ({\bf \Omega}({\bf {a}}_2)\nabla_2){\bf R}({\bf {a}}_2))} {|{\bf {R}}({\bf {a}}_1)-{\bf {R}}({\bf {a}}_2)|}d^3{\bf {a}}_1d^3{\bf {a}}% _2, \end{equation} where we introduce the notation \[ {\bf {\Omega}}({\bf {a}},t)={\bf \Omega}_0({\bf a})+ \mbox{curl}_{{\bf a}}[{\bf % h}_0({\bf a})\times{\bf U}({\bf a},t)]. \] Equations of motion for ${\bf {U}}$ and ${\bf {R}}$ follow from the variational principle for action with Lagrangian (\ref{LAGRANGIAN1}): \begin{equation} \label{var_U} \left[({\bf h}_0\nabla_{{\bf a}}){\bf R}\times {\bf R}_t\right]\cdot ({% \partial{\bf R}}/{\partial a_{\lambda}})= -{\delta {\cal H}}/{\delta U_{\lambda}}, \end{equation} \begin{equation} \label{var_R} [({\bf \Omega(a,}t) \nabla_{% {\bf a}}){\bf R}\times {\bf R}_t] -[({\bf h}_0\nabla_{{\bf a}}){\bf R}\times ({\bf U}_t\nabla_{{\bf a}}){\bf R}] ={\delta {\cal H}}/{\delta{\bf R}}. \end{equation} These equations can be obtained also directly from the MHD system (\ref{rho-t}-\ref{h-t}) by the same scheme as it was done for ideal hydrodynamics. Thus, we have variational principle for the MHD-type equations for two solenoidal vector fields. Their topological properties are fixed by ${\bf \Omega}_0({\bf % a})$ and ${\bf h}_0({\bf a})$. These quantities represent Casimirs for the initial Poisson bracket (\ref{brack}). It is worth noting that the obtained equations of motion have the gauge invariant form. This gauge invariance is a remaining symmetry connected with relabeling of Lagrangian markers of magnetic lines in two-dimensional manifold which can be specified always locally. Coordinates of this manifold enumerate magnetic lines. This symmetry leads to conservation of volume of magnetic tubes including infinitesimally small magnetic tubes, namely, magnetic lines. This property explains why the Jacobian of the mapping ${\bf r=R(a,}t)$ can be not equal identically to unity. \section*{Acknowledgments} Authors thank A.B.Shabat for useful discussion of the connection between NLSE and equations (\ref{dOmega_dt}), that resulted in integrability declaration for (\ref{hamilt_chi}). This work was supported by the Russian Foundation of Basic Research under Grant no. 97-01-00093 and by the Russian Program for Leading Scientific Schools (grant no.96-15-96093). Partially the work of E.K. was supported by the Grant INTAS 96-0413, and the work of V.R. by the Grant of Landau Scholarship. \newpage
\section{ A lower limit to the production rate of Be-neutrinos} The B-neutrinos from \begin{eqnarray} \label{pcap} ^7Be+ p \rightarrow & ^8B + \gamma \quad \qquad \qquad \qquad\nonumber \\ & ^8B \rightarrow \, 2\alpha + e^+ + \nu_e \quad , \end{eqnarray} have been observed by Kamiokande\cite{kam} and Superkamiokande \cite{skam}. Since B-neutrinos and Be-neutrinos are both sons of $^7Be$ nuclei, one expects that detection of the former gives information on the latter. Our aim is to determine a lower limit on the production rate of Be-neutrinos, $L(Be)$, starting from this consideration. As the rates of (\ref{ecap}) and (\ref{pcap}) depend differently on the solar temperature, we need some information on it. This is (indirectly) provided by helioseismology, which determines the sound speed with an accuracy of one per cent or better, even close to the solar center, see e.g. \cite{eliosnoi}. Temperature is obtained from the sound speed if the chemical composition of the solar plasma is known. This is the only information which we shall take from SSMs, in the form of the hydrogen abundance at the solar centre $X_c$, a quantity which is largely independent of solar models, since it reflects the amount of hydrogen burnt all along the sun history. Let us make this argument in some detail. The $^7Be$ and $^8B$ luminosities, at production, can be written as: \begin{eqnarray} \label{eq1} L(B)&=& \int d^3r\, n_1 n_7 <\sigma v>_{17} = \lambda_{17}/ T_o^{13} \int d^3r \, n_1 n_7 T^{13} \\ \label{eq2} L(Be)&=& \int d^3r\, n_e n_7 <\sigma v>_{e7} = \lambda_{e7}/ T_o^{-0.5} \int d^3r \, n_e n_7 T^{-0.5} \end{eqnarray} where we have used a parametrization of the form $<\sigma v>_{ij}=\lambda_{ij}(T/ T_o)^{\alpha_{ij}}$, the temperature scale $T_o$ is chosen as the central temperature of the model in \cite{BP98}, hereafter BP98, $T_o=1.5697\cdot 10^7$ K, and according to \cite{adelberger}: \begin{eqnarray} \label{eql1} \lambda_{e7}&=& 2.34 \cdot 10^{-33} \, (1 \pm 2\%) {\mbox{cm$^3$ s$^{-1}$}} \quad \\ \label{eql2} \lambda_{17}&=& 1.04 \cdot 10^{-35} \, (1 \pm 16\%) {\mbox{cm$^3$ s$^{-1}$}}\quad \end{eqnarray} (here and in the following errors shown are combinations in quadrature of systematic and statistical $1\sigma$ errors). By using eqs. (\ref{eq1}) and (\ref{eq2}) one can relate the production rates of Boron and Beryllium neutrinos. From eq. (\ref{eq1}) one has: \begin{equation} \label{eqlb} L(B) = \frac{\lambda_{17}}{ \lambda_{e7}} T_o^{-13.5} \int d^3r n_e n_7 <\sigma v>_{e7} T^{13.5} x \end{equation} where $x=n_1/n_e$ is the ratio of free protons to electrons. Since temperature decreases sharply when moving away from the solar center, one can assume that at any point in the solar interior $x(r)T(r)^{13.5}\leq x_c T_c^{13.5}$, where, here and in the following, the suffix $c$ refers to the solar center. In this way one has: \begin{equation} \label{eqlb2} L(B) \leq \frac{\lambda_{17}}{ \lambda_{e7}} \left ( \frac{T_c}{T_o} \right )^{13.5} x_c \, L(Be) \, . \end{equation} As the produced $\nu_e$ can oscillate into species with a smaller or vanishing cross section in the detector, the observed luminosity $L(B)_{obs}$ in Kamiokande and Superkamiokande cannot exceed the produced luminosity, $L(B)_{obs} \leq L(B)$, so that one has the following lower limit for $L_{Be}$: \begin{equation} \label{eqlbe} L(Be) \geq L(B)_{obs} \frac{\lambda_{e7}}{\lambda_{17}} \left ( \frac{T_c}{T_o} \right )^{13.5} \frac{1}{x_c} \,. \end{equation} Now we use the fact that the solar center can be described as perfect gas of fully ionized H and He, to a very good approximation. In terms of the isothermal squared sound speed $u=P/\rho$ and of the hydrogen mass fraction $X$, this gives: $ kT_c = u_c m_p /(3/4 +5/4 X_c)$. One also has: $ x_c = 2X_c/(X_c + 1) $. In this way one gets: \begin{equation} \label{eqlbe2} L(Be) \geq \frac{\lambda_{e7}}{\lambda_{17}} \left (\frac{kT_o}{m_pu_c} \right )^{13.5} \frac{(X_c+1)}{(2X_c)} (5/4 X_c +3/4)^{13.5} L(B)_{obs} \end{equation} The equation above can of course be translated in terms of fluxes. We take $ \Phi(B)_{obs}=2.42 \cdot (1 \pm 3\% )\cdot 10^6$ cm$^{-2}$ s$^{-1}$ from \cite{skam} and $u_c=1.53 \cdot (1\pm 1\%) \cdot 10^{15} cm^2/s^2$, in agreement with helioseismic determinations, see \cite{eliosnoi}. For the central hydrogen abundance we have to rely on solar model calculations. Recent SSM calculations all yield $X_c$ in the narrow range $0.333<X_c<0.347$, with a mean value close to the BP98 estimate, $X_c^{BP98}=0.339$, see table \ref{tabxc}. The calculated value of $X_c$ is sensitive to opacity, metal abundance and nuclear cross sections, see table \ref{tabxc2}. The $1\sigma$ uncertainty on opacity and metal abundance are respectively 5\% and 6\%, according to \cite{report} and \cite{BP95}, and that on $S_{pp}$ is 1.7\%, from \cite{adelberger}. By computing suitable solar models and adding errors in quadrature, we conclude: \begin{equation} \label{xcerror} X_c=0.339 \pm 0.010 \quad . \end{equation} We have also computed $X_c$ for a series of ``non standard solar models'', where some input parameters have been varied, one at the time, by about $\pm 3 \sigma$ with respect to the SSM reference input, see table \ref{tabxc}. Even in this case, $X_c$ stays in the range 0.329 -- 0.358, i.e. within about $\pm 5\%$ from BP98. We remark that $X_c$ is essentially an indicator of how much hydrogen has been burnt so far, starting from an initial value $X_{in}$ about 0.7. The stability of $ X_c$ corresponds to the fact that any solar model has to account for an integrated solar luminosity of about $L_{\odot}t_{\odot} =5.5 \cdot 10^{50}$ erg. On these grounds, we consider the adopted value of $X_c$ as rather safe. In this way we get: \begin{equation} \label{eqfibe} \Phi (Be) \geq (1\pm 0.24) \cdot 10^9 {\mbox{cm$^{-2}$ s$^{-1}$}} \, . \end{equation} where the error include, in quadrature, all uncertainties mentioned above. The uncertainty on $\lambda_{17}$, $X_c$ and $u_c$ contribute to the total error 16\%, 13\% and 12\% respectively. The inequality (\ref{eqfibe}) defines a minimum flux $\Phi(Be)_{min} \simeq 1\cdot 10^9$ cm$^{-2}$ s$^{-1}$, which is one fifth of the SSM prediction, see eq. (\ref{eq_bessm}). We note that the only input from SSM is the value of $X_c$, whereas all other inputs, $\Phi(B)_{obs}$ and $u_c$, are from observational data. We have obtained this minimal flux using only the physical information that is relevant to the ratio between the Beryllium and Boron luminosities $L(Be) / L(B)$; for instance, the actual value of $n_7(r)$ never matters for our result. However, the additional physical information that determines the two fluxes separately, in particular the measured $\Phi(B)$, can only strengthen this limit. In fact, a solar-model-independent analysis of the Beryllium and Boron neutrino flux production ($0<T_{c}<\infty$, cross sections more than $3\sigma$'s away from the central values, profiles of densities that are not constrained by helioseismology varied by factors larger than 30) shows that the lower limit to $\Phi (Be)$ is $1.6\times 10^{9}$~cm$^{-2}$~s$^{-1}$, as can be inferred from Fig.~3 of Ref.~\cite{hemix} and the measured $\Phi(B)$. \section{ Implications for solar neutrino experiments} \subsection{Borexino and LENS} The relevance of the bound (\ref{eqfibe}) can be appreciated when discussing the complementarity between Borexino and LENS. We remind that LENS is sensitive to $\nu_e$ only, whereas the signal of Borexino can get contribution also from $\nu_\mu$ or $\nu_\tau$, their cross section being about 1/5 than that of $\nu_e$. For both experiments, a signal well below 1/5 of the SSM prediction will be a definite proof of neutrino oscillations, since it leads to a violation of eq. (\ref{eqfibe}). A signal at the level of 1/5 of the SSM prediction in Borexino could be interpreted as due to Small Mixing (SM) angle oscillations into active neutrinos, where one expects that all $\nu_e$ from Beryllium have been transformed into $\nu_\mu$. However, one could still insist on standard neutrinos, arguing for some drastic (maybe desperate) modification of the solar model. In this situation, a clear discrimination will be provided by LENS: for the SM case, the LENS signal, barrying the background, has to vanish, so that the bound (\ref{eqfibe}) will be violated, giving a definite proof of neutrino oscillations (furthermore, the comparison with Borexino will show the presence of $\nu_{\mu}$ or $\nu_{\tau}$). \subsection{Gallium experiments} As well known, the solar luminosity essentially fixes the total production rate of neutrinos. Since neutrino cross sections increase with energy, the minimal Gallium signal, in the absence of oscillations, can be estimated by assuming that the total flux consists of pp neutrinos only. The pep-neutrinos can be safely included in this estimate, as the ratio of pep to pp-neutrinos is well known and essentially unsensitive to solar physics details, see \cite{primo} and \cite{report}. By using updated cross sections from \cite{cross}, this arguments gives as a minimal Gallium signal in the absence of oscillations: \begin{equation} \label{eqsgalum} G_{Lum} = (79.5 \pm 2.0) \, {\mbox{SNU}} \end{equation} where the error arises mainly from the capture cross section of pp neutrinos. As well known, also B-neutrinos contribute to the Gallium signal. Their contribution is best estimated by using experimental data. If one takes into account the flux measured by Superkamiokande, with the capture cross section of \cite{cross}, this contributes an additional $(5.8 \pm 1.5)$ SNU, where most of the error comes again from the capture cross section. All this results in: \begin{equation} \label{sgalum2} G_{Lum+SK}= (85.3 \pm 2.5) \, {\mbox{SNU}} \, . \end{equation} According to the previous discussion, one has to include now the minimal contribution of $^7Be$ neutrinos. For $\Phi(Be)=\Phi(Be)_{min}$ by using the luminosity constraint (see section 2.4 of \cite{report}) one has an additional contribution of $(5.9 \pm 1.4)$ SNU, where most of the error comes from $\lambda_{17}$, so that in conclusion the minimal Gallium signal is now: \begin{equation} \label{eqsgamin} G_{min}= (91 \pm 3) \,{\mbox{SNU}} \,. \end{equation} This has to be compared with the Gallex \cite{gallex} and Sage\cite{sage} average: \begin{equation} \label{eqsgaexp} G_{exp}= (72 \pm 6) \, {\mbox{SNU}} \,. \end{equation} All this means that the present experimental result is about three sigmas below the minimal expectation in the absence of oscillations. This also illustrates the potential of GNO \cite{gno}, the successor of Gallex, which should reduce the total error down to about 4 SNU. If the present central value is mantained, the discrepancy with the minimal prediction will be at the level of about $5\sigma$, thus providing a clean signature of neutrino oscillations. \subsection{The Chlorine result} The solar luminosity constraint, together with the assumption that the ratio of pep-neutrinos over pp-neutrinos $\psi=\Phi(pep)/ \Phi(pp)=0.0023$ is correctly determined by SSM calculation, can be used to provide a lower limit also for the Chlorine signal. As well known, due to the fact that cross sections increase with neutrino energy, the minimal signal is obtained by maximixing the number of lowest-energy neutrinos, consistent with the luminosity constraint. Since a flux $\Phi_{pp}$ of pp-neutrinos is anyhow accompanied by a flux $\Phi(pep)= \psi \Phi(pp)$, this implies a minimal Chlorine signal: \begin{equation} \label{cl1} C_{Lum}=\frac{K_\odot}{Q_{pp}/\psi +Q_{pep}} \, \sigma_{pep} \, , \end{equation} where $K_{\odot}$ is the solar constant, $Q_i$ is the average electromagnetic energy released for emitted $i$-neutrino (see \cite{report}) and $\sigma_{i}$ is the averaged $i$-neutrino cross section on Chlorine detector. By using the cross section from \cite{libro}, but for the absorption cross section of $^{8}B$ neutrinos from \cite{sboro}, one has: \begin{equation} \label{cl2} C_{Lum}= (0.243 \pm 0.005 )\, {\mbox{SNU}} \, . \end{equation} From the Superkamiokande result one can deduce the B-neutrinos contribution of $(2.76 \pm 0.12)$ SNU, so that: \begin{equation} \label{cl3} C_{Lum+SK}= (3.0 \pm 0.1 )\, {\mbox{SNU}} \, . \end{equation} The minimal Be-neutrino flux implies the additional contribution of \begin{eqnarray} \label{cl4} C_{Be_{min}}&=& \Phi(Be)_{min} [ \sigma_{Be} - \sigma_{pep} \frac{Q_{Be}}{Q_{pp}/\psi +Q_{pep}}] \nonumber \\ &=& (0.24 \pm 0.06) \, {\mbox{SNU}} \ . \end{eqnarray} In this way the minimal Chlorine signal becomes: \begin{equation} C_{min}=(3.24 \pm 0.14) \, {\mbox{SNU,}} \end{equation} to be compared with the experimental result \cite{homestake} \begin{equation} C_{exp}=(2.56 \pm 0.22) \, {\mbox{SNU.}} \end{equation} Again the Be-neutrinos contributions, eq. (\ref{cl4}), corresponds to the ``$1\sigma$`` uncertainty of the experimental result, and again the experimental signal is about three sigmas below the minimal prediction. \subsection{Combining experimental results} A global view of the ``solar neutrino puzzle`` is presented in Fig. \ref{fig1} which updates Fig. 7 of \cite{report}. As a generalization of eq. (\ref{eqfibe}) for an arbitrary value of the observed $^8B$ flux $\Phi(B)$ one has: \begin{equation} \label{eqbebo} \Phi(Be)_{min}=4\cdot 10^2 \Phi(B) \, . \end{equation} The corresponding thick ``diagonal`` line, in Fig. 1 defines thus the lower border of the physical region. The shaded area, corresponding to the region within $3 \sigma$ from each experimental result, is almost completely out of the physical region. \section{Conclusions} As a summary, in the absence of oscillations we predict: \begin{eqnarray} \Phi(Be)_{min} &=& (1\pm0.24)\cdot 10^9\, {\mbox{cm$^{-2}$ s$^{-1}$}} \nonumber \\ G_{min}&=&(91 \pm 3 )\, {\mbox{SNU}} \nonumber \\ C_{min}&=&( 3.24\pm 0.14 )\, {\mbox{SNU}} \nonumber \, . \end{eqnarray} Let us list the information and assumptions behind these results: \noindent i) the measured $^8B$ flux by Superkamiokande; \noindent ii) the helioseismically determined sound speed, near or at the solar center, $u_c$; \noindent iii) the measured value of $\lambda_{17}$; \noindent iv) the value of $\lambda_{e7}$, derived from the lifetime of $^7Be$ in the laboratory; \noindent v) the luminosity constraint, i.e. the present observed solar luminosity equals the presently generated nuclear power in the sun; \noindent v) the central Hydrogen abundance, the only information we take from SSM calculations. We remark that we do not need to know the central solar temperature, nor the values of the astrophysical S-factors for the He+He reactions. \acknowledgments We are extremely grateful to V. Berezinsky and G. Fiorentini for suggesting us the problem and for useful discussions and comments.
\section{INTRODUCTION} Since starburst phenomena affect both the chemical evolution of galaxies and the physical conditions of the interstellar medium in galaxies, an understanding of starburst activity is one of the important issues in astrophysics. Starbursts occur in dense molecular gas clouds which are generally located in the central regions of galaxies. Massive stars formed in the starbursts provide a negative feedback to the parent and ambient molecular gas clouds in the form of intense radiation fields and strong stellar winds as well as subsequent supernova explosions. Therefore it is important to investigate the molecular gas properties of starburst galaxies (Young \& Scoville 1991; Henkel, Mauersberger, \& Baan 1991; Sanders \& Mirabel 1996 and references therein). One of the interesting properties of molecular gas in starburst galaxies is that galaxy mergers with luminous starbursts (i.e., $L$(FIR){\thinspace}$\gtrsim 10^{11}L_{\odot}$: hereafter luminous starburst mergers) tend to have higher $R_{1-0}$ ($\equiv I[^{12}$CO($J$=1--0)]/$I[^{13}$CO($J$=1--0)]) integrated line intensity ratios than normal spiral galaxies (Aalto et al. 1991, 1995, 1997; Casoli et al. 1991; Casoli, Dupraz, \& Combes 1992a, 1992b; Hurt \& Turner 1991; Turner \& Hurt 1992; Garay, Mardones, \& Mirabel 1993; Henkel \& Mauersberger 1993; Henkel et al. 1998). One possible explanation for the higher $R_{1-0}$ values is that they are due to the inflow of disk gas with high $^{12}$C$/^{13}$C abundance ratios, possibly combined with a $^{12}$C enhancement caused by nucleosynthesis in massive stars (e.g., Henkel et al. 1998). However, recently, Taniguchi \& Ohyama (1998a; hereafter TO98) have compared far-infrared luminosities [$L$(FIR)] with CO luminosities, $L[^{12}$CO($J$=1--0)$]$ and $L[^{13}$CO($J$=1--0)$]$, for a sample of normal and starburst galaxies, and have found that the observed higher $R_{1-0}$ values are associated almost exclusively in the luminous starburst mergers and appear to be attributed to a lower intensity of $^{13}$CO($J$=1--0) with respect to $^{12}$CO($J$=1--0) when compared to normal galaxies. TO98 suggested either that $^{13}$CO is underabundant with respect to $^{12}$CO, or that the $^{13}$CO(1--0) level population is more depressed relative to $^{12}$CO(1--0) due to excitation and/or optical depth effects, leading to the high $R_{1-0}$ in the luminous starburst mergers studied in their paper. In this paper, we investigate the second possibility using available higher transition data of both $^{12}$CO and $^{13}$CO emission lines. \section{$^{12}$CO($J$=2--1)$/^{13}$CO($J$=2--1) INTEGRATED LINE INTENSITY RATIO} TO98 demonstrated that the comparison of $L$(FIR) with both $L$[$^{12}$CO] and $L$[$^{13}$CO] provides a powerful tool for understanding the origin of the high-$R_{1-0}$ in luminous starburst mergers (see also Taniguchi \& Ohyama 1998b). If the observed high $R_{1-0}$ values are attributed to exication and optical depth effects, for example higher gas kinetic temperatures and/or denser gas clouds, then this could possibly be discerned in the measured values of $R_{2-1}$ or, if necessary, in even higher transition line ratios. In order to examine if this is the case, we investigate the excitation properties of both $^{12}$CO and $^{13}$CO molecules for a sample of normal and starburst galaxies using data currently available from the literature. We have compiled $^{12}$CO($J$=2--1) and $^{13}$CO($J$=2--1) integrated intensities from Aalto et al. (1995) and Casoli et al. (1992b). The integrated intensity ratio $I$[$^{12}$CO($J$=2--1)]$/I$[$^{13}$CO($J$=2--1)] is referred as $R_{2-1}$. Our sample consists of 24 galaxies and includes objects with extreme infrared luminosities such as the ultraluminous infrared galaxy Arp 220. These integreated intensities are then used to compute CO luminosities; $L$(CO) is defined as $L$(CO)$=A\times I$(CO) K km s$^{-1}$ pc$^2$ where $A$ is the observed area in units of pc$^2$ and $I$(CO)$=\int T_{\rm A}^* \eta^{-1} dv$ K km s$^{-1}$ where $T_{\rm A}^*$ is the observed antenna temperature corrected for atmospheric extinction and $\eta$ is the main beam efficiency. The FIR data are compiled from the {\it IRAS} Faint Source Catalog (Moshir et al. 1992). The FIR luminosities are estimated using $L$(FIR)$=4\pi D^2 1.26\times 10^{-11} [2.58\times S(60) + S(100)]$ (ergs s$^{-1}$) where $S$(60) and $S$(100) are the {\it IRAS} 60{\thinspace}$\mu$m and 100{\thinspace}$\mu$m fluxes in units of Jy and $D$ is the distance (Helou, Soifer, \& Rowan-Robinson 1985). Distances of nearby galaxies are taken from the Nearby Galaxies Catalog (Tully 1988); distances of other galaxies are estimated using a Hubble constant $H_0$ = 75 km s$^{-1}$ Mpc$^{-1}$ with $V_{\rm GSR}$ (recession velocity with respect to the Galactic Standard of Rest) given in de Vaucouleurs et al. (1991). The compiled data are given in Table 1. All of the data presented here have been corected for beam-size (see Aalto et al. 1995; Casoli et al. 1992b). Although our sample is not statistically complete, it is the largest sample compiled so far. TO98 defined the class of high-$R_{1-0}$ galaxies by adopting the criterion of $R_{1-0} \geq 20$. Using this limit, the present sample contains the following seven high-$R_{1-0}$ objects; NGC 1614, NGC 3256, NGC 4194, NGC 6240, Arp 220, Arp 299, and IRAS 18293$-$3413. In Figure 1, we compare $L$[$^{12}$CO($J$=2--1)] with $L$[$^{13}$CO($J$=2--1)]. For reference, we also show the comparison between $L$[$^{12}$CO($J$=1--0)] and $L$[$^{13}$CO($J$=1--0)] in the left panel which is taken from TO98. Although two high-$R_{1-0}$ galaxies, NGC 4194 and NGC 6240, have significantly lower $L$[$^{13}$CO($J$=2--1)] with respect to $L$[$^{12}$CO($J$=2--1)], the remaining galaxies have $R_{2-1}$ ratios within the upper range (i.e. $\sim$10--30) found for $R_{1-0}$. In Figure 2, we compare $L$(FIR) with both $L$[$^{12}$CO($J$=2--1)] and $L$[$^{13}$CO($J$=2--1)]. $L$[$^{12}$CO($J$=2--1)] appears to be correlated with $L$(FIR) (in an integrated intensity versus flux plot, i.e. after removing the the $D^2$ effect from Figure 2, there is a good correlation), but the scatter is larger than observed for the correlation between $L$[$^{12}$CO($J$=1--0)] and $L$(FIR) (TO98). On the other hand, the correlation between $L$(FIR) and $L$[$^{13}$CO($J$=2--1)] is poorer than that between $L$(FIR) and $L$[$^{12}$CO($J$=2--1)] because the majority of the high-$R_{1-0}$ galaxies have lower $L$[$^{13}$CO($J$=2--1)] by a factor of 3 than what would be expected from the correlation for the normal-$R_{1-0}$ galaxies. Thus, we find that both $^{12}$CO($J$=2--1) and $^{13}$CO($J$=2--1) show similar behavior as observed in the $J$=1--0 transition (TO98). In order to show this more clearly, we present a diagram of $R_{1-0}$ versus $R_{2-1}$ for our sample in Figure 3. In Table 2, we summarize the statistical properties for both the high-$R_{1-0}$ and the ``normal-$R_{1-0}$" galaxies. \section{$^{12}$CO($J$=3--2)$/^{13}$CO($J$=3--2) INTEGRATED LINE INTENSITY RATIO} Published measurements of extragalactic $^{13}$CO($J$=3--2) emission are available only for M82 (Tilanus et al. 1991; Wild et al. 1992) and IC 342 (Wall \& Jaffe 1990). Tilanus et al. (1991) obtained $^{13}$CO($J$=3--2) spectra at three positions in M82; the center and the two peaks of the circumnuclear ring located $\pm$12 arcsec from the center using the 15{\thinspace}m JCMT with a beam size of 14 arcsec (FWHM). They obtained $R_{3-2} \simeq 15$ for the nucleus while $\simeq 10$ for the circumnuclear ring. On the other hand, Wild et al. (1992) measured both $^{12}$CO($J$=3--2) and $^{13}$CO($J$=3--2) emission lines using the IRAM 30{\thinspace}m radio telescope in February 1992 and obtained both $I$[$^{12}$CO($J$=3--2)] = 1334 K km s$^{-1}$ at ($\Delta\alpha$, $\Delta\delta$) = ($-$5, $-$5) and $I$[$^{13}$CO($J$=3--2)] = 70.4 K km s$^{-1}$ at ($\Delta\alpha$, $\Delta\delta$) = ($-$7, $-$5) where $\Delta\alpha$ and $\Delta\delta$ are offsets from the nucleus position in right ascension and declination, respectively, in units of arcsec. Although the measured positions are slightly different, these measurements give an integrated intensity ratio, $R_{3-2} \simeq 18.9$. This together with the results by Tilanus et al. (1991) indicates that $R_{3-2}$ in the nuclear region is $\simeq$ 15 -- 20. Since this value is nearly the same as the threshold value which defines the class of high-$R_{1-0}$ galaxies (TO98), it suggests that the $J = 3$ transition is still not high enough to allow dissentanglement of radiative transfer and abundance effects in M82. As suggested by Taniguchi \& Ohyama (1998b), it is possible that a large value of $R$ can be attributed to the effect of superwind activity (i.e., the possible destruction of dense gas as well as dust grains, and the large velocity widths observed in the CO outflow). In fact, in M82, $R_{3-2}$ is higher in the nuclear region than in the starburst ring (Tilanus et al. 1991). Since M82 is indeed a superwind-starburst galaxy (Bland \& Tully 1988), the higher $R_{3-2}$ value in the nuclear region of M82 could possible be due to superwind activity. However, Tilanus et al. (1991) suggested that intensity ratios of the three lowest transition lines of $^{12}$CO and $^{13}$CO can be explained if $^{13}$CO is overabundant with respect to $^{12}$CO just like what has been measured in the Galactic center, being contrary to our interpretation. Another measurement of the $^{13}$CO($J$=3--2) line was obtained at the central region of IC 342; $R_{3-2} \simeq 7.7$ (Wall \& Jaffe 1990). However, since the beam size of $^{13}$CO($J$=3--2) observation (24 arcsec) is different from that of $^{12}$CO($J$=3--2) one (15 arcsec), the above value may not be reliable. Until more extragalactic CO($J$=3 -- 2) data is obtained it is clearly impossible to draw any firm conclusions about $R_{3-2}$ in luminous starburst mergers as well as normal galaxies. \section{DISCUSSION} We have shown that the $R_{2-1}$ ratio is also high (typically $> 20$) in the high-$R_{1-0}$ galaxies. Furthermore, the $R_{3-2}$ value in the nuclear region of M82 suggests that this ratio may also be high in luminous starbursts, but this is only for one object. However, it is noted that the $^{12}$CO($J$=3--2)/$^{12}$CO($J$=1--0) integrated intensity ratio of starburst galaxies is often found to be higher than that in normal galaxies (e.g., Devereux et al. 1994 and references therein). Further, some nearby starburst galaxies such as M82 are detected in CO($J$=4--3) (G\"usten et al. 1993) and in CO($J$=6--5) (Harris et al. 1992). The detection of these higher-transition CO lines suggests the presence of warm and dense gas clouds in starburst galaxies. Although the kinetic gas temperature is not necessarily comparable to the dust temperature, it is interesting to compare molecular gas properties with the dust temperature which is measured from the {\it IRAS} 60 $\mu$m to 100 $\mu$m flux ratio, $S(60)/S(100)$. In Figure 4, we compare $R_{2-1}$ with $S(60)/S(100)$ for the galaxies studied here. We also compare $R_{1-0}$ with $S(60)/S(100)$ for reference (the left panel). Though no tight correlation can be seen in either of these diagrams, we find that both $R_{1-0}$ and $R_{2-1}$ tend to increase with increasing $S(60)/S(100)$. This tendency suggests that the galaxies with higher dust temperatures have higher $R$ values on the average. In the lower panels of Figure 4, we show comparisons of $S(60)/S(100)$ with both $L$[$^{12}$CO($J$=2--1)]$/L$[$^{12}$CO($J$=1--0)] and $L$[$^{13}$CO($J$=2--1)]$/$L$[^{13}$CO($J$=1--0)] ratios. Interestingly, we find no correlation in both the diagrams. Average ratios of both $L$[$^{12}$CO($J$=2--1)]$/L$[$^{12}$CO($J$=1--0)] and $L$[$^{13}$CO($J$=2--1)]$/L$[$^{13}$CO($J$=1--0)] are $0.81 \pm 0.17$ and $1.15 \pm 0.71$, respectively for the high $R_{1-0}$ starburst mergers. Therefore, there seems to be no significant difference in the excitation toward $J$=2 between $^{12}$CO and $^{13}$CO. We also investigate whether a correlation exists between $R_{1-0}$ and the luminosity ratio $L$[$^{12}$CO($J$=2--1)]$/L$[$^{12}$CO($J$=1--0)] for our sample (Figure 5). Since the high-$R_{1-0}$ mergers tend to have higher dust temperatures (see Table 2), it is likely that their CO kinetic temperatures are also higher than those of the normal-$R_{1-0}$ galaxies. These high temperatures could lead to both higher $R_{1-0}$ and to higher $L$[$^{12}$CO($J$=2--1)]$/L$[$^{12}$CO($J$=1--0)] as demonstrated by large velocity gradient models (Sakamoto et al. 1994, 1997). However, there is no such tendency as shown in Figure 5. Finally, we investigate whether there is any relationship between $R_{1-0}$, $R_{2-1}$ and the $L$(FIR)/$L$(CO) ratio which is generally considered to give a measure of star formation efficiency. Figure 6 shows that the high-$R_{1-0}$ mergers as a group have a higher mean $L$(FIR)/$L$(CO) ratio as well as a higher mean $R_{2-1}$ ratio than the ``normal-$R_{1-0}$" galaxies. But, other than that there appears to be no clear correlation between either $R_{1-0}$ or $R_{2-1}$ with respect to $L$(FIR)/$L$(CO). Therefore, we are led once again to suggest that ``superwinds" may be the best explanation for what produces the high-$R_{1-0}$ values. Indeed, it should be noted that nearly all of the high-$R_{1-0}$ galaxies show morphological and/or spectroscopic evidence for superwinds (Taniguchi \& Ohyama 1998b). Both the abnormally large velocity gradients assocated with these superwinds and the possible destruction of dense gas clouds by the dynamical effect of the superwind activity, could possibly combine to reduce the observed intensity of the much more optically thin $^{13}$CO in the lower-$J$ transitions relative to the $^{12}$CO emission. Further tests of this hypothesis will require measurements of the $J \geq 3$ transitions of both $^{12}$CO and $^{13}$CO for the galaxies in Table 1. \vspace{1ex} We would like to thank Seiichi Sakamoto for useful discussions. YO was supported by a Grant-in-Aid for JSPS Fellows by the Ministry of Education, Science, Sports and Culture. This work was supported in part by the Ministry of Education, Science, Sports and Culture in Japan under Grant Nos. 07055044, 10044052, and 10304013.
\section{Introduction} It is generally agreed that a key structural element of the high-$T_c$ superconductors is the quasi-two dimensional copper-oxygen plane. In most of the cuprates this plane has nearly a tetragonal symmetry with a small orthorhombic distortion. \cite{1} The unit cells of some of these materials contain other elements, like the one-dimensional copper-oxygen chains in ${\rm YBa_2Cu_3O_{7-\delta}}$ (YBCO) which reduce their symmetry to the orthorhombic one ($C_{2v}$). It is possible that the Cu-O chains may also contribute to the superconducting behavior of YBCO as the in-plane transport properties display an appreciable anisotropy significantly larger than that expected from a slight orthorhombicity of the ${\rm CuO_2}$ planes. Several experiments \cite{2,3,5,6,7,4,11,8} were designed to measure the transport properties separately for the $a$ and $b$ directions of the copper-oxygen planes. In YBCO, a large anisotropy of about $50\%$ between the $a$ and $b$ directions has been observed \cite{2,3} in the zero temperature penetration depth ($\lambda_a(0)/\lambda_b(0)\approx 1.55$) and a smaller anisotropy has been seen in the thermal conductivity (maximum value of $\kappa_a/\kappa_b\approx 1.15$ occurring near 40K). \cite{5,6} Also the microwave absorption measurements \cite{3,7} in YBCO show a significant in-plane anisotropy in the surface resistance and in the infrared conductivity, the later remaining about a factor of 2 larger in the $b$ direction for the temperature range of 10-90K. The in-plane anisotropy of the surface resistance \cite{3} in the normal state of YBCO agrees with the observed dc resistivity anisotropy \cite{4,11} $\varrho_a/\varrho_b\sim 2$, which was indicated by the band-structure calculations \cite{9} as well. Although very weak orthorhombicity due to a superlattice distortion in the Bi-O layer, \cite{1} ${\rm Bi_2Sr_2CaCu_2O_8}$ (BSCCO) displays $ab$-plane anisotropy \cite{8} of about $10\%$ in the optical conductivity and the dc resistivity. It has been pointed out, \cite{10,12} that in an orthorhombic system the superconducting order parameter becomes a mixture of $d$-wave and $s$-wave components as they belong to the same irreducible representation of the crystal point group. This leads to an interesting feature in the specific heat jump at the phase transition in the presence of a potential impurity scattering. \cite{21}\\ \indent Using a simplified single-band model simulating the orthorhombic anisotropy of the Fermi surface for YBCO and BSCCO, we have analyzed a group of superconducting states in the untwinned systems. We have studied the effect of potential scattering on the stability of these states and also on the jump in the electronic specific heat at the phase transition. \section{Order parameter and Fermi surface anisotropy} We consider a singlet superconducting order parameter with its orbital part defined as follows \begin{equation} \label{e2} \Delta\left({\bf k}\right)=\Delta e\left({\bf k}\right) \end{equation} \noindent where $e\left({\bf k}\right)$ is a basis real function of a one-dimensional (1D) irreducible representation of $C_{4v}$ point group. We normalize $e\!\left({\bf k}\right)$ by taking its average value over the Fermi surface (FS) $\left<e^{2}\right>=\int_{FS}dS_k n\left({\bf k}\right) e^2\left({\bf k}\right)=1$, where $\int_{FS}dS_{k}$ represents the integration over the Fermi surface and $n\left({\bf k}\right)$ is the angle resolved FS density of states, which obeys $\int_{FS}dS_k n\left({\bf k}\right)=1$. This normalization gives $\Delta$ the meaning of the absolute magnitude of the order parameter. Our discussion is limited to the functions $e\left({\bf k}\right)$ confined to the XY plane only, which seems to be appropriate for the high $T_{c}$ compounds.\cite{12} These functions are listed in Tab. 1, with the symmetry group notation after Ref. 14. The basis functions of the $C_{4v}$ irreducible representations $\Gamma_{1}^{+}$, $\Gamma_{3}^{+}$ and $\Gamma_{4}^{+}$ are taken as three linearly independent second order polynomials. A fourth order polynomial is used then as a basis function of the remaining $\Gamma_{2}^{+}$ representation. Finally, for the sake of comparison, we consider two more fourth order polynomials belonging to the identity irreducible representation $\Gamma_{1}^{+}$. The superconducting states given by the functions $e\left({\bf k}\right)$ constructed from the angular momentum eigenfunctions corresponding to a quantum number L=2 are called the $d$-wave states (second order polynomials) and those for which the L=4 eigenfunctions were used are called the $g$-wave states (fourth order polynomials). Unless an ambiguity arises we will use a more informal name of extended s-wave for the states which are invariant under rotation through $\pi/2$ about the $z$ axis, that is for $\Gamma_{1}^{+}$ and $\Gamma_{2}^{+}$ of $C_{4v}$ irreducible representations given in Tab. 1. The presence of the orthorhombic anisotropy means that the relevant point group is a subgroup of the square, which is not the same for all the cuprate superconductors. \cite{12} In the case of YBCO, the $a$- and $b$- crystal axes become inequivalent while in BSCCO the two orthogonal $45^{\circ}$ axes do so. Therefore the group classification of the superconducting states differs in YBCO and BSCCO compounds (see Tab. 1). A rotation of a coordinate system through $45^{\circ}$ about the z-axis transforms a YBCO-type geometry into a BSCCO-type one and also shows an equivalence of the superconducting states, which we summarize in Tab. 2. Distinguishing between these two symmetries, we make a simple approximation of the orthorhombic anisotropy by assuming the following form of the electron band \cite{10,14} energy $\xi_{\bf k}$ measured from the Fermi energy $\varepsilon_{F}$ level \begin{equation} \label{e2a} \xi_{\bf k}=c_xk_{x}^{2}+c_yk_{y}^{2}-\varepsilon_{F},\;\;\;\;\;\; {\rm for\;\;YBCO} \end{equation} \begin{equation} \label{e2b} \xi_{\bf k}=\frac{1}{2}\left(c_x+c_y\right)\left(k_{x}^{2}+k_{y}^{2}\right) +\left(c_x-c_y\right)k_{x}k_{y}-\varepsilon_{F},\;\;\;\;\;\; {\rm for\;\;BSCCO} \end{equation} \noindent A dimensionless ratio of the effective masses $c_x/c_y$ becomes a parameter describing the orthorhombic anisotropy of the Fermi surface. It changes from 0 to 1, with $c_x/c_y=1$ describing a circular FS and $c_x/c_y=0$ corresponding to a one-dimensional limit, non physical for the cuprates. For the sake of simplicity the orthorhombic symmetry is introduced as a deviation from a cylindrical FS not from a tetragonal one. We assume that this simplified single-band model \cite{10} given by Eqs. (\ref{e2a}) and (\ref{e2b}) captures some essential anisotropic features of both types of orthorhombicity i.e. that of YBCO and BSCCO. Particularly in the case of YBCO, where the symmetry lowering is mainly due to the existence of the chains, limiting the superconducting electrons to the ${\rm CuO_2}$ planes only and introducing the crystal field anisotropy by an elliptical Fermi surface is to be understood as a first approximation which reflects the basic properties related to the orthorhombicity of the compound. We employ the above single-band model to study the effect of orthorhombic anisotropy on the critical temperature $T_c$ and the specific heat jump at the phase transition in the presence of nonmagnetic impurity scattering. \section{Nonmagnetic impurity scattering near ${\bf T_c}$} The single-particle Green's function in the presence of nonmagnetic, noninteracting impurities is expressed in Nambu space as \begin{equation} \label{e4} \hat{G}\left(\omega,{\bf k}\right)=-\frac{1} {\tilde{\omega}^{2}+{\xi_{k}}^{2}+|\tilde{\Delta}\left( {\bf k}\right)|^{2}}\left(i\tilde{\omega}\hat{\tau}_0 +\xi_{k}\hat{\tau}_3+\tilde{\Delta}\left({\bf k}\right)\hat{\tau}_2\right) \end{equation} \noindent where $\hat{\tau}_0$ and $\hat{\tau}_i$ ($i=1,2,3$) represent the unit and Pauli matrices in particle-hole space respectively and $\xi_{k}$ is the quasiparticle energy (Eqs. (\ref{e2a}), (\ref{e2b})). The renormalized Matsubara frequency $\tilde{\omega}\left({\bf k}\right)$ and the renormalized order parameter $\tilde{\Delta}\left({\bf k}\right)$ are given by \begin{eqnarray} \label{e6} \tilde{\omega}=\omega-\Sigma_0,&\;\;\;\;& \tilde{\Delta}\left({\bf k}\right)=\Delta\left({\bf k}\right)+\Sigma_1 \end{eqnarray} \noindent with $\omega=\pi T(2n+1)$ (T is the temperature, n is an integer). $\Sigma_0$, $\Sigma_1$ are the self-energies due to the electron-impurity scattering obtained in the t-matrix approximation. \cite{15,16} This approach introduces two parameters describing the scattering process: $c=1/(\pi N_0 V_i)$ and $\Gamma=n_i/\pi N_0$, where $N_{0}$, $V_i$ and $n_i$ are respectively the overall density of states at the Fermi surface (FS), the impurity (defect) potential and the impurity concentration. We assume $s$-wave scattering by the impurities, that is, $V_i$ does not have an internal momentum-dependence. \cite{27} It is particularly convenient to think of $c$ as a measure of the scattering strength, with $c=0$ in the unitary limit and $c\gg 1$ for weak scattering that is the Born limit. Assuming a particle-hole symmetry of the quasiparticle spectrum we get the self-energies defined as follows \begin{eqnarray} \label{e6a} \displaystyle\Sigma_0=-\Gamma\frac{g_0}{c^2+g^{2}_{0}+g^{2}_{1}},&\;\:\;\;& \displaystyle\Sigma_1=\Gamma\frac{g_1}{c^2+g^{2}_{0}+g^{2}_{1}} \end{eqnarray} \noindent with $g_0$, $g_1$ functions determined by the self-consistent equations \begin{equation} \label{e6b} \displaystyle g_0=\frac{1}{N_0\pi}\sum_{\bf k}\frac{\tilde{\omega}} {\tilde{\omega}^{2}+{\xi_{k}}^{2}+|\tilde{\Delta}\left({\bf k}\right)|^{2}} \end{equation} \begin{equation} \label{e6c} \displaystyle g_1=\frac{1}{N_0\pi}\sum_{\bf k}\frac{\tilde{\Delta}\left({\bf k}\right)} {\tilde{\omega}^{2}+{\xi_{k}}^{2}+|\tilde{\Delta}\left({\bf k}\right)|^{2}} \end{equation} \noindent The self-consistency equation for the order parameter reads \begin{equation} \label{e7} \Delta\left({\bf k}\right)=-T\sum_{\omega}\sum_{{\bf k'}} V\left({\bf k}, {\bf k'}\right) \frac{\tilde{\Delta}\left({\bf k'}\right)} {\tilde{\omega}^{2}+{\xi_{k'}}^{2}+|\tilde{\Delta}\left( {\bf k'}\right)|^{2}} \end{equation} \noindent where $V\left({\bf k}, {\bf k'}\right)$ is the phenomenological pair potential taken as \begin{equation} \label{e8} V\left({\bf k}, {\bf k'}\right)=-V_{0}e\left({\bf k}\right) e\left({\bf k'}\right) \end{equation} \noindent To proceed further, we restrict the wave vectors of the electron self-energy and pairing potential to the Fermi surface and replace $\sum_{\bf k}$ by $N_{0}\int_{FS}dS_{k}n\left({\bf k}\right)\int d\xi_{k}$. Integrating over $\xi_{k}$, the gap equation (\ref{e7}) can be transformed after a standard procedure \cite{17} into \begin{equation} \label{e8a} \ln\left(\frac{T}{T_{c_{0}}}\right)=2\pi T\sum_{\omega\ge 0} \left(\int_{FS}dS_{k}n\left({\bf k}\right) f\left(\omega,{\bf k}\right)-\frac{1}{\omega}\right) \end{equation} \noindent where the $f\left(\omega,{\bf k}\right)$ function is defined as follows \begin{equation} \label{e8b} \displaystyle f\left(\omega,{\bf k}\right)= \frac{e\left({\bf k}\right)\tilde{\Delta}\left({\bf k}\right)} {\Delta\left[\tilde{\omega}^{2}+|\tilde{\Delta}\left( {\bf k}\right)|^{2}\right]^{\frac{1}{2}}} \end{equation} \noindent and $T_{c_{0}}$ is the critical temperature in the absence of impurities. We expand the gap equation (\ref{e8a}) in powers of $\Delta^{2}$ around $\Delta=0$ taking into account that $\tilde{\omega}$ and $\tilde{\Delta}$ are functions of $\Delta^{2}$ as given by Eqs. (\ref{e6})-(\ref{e6c}). Keeping up to the second power terms in $\Delta$ we get the Ginzburg-Landau approximation of the gap equation \begin{equation} \label{e9} \ln\left(\frac{T}{T_{c_{0}}}\right)=-f_{0}-\frac{1}{2}f_{1} \left(\frac{\Delta}{2\pi T}\right)^{2} \end{equation} \noindent where the coefficients are given by \begin{equation} \label{e10} f_{0}=-2\pi T\sum_{\omega>0}\left(\int_{FS}dS_{k}n\left({\bf k}\right) \left(f\left(\omega,{\bf k}\right)\right)_{\Delta=0} -\frac{1}{\omega}\right) \end{equation} \begin{equation} \label{e11} f_{1}=-\left(2\pi T\right)^{3}\sum_{\omega} \int_{FS}dS_{k}n\left({\bf k}\right) \left(\frac{df\left(\omega,{\bf k}\right)}{d\Delta^{2}}\right)_{\Delta=0} \end{equation} \noindent Taking the derivatives with respect to $\Delta^2$ \begin{equation} \label{e25} \displaystyle\frac{d\;\;\;}{d\Delta^{2}}= \frac{\partial\;\;\;}{\partial\Delta^{2}}+ \sum_{\omega}\left\{\frac{d\tilde{\omega}} {d\Delta^{2}}\frac{\partial\;\;}{\partial\tilde{\omega}} +\frac{d\tilde{\Delta}\left({\bf k}\right)} {d\Delta^{2}}\frac{\partial\;\;\;\;\;\;} {\partial\tilde{\Delta}\left({\bf k}\right)}\right\} \end{equation} \noindent and with a use of the relations given in Eqs. (\ref{e6})-(\ref{e6c}) and (\ref{e8b}) we obtain \begin{equation} \label{e13} \begin{array}{l} \displaystyle f_{0}=\left(1-\left<e\right>^{2}\right) \left(\psi\left(\frac{1}{2}+\varrho\right) -\psi\left(\frac{1}{2}\right)\right)\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \;\;\;\;\;\;\;\;\;\;\;\;\; \end{array} \end{equation} \vspace{2ex} \begin{equation} \label{e14a} \begin{array}{l} \displaystyle f_{1}=2\left<e\right>\left[2\left<e^3\right>+5\left<e\right>^3 -7\left<e\right>\right]\varrho^{-2}\left(\psi\left(\frac{1}{2} +\varrho\right)-\psi\left(\frac{1}{2}\right)\right)\\ \\ \displaystyle +2\left<e\right>\left[-2\left<e^{3}\right>-3\left<e\right>^3 +5\left<e\right>\right] \varrho^{-1}\psi^{(1)}\left(\frac{1}{2}+\varrho\right) +4\left<e\right>^2\left[1-\left<e\right>^{2}\right]\varrho^{-1} \psi^{(1)}\left(\frac{1}{2}\right)\\ \\ \displaystyle +\frac{1}{2}\left[-\left<e^{4}\right>+3\left<e\right>^{4} +4\left<e\right>\left<e^{3}\right>-6\left<e\right>^{2}\right] \psi^{(2)}\left(\frac{1}{2}+\varrho\right) -\frac{1}{2}\left<e\right>^{4}\psi^{(2)}\left(\frac{1}{2}\right)\\ \\ \displaystyle +\frac{1}{6}\left[2\left(\left<e\right>^{2}-1\right)^2\frac{1}{c^2+1} -\left<e\right>^{4}+2\left<e\right>^{2}-1\right] \varrho\psi^{(3)}\left(\frac{1}{2}+\varrho\right) \end{array} \end{equation} \noindent where $\varrho=\Gamma/\left[(c^2+1)2\pi T\right]$ and $\psi$, $\psi^{(n)} \;(n=1,2,3)$ are the polygamma functions. \cite{19} In the unitary limit $c=0$ and $\varrho=\Gamma/(2\pi T)$. Alternatively for weak scattering ($c\gg 1$) we obtain the Born scattering rate $\varrho=\pi N_0 n_i V^2_i/(2\pi T)$ and also neglect $\varrho/(c^2+1)$ in the last term of Eq. (\ref{e14a}). Coefficients $f_0$ and $f_1$ involve three different types of the Fermi surface averages of the superconducting order parameter namely, $\left<e\right>$, $\left<e^{3}\right>$, and $\left<e^{4}\right>$, which depend on the orthorhombic anisotropy parameter $c_x/c_y$. \cite{28} These averages enter the free energy and determine the thermodynamic properties at the phase transition. While the critical temperature $T_c$ is determined by the $f_0$ function and therefore is characterized only by $\left<e\right>$, the other thermodynamic quantities like the entropy or the specific heat for instance, involve, through the $f_1$ function (Eq. (\ref{e14a})), the Fermi surface average values of higher powers of $e\left({\bf k}\right)$. It is illustrative for the purpose of this paper to show these averages as functions of the orthorhombic anisotropy parameter $c_x/c_y$. We present $\left<e\right>$, $\left<e^{3}\right>$, and $\left<e^{4}\right>$ in Figs. 1a - 1c respectively, where the curve numbers correspond to the states listed in Tab. 2. The analytical expressions of these averages for the YBCO symmetry (Eq. (\ref{e2a})) are given in the Appendix. It is worth mentioning here, that as the $k^2_x-k^2_y$ state ($d_{x^2-y^2}$) belongs to different irreducible representations (Tab. 1) in YBCO ($\Gamma^{+}_1$) and BSCCO ($\Gamma^{+}_3$), its FS average value $\left<e\right>$ is non zero in the first compound ($c_x/c_y<1$), but $\left<e\right>$ is unchanged by the orthorhombic crystal field and is zero in the second one. Further, for the $k_xk_y$ state ($d_{xy}$) $\left<e\right>=0$ in YBCO ($\Gamma^{+}_3$) and $\left<e\right>\neq 0$ in BSCCO ($\Gamma^{+}_1$). These facts are of crucial importance for the critical temperature and the jump in the electronic specific heat at the phase transition, which we discuss in the following sections. \section{Critical temperature ${\bf T_c}$} We analyze in the Ginzburg-Landau (G-L) regime the stability of different superconducting states in the presence of impurity potential scattering and the orthorhombic crystal field. For any FS we consider the states of the same critical temperature in the absence of impurities ($T_{c_{0}}$), which means that we normalize the order parameter eigenfunctions as $\left<e^2\right>=1$. \cite{29} Then we look at the impurity effect on $T_c$ of these states. It should be noted that since the orthorhombic anisotropy changes the density of states at the Fermi level, $T_{c_{0}}$ is different for different values of $c_x/c_y$ even with the condition $\left<e^2\right>=1$. We are interested in the influence of impurity scattering on the states of different symmetry but with the same $T_{c_{0}}$ in the presence of a given orthorhombic crystal field i.e. for a given $c_x/c_y$ value. The G-L free energy difference $\Delta F=F_s-F_n$ between the superconducting ($F_s$) and the normal ($F_n$) phase written up to the fourth order terms in $\Delta$ \begin{equation} \label{e14c} \Delta F=\alpha\Delta^2+\frac{1}{2}\beta\Delta^4 \end{equation} \noindent with the coefficients $\alpha$ and $\beta$ determined from Eqs. (\ref{e9}), (\ref{e13}) and (\ref{e14a}) i.e. \begin{equation} \label{e14d} \alpha=N_0 \left(\ln\left(\frac{T}{T_{c_{0}}}\right)+f_0\right) \end{equation} \begin{equation} \label{e14e} \beta=\frac{N_0 f_1}{2\left(2\pi T_c\right)^2} \end{equation} \noindent leads to the free energy minimum given by \begin{equation} \label{e14f} \left(\Delta F\right)_{min}=-\frac{1}{2}\frac{\alpha^2}{\beta} \end{equation} \noindent One can see from Eqs. (\ref{e13}) and (\ref{e14d}) that when $\alpha=0$ for a state with $\left<e\right>=0$ then its value is less than zero for a nonzero $\left<e\right>$. Therefore according to Eq. (\ref{e14f}) even an infinitesimal impurity scattering rate stabilizes states with $\left<e\right>\neq 0$ over $\left<e\right>=0$ states. In other words the critical temperature in the presence of nonmagnetic impurities given by \cite{22,23} \begin{equation} \label{e14b} \ln\left(\frac{T_c}{T_{c_{0}}}\right)=\left(\left<e\right>^{2}-1\right) \left(\psi\left(\frac{1}{2}+\varrho_c\right) -\psi\left(\frac{1}{2}\right)\right) \end{equation} \noindent with $\varrho_c$ as the value of $\varrho$ at $T_c$, is higher for the states with a nonzero $\left<e\right>$ value than those characterized by $\left<e\right>=0$. It is also worth mentioning that the states characterized by the same value of $\left<e\right>$ are degenerate. We show the solutions of Eq. (\ref{e14b}) for the states from Tab. 2 in the case of a not broken tetragonal symmetry ($c_x/c_y=1$) in Fig. 2, and for an orthorhombic symmetry with $c_x/c_y=0.8$ in Fig. 3. Comparison of Fig. 1a (at $c_x/c_y=1$) and Fig. 2 shows a degeneracy of the states with the same $\left<e\right>$ value. When the FS symmetry is lowered to $C_{2v}$ the states fall into two irreducible representations $\Gamma^{+}_1$ and $\Gamma^{+}_3$ (Tab. 1). The two $\Gamma^{+}_3$ states (curves 3 and 6 in Fig. 3) are still degenerate as $\left<e\right>=0$ in that case (Fig. 1a). On the other hand most of $\Gamma^{+}_1$ states are influenced by the orthorhombicity of the system and their critical temperatures are split. We present the evaluation of $T_c$ for the states under consideration at the impurity scattering level \cite{26} $\varrho_c T_c/T_{c_{0}}=0.1$ as a function of orthorhombic anisotropy parameter $c_x/c_y$ in Fig. 4. As mentioned earlier, the critical temperature depends on the order parameter and FS symmetries through $\left<e\right>$ only. This leads to a remarkable similarity between the effect of FS orthorhombicity on $T_c$ (Fig. 4) and on $\left<e\right>$ (Fig. 1a). \section{Specific heat jump at the phase transition} The specific heat jump at $T_c$, $\Delta C(T_c)$, normalized by the normal state specific heat $C_N(T_c)$ is given by \begin{equation} \label{e14} \displaystyle\frac{\Delta C(T_c)}{C_N(T_c)}=\frac{12}{\left(f_1\right)_{T=T_c}} \left[1+T_c\left(\frac{df_0}{dT}\right)_{T=T_c}\right]^{2} \end{equation} \noindent and $f_0$ from Eq. (\ref{e13}) yields \begin{equation} \label{e15} \displaystyle\frac{\Delta C(T_c)}{C_{N}(T_c)}=\frac{12}{f_{1}\left(\varrho_{c}\right)} \left[1+\left(\left<e\right>^2-1\right)\varrho_{c}\psi^{(1)}\left(\frac{1}{2}+ \varrho_{c}\right)\right]^{2} \end{equation} \noindent where $\varrho_c$ is $\varrho$ at $T=T_c$. This rather cumbersome formula, when considered along with Eq. (\ref{e14a}), reduces significantly for $\left<e\right>=0$ case \begin{equation} \label{e15a} \displaystyle\frac{\Delta C(T_c)}{C_{N}(T_c)}=\frac{\displaystyle 12\left[1-\varrho_c\psi^{(1)} \left(\frac{1}{2}+\varrho_c\right)\right]^{2}} {\displaystyle \frac{\mu}{6}\varrho\psi^{(3)}\left(\frac{1}{2}+\varrho_c\right) -\frac{1}{2}\left<e^{4}\right> \psi^{(2)}\left(\frac{1}{2}+\varrho_c\right)} \end{equation} \noindent where $\mu=(1-c^2)/(1+c^2)$. For an appropriate choice of $\left<e^{4}\right>$ value $\Delta C(T_c)/C_{N}(T_c)$ from Eq. (\ref{e15a}) agrees with the result obtained by Hirschfeld et al. \cite{16} as well as with that obtained by Suzumura and Schulz \cite{20} in the Born limit. For a pure system \cite{21} where $\varrho_c=0$ \begin{equation} \label{e16} \displaystyle\left(\frac{\Delta C\left(T_c\right)}{C_{N}\left(T_c\right)}\right)_{\varrho_c=0} =-\frac{24}{\displaystyle\psi^{(2)}\left(\frac{1}{2}\right)\left<e^{4}\right>} \approx\frac{1.426}{\left<e^{4}\right>} \end{equation} \noindent Alternatively, in a highly impure superconductor with $\varrho_c\rightarrow\infty$ i.e. $T_c\rightarrow 0$ due to suppression by the impurities, the two cases, depending on $\left<e\right>$ value, are to be distinguished. \cite{21} First, when $\left<e\right>\neq 0$ leads to \begin{equation} \label{e17} \displaystyle\left(\frac{\Delta C\left(T_c\right)}{C_{N}\left(T_c\right)}\right) _{\varrho_c\rightarrow\infty}= -\frac{24}{\displaystyle\psi^{(2)}\left(\frac{1}{2}\right)}\approx 1.426 \end{equation} \noindent and the second, with $\left<e\right>=0$ yields \begin{equation} \label{e18} \displaystyle\left(\frac{\Delta C\left(T_c\right)}{C_{N}\left(T_c\right)}\right) _{\varrho_c\rightarrow\infty}=0 \end{equation} \noindent We note, that the specific heat jump value in $\varrho_c\rightarrow\infty$ limit for a nonzero value of $\left<e\right>$ given by Eq. (\ref{e17}) agrees with that of an isotropic $s$-wave superconductor. This fact has a simple intuitive interpretation. A nonzero Fermi surface average of the order parameter leads to an asymptotic power-law critical temperature suppression for large impurity concentration \cite{14} $T_c\sim\left(T_{c_{0}}\right)^{1/\left<e\right>^2}\left[\Gamma/ \left(c^2+1\right)\right]^{(1-1/\left<e\right>^2)}$, therefore $T_c$ is almost constant for large $\Gamma$ values. The impurity effect, then, in the high impurity concentration range is the same as in the case of $s$-wave superconductivity, where $T_c$ is not changed by the nonmagnetic impurities. Alternatively, for $\left<e\right>=0$ we observe a strong impurity-induced suppression of the critical temperature \cite{22,23} leading to its zero value at finite impurity concentration, which is reflected by a zero specific heat jump limit value in Eq. (\ref{e18}). Therefore the jump in the specific heat at $T_c\rightarrow 0$ approaches a value of 1.426 for the states belonging to $\Gamma^{+}_1$ representation if their $\left<e\right>$ is nonzero, whereas it decreases to zero for all the other states described by the non identity irreducible representations for which $\left<e\right>=0$. Indeed, as it has been shown for the representative order parameters, \cite{24,25} the gap anisotropy is smeared out by the isotropic impurity scattering in the first case ($\left<e\right>\neq 0$) and the density of states approaches that of an isotropic $s$-wave superconductor. The nonmagnetic impurities, however, are severe pair-breakers in a state with $\left<e\right>=0$ and lead to a finite density of states at the Fermi energy. This qualitative difference is expected to be reflected in the quantities proportional to the density of states, for instance in the specific heat. According to the classification given in Tab. 2, in the limit of $T_c\rightarrow 0$ one should observe the BCS specific heat jump value for a $d_{x^2-y^2}$ state in YBCO ($\Gamma^{+}_1$ representation), but the jump should vanish in BSCCO ($\Gamma^{+}_3$ representation). On the other hand $d_{xy}$ state in YBCO should lead to a zero jump in $T_c\rightarrow 0$ limit, but the BCS-like jump in BSCCO. In the extended $s$-wave state $k_xk_y(k_x^2-k_y^2)$, which belongs to $\Gamma^{+}_3$ representation in both YBCO and BSCCO structures, the specific heat jump should decrease to zero with $T_c\rightarrow 0$ in both compounds. The results for specific heat jump in Born and unitary scattering limits are shown in Figs. 5 and 6 respectively for a circular Fermi surface ($c_x/c_y=1$). The YBCO-type FS with a small orthorhombic anisotropy ($c_x/c_y=0.8$) leads to the solutions presented in Fig. 7 (Born scattering) and Fig. 8 (unitary scattering). Finally, Figs. 9 and 10 correspond to both considered impurity scattering limits in a system with a Fermi surface of a large orthorhombicity given by $c_x/c_y=0.2$. In the case of a not broken tetragonal symmetry (Figs. 5 and 6) the specific heat jump of the identity representations $\Gamma^{+}_1$ with $\left<e\right>\neq 0$ (curves 1,4,5), is almost constant and nonzero, but that of the other $\Gamma^{+}_2$, $\Gamma^{+}_3$ and $\Gamma^{+}_4$ irreducible representations with $\left<e\right>=0$ (curves 6,2 and 3 respectively) is suppressed to zero by the impurities. It is also remarkable in Figs. 5 and 6 that the states having the same values of $\left<e\right>$, $\left<e^3\right>$ and $\left<e^4\right>$ averages (compare Figs. 1a-1c at $c_x/c_y=1$) display the same values of the jump in the specific heat. When the FS symmetry changes into the orthorhombic there are four $\Gamma^{+}_1$ representations in the set of considered functions (Tabs. 1, 2). We look at YBCO symmetry first. Except for two states of $\Gamma^{+}_3$ symmetry ($k_xk_y$, $k_xk_y(k^2_x-k^2_y)$) which are strongly suppressed by the nonmagnetic impurity scattering, the specific heat jump of all the other states goes to a BCS value of 1.426 in $T_c\rightarrow 0$ limit. For a small orthorhombic anisotropy a dramatic rise of the specific heat jump is seen only in the $d_{x^2-y^2}$ state (Figs. 7 and 8), whereas for a large FS anisotropy this increase, however not so abrupt, is observed for all four identity representations (Figs. 9 and 10). In the case of BSCCO symmetry (see Tab. 2) the sharp rise in the specific heat jump for a small orthorhombic FS anisotropy (Figs. 7 and 8) will be observed only for the $d_{xy}$ state. The $d_{x^2-y^2}$ state will not mix with the identity representations in BSCCO, which will lead to a zero value of the specific heat jump in the large impurity scattering limit. Therefore observation of a sharp rise in the specific heat jump at the phase transition with the transition temperature $T_c\rightarrow 0$ in dirty BSCCO would provide information about the possible presence of $d_{xy}$ state in the condensate. Note that in impure YBCO the same observation would signal a realization of the $d_{x^2-y^2}$ state. We do not consider the mixed states since it would require an analysis of many cases even if we confine ourselves to the representations from Tab. 1. However, one should take into account that for a small FS anisotropy (Figs. 7 and 8) the rise in the specific heat jump at $T_c\rightarrow 0$ will be less pronounced for a superposition of some $\Gamma^{+}_1$ states than for a single $d_{x^2-y^2}$ state in the case of YBCO or a $d_{xy}$ one in BSCCO. Comparing the effect of Born scattering on the jump in the specific heat (Figs. 5, 7, 9) with that of the unitary impurity scattering (Figs. 6, 8, 10), we notice a difference in the range of medium impurity concentration for the states with the small $\left<e\right>$ values (see Fig. 1a). It means that the states suppressed by the impurity scattering the most, display the largest difference between these two scattering limits, which is particularly seen in the states of $\left<e\right>=0$, where the pair-breaking effect of the impurities is the strongest (Eq. (\ref{e14b})). \section{Conclusion} The orthorhombic anisotropy of the cuprates results in a number of states belonging to the identity irreducible representation ($\Gamma^{+}_1$). For these states the average value of the order parameter over the Fermi surface ($\left<e\right>$) may not vanish. The behavior for states with $\left<e\right>\neq 0$ is qualitatively different than for states with $\left<e\right>=0$. We have studied the effect of nonmagnetic impurity scattering on the superconducting states which may be realized in the high temperature superconductors, distinguishing between the orthorhombicity of YBCO and BSCCO compounds. It has been pointed out, that the potential scattering stabilizes the $\Gamma^{+}_1$ states (with $\left<e\right>\neq 0$) against those with $\left<e\right>=0$. A very interesting feature may be observed in the electronic specific heat jump at the phase transition for a small crystal orthorhombicity. The $d_{x^2-y^2}$ state in YBCO and the $d_{xy}$ state in BSCCO are predicted to lead to a sharp rise in the jump in specific heat at $T_c$ in the limit $T_c\rightarrow 0$ when the critical temperature is suppressed by the potential impurity scattering. This effect may be helpful in the identification of the superconducting states in the cuprates. \medskip \section*{Acknowledgment} \noindent Work supported by the Natural Sciences and Engineering Research Council of Canada. \newpage
\section{Introduction} In many studies of the interstellar medium, it is essential to determine accurately the amount of molecular gas along the line of sight. However, the most abundant molecule, H$_2$, not only has a large excitation temperature ($T_{\rm ex}\ge$509~K), which is not usually attainable in a cold interstellar medium, but also has very small rotational transition probabilities. In order to infer the H$_2$ column density $N({\rm H_2})$, therefore, we generally rely on indirect methods. The most well-known indirect method for estimating the H$_2$ column density is to observe other molecules such as CO, CS, or NH$_3$. However, the fractional abundance of a molecule relative to H$_2$ is unlikely to be uniform from one cloud to another, or even sometimes within a cloud. In addition, there may be some differences among the distributions of individual molecules because of their different photochemical properties. In the case of CO, which has been widely used as a tracer of H$_2$, the abundance varies from cloud to cloud by up to three orders of magnitude depending on the visual extinction and astrochemical properties (Scoville $\&$ Sanders 1987; van Dishoeck $\&$ Black 1988; Magnani $\&$ Onello 1995). The infrared all-sky maps produced by the {\it InfraRed Astronomical Satellite} ($IRAS$) mission presented a new opportunity to study the distribution of interstellar molecular gas. The Galactic radiation in the far-infrared (far-IR) appears to arise mostly from dust grains well-mixed with interstellar gas (Mathis, Mezger, $\&$ Panagia 1983). The $IRAS$ 100~$\mu$m\ emission intensity, $I_{100}$, has been found to be tightly correlated with the total column density of interstellar gas at $|b| \geq$5$^\circ$, and the 100~$\mu$m\ emissivity per hydrogen nucleus, $I_{100}$/$N({\rm H})$, seems to be fairly uniform in the solar neighborhood : $I_{100}$/$N({\rm H})$$\simeq$1~MJy~sr$^{-1}$~{(10$^{20}$~cm$^{-2}$)}$^{-1}$~(de Vries et al. 1987; Boulanger $\&$ P\'{e}rault 1988; Heiles, Reach, $\&$ Koo 1988; Deul $\&$ Burton 1990, 1992). Therefore, \ihund~excess, i.e., the $IRAS$ 100~$\mu$m\ emission in excess of what is expected from the HI column density, could be used as a tracer of H$_2$ in regions where the amount of ionized gas is negligible. Based on such an idea, several groups have identified ``infrared-excess (IR-excess)'' clouds in order to investigate the distribution of molecular gas. However CO observational studies on the IR-excess\ clouds showed that there is a substantial difference between the distribution of IR-excess\ clouds and that of CO-emitting clouds (D\'{e}sert, Bazell, $\&$ Boulanger 1988; Blitz, Bazell, $\&$ D\'{e}sert 1990; Heithausen et al. 1993; Reach, Koo, $\&$ Heiles 1994; Meyerdierks $\&$ Heithausen 1996; Reach, Wall, $\&$ Odegard 1998). This discrepancy between the two distributions is in a sense expected, because firstly the dust abundance and radiation-field strength vary, and secondly the CO abundance is low in diffuse molecular clouds, and finally the infrared emissivity of the dust associated with molecular gas is low. Another $IRAS$ method for tracing molecular gas is to use 100~$\mu$m\ optical depth, $\tau_{100}$. This might be more accurate than using 100~$\mu$m\ intensity, at least in principle, because optical depth is simply proportional to the amount of dust along the line of sight if dust properties are the same. Indeed, for several high-latitude clouds, 100~$\mu$m\ (or 60~$\mu$m) optical depth has been found to be correlated with molecular gas column density better than the 100~$\mu$m\ intensity alone (Langer et al. 1989; Snell, Heyer, $\&$ Schloerb 1989; Jarrett, Dickman, $\&$ Herbst 1989). `IRAS' clouds similar to the IR-excess\ clouds have been identified using this method, too (Wood et al. 1994). These previous studies, which were limited to high latitudes ($|b|$$\gtrsim$10$^\circ$), however, simply calculated dust optical depth after subtracting flat background emission from the 60 and 100~$\mu$m\ intensity maps and showed that it had a very good correlation with the integrated intensities of CO isotopes. In this paper, we make a detailed comparison of 100~$\mu$m\ intensity, 100~$\mu$m\ optical depth, and the distribution of CO emission along several one-dimensional cuts. Our study differs from most previous studies in that we use both high-resolution ($\sim$3$'$) HI and CO data in the analysis (cf. Reach et al. 1994). The HI data make it possible to derive accurate correlations between $\tau_{100}$\ (or $I_{100}$) and $N({\rm HI})$, which is used to identify `\thund-excess' (or \ihund-excess) regions and to estimate the excess column densities of H nuclei. This excess H column density is compared with the distribution of CO integrated intensity, $W_{\rm CO}$, to derive the important conversion factor $X$$\equiv$$N({\rm H_2})$/$W_{\rm CO}$. For comparison, most previous studies (e.g., de Vries et al. 1987; Heithausen $\&$ Thaddeus 1990) that derived the conversion factor from the $IRAS$ data have been based on the comparison of \ihund~excess\ and $W_{\rm CO}$, which may significantly underestimate $X$ (Magnani $\&$ Onello 1995; Reach et al. 1998). In this paper we compare these two approaches in detail and discuss their limitations. The object that we study in this paper is the galactic worm GW46.4+5.5. Galactic worms are wiggly, vertical structures which look like worms crawling away from the Galactic plane in median-filtered HI maps (Heiles 1984; Koo, Heiles, $\&$ Reach 1992). GW46.4+5.5\ is a $\sim$8$^\circ$ long, filamentary structure extending vertically from the Galactic plane in both far-IR (or HI) and radio continuum emission (Figure 1). It is possible that GW46.4+5.5\ is the wall of a supershell similar to the North Polar Spur, but at a greater distance (Kim, Koo $\&$ Heiles 1999). We describe the HI and CO line observations in Section 2. In Section 3 we evaluate $I_{100}$\ and $\tau_{100}$\ excesses using the $IRAS$ and HI data, and compare their distributions with that of $W_{\rm CO}$. We discuss the nature of \ihund~excess\ in GW46.4+5.5\ and some potential problems of \thund-excess\ method in Section 4. The main conclusions are summarized in Section 5. \section{Observations} HI 21~cm line observations were carried out using the 305 m telescope at Arecibo Observatory in 1990 October. The telescope had a HPBW of 3.$'$3 and a beam efficiency of 0.8 at 1.4 GHz (Reach, Koo, $\&$ Heiles 1994). We observed both circular polarizations simultaneously using two 1024 channel correlators with 5 MHz bandwidth each, so that the velocity resolution was 2.03~km~s$^{-1}$\ after Hanning smoothing. Each spectrum was obtained by integrating for 1 minute using frequency switching. We made a total of 9 one-dimensional cuts through GW46.4+5.5\ at a set of constant galactic latitudes. The beam separation was 3$'$. The positions of our one-dimensional cuts are listed in Table 1. CO J=1$-$0 line observations were made in 1994 February using the 4~m telescope (HPBW=2.$'$5) at Nagoya University in Japan. We obtained 8 one-dimensional cuts in the same way as in the HI line observations. The observed positions together with their reference positions are summarized in Table 1. An SIS mixer receiver and a 1664 channel Acousto-Optical Spectrometer (AOS) with 40 MHz bandwidth were used. The velocity resolution was 0.67~km~s$^{-1}$\ after Gaussian smoothing. The system temperature varied in the range 480$-$700~K during the observing sessions depending on weather conditions and elevation of the source. Absolute-position switching instead of frequency switching was used in order to prevent the contamination of spectra by the atmospheric CO emission. The reference positions were checked to be free of appreciable ($T^*_{\rm R}$$<$0.1~K) CO emission. The velocity was centered at $v_{\rm LSR}$=40~km~s$^{-1}$\ and the velocity coverage was from $-$19 to +91~km~s$^{-1}$. The on-source integration time was 2 minutes and the typical rms noise level was 0.2~K per channel after Gaussian smoothing. The intensity scale was calibrated with respect to the standard source S140 which was assumed to have $T^*_{\rm R}$=20~K (Yang $\&$ Fukui 1992). \section{Infrared Excess and Molecular Gas} \subsection{HI Gas and Infrared Excess} Figure 2 shows a sample from our HI spectrum at ($l$, $b$)=(45.$^\circ$10, 4.$^\circ$00) where the strongest CO line emission was detected. There are three peaks at positive velocities in the HI spectrum and the peaks in the velocity range $v_{\rm LSR}$$\simeq$18$-$40~km~s$^{-1}$\ are the components associated with GW46.4+5.5\ (Kim et al. 1999). The components at negative velocities might be from the warped Galactic plane outside of the solar circle. The physical and dynamical properties of the worm will be discussed in a separate paper (Kim et al. 1999). We assume that the HI emission is optically thin and compute the HI column density along a given line of sight, $N$(HI), from \begin{equation} N({\rm HI})~=~1.822~\times~10^{18}~\int{T_{\rm b}}~{\rm d}v~~~~{\rm cm}^{-2}, \end{equation} \noindent where $T_{\rm b}$ is the brightness temperature in K and $v$ is the velocity in km~s$^{-1}$. Our assumption seems to be valid because the peak temperature of the HI line is much lower than the typical spin temperature of HI gas, $T_{\rm s}$=125~K, for most sight lines. The integral range is from $v_{\rm LSR}$=$-$100 to +150~km~s$^{-1}$. Figure 3a is the plot of the 100 $\mu$m\ intensity, $I_{100}$, against $N$(HI) in the observed regions. We can see a strong correlation between the two physical parameters. Different beam sizes of the HI and $IRAS$ observations could introduce some scattered data points in the figure. But the effect must be small because the beam sizes are roughly the same ($\sim$3$'$). The figure shows only the data from the regions at $b \ge$3$^\circ$\ where both HI and CO line observations were made. The data at lower galactic latitudes are not used because of the possible contribution from dust associated with ionized gas. The open circles, squares, and pentagons represent the points with detectable CO emission at $b$~=~3$^\circ$, 4$^\circ$, and 5$^\circ$, respectively. The area of each symbol is proportional to the integrated CO intensity. The crosses represent the points without detectable CO. A least-squares fit excluding the points with detectable CO yields ($I_{100}$/MJy~sr$^{-1}$)~=~(1.32$\pm$0.02) ($N({\rm HI})$/10$^{20}$~cm$^{-2}$)~$-$~3.37$\pm$0.7. Our fitting procedure involves two steps: (1) All the crosses are fitted to a straight line, and (2) the crosses within $\pm$2$\sigma$ deviation from the first fit are fitted to a new straight line. The negative zero-intercept could be from errors in the subtraction of zodiacal light. The derived $I_{100}$/$N$(HI) agrees with the results of previous observational studies. For example, Boulanger et al. (1986), de Vries et al (1987), and Heiles et al. (1988) obtained 1.4$\pm$0.3, 1.0$\pm$0.4, and $\sim$1.3 MJy~sr$^{-1}$~{(10$^{20}$~cm$^{-2}$)}$^{-1}$, respectively. We now compare the HI column density and 100~$\mu$m\ optical depth. Assuming that the far-IR emission is optically thin and that the temperature of dust grains is constant along the line of sight, the 60/100 $\mu$m\ color temperature, $T_{\rm d}$, can be derived according to the formula \begin{equation} T_{\rm d} = {c_1 \biggl[\frac{1}{\lambda_{\rm 60}} - \frac{1}{\lambda_{\rm 100}} \biggr]}/ {{\rm ln} \biggl[\frac{I_{\rm 100}}{I_{\rm 60}} {\biggl(\frac{\lambda_{\rm 100}}{\lambda_{\rm 60}} \biggr)}^{n+3} \biggr]}, \end{equation} \noindent where $c_1$=$hc/k_{\rm B}$=1.441 and $n$ is the index in the emissivity law, $Q_{\rm abs}(\lambda) \sim \lambda^{-n}$. For the graphite-silicate dust grain model of Mathis, Rumpl, $\&$ Norsdsieck (1977; hereafter MRN dust model), $n$ lies between 1 and 2 (Draine $\&$ Lee 1984). We take $n$=1.5 in this paper. The 100 $\mu$m\ optical depth, $\tau_{100}$, is derived from $\tau_{100}$=$I_{100}$/$B_{\rm \lambda}$($T_{\rm d}$), where $B_{\rm \lambda}$(T) is the Planck function. Figure 3b displays the 100 $\mu$m\ optical depth versus the HI column density. A strong correlation also exists between these two quantities. A least-squares fit has been performed in a similar way to yield ($\tau_{100}$/10$^{-5}$) = (1.00$\pm$0.02) ($N({\rm HI})$/10$^{20}$ cm$^{-2}$) + 0.01$\pm$0.6. The estimated $\tau_{100}$/$N({\rm H})$\ ratio is much smaller than the value $<\tau_{\rm 100}/N({\rm H})>$ = 6.3 $\times10^{-5}$ {(10$^{20}$ cm$^{-2}$)}$^{-1}$ obtained by Boulanger et al. (1996) from an analysis of the {\it Cosmic Background Explorer} ($COBE$) mission and the Leiden-Dwingeloo HI survey data. This difference between the two values cannot be entirely due to the calibration difference between the $IRAS$ and $COBE$ observations. Instead, the difference may be attributed to the effects of small, transiently heated dust grains (Langer et al. 1989) and to the insensitivity of the $IRAS$ observations to cold ($T_{\rm d}$$ <$15~K) dust grains (Snell et al. 1989; Jarrett et al. 1989; Wood et al. 1994). If the dust-to-gas ratio and the infrared emissivity per hydrogen nucleus are uniform over the atomic and molecular regions, the infrared excess can be used to estimate the H$_{\rm 2}$ column density. The infrared excess may be determined either from $I_{100}$~or $\tau_{100}$\ using the following formulae: \begin{equation} N({\rm H_{\rm 2}})_{I_{100}}~\equiv~\frac{1}{2} \bigg[ \frac{I_{\rm 100,c}}{<I_{\rm 100,c}/N({\rm H})>}~-~N({\rm HI}) \biggr] ~~~~{\rm cm}^{-2} \end{equation} \noindent and \begin{equation} N({\rm H_{\rm 2}})_{\tau_{100}}~\equiv~\frac{1}{2} \biggl[ \frac{\tau_{\rm 100,c}}{<\tau_{\rm 100,c}/N({\rm H})>}~-~N({\rm HI}) \biggr] ~~~~{\rm cm}^{-2}, \end{equation} \noindent where the subscript `c' indicates the quantity corrected for offset and the angle bracket indicates an average ratio. Since we excluded the data points with detectable CO, we may use the derived ratios as the ratios with respect to the column density of {\it hydrogen nuclei}, i.e., $<I_{\rm 100}/N({\rm H})>$=1.32 MJy~sr$^{-1}$~{(10$^{20}$~cm$^{-2}$)}$^{-1}$ and $<\tau_{\rm 100}/N({\rm H})>$=1.00$\times 10^{-5}$ {(10$^{20}$~cm$^{-2}$)}$^{-1}$, respectively. \subsection{Comparison of Infrared Excess and CO Line Intensity} We have detected CO emission in five regions (Table 2). Figure 2 shows the spectrum of the strongest CO line and compares it with the corresponding HI spectrum. Note that the velocity range of our CO observations was from $-$19 to +91~km~s$^{-1}$, so that we could not detect the CO emission, if any, associated with the HI gas at negative velocities. It is, however, very unlikely that there is appreciable CO emission in this outer part of the Galaxy. We, in fact, made some test observations with enough velocity coverage, but could not detect any emission ($T^*_{\rm R}$$<$0.2~K). In any case, we believe that we have detected the emission from most of the CO gas. The CO emission from regions C and D is almost certainly associated with GW46.4+5.5, while that from regions A and B is probably not because the velocities are very different. The CO emission from region E is also possibly related to the worm although its central velocity is somewhat lower than the velocity of the worm. The physical association of molecular gas with GW46.4+5.5\ will be discussed in a separate paper (Kim et al. 1999). As we have mentioned in Section 3.1, the regions at $b$=1$^\circ$\ (A and B) are not included in our analysis because of the complication due to the presence of ionized gas. Region E is not included because it has only a few observed points. We thus limit our subsequent analysis to regions C and D. Figures 4a and 4b compare the values of $N$(H$_{\rm 2}$) derived from equations (3) and (4) with the integrated CO line intensity in regions C and D, respectively. The integral ranges are from $v_{\rm LSR}$=+20 to +40~km~s$^{-1}$\ for region C and from $v_{\rm LSR}$=+15 to +30~km~s$^{-1}$\ for region D. These are velocity ranges where the CO emission is detected. In region C, the $N({\rm H_2})$$_{\tau_{100}}$ distribution matches very well that of $W_{\rm CO}$, but the $N({\rm H_2})$$_{I_{100}}$ distribution does {\it not}. The $N({\rm H_2})$$_{I_{100}}$ distribution differs from that of $W_{\rm CO}$\ significantly at $b$$\simeq$46$^\circ$$-$47$^\circ$. In region D, on the other hand, both $N({\rm H_2})$$_{\tau_{100}}$ and $N({\rm H_2})$$_{I_{100}}$ match very well with $W_{\rm CO}$\ although their absolute scales differ by about a factor of 2. Hence, \thund~excess\ appears to trace molecular gas better than \ihund~excess, which is also obvious from Figures 3a and 3b since the points with CO are separated more clearly from those without CO in the latter. This is more clearly shown in Figure 5, where $N({\rm H_2})$$_{\tau_{100}}$ and $N({\rm H_2})$$_{I_{100}}$ are compared with $W_{\rm CO}$. $N({\rm H_2})$$_{\tau_{100}}$ is obviously proportional to $W_{\rm CO}$\ whereas $N({\rm H_2})$$_{I_{100}}$ has much weaker correlation with $W_{\rm CO}$. We determined $N({\rm H_2})$$_{\tau_{100}}$/$W_{\rm CO}$=(0.70$\pm$0.05)$\times 10^{20}$ cm$^{-2}$~{(K~\kms)}$^{-1}$\ by a least-squares fit. The error quoted represents only a statistical error in the fit. This conversion factor is much smaller than the estimated value for molecular clouds in the Galactic plane, $X$=(1.8$-$4.8) $\times$ 10$^{20}$ cm$^{-2}$~{(K~\kms)}$^{-1}$\ (Scoville $\&$ Sanders 1987), but comparable to that of high latitude clouds, $X$ $\simeq$ 0.5 $\times$ 10$^{20}$ cm$^{-2}$~{(K~\kms)}$^{-1}$\ (e.g., de Vries et al. 1987; Heithausen $\&$ Thaddeus 1990; Reach et al. 1998). Since there seems to be a systematic difference between the correlations in regions C and D, we also determined the slopes of the correlations separately. They are (0.40$\pm$0.05) and (0.93$\pm$0.06)$\times 10^{20}$ cm$^{-2}$~{(K~\kms)}$^{-1}$, respectively. This difference could be due to variations in the dust-to-gas ratio (Magnani $\&$ Onello 1995) or due to variations of the gas-phase carbon abundance (Heithausen $\&$ Mebold 1989). Further study is needed to address this question. As we will discuss in more detail in Section 4.2, the correlation between $W_{\rm CO}$\ and $N({\rm H_2})$$_{I_{100}}$ is not apparent partly because of the existence of a local heating source in region C. If we limit the analysis to region D, there is a weak correlation between the two quantities, $N({\rm H_2})$$_{I_{100}}$/$W_{\rm CO}$ = (0.59$\pm$0.05) $\times 10^{20}$ cm$^{-2}$~{(K~\kms)}$^{-1}$. Note that the coefficient of proportionality is about 1/2 that of the $W_{\rm CO}$-$N({\rm H_2})$$_{\tau_{100}}$ relationship, the significance of which will be discussed in next section. \section{Discussions} \subsection{IR~excess\ as a Molecular Tracer} \noindent \centerline{\it 4.1.1~~\ihund~excess\ as a molecular tracer} We have found that the \ihund-excess\ distribution does not correlate well with the CO distribution. The discrepancy between the two distributions has also been noted in several previous studies (D\'{e}sert et al. 1988; Blitz et al. 1990; Heithausen et al. 1993; Reach et al. 1994; Meyerdierks $\&$ Heithausen 1996; Reach et al. 1998). In general, the regions with such discrepancies are divided into two categories: (a) \ihund-excess\ regions without CO emission, and (b) CO-emitting regions without \ihund~excess. The sources that fall into the first category could be regions with enhanced 100~$\mu$m\ emissivity. The 100~$\mu$m\ emissivity is generally dependent on the dust-to-gas ratio and dust temperature. It has been found that the dust-to-gas ratio changes with position on an angular scale of $\sim$10$^\circ$\ and by up to a factor of 4 (Burstein $\&$ Heiles 1978; D\'{e}sert et al. 1988). Variation in the dust-to-gas ratios on smaller scales is expected, too. The steady-state temperature of dust grains is determined by their optical properties and the intensity of radiation field (e.g., Draine $\&$ Anderson 1985). The \ihund-excess\ region at $l$$\simeq$46$^\circ$\ in region C is due to the strong radiation field from a nearby O star (see Section 4.2). Alternatively, the sources could be diffuse molecular clouds without detectable CO (Blitz et al. 1990; Reach et al. 1994; Meyerdierks $\&$ Heithausen 1996; Reach et al. 1998). This is possible because CO may not be shielded by HI, H$_2$, and dust from photodissociation, and is unlikely to be collisionally excited in the low-density region. On the other hand, the sources in the second category, i.e., molecular clouds without \ihund~excess, could be present because of the low emissivity of dust grains associated with molecular gas. If there were no embedded heating source in the molecular cloud, the dust temperature drops from the surface to the center of the cloud. According to Mathis et al. (1983), for example, the dust temperature drops by about 3~K from the edge to the center for a cloud with $A_{\rm v}$=5~mag. The 100~$\mu$m\ emissivity of dust grains is very sensitive to dust temperature and changes by about 40\% as dust temperature changes by 1~K over the range 18$\leq T_{\rm d} \leq$25~K. At $l$$\simeq$46.\degrs3 in region C, the low emissivity almost completely counterbalances the significant fraction of \ihund~excess\ from dust grains associated with molecular gas. The low far-IR emissivity of molecular clouds has another important implication for using \ihund~excess\ as a molecular tracer. {\it The \ihund~excess\ significantly underestimates the H$_2$ column density.} In region D, the H$_2$ column density derived from \ihund~excess\ is smaller than that derived from \thund~excess\ by about a factor of 2. This is because in equation (3) we used $<I_{100}/N{\rm (H)}>$ obtained from HI line observations. If there is a molecular cloud along the line of sight, a correct expression for the observed 100~$\mu$m\ emission would be \begin{equation} I_{100} = {<I_{100}/N{\rm (H)}>}_{\rm HI} N{\rm (HI)} + 2{<I_{100}/N{\rm (H)}>}_{\rm H_2} N{\rm (H_2)}, \end{equation} \noindent where ${<I_{100}/N{\rm (H)}>}_{\rm HI}$ and ${<I_{100}/N{\rm (H)}>}_{\rm H_2}$ are, respectively, 100~$\mu$m\ emissivities per hydrogen nucleus in atomic and molecular regions. Hence, we need to apply a correction factor $\xi_{\rm c}$ to $N({\rm H_2})$$_{I_{100}}$ in equation (3) in order to obtain an accurate H$_2$ column density: \begin{equation} \xi_{\rm c} \equiv {<I_{100}/N{\rm (H)}>}_{\rm HI} / {<I_{100}/N{\rm (H)}>}_{\rm H_2}. \end{equation} \noindent Since ${<I_{100}/N{\rm (H)}>}_{\rm HI}$ is in general greater than ${<I_{100}/N{\rm (H)}>}_{\rm H_2}$, $\xi_{\rm c} \ge$1 and, in GW46.4+5.5, $\xi_{\rm c} \simeq$2. Our correction factor $\xi_{\rm c}$ is equivalent to ${<B_\nu (T_{\rm d})>}_{\rm HI}$ /${<B_\nu (T_{\rm d})>}_{\rm H_2}$ used by Magnani $\&$ Onello (1995) and Reach et al. (1998). From their study of \ihund-excess\ clouds, Reach et al. (1998), for example, found $\xi_{\rm c}$=3.8 as an average value for the solar neighborhood. For the inner Galaxy, if we use the mean temperature of ${<T_{\rm d}>}_{\rm HI}$=21.0$\pm$1.0~K and ${<T_{\rm d}>}_{\rm H_2}$=19.0$\pm$1.0~K derived by Sodroski et al. (1994) using the Diffuse Infrared Background Experiment (DIRBE) 140 and 240~$\mu$m\ observations, we obtain $\xi_{\rm c}$=2.1. Our result for $\xi_{\rm c}$, however, cannot be compared directly with these results because our study is based on the $IRAS$ 60~$\mu$m\ brightness which is largely affected by the emission from small grains. Since $B_\nu (T_{\rm d})$ is very sensitive to $T_{\rm d}$, it is necessary to estimate the correction factor independently in studying specific regions (Lee, Kim, $\&$ Koo 1999). \noindent \centerline{\it 4.1.2~~\thund~excess\ as a molecular tracer} According to our results, \thund~excess\ appears to be an accurate indicator of the molecular content along the line of sight; it not only traces the presence of molecular gas correctly but also increases linearly with the amount of molecular gas. There are two reasons for this. First it separates the intensity peaks produced by enhanced infrared emissivity. Second it corrects for the low infrared emissivity of molecular clouds. In the present work, however, there are two limitations as discussed below. The first limitation stems from using a mean temperature in deriving $\tau_{100}$\ along the line of sight. The expression for $N({\rm H_2})$\ in equation (4) is, in fact, an exact one if dust properties and the dust-to-gas ratio remain the same in atomic and molecular regions. In practice, however, there is an error because, in deriving $\tau_{100}$\ on the right hand side of equation (4), we have used an average color temperature along the line of sight (defined in equation (2)) even if we need to use different color temperatures in deriving the optical depths of atomic and molecular regions. It is straightforward to estimate the error when there are only two dust components with different emissivities along the line of sight. If there is atomic gas of optical depth $\tau_{100}$(HI) at color temperature $T_{\rm HI}$\ and molecular gas of optical depth $\tau_{100}$(H$_2$) at $T_{\rm H_2}$($<$$T_{\rm HI}$), the ratio of the optical depth $\tau_{100}'$, derived by using a mean color temperature, to the true optical depth $\tau_{100}$ is given by \begin{equation} \frac{\tau_{100}'}{\tau_{100}} = { \biggl [ \frac{1+r~{\rm exp}(-\triangle t)} {1+r~{\rm exp}(-\frac{\triangle t}{\lambda_{60}/\lambda_{100}})} \biggr ] }^{\frac{\lambda_{60}/\lambda_{100}} {1 - \lambda_{60}/\lambda_{100}}} \frac{1+r~{\rm exp}(-\triangle t)}{1+r}, \end{equation} \noindent where $r \equiv \tau_{100}({\rm H_2})/\tau_{100}({\rm HI})$, $\triangle t \equiv$ ($c_1/\lambda_{\rm 100}$)(1/${T_{\rm H_2}}$ $-$ 1/${T_{\rm HI}}$), and the other parameters have the same meaning as in Section 3. The derived optical depth is always less than the true value and, as expected, the difference converges to zero as $r$ approaches zero or infinity. Figure 6 shows, as solid lines, how $\tau_{100}' / \tau_{100}$ varies in the ($r$, $\triangle t$) plane. The ratio $\tau_{100}' / \tau_{100}$ is affected more strongly by the warmer dust associated with atomic gas in the sense that the error in optical depth is larger at $\tau_{100}({\rm H_2})/\tau_{100}({\rm HI})$=$r_{\rm o}$ ($>$1) than that at $\tau_{100}({\rm H_2})/\tau_{100}({\rm HI})$=1/$r_{\rm o}$ for a given value of $\triangle t$. Similar figures can be found in Langer et al. (1989) and Snell et al. (1989), but Figure 6 is a generalized one. The error in optical depth would induce an error in $N({\rm H_2})$$_{\tau_{100}}$. The error can be determined by the following formula: \begin{equation} \frac{N({\rm H_2})_{\tau_{100}'}}{N({\rm H_2})_{\tau_{100}}} = \frac{\tau_{100}'/\tau_{100}-1/(1+r)} {r/(1+r)}. \end{equation} \noindent The dependence of $N({\rm H_2})$$_{\tau_{100}'}$/$N({\rm H_2})$$_{\tau_{100}}$ on ($r$, $\triangle t$) is displayed by dotted lines in Figure 6. The ratio $N({\rm H_2})$$_{\tau_{100}'}$/$N({\rm H_2})$$_{\tau_{100}}$ is in general smaller than $\tau_{100}' / \tau_{100}$ for a given ($r$, $\triangle t$) pair. The most striking difference between the two quantities is that, as $r$ approaches zero, the error in $N({\rm H_2})$$_{\tau_{100}}$ remains constant for given $\triangle t$ whereas that in $\tau_{100}$ tends to zero. Since the visual extinction is $A_V$$\simeq$1.5~mag at the CO peak in region D using $N({\rm H})$/$A_V$=1.9$\times$10$^{21}$~cm$^{-21}$~mag$^{-1}$ (Bohlin, Savage, $\&$ Drake 1978), $T_{\rm H_2}$\ is expected to be $<$3~K lower than $T_{\rm HI}$\ (Mathis et al. 1983). Therefore, the error in $N({\rm H_2})$$_{\tau_{100}}$ may be $\lesssim$20\% in region D where $\tau_{100}({\rm H_2})/\tau_{100}({\rm HI}) \simeq$1/2 (See Figure 6). The error in the correction factor $\xi_{\rm c}$ is directly related to that in $N({\rm H_2})$$_{\tau_{100}}$ because it is actually derived from $N({\rm H_2})$$_{\tau_{100}'}$/$N({\rm H_2})$$_{I_{100}}$: $\xi_{\rm c}' / \xi_{\rm c}$=$N({\rm H_2})$$_{\tau_{100}'}$/$N({\rm H_2})$$_{\tau_{100}}$. The second limitation derives from using $I_{60}$\ in estimating the dust optical depth. Interstellar dust grains are known to be composed of small, transiently heated grains and classical, large grains in thermal equilibrium (Draine $\&$ Anderson 1985; D\'{e}sert, Boulanger, $\&$ Puget 1990). The emission at 60~$\mu$m\ comes from both small grains and large grains, while the emission at 100~$\mu$m\ originates mainly from large grains. D\'{e}sert et al. (1990) divided dust grains into three components: Polycyclic Aromatic Hydrocarbins (PAHs), Very Small Grains (VSGs), and Big Grains (BGs). In their model, the contributions of VSGs are $\sim$60\% at 60~$\mu$m\ and $\sim$15\% at 100~$\mu$m\ for the solar neighborhood. Sodroski et al. (1994) found from an analysis of the DIRBE and IRAS data that the average values of the contribution of small grains at 60~$\mu$m\ are $\sim$40\% and $\sim$60\% in the inner and outer Galaxy, respectively. $IRAS$ studies of nearby molecular clouds suggest that the small grains are present largely in the halo surrounding the clouds and that their relative abundances vary (Boulanger et al. 1990; Laureijs, Clark, $\&$ Prusti 1991; Bernard, Boulanger, Puget 1993). Therefore, $\tau_{100}$\ that we have derived using $I_{60}$/$I_{100}$\ does {\it not} represent the usual ``optical depth" of dust grains. Instead it might be largely affected by small grains with varying abundances. However, the fact that we do observe a very good correlation between \thund~excess\ and $W_{\rm CO}$\ (or $\tau_{100}$\ and $N({\rm HI})$) seems to indicate that $\tau_{100}$\ is still proportional to the amount of dust along the line of sight. Compared to $I_{100}$, $\tau_{100}$\ can be interpreted as a quantity corrected for variation in the abundance of small grains as well as variation in the temperature of large grains. If one uses far-IR data at wavelengths longer than 100~$\mu$m, such as the DIRBE data for the 100, 140, and 240~$\mu$m\ wavebands, one can resolve this problem. In the present days, however, the IRAS data have the highest angular resolution among the available infrared survey data. In the past decade, several groups have found for high-latitude clouds that $\tau_{100}$\ (or $\tau_{60}$) is better correlated with the gas column density than $I_{100}$\ (Langer et al. 1989; Snell et al. 1989; Jarrett et al. 1989; Wood et al. 1994). At high latitudes, however, it is unlikely that the \thund-excess\ distribution is markedly different from the \ihund-excess\ distribution, because most high-latitude clouds are diffuse and quiescent compared to the molecular clouds near the Galactic plane. In addition, the low signal-to-noise ratio of $I_{60}$\ could introduce a large error in $\tau_{100}$. At low latitudes, on the contrary, the \thund-excess\ distribution is expected to represent the distribution of molecular gas much better than the \ihund-excess\ distribution, because the number density of luminous stars increases and most molecular clouds contain high-extinction regions and/or embedded luminous stars. \subsection{Local Heating Source in GW46.4+5.5} In the case of region C, the \ihund-excess\ distribution is largely affected by the variation of dust temperature (Figure 4a). The \ihund-excess\ peak at $b$=46.\degrs0 appears not to be due to dust grains associated with molecular gas but to an increase in dust temperature. We found a massive star, LSII +12.\degrs3, near the position of interest, ($l$, $b$)~=~(46.\degrs0, 3.\degrs0) (Drilling 1975). The star was found to be a luminous $blue$ giant star, O9~III (Vijapurkar $\&$ Drilling 1993). The coordinates of the star are ($\alpha$, $\beta$)$_{1950}$=(19:03:14.3, 12:46:21) or ($l$, $b$)=(45.\degrs97, 2.\degrs75). An infrared point source (IRAS19031+1247) and a radio point source with a diffuse extended envelope (F$\ddot{\rm u}$rst et al. 1990) lie in the vicinity of the star. IRAS19031+1247 seems to be an OH/IR star on the basis of its position in the (60$-$25)~$\mu$m\ versus (25$-$12)~$\mu$m\ color-color diagram (cf. Lewis 1994). In order to reveal the physical relationship between the two sources and the star, further studies are required. We estimate the distance of the massive star to be $d$=2.2$\pm$0.8~kpc, using the spectroscopic data mentioned above and the $UBV$ photometric data of Drilling (1975), $V$=10.$\!\!^m$72, $B-V$=1.$\!\!^m$00. Here we adopt $M_V$=$-$5.$\!\!^m$3$\pm$0.$\!\!^m$7 for O9~III (Conti et al. 1983) and $R \equiv A_V$/$E_{B-V}$=3.3 because $R$ for the region is likely to be somewhat higher than the average value for the Galactic plane (Turner 1976). The estimated value is barely in agreement with the kinematic distance of HI gas associated with GW46.4+5.5, $d_{\rm kin}$=1.4~kpc (Kim et al. 1999). Figure 7 displays the large-scale distribution of the $I_{60}$/$I_{100}$~ratio around region C. The $I_{60}$/$I_{100}$~ratio is enhanced around the massive star, which is located where two arrows meet, and decreases monotonically from the star to the position of interest. Assuming that $Q_{\rm abs}(\lambda)$ varies as $\lambda^{-n}$, the total excess infrared luminosity from the $I_{60}$/$I_{100}$~ enhanced region can be derived from the following formula: \begin{equation} L_{\rm IR}^{\rm ex}~=~4 \sigma d^2 \int_{\Omega_{\rm s}} \Bigl[ <\tau>_{\rm T_{\rm d}} T_{\rm d}^4 - <\tau>_{\rm T{\rm d,B}} T_{\rm d,B}^4 \Bigr] d\Omega, \end{equation} \noindent where $\sigma$ is the Stephen-Boltzmann constant, $d$ is the distance to the massive star, and the subscript B indicates quantities in the absence of the star. The Planck-averaged optical depth is defined as $<\tau>_{\rm T}$=${ \int \tau_{\lambda} B_{\lambda}({\rm T}) d\lambda} / {\int B_{\lambda}({\rm T}) d\lambda}$ ($\propto$~$T_{\rm d}$$^{1.5}$ for $n$=1.5). Using $d$=2.2~kpc and $T_{\rm d,B}$=25.2~K ($I_{60}$/$I_{100}$=0.22 with $n$=1.5), the total excess infrared luminosity is $L_{\rm IR}^{\rm ex}$=6500~$L_\odot$\ over a solid angle of $\Omega_{\rm s}$=2.4$\times$10$^{-4}$~sr. This value is an order of magnitude smaller than the stellar luminosity, $L_*$=2.2$\times$10$^5$~$L_\odot$\ (Panagia 1973). Hence, the star has enough luminosity to produce the observed infrared luminosity, and the \ihund-excess\ peak at $b$=46.\degrs0 in region C is probably due to heating by an O-type giant star. \section{Conclusions} Infrared (IR) excess, defined as the emission in excess of what would be expected from HI emission, has been used for estimating the amount of molecular gas along the line of sight. However, at the same time, there have been molecular, particularly CO, line studies showing that the method does not necessarily yield reliable results. That is, there are both IR-excess\ regions without detectable CO and CO-emitting regions without IR excess. Region C in our study is a good example. The sources that fall into the first category could be either regions with enhanced IR emissivity or diffuse molecular clouds without CO emission. In region C, it is excess heating from a nearby O star that produces an IR-excess region. On the other hand, the sources that fall into the second category could be molecular clouds with low IR emissivity. In region C, the low emissivity completely hides the presence of a molecular cloud in the infrared. Our result shows that we can avoid the above confusion problem by using the optical-depth excess instead of IR excess, i.e., the optical depth in excess of what would be expected from HI emission. We have found a very good correlation between 100~$\mu$m\ optical-depth ($\tau_{100}$) excess and integrated intensity of CO emission, $W_{\rm CO}$. We derive the conversion factor between $N({\rm H_2})$\ and $W_{\rm CO}$, $X$$\simeq$0.7$\times$10$^{20}$ for the galactic worm GW46.4+5.5. According to our result, however, the conversion factors in regions C and D differ by about a factor of 2, although both regions are parts of the worm. A more complete observation is necessary for the study of the variation of the conversion factor. Another merit of using optical-depth excess, which may be more important, is that it gives an {\it accurate} value for the amount of molecular gas. IR excess would significantly underestimate $N({\rm H_2})$\ if one uses an IR emissivity estimated from HI emission, ${<I_{100}/N{\rm (H)}>}_{\rm HI}$. We have introduced a correction factor $\xi_{\rm c} \equiv {<I_{100}/N{\rm (H)}>}_{\rm HI} / {<I_{100}/N{\rm (H)}>}_{\rm H_2}$, which should be applied in order to account for the different IR emissivity of molecular clouds. In general $\xi_c \ge 1$ because dust grains in molecular clouds are usually colder than dust grains in atomic gas. In region D in our study, $\xi_c \simeq 2$. We should therefore be rather cautious in converting IR excess to $N({\rm H_2})$. Two potential problems in using optical-depth excess, however, would be the low surface brightness and the contribution of small, transiently heated grains. If the surface brightness is low, the noise and/or the systematic errors produced in the subtraction of zodiacal emission may make it impossible to derive an accurate optical depth. It would be worthwhile to check the applicability of the method at high galactic latitudes. The second problem may be resolved by using far-IR data at wavelengths longer than 100~$\mu$m. \acknowledgements We wish to thank Dr.~Y. Fukui, Dr.~A. Mizuno, and the graduate students of the Department of Physics and Astrophysics, Nagoya University for their help with the CO line observations. We are very grateful to Dr.~W. T. Reach and the anonymous referee for their comments and suggestions. We also thank Dr.~Hwankyung Sung for helpful discussions, and Mr.~Hyo-Ryung Kim for providing some useful CO line spectra for this work. This work has been supported in part by the 1994 KOSEF International Cooperative Research Fund. \vskip 1.5cm
\section{Introduction} The roughening transition has been studied in great detail, both theoretically and experimentally \cite{Nozieres,BGR}. Direct analogies with the (two dimensional) $XY$-model or the Coulomb gas furthermore make this problem particularly enticing \cite{mappings}. More recently, the role of disorder on the roughening transition or on the properties of the $XY$ model, has attracted considerable interest \cite{Cardy,BG,OS,CLD,EN}. In particular, replica calculations and Functional Renormalization Group ({\sc frg}) methods have been applied to this problem, with sometimes conflicting results \cite{Comment}. In this paper, we wish to reconsider the problem of the roughening transition in the absence of disorder, from a {\sc frg} point of view, where the flow is not {\it a priori} projected onto the first harmonic of the periodic potential. Within a local renormalization scheme, we establish exact equations for the evolution of the full periodic potential $V(\varphi)$, and the surface tension $\gamma$ with the length scale $L=e^\ell$, which we analyze both numerically and analytically, in the low temperature phase. If we start with a sinusoidal periodic potential, the shape of the fixed point potential $V^*(\varphi)$ evolves to a nearly parabolic shape with matching points becoming more and more singular as the length scale increases. The nature of the singularity is investigated in detail close to the fixed point, that is for small values of the rescaled temperature $\displaystyle \overline T = \frac{T}{2\pi \gamma \lambda^2}$ where $\lambda$ is the periodicity of the potential and $\gamma$ the elastic stiffness. We find that the width $\Delta \varphi$ of the singular region scales as $L^{-3g(\overline T)/5}$, where $g(\overline T)$ governs the scaling of the surface stiffness with the length scale according to $\gamma(L)\sim L^{g(\overline T)}$. The exponent $g(\overline T)$ tends towards $2$ with negative corrections which we calculate, when $\overline T$ goes to $0$ (i.e. for $L \to \infty$). The paper is organized as follows. In section $\mathbf 2$, we introduce the model: we outline the calculations involved and discuss the differences with the approach of Nozi\`eres and Gallet ({\sc ng}), and briefly examine the problem for $d<2$. We then explain in section $\mathbf 3$, by a mean field argument the origin of the singularity that develops during the renormalization flow. In section $\mathbf 4$, we present a scaling form for the renormalized potential, around its maxima and close to the fixed point, which accounts for the nature of the singularity. Using our renormalization group flow, we compute in section $\mathbf 5$ the step energy as a function of temperature. Finally, in section $\mathbf 6$, we look at the case of a contact line in a periodic potential, as this is a physical realization of a non local elastic stiffness. \section{Model and functional renormalization group} We consider an elastic interface whose height fluctuations are described by a profile $\Phi(x)$, where $x$ is a $d$-dimensional vector, in the presence of a deterministic periodic potential $V$. Supposing that the slope of the interface is everywhere small, the energy of the system is: \begin{equation}\label{energie} H[\Phi]=\frac{\gamma}{2}\int d^dx\ (\nabla \Phi(x))^2 + \displaystyle \int d^dx\ V\biggl(\frac{\Phi(x)}{\lambda}\biggr) \end{equation} \noindent where $\gamma$ is the elastic stiffness and $\lambda$ the periodicity of the potential. In the absence of periodic potential, the height fluctuations of the surface on a length scale $L$ scale as $L^{2-d}$. For $d>2$, the interface is therefore always flat. For the critical dimension $d=2$, the interface is rough only if the temperature exceeds a certain critical temperature $T_R$. When the potential $V$ is harmonic, this model is the continuous version of the Sine-Gordon model. {\sc ng} have studied the statics of this problem using a two-parameter renormalization group scheme, and have written flow equations for $\gamma(L)$ and the amplitude $v_o(L)$ of the periodic potential. They suppose that during the flow, $v_o$ remains small compared with the temperature and neglect all higher harmonics of the potential. Correspondingly, within this procedure, the renormalization scheme ceases to be valid when $v_o$ becomes of the order of the temperature. In the low temperature `flat' phase, this occurs after a finite renormalization since $v_o$ grows with distance. In our calculation, we consider a general periodic function with the only constraint that it should be sufficiently smooth (we shall explain this more quantitatively in the following). Since we re-sum the whole perturbation expansion in $v_o/T$, there is however no constraint on the amplitude of the potential, and the renormalization procedure can be carried on any length scale without interruption. The relevant coupling constant appears to be $v_o/\gamma$ rather than $v_o/T$. During the renormalization flow, we keep track of the whole function $V(\varphi)$ instead of projecting onto the first harmonic, so that we have a more quantitative knowledge of the behaviour of the potential for low temperatures. Technically, we proceed by considering the partition function: \begin{equation}\label{partition1} \displaystyle Z=\int d\Phi(x) \ e^{-\beta H[\Phi(x)]} \end{equation} \noindent We perform the renormalization procedure by splitting the field $\Phi$ into a slowly-varying and a rapidly-varying part as: \begin{equation}\label{split} \Phi(x)=\Phi^<(x)+\Phi^>(x) \end{equation} \noindent The Fourier modes $k$ of $\Phi^<$ are such that $0 \leq |k| \leq |\Lambda|/s$, and those of $\Phi^>$, such that $|\Lambda|/s \leq |k| \leq |\Lambda|$, where $s=e^{d\ell}$, $\displaystyle |\Lambda|$ being a high momentum cut-off, of the order of $1/a$, where $a$ is the lattice spacing. We integrate over the fast modes in the partition function and retain only the terms that renormalize the gradient term and the potential term. The other terms that are generated are discarded as irrelevant. Within this renormalization scheme, our calculation is exact. After some algebra detailed in appendix ${\mathbf A}$, we obtain a set of flow equations for $d=2$, for $\displaystyle \overline V=\frac{V}{\gamma \lambda^2 |\Lambda|^2}$ and $\displaystyle \overline T=\frac{T}{2\pi \gamma \lambda^2 }$, the rescaled potential and temperature (note that $\overline V$ and $\overline T$ are dimensionless): \begin{equation}\label{flow} \begin{array}{l} \displaystyle \frac{d \overline V}{d\ell}=\displaystyle (2-g)\overline V - \pi \frac{{\overline V'}^2}{(1+{\overline V}'')} + \frac{\overline T}{2} \ln(1+{\overline V}'') \\ \\ \displaystyle \frac{d \gamma}{d\ell}= g \gamma \\ \\ \displaystyle \frac{d \overline T}{d\ell}= - g \overline T \end{array} \end{equation} \noindent where $g$ is given by: \begin{equation} \label{Exp_g} g=4\pi \ \displaystyle \int_0^{1} d\varphi \ \frac{\displaystyle{\overline V}'^2(\varphi){\overline V}'''^2(\varphi)} {\biggl(1+{\overline V}''(\varphi)\biggr)^5} + \frac{\overline T}{4} \int_0^{1} d\varphi \ \frac{{\overline V}'''^2(\varphi)}{\biggl(1+{\overline V}''(\varphi)\biggr)^4}\end{equation} \noindent These equations call for some comments. \begin{itemize} \item The relevant perturbative parameter appears to be $\overline V$, rather than $V/T$. In the limit $\overline V \ll 1$, and in the case where the potential is purely harmonic (i.e. $V(\varphi)=v_o \cos(2\pi\varphi)$), the {\sc rg} equations read: \begin{equation}\label{NG} \begin{array}{l} \displaystyle \frac{du_o}{d\ell}= \biggl(2- \frac{\pi T}{\gamma \lambda ^2}\biggr)u_o \\ \\ \displaystyle \frac{d\gamma}{d\ell}= 2 \pi^4 \left(\frac{2 \pi T}{\gamma \lambda^2}\right) \frac{u_o^2}{\gamma^2 \lambda^4} \end{array} \end{equation} \noindent where $u_o=v_o/|\Lambda|^2$. The first equation is trivial and identical to the one in {\sc ng}, and immediately leads to the value of the roughening transition temperature: $T_R=2\gamma_\infty \lambda^2/\pi$, where $\gamma_\infty$ is the renormalized value of $\gamma$. The second is close to, but different from the one obtained in the particular renormalization scheme used by {\sc ng}: near the critical temperature $T_R$, the coefficient between parenthesis is equal to $4$ in our case and to $0.4$ according to {\sc ng}. \item The renormalization of the surface tension, as measured by $g$, is always {\it positive}. One can check that, as has been pointed out by {\sc ng}, if the initial potential is parabolic (i.e. $V(\phi)=v_0 \phi^2$), then the coefficient $g$ vanishes identically, and there is no renormalization of the surface tension. This is indeed expected since in this (quadratic) case, all modes are decoupled. \item The flow equations only make sense if $\overline V''>-1$. We have checked numerically that if this condition is satisfied at the beginning, it prevails throughout the flow. On the other hand, if the initial potential is so steep that this condition is violated, the perturbative calculation is meaningless. This comes from the fact that metastable states, where the surface zig-zags between nearby minima of the potential, appear at the smallest length scales. In this respect, it is useful to note that the last term of the flow equation on $\overline V$ comes from the integration of the Gaussian fluctuations of the fast field around the slow field. The condition $\overline V''>-1$ is a stability condition for these fast modes. If the unrenormalized potential is harmonic (i.e. $V(\varphi)=v_o \cos(2\pi\varphi)$), and the unrenormalized surface tension given by $\gamma_o$, then this condition reads $\displaystyle \frac{v_o}{\gamma} \biggl(\frac{2\pi}{\lambda|\Lambda|}\biggr)^2 < 1$, which simplifies to $\displaystyle \frac{v_o}{\gamma} < 1$ in the case where $\lambda = a$. If the initial value of the potential is too large, one actually expects the transition to become first order (but see \cite{Nozieres}). Actually, a variational calculation indeed predicts the transition to become first order when $\displaystyle \frac{v_o}{\gamma} \geq 1$ \cite{Saito}. \item The fundamentally new term in the above equation is the second term, proportional to $\overline V'^2$, and independent of temperature. This term leads to the appearance of singularities in the flow equation: up to second order in $\overline V$, this equation is close to the Burgers' equation (see below) for which it is well known that shocks develop in time. The fact that this term survives even in the zero temperature limit is at first sight strange, since one could argue that for $T=0^+$, there are no longer any thermal fluctuations, and thus no renormalization. This argument is not correct because we are computing a {\it partition function}, thereby implicitly assuming that the infinite time limit is taken {\it before} the zero temperature limit. Such a non trivial renormalization has also been found in the context of pinned manifolds \cite{BF,BBM}, and can be understood very simply using a mean-field approximation, which we detail in the next section. \item For completeness let us consider the case $1<d<2$. By simple scaling arguments, we can see that for $d<1$ the interface is always rough. For $1<d<2$, we have a roughening transition between a flat phase and a rough phase. In this case, we allow $\Phi$ to renormalize and suppose that $\lambda$ renormalizes in the same way according to: \begin{equation}\label{ren_lambda} \displaystyle \frac{d\lambda}{d\ell}=\zeta \lambda \end{equation} \noindent The other flow equations in terms of the rescaled parameters $\displaystyle \overline V=\frac{V}{\gamma \lambda^2 |\Lambda|^2}$ and $\displaystyle \overline T=\frac{K_d |\Lambda|^{d-2}T}{\gamma \lambda^2 }$ (with $K_d=\displaystyle S_d/(2\pi)^d$ where $S_d$ is the $d$-dimensional sphere) now read: \begin{equation}\label{flow_d} \begin{array}{l} \displaystyle \frac{d \overline V}{d\ell}=\displaystyle (d-g_d-2\zeta)\overline V - \pi \frac{{\overline V'}^2}{(1+{\overline V}'')} + \frac{\overline T}{2} \ln(1+{\overline V}'') \\ \\ \displaystyle \frac{d \gamma}{d\ell}= (g_d+d-2) \gamma \\ \\ \displaystyle \frac{d \overline T}{d\ell}= - (g_d+d-2+2\zeta) \overline T \end{array} \end{equation} \noindent where $g_d$ is given by: \begin{equation} \label{Exp_g_d} \displaystyle g_d=\frac{2}{d} g, \end{equation} \noindent with $g$ given by equation (\ref{Exp_g}) above. These equations have a non trivial fixed point for $g_d=2-d$ and $\zeta=0$. This corresponds to a rescaled temperature $T_R$ and a renormalized rescaled potential $\overline V$ such that equations (\ref{flow_d}) and (\ref{Exp_g_d}) are satisfied. For $1<d<2$, we obtain $T_R$ numerically by proceeding as follows: we self-consistently solve the differential equation on $\overline V$ obtained by putting $\displaystyle \frac{d\overline V}{d\ell}=0$ for different fixed rescaled temperatures $\overline T$, imposing that $g$ is given by equation (\ref{Exp_g}). This enables us to plot $g$ as a function of $\overline T$. The rescaled temperature $\overline T_R$ corresponding to the transition temperature is such that $g_d=2-d$. In the figure (\ref{fig1}), we have plotted the result for $d=3/2$, with $\gamma_o=1$. In that case, $\overline T \simeq 1.2$. \end{itemize} \begin{figure} \centerline{\hbox{\epsfig{figure=Fig1.ps,width=8cm}}} \vskip 0.8cm \caption{$g$ as a function of the rescaled temperature $\overline T$ for d=3/2. The point where $g=2-d=1/2$ determines the transition temperature.} \label{fig1} \end{figure} \section{Mean field analysis and effective potential} In this section, we show on a simplified mean field version of the model how the non linear term $V'^2$ arises in the flow equation of the potential. Using a discrete formulation of the problem and replacing the local elasticity modeled by the surface tension term by a coupling to all neighbours, we can rewrite the energy as: \begin{equation}\label{H_mf} \displaystyle H_{\it mf}(\{\Phi\}_i,\overline \Phi)=\frac{\gamma}{2a} \sum_i^N (\Phi_i-\overline \Phi)^2 +a \sum_i^N V(\Phi_i) \end{equation} \noindent where $\overline \Phi$ is the center of mass of the system, $a$ the lattice spacing and $L=Na$. Implementing the constraint $\overline \Phi =\displaystyle 1/N \sum \Phi_i$ by means of a Lagrange multiplier in the partition function, we have: \begin{equation}\label{Z_mf} \displaystyle Z[\overline \Phi] = \displaystyle \int d\eta \int \prod_i d\Phi_i \ e^{-\displaystyle \beta H_{\it mf}(\{\Phi\}_i,\overline \Phi)- \eta \biggl(N\overline \Phi-\displaystyle \sum_i \Phi_i\biggr)} \end{equation} \noindent which can also be expressed as: \begin{equation}\label{Zfact} \displaystyle Z[\overline \Phi] = \int d\eta \ e^{\displaystyle N \log z(\overline \Phi,\eta)+\displaystyle N \frac{\eta ^2 a}{2\beta \gamma}} \end{equation} \noindent where \begin{equation}\label{zo} \displaystyle z(\overline \Phi,\eta)= \int d\Phi \ e^{\displaystyle -\frac{\beta \gamma}{2a} \biggl(\overline \Phi - \Phi+ \displaystyle \frac{\eta a}{\beta \gamma}\biggr)^2 -\beta aV(\Phi)}= \displaystyle z\biggl(\displaystyle \overline \Phi+\frac{\eta a}{\beta \gamma}\biggr) \end{equation} \noindent We are left with a simpler problem since we now have a one-body problem. We introduce the auxiliary partition function $z_R(\Psi, \tau)$ defined as: \begin{equation}\label{diff} \displaystyle z_R(\Psi, \tau)=\displaystyle \sqrt \frac{\beta \gamma}{2\pi a \tau} \displaystyle \int d\Phi \ e^{\displaystyle -\frac{\beta \gamma}{2a} \frac{(\Psi-\Phi)^2}{\tau} -\beta a V(\Phi)} \end{equation} \noindent Up to a multiplicative constant, one has $\displaystyle z\biggl(\displaystyle \overline \Phi+\frac{\eta a} {\beta \gamma}\biggr)= \displaystyle z_R \biggl(\displaystyle \overline \Phi + \frac{\eta a}{\beta \gamma}, \tau=1\biggr)$, where $z_R(\Psi, \tau)$ verifies the diffusion equation: \begin{equation}\label{equadiff} \displaystyle \frac{\partial z_R}{\partial \tau}= \frac{a}{\beta \gamma} \frac{\partial ^2 z_R}{\partial \Psi ^2} \end{equation} with an initial condition given by: \begin{equation}\label{init} \displaystyle z_R(\Psi, \tau=0)= e^{-\displaystyle \beta aV(\Psi)} \end{equation} \noindent Defining now the effective pinning potential $V_R$ as: \begin{equation}\label{V_R} aV_R\biggl(\overline \Phi + \displaystyle \frac{\eta a}{\beta \gamma},\tau\biggr)= -T\log z_R \biggl(\overline \Phi + \frac{\eta a}{\beta \gamma}, \tau\biggr) \end{equation} \noindent we can then easily show that $V_R$ is the Hopf-Cole solution of the non linear Burgers' equation \cite{BBM}: \begin{equation}\label{diffVren} \displaystyle \frac{\partial V_R}{\partial (\tau a)}=\frac{T}{\gamma} \frac{\partial ^2 V_R}{\partial \Psi ^2} - \frac{a}{\gamma} \biggl(\frac{\partial V_R}{\partial \Psi}\biggr)^2 \end{equation} \noindent where the temperature independent non linear term $V_R'^2$ indeed appears. It is easy to show that when $N \to \infty$ or $T \to 0$, the original partition function can be solved by a saddle point method, leading after a change of variables to an effective potential per unit length: \begin{equation}\label{V_eff_final} \displaystyle V_{\it eff}(\overline \Phi)=V_R (u,\tau=1)-\displaystyle \frac{\gamma}{2a^2}(\overline \Phi-u)^2 \end{equation} \noindent where $u$ is given by: \begin{equation}\label{eta*} \displaystyle \frac{\gamma}{2a}(u-\overline \Phi)= \frac{\partial V_R}{\partial \varphi}(\varphi,\tau=1) \displaystyle \biggr|_{\varphi=u} \end{equation} \noindent Now, by changing $V_R$ to $-V_R$, one can see that $-V_{\it eff}$ can also be written as the solution of a Burgers' equation. It is known from results on the Burgers' equation, that, with `time' $\tau$, the effective potential $V_R$ develops shocks, smoothed out at finite temperature, between which it has a parabolic shape. The appearance of singularities is due to the non linear term in the partial differential equation, {\it which indeed survives in the limit} $T=0$. It is interesting to see how this `toy' renormalization group captures some important features of the full scheme, such as the one shown above for a non disordered potential. \section{Analysis for small $\overline T$} In this section, we go back to the model introduced in section $\mathbf 2$ and analyze the nature of $\overline V$ close to the low temperature fixed point, that is for small values of the rescaled temperature ${\overline T}$. Since $g >0$ in the low temperature phase, this corresponds to the large scale structure of the renormalized potential for all temperatures $T<T_R$. Expanding ${\overline V}$ around one of its minima as ${\overline V}(\varphi)={\overline V}_{m}+\frac{1}{2}\kappa (\varphi-\varphi^*)^2$, and replacing ${\overline V}$ in the flow equation(\ref{flow}), we have \begin{equation}\label{flow_min} \begin{array}{l} \displaystyle \frac{d \overline V_m}{d\ell}=\displaystyle (2-g)\overline V_m + \frac{\overline T}{2} \ln(1+\kappa) \\ \\ \displaystyle \frac{d \kappa}{d\ell}=\displaystyle (2-g)\kappa -2\pi \frac{\kappa ^2}{1+\kappa} \end{array} \end{equation} \noindent One can actually check that a parabolic shape for $\overline V$ is exactly preserved by the renormalization flow. However, since the potential has to be periodic, these parabolas should match periodically around each maximum that is for $\varphi-\varphi^*=0, 1, 2,...$. The region of the maximum is therefore expected to be singular. To investigate the nature of the renormalized periodic potential around its maximum value, we will thus make a scaling ansatz on ${\overline V}''$ for small $\overline T$. For our perturbative calculation to be valid, we expect $\overline V''(0)$ to be $> -1$. Now since we expect a singularity to develop as $\overline T$ goes to $0$, it is probable (and actually self-consistently checked) that $\overline V''(0)$ should tend towards $-1$. As $\overline T$ goes to $0$, we thus make the scaling ansatz: \begin{equation}\label{ansatz_V''} \displaystyle 1+{\overline V}''(\varphi)=\displaystyle {\overline T}^{\delta}{\cal F}' \biggl(\frac{\varphi}{{\overline T}^{\alpha}}\biggr) \end{equation} \noindent where ${\cal F}'(0) > 0$. This means that the width of the singular region behaves as $\Delta \varphi \sim {\overline T}^{\alpha}$. Hence, in the scaling region: \begin{equation}\label{ansatz_V'} \displaystyle {\overline V}'(\varphi)=\displaystyle -\varphi+{\overline T}^{\delta+\alpha}{\cal F} \biggl(\frac{\varphi}{{\overline T}^{\alpha}}\biggr) \end{equation} with ${\cal F}(0) = 0$ to ensure that $\varphi=0$ is a maximum of $\overline V$. Integrating once more the above equation, one finds: \begin{equation}\label{ansatz_V} \displaystyle {\overline V}^{}(\varphi)=\displaystyle {\overline V}_M-\frac{\varphi^2}{2}+ {\overline T}^{\delta+2\alpha}{\cal G}\biggl(\frac{\varphi}{{\overline T}^{\alpha}}\biggr) \end{equation} \noindent with ${\cal G}'= {\cal F}$. Replacing this last equation in the flow equation for $\overline V$, we obtain: \begin{equation}\label{V_max} \frac{d \overline V_M}{d\ell}=(2-g){\overline V}_M + \frac{\overline T}{2}\log{\overline T}^{\delta} \end{equation} \noindent Suppose that equations (\ref{flow_min}) have a fixed point as $\overline T$ goes to zero, and that close to the fixed point one can neglect the left hand side of these equations. This leads to the relation $\kappa=(2-g)/(2\pi-2+g)$. Supposing moreover that the parabolic solution extends almost over a whole period and that the correction brought about by the rounding off of the singularity around the maxima of the potential is negligible, we also have \begin{equation}\label{Eq_1} \displaystyle ({\overline V}_M-{\overline V}_m) \simeq \frac{\kappa}{2} \displaystyle = \frac{2-g}{4\pi-4+2g} \end{equation} \noindent Now, subtracting equation (\ref{V_max}) from equation (\ref{flow_min}), we find in the limit $\overline T \to 0$: \begin{equation}\label{Eq_2} \displaystyle \frac{d}{d\ell}({\overline V}_M-{\overline V}_m)= (2-g)({\overline V}_M-{\overline V}_m) + \frac{\delta}{2} \overline T \log {\overline T} \end{equation} \noindent Combining equations (\ref{Eq_1}) and (\ref{Eq_2}), we find that for the previous equation to have a fixed point as $\overline T \to 0$, $g \to 2$ with negative corrections as: \begin{equation}\label{2-g} (2-g) \simeq \displaystyle \sqrt{2\pi \delta{\overline T}\log \frac{1}{\overline T}} \end{equation} \noindent This result is independent of the way we calculate $g$, the correction to the surface tension. In particular, it shows that at zero temperature, the surface tension diverges as $(L/a)^2$, where $L$ is the size of the system. We can deduce an equation satisfied by ${\cal F}'$, by plugging the ansatz for the derivatives of $\overline V$ in the flow equation for $\overline V'$: \begin{equation}\label{eq_renV'} \displaystyle \frac{d\overline V'}{d\ell}= (2-g-2\pi)\overline V' + 2\pi \frac{\overline V'}{1+\overline V'} + \pi {\overline V'}^2 \frac{\overline V'''}{(1+\overline V'')^2} +\frac{\overline T}{2} \frac{\overline V'''}{1+\overline V''} \end{equation} \noindent Close to the fixed point, we again suppose that to leading order in $\overline T$, $\displaystyle \frac{d\overline V}{d\ell}=0$ in the above equation. Plugging in the ansatz for ${\overline V}'$ and ${\overline V}''$, and considering the leading term in ${\overline T}$, we get to lowest order in $\overline T$: \begin{equation}\label{eq_calF} -\displaystyle \frac{2\pi u}{{\cal F}'(u)} \ {\overline T}^{\alpha-\delta} +\frac{\pi u^2{\cal F}''(u)}{{\cal F}'(u)} \ {\overline T}^{\alpha-\delta} +\frac{{\cal F}''(u)}{2{\cal F}'(u)} \ {\overline T}^{1-\alpha} =0 \end{equation} \noindent We can show that necessarily $\alpha-\delta=1-\alpha$. Indeed, if $1-\alpha < \alpha-\delta$, ${\cal F}''$ would be equal to zero while if $1-\alpha > \alpha-\delta$, ${\cal F}'(0)$ would be equal to zero, both alternatives being thus impossible. Hence, equation (\ref{eq_calF}) can be rewritten as \begin{equation}\label{eq_calF2} \displaystyle \frac{1}{2} \frac{d}{du} \log {\cal F}' - \pi \frac{d}{du} \biggl(\frac{u^2}{{\cal F}'}\biggr) =0 \end{equation} \noindent which yields after integration: \begin{equation}\label{Eq_calF3} \displaystyle {\cal F}'(u)\log \biggl( \frac{{\cal F}'(u)}{{\cal F}'(0)}\biggr)=2\pi u^2 \end{equation} \noindent At this stage, we can note that the exponent relation \begin{equation}\label{Rel_exp1} 2\alpha-\delta=1 \end{equation} \noindent is independent of the scheme used to calculate of $g$ (see Appendix). In the rest of this section, we calculate the exponents $\alpha$ and $\gamma$, and using $g(\overline T \to 0)=2$, we also obtain ${\cal F}'(0)$. These results now somewhat depend on the precise renormalization scheme we use to calculate the correction $g$ to the surface tension. Replacing the derivatives of $\overline V$ by their expressions in terms of ${\cal F}'$ and ${\cal F}''$, in equation (\ref{Exp_g}), and changing variables from $\varphi$ to $\displaystyle u=\frac{\varphi}{{\overline T}^{\alpha}}$, we get to lowest order in $\overline T$: \begin{equation}\label{eq_g(T)} \displaystyle g(\overline T) \simeq {\overline T}^{\alpha-3\delta} \ 8\pi \int_0^{\infty} du \ u^2 \frac{{\cal F}''^2(u)}{{\cal F}'^5(u)} + {\overline T}^{1-\alpha-2\delta} \ \frac{1}{2} \int_0^{\infty} du \ \frac{{\cal F}''^2(u)}{{\cal F}'^4(u)} \end{equation} \noindent From the exponent relation $2\alpha-\delta=1$ derived previously, and the fact that $g(\overline T \to 0)$ is finite, we have another exponent relation $\alpha=3\delta$, so that $\alpha=3/5$ and $\delta=1/5$. We can also show that $g(\overline T \to 0)$ can be expressed in terms of ${\cal F}'(0)$. From expression (\ref{Eq_calF3}), one can see that ${\cal F}'$ is a strictly increasing function on $[0,\infty]$ taking its values in $[e,\infty]$, and so we can change variables from ${\cal F}'$ to its inverse function. Defining a new variable $x$ as \begin{equation}\label{cgt_var} \displaystyle x = \frac{e {\cal F}'(u)}{{\cal F}'(0)} \end{equation} \noindent we have: \begin{equation}\label{eq_u(x)} \displaystyle u= \biggl( \frac{{\cal F}'(0)}{2\pi e} \biggr)^{1/2} (x\log (x))^{1/2} \end{equation} \noindent and \begin{equation}\label{eq_du/dx} \displaystyle \frac{du}{dx}= \frac{1}{2} \frac{\log(x)}{(x\log(x))^{1/2}} \biggl( \frac{2\pi e}{{\cal F}'(0)} \biggr)^{1/2} \end{equation} \noindent We can now express $g(\overline T \to 0)$ in terms of an integral over $x$ from $e$ to $\infty$ as \begin{equation}\label{eq_g(0)} \displaystyle g(0)= \frac{2}{\pi ^2} \biggl(\frac{2\pi e}{{\cal F}'(0)} \biggr)^{5/2} \biggl\{ \int_{e}^{\infty} dx \ \frac{(x \log (x/e))^{3/2}}{x^5\log (x)} + \int_{e}^{\infty} dx \ \frac{(x \log (x/e))^{1/2}}{x^4\log (x)} \biggr\} \end{equation} Hence, using the fact that $g(\overline T \to 0)=2$, we finally find the constant ${\cal F}'(0) \simeq 1.38$. \section{Step energy as a function of temperature} From physical considerations we know that below the roughnening temperature, the interface grows by forming terraces. An important quantity governing the kinetics of growth is therefore the step energy. The width $\xi$ of a step and its energy per unit length $\beta_S$ can be obtained by comparing the elastic energy and the potential energy of a profile $\Phi(x)$ which changes by one period over the length $\xi$. Requiring that these two energies are of the same order of magnitude leads to: $\displaystyle \gamma/\xi ^2 \sim v_o$ where $v_o$ is the amplitude of the periodic potential, or $\displaystyle \xi \sim \sqrt {\frac{\gamma}{v_o}}$, and a step energy which scales as $\displaystyle \beta_S \propto \sqrt{{v_o}{\gamma}}$. Since a step profile include Fourier modes such that $\xi^{-1} < k < |\Lambda|$, it is natural to use in the above equations the values of $\gamma$ and $v_o$ calculated for the length $L = a e^{\ell} = \xi$. Since $\xi(L) \sim L/\sqrt{\overline{v}_o(L)}$, one sees that this corresponds to stopping the renormalization procedure when $\overline{v}_o(L) \sim 1$. We have integrated numerically the {\sc rg} flow, starting from $\gamma=1$ and from harmonic potentials of various amplitudes $v_o \ll 1$, and stopping for an arbitrary value $\overline V$, chosen here to be $\overline v_c=0.4$. \footnote{Other values of $\overline v_c$ would not change the qualitative features reported below, provided $\overline v_c$ is not too large.} The resulting step energy as a function of temperature is plotted in Figure (\ref{figure2}). For $T$ close to $T_R$, one finds that $\xi$ diverges as $\displaystyle e^{1/\sqrt{T_R-T}}$, as it should since our {\sc rg} flow essentially boils down to the standard one \cite{Nozieres}. For small temperatures, however, we find that $\beta_S$ tends to a finite value with a linear slope in temperature. This slope is seen to decrease as the initial amplitude of the potential $\overline v_o$ increases. For $\overline v_o=0.01$, $\beta_S$ decreases by $\sim 30 \%$ when $T$ increases from $0$ to $0.25\ T_R$. This decrease falls to $\sim 10 \%$ for $\overline v_o=0.1$. Experiments on Helium 4, on the other hand, have established that the step energy does only depend very weakly on temperature at small temperature, by not more than $5\%$ when the temperatures varies from $0.05 \ T_R$ to $0.25 \ T_R$. This suggests that the initial amplitude of the potential is of the same order as $\gamma_o$: in this case, the width of the step is of order $a$, and the bare parameters are not renormalized except possibly very close to $T_R$. The conclusion that experiments must be in the regime $\overline v_o \sim 1$ is in agreement with \cite{BGR}, where $\overline v_o$ is called $t_c$ (up to a numerical prefactor); $v_o/\gamma_o$ was estimated to be $\sim 0.05$. Since our {\sc rg} flow is different from the one obtained by Nozi\`eres and Gallet, the values of the physical parameters obtained by a fit of our theory to the experiments will actually differ. \begin{figure} \centerline{\hbox{\epsfig{figure=Fig2.ps,width=8cm}}} \vskip 0.8cm \caption{$\beta_S/\protect\sqrt{\overline v_o}$ as a function of the rescaled temperature for three different values of the bare periodic potential $\overline v_o$, with $\gamma_o=1$.} \label{figure2} \end{figure} \section{Case of the contact line} In this section, we repeat the previous analysis for the case of a contact line on a periodic substrate \cite{DG,JDG}. The roughness of a contact line on a disordered substrate, at zero temperature, has been studied analytically and compared with the experimental predictions for the case of superfluid helium on a disordered cesium substrate, where the disorder arises from randomly distributed wettable heterogeneities which are oxydized areas of the substrate \cite{Rolley,HM}. A physical realization of the theoretical situation we consider here could be achieved by preparing a substrate with equally spaced oxydized lines which would act as periodic pinning grooves. In this case the critical dimension is $d=1$. We denote by $\Phi$ the position of the line with respect to a mean position. The energy of the system is the sum of an elastic term and a potential term given by: \begin{equation}\label{energie_ligne} \displaystyle H[\Phi]=\frac{\gamma}{2} \int \ \frac{dk}{2\pi} |k||\Phi(k)|^2+ \int_0^L dx \ V(\Phi(x)) \end{equation} \noindent where $L$ is the length of the substrate and $\gamma$ the stiffness. The renormalization procedure is carried as before except that the propagator is now given by $\displaystyle G(k)=\frac{1}{\beta \gamma |k|}$. Moreover the renormalization of the stiffness now only comes from the scale change leading to the much simpler flow equation for $\gamma$: \begin{equation}\label{ren_gamma} \displaystyle \frac{d\gamma}{d\ell}=\gamma \end{equation} \noindent Defining as before the rescaled parameters $\overline V$ and $\overline T$ with $\displaystyle \overline V=\frac{V}{\gamma \lambda^2 |\Lambda|}$ and $\displaystyle \overline T=\frac{2T}{\gamma \lambda^2}$, the flow equations for $\overline V$ and $\overline T$ read: \begin{equation}\label{ren_V} \displaystyle \frac{d\overline V}{d\ell}= \overline V -\frac{\overline V'^2}{1+\overline V''} +\frac{\overline T}{2}\log(1+\overline V'') \end{equation} \noindent and \begin{equation}\label{ren_T} \displaystyle \frac{d\overline T}{d\ell}=-\overline T \end{equation} \noindent During the flow, $\overline T$ flows to zero and the renormalized rescaled potential $\overline V$ develops shocks between which it has a parabolic shape. We characterize the singularities that develop around the maxima of $\overline V$ by the following scaling ansatz: \begin{equation}\label{scale_ligne1} \displaystyle 1+\overline V''(\varphi)= \displaystyle T^{\delta} \displaystyle e^{-\frac{A}{\overline T}} {\cal F}' \biggl(\frac{\varphi}{T^{\alpha} \displaystyle e^{-\frac{B}{\overline T}}}\biggr) \end{equation} \noindent where ${\cal F}'(0) > 0$ and ${\cal F}(0) = 0$. Putting $\displaystyle u=\frac{\varphi}{{\overline T}^{\alpha} \displaystyle e^{ -\frac{B}{\overline T}}}$, this implies that for $u \sim 1$, \begin{equation}\label{scale_ligne2} \displaystyle {\overline V}'(u)=\displaystyle -{\overline T}^{\alpha} \displaystyle e^{ -\frac{B}{\overline T}} u +{\overline T}^{\delta+\alpha} \displaystyle e^{ -\frac{A+B}{\overline T}} {\cal F}(u) \end{equation} \noindent and \begin{equation}\label{scale_ligne3} \displaystyle {\overline V}(u)=\displaystyle {\overline V}_M- {\overline T}^{2\alpha} \displaystyle e^{ -\frac{2B}{\overline T}} \frac{u^2}{2} + {\overline T}^{\delta+2\alpha} \displaystyle e^{ -\frac{A+2B}{\overline T}} {\cal G}(u) \end{equation} \noindent with ${\cal G}'= {\cal F}$. Plugging the previous expressions into the flow equation $\overline V'$: \begin{equation}\label{ligne_V'} \displaystyle \frac{d\overline V'}{d\ell}= -\overline V' + 2\overline V' \frac{\overline V''}{1+\overline V'} + {\overline V'}^2 \frac{\overline V'''}{(1+\overline V'')^2} +\frac{\overline T}{2} \frac{\overline V'''}{1+\overline V''} \end{equation} \noindent and supposing that $\displaystyle \frac{d\overline V'}{d\ell}=0$, to leading order as $\overline T$ goes to zero, we have to leading order in $\overline T$: \begin{equation}\label{eq_calFbis} -\displaystyle \frac{2u}{{\cal F}'(u)} \ {\overline T}^{\alpha-\delta} \displaystyle e^{\frac{A-B}{\overline T}} +\frac{u^2{\cal F}''(u)}{{\cal F}'^2(u)} \ {\overline T}^{\alpha-\delta} \displaystyle e^{\frac{A-B}{\overline T}} +\frac{{\cal F}''(u)}{2{\cal F}'(u)} \ {\overline T}^{1-\alpha} \displaystyle e^{\frac{B}{\overline T}} =0 \end{equation} \noindent Since ${\cal F}'(0)>0$, we obtain a non trivial solution only if \begin{equation}\label{ligne_rel_exp} \displaystyle 2\alpha-\delta=1 \quad {\rm and} \quad A-B=B \end{equation} \noindent and ${\cal F}'$ is again solution of equation (\ref{Eq_calF3}). We can obtain the values of the parameters $A$ and $B$ by considering separately the singular part and the regular part of the renormalized rescaled potential $\overline V$ close to the fixed point. We expand $\overline V$ around one of its minima $\varphi^*$ as: \begin{equation}\label{ligne_exp_min} \displaystyle \overline V(\varphi)=\overline V_m + \frac{\kappa}{2}(\varphi-\varphi^*)^2 \end{equation} \noindent and plug the resulting expression in the flow equation for $\overline V$. This yields: \begin{equation}\label{ligne_flow_min} \begin{array}{ll} \displaystyle \frac{dV_m}{d\ell} & = \displaystyle V_m+\frac{\overline T}{2}\log(1+\kappa) \\ \\ \displaystyle \frac{d\kappa}{d\ell} & = \displaystyle \kappa - \frac{2\kappa ^2}{1+\kappa} \end{array} \end{equation} \noindent We note that the fixed point value for $\kappa$ is now finite as $\overline T$ goes to zero and is given by $\kappa^*=1$. Similarly the flow equation for $\overline V(0)=\overline V_M$ is: \begin{equation}\label{ligne_flow_max} \displaystyle \frac{d\overline V_M}{d\ell}=\overline V_M+\frac{\overline T}{2} \log \biggl(T^{\delta}\ e^{-\frac{A}{\overline T}}{\cal F}'(0)\biggr) \end{equation} \noindent Combining equations (\ref{ligne_flow_min}) and (\ref{ligne_flow_max}), the flow equation for the amplitude of $\overline V$ is given to leading order as $\overline T \to 0$ by: \begin{equation}\label{ligne_amp} \displaystyle \frac{d}{d\ell}(\overline V_M- \overline V_m)=(\overline V_M- \overline V_m)-\frac{A}{2} \end{equation} \noindent Now, we expect that the singularity brings but a small correction to the parabolic part of the rescaled potential $\overline V$, so that at the fixed point the amplitude of $\overline V$ is given by $\displaystyle \frac{\kappa^*}{2}$, leading to $A=1$. The value of the exponents $\alpha$ and $\delta$ would require the analysis of subdominant terms. The conclusion of this section is that in the case of the contact line, the width of the singular region of the renormalized potential decreases exponentially with length scale: the potential quickly becomes a succession of matched parabolas. \section{Conclusion} In this paper, we studied the problem of the thermal roughening transition using a {\sc frg} formalism. We have shown that below the roughening temperature, the periodic potential on large length scales cannot be described by its lowest harmonic and that during the flow shocks are generated in the effective pinning potential. We expect that this result is more generally valid, and also holds in the case of a disordered pinning potential \cite{BF,BBM}. By performing a resummation of our perturbation expansion, our results are in principle valid in the strong coupling regime, where the coupling constant is proportional to $V/\gamma$ (rather than $V/T$). Correspondingly, we stop the renormalisation procedure not when $V(L) \sim T$ (as in {\sc ng}), but rather when $L$ reaches the size of the objects under investigation (for example the width of the steps). By comparing our numerical result with the experimental determination of the step energy of liquid Helium 4, we have concluded that the surface of Helium 4 crystals are such that the coupling to the lattice is of the same order of magnitude as the surface tension. This is in qualitative agreement with Balibar et al. \cite{BGR}, who estimate $v_o/\gamma_o \sim 0.05$. \section*{Acknowledgements} We wish to thank Sebastien Balibar, T. Emig and M. M\'ezard for very interesting discussions. \newpage
\section{Introduction} The search for planets around nearby stars is a major objective of modern astronomy. Recently, astronomers have discovered Jupiter-mass planets orbiting nearby stars (see e.g., Mayor \& Queloz~1995; Butler \& Marcy~1996; Cochran, \etal~1997; Marcy, \etal~1997; Noyes, \etal~1997; Delfosse, \etal~1998; Marcy, \etal~1998; Delfosse, \etal~1999). They observed the periodic variation in the central stars' velocities due to their motion around the center-of-mass of the star-planet systems. However, they were unable to image the planets directly---the planets are too close to the stars that they orbit, so diffraction in the telescope and atmospheric seeing cause the starlight to wash out the planets. This problem is generic to observations of compact, compound sources in which the range of brightnesses of the individual components is large. On another forefront, many astrophysical systems have structure on milli-arcsecond or sub-milli-arcsecond angular scales. Examples of this include image splitting of background stars in microlensing events by massive compact halo objects (MACHOs), binary stars, globular cluster cores, Galactic supernovae, the accretion disk around a massive black hole at the center of the Milky Way, cores of nearby galaxies, lensed images of galaxies, and distant galaxies and clusters all have structure on the milli-arcsecond scale. Apparent angular displacements due to parallax and proper motions are also on this angular scale for objects in the nearest few kpc of the Galaxy. The proper motion of a star at $10\kpc$, with a velocity relative to the Sun of $200\kmps$, is approximately $5\mas\yr^{-1}$. The separation of dim sources from nearby bright ones, and the resolution of images at milli-arcsecond scales in the UV, optical, and near IR are limited principally by diffraction, $\Delta\theta\geq\theta_{\rm diff}\simeq 1.2\lambda/D$. Here $\lambda$ is the wavelength of the observations and $D$ is the diameter of the telescope or the baseline of the interferometer. On the ground the effects of seeing on resolution can be reduced using adaptive optics (AO) to nearly the diffractive limit in the IR, not yet at shorter wavelengths. However, the diffractive limit at $1\unit{\mu m}$ on a 10-m telescope is still $25\mas$, too large to resolve structures like Galactic microlenses. Moreover, while AO is typically very good at squeezing the central core of the point spread function (PSF) to near the diffraction limit, it still leaves the halo of the PSF\@. Getting all the way to the brightness contrasts of interest for planet searches ($>10^8$) strictly on the basis of AO, is considered technologically challenging in the IR, and daunting at shorter wavelengths. Atmospheric seeing can be eliminated entirely by observing from space, though scattering inside the telescope and diffraction persist. These combined effects make it difficult to observe most expected planetary systems from the existing Hubble Space Telescope, and even for the Next Generation Space Telescope (NGST). Greater resolution can be obtained using interferometry, and plans exist to do both ground-based and space-based IR and optical interferometry. The Space Interferometry Mission, a $10\meter$ baseline interferometer expects to do astrometry at the microarcsecond level, and so directly detect the wobble in a star due to an orbiting Jovian planet, but will not separate the light of the planet from that of the star. A space-based interferometer the same size as BOSS, such as the proposed Terrestrial Planet Finder, would be superior to BOSS for high resolution imaging of complex sources, though for high-contrast observations or resolution of simple multi-component sources, it too might benefit if combined with a scaled up version of BOSS. Such very-large space-base interferometers will be more challenging technologically, more costly, and require more time to build than BOSS. Lunar occultations have long been used to resolve small-scale angular structure (see e.g., Han, Narayanan, \& Gould~1996; Simon \etal~1996; Mason~1996; Richichi \etal~1996; Cernicharo \etal~1994; Adams \etal~1988). As the Moon orbits the Earth, it sweeps eastward across the night sky. If it occults another source, by monitoring the light-curve from that source, one can deduce the integrals of the source surface-brightness perpendicular to the apparent lunar velocity as a function of angular position parallel to the lunar velocity. For a binary source this means that one can measure the separation of the source components projected along the moon's velocity vector, and the relative intensity of the two sources. There are however several problems with using the Moon this way: \noindent (1) The lunar orbit is fixed; sources cannot be scheduled for occultation. {}From the ground, sources at ecliptic latitudes greater than about $5^\circ$ are never occulted; from space, the fraction of the sky which is occulted is similarly small. \noindent (2) The apparent angular velocity of the Moon with respect to the background stars ($0.55\as\second^{-1}$) is large and fixed by the orbital speed of the Moon. \noindent (3) The Moon is bright despite its low albedo. Even the dark side is visible to the naked eye. To avoid these difficulties, we are investigating the scientific merits of launching a large occulting satellite. The Big Occulting Steerable Satellite (BOSS) will consist of a large occulting mask. Patch geometry and transmission function can be optimized for either high-resolution image reconstruction or separation of bright from dim sources. In this paper we will concentrate on the latter. The configuration that we will consider is a square structure with a circularly symmetric transmission function that rises smoothly from zero (opaque) at the center to maximally transmissive where the inscribed circle osculates the edge of the square. The size of the inscribed circle, as will be made clear below, needs to be large, approximately $35\meter$ in radius. Similar ideas have been proposed independently in the past (see eg.\ Schneider~1995). However here we more thoroughly discuss the optical and mechanical properties as well as the capabilities and challenges of such a satellite. Furthermore we have developed a configuration for BOSS that would clearly contribute to several outstanding astronomical challenges, and that is viable for launch in the next 5 years. Given BOSS's size, one needs to do whatever possible to minimize its mass. The most likely approach is to fashion BOSS from a thin film supported by a framework of inflatable or deployable struts. Appropriate films with surface densities of $\sigma \simeq 10-20\gram\meter^{-2}$ are available, and the technology for deploying large structures composed of them reliably is under active development. In the near term there are two possible configurations for the satellite. The first is to deploy BOSS at L2, the second Lagrangian point of the Earth-Sun system $0.01\AU$ past the Earth along the Sun-Earth line, in conjunction with a large space telescope such as NGST (8-m class telescope). The telescope and BOSS would orbit the L2 point, with BOSS being steered to cause the desired occultations. We term this the space-space configuration. The alternative is to place BOSS in a high apogee eccentric Earth orbit, with observations done using ground-based telescopes. We term this the space-ground configuration. The space-space configuration has certain definite advantages. Problems of reflected sunlight entering the telescope would be considerably reduced because of greater flexibility in the telescope-satellite-source-sun-earth geometry; orbital dynamics would likely be simpler, more stable and more predictable. By using a combination of ion engines and sailing in the solar radiation pressure, one should also be able to obtain very long integration times. A main drawback is that one would incur the the added expense of a space telescope, however current plans are for NGST to be located in a halo orbit around L2. The space-ground has advantages in cost and risk since no astronomical instruments are deployed in space. Moreover, the satellite could be used from any telescope, including those of amateur astronomers, by making satellite coordinates publicly available over the Internet. Furthermore, since one is using ground-based telescopes, they can be quite large, and moreover any improvements in telescope or detector technology, such as advancements in adaptive optics systems, can be immediately used in conjunction with the satellite. Aside from observing issues such as atmospheric seeing and terrestrial thermal backgrounds, the major complications in the space-ground configuration are likely to be preventing excessive sunlight from being reflected into the telescope, and ensuring adequate stability and predictability of the orbital dynamics given solar radiation pressure, magnetospheric and atmospheric drag, and gravity gradients. Neither of these issues has yet been fully addressed, though a brief discussion of the reflection issue is included below. In this work we will focus on the space-space configuration. Although solar sailing will likely be important in either configuration, the satellite will undoubtedly not be able to rely entirely on solar sailing for maneuverability and ion engines will be required. As will be discussed below, the requirements on the propulsion systems are not unreasonable, and can probably be met with currently available systems. \section{Location and Size of BOSS} Before directly addressing BOSS's scientific capabilities we must address some important logistical questions: Where should the observer be located? Where should BOSS be located? How big does BOSS have to be? These are, of course, related questions. In this paper, we will address them in the context of the space-space configuration; similar questions will arise in the space-ground mode. \subsection{Locating BOSS} The criteria determining where BOSS should be located are: (1) longest possible duration occultations, (2) easiest possible multiple target acquisition, (3) highest possible duty cycle, and (4) lowest possible near-IR backgrounds. The first objective---long duration occultations---is both the most important and the most difficult to accommodate. In the occultation of a bright star to allow the observation of an associated planet, the brightness contrast between the sources is usually very large, typically $> 10^8$. In the geometric optic limit, the satellite completely blocks the bright source during occultations. However, because of diffraction by the satellite, some of the light from the bright source reaches the telescope (see the discussion in section \ref{sect:diffract}). Of course, the larger the diameter of the occulter, the greater the fraction of light which is blocked; however, the further away the planet must be from the occulted star not to itself be occulted. Thus there is a competition between the need to block as much stellar light as possible and the desire to detect planets as near as possible to the occulted star. The matter is complicated by the fact that for a fixed solid angle subtended by the occulter, the fraction of light which arrives at the telescope is a decreasing function of the telescope-occulter separation. Consequently, one would like to place the occulter as far away from the telescope as possible, while making the occulter as large as possible. Current technology for constructing and deploying large, light structures in space suggests that an occulter size of approximately $70\meter$ is achievable. The relevant angular separation for planets around nearby stars is approximately $1 \AU/10\parsec = 0.1\as$. Thus ideally one would like a $70\meter$ diameter BOSS located at a distance of order $10^4\hbox{--}10^5\km$. Since the angular size of the satellite is chosen to be comparable to the angular separation of the bright and dim sources, the time over which one can integrate on the dim source is approximately the time during which the satellite occults the bright source---essentially the time one can keep the satellite from crossing its own length. This must be long enough to allow the dim source to be detected above both the associated bright source and any background. We have identified two principle options for making long duration occultations, one for a satellite in orbit around the Earth occulting ground based telescopes (the space-ground configuration), one for a space based satellite occulting a space based telescope (the space-space configuration). \subsection{Space-Ground Configuration} Although we will not focus on the space-ground configuration we note that it is possible to construct orbits with long integration times over most of the sky for most telescopes at tropical sites (such as Keck). This is done by placing BOSS in a high apogee orbit where the instantaneous tangential velocity at apogee is equal to the rotational speed of the Earth (and hence the telescope). Thus at apogee BOSS will appear to hang, or even undergo retrograde motion, over the star. For sources over 95\% of the sky orbits with 4 or 5 day periods can be constructed that lead to $100$--$2300$ second occultations. \subsection{Space-Space Configuration} A number of factors persuade us to consider more carefully the possibility of a space-space configuration for the BOSS. First, although long occultations are possible in a ground-space configuration, very careful orbit selection is necessary---one must insert the satellite at precisely the right place, with precisely the correct velocity at precisely the right time in order to have the velocity of BOSS cancel the velocity of the telescope. Second, the satellite period is of order days and only one long occultation occurs per period, so the duty cycle is small. Third, changing from one target star to another is a complicated exercise in orbital dynamics, which can be costly in terms of both time and propellant. Fourth, space based telescopes are not subject to atmospheric seeing, nor to the same background levels as ground-based telescopes in the infrared wavelengths most interesting to planet discovery. Finally, space-based optical-IR telescopes to follow the Hubble Space Telescope, such as NGST, are being planned. The Lagrangian points of either the Earth orbit or any other planetary orbit hold particular attraction for the deployment of a BOSS in conjunction with a space telescope. This is because the magnitude of the local effective gravity is quite small, vanishing exactly at the Lagrange point (by definition). The unstable L2 point holds extra attraction because it is kept relatively clean of destructive debris. Indeed, NGST is likely to be deployed in a halo orbit around L2. This point is located along the line from the Sun to the Earth approximately $1.5\times10^6\km$ beyond the Earth. Deployment of BOSS in conjunction with a space telescope at or near L2 is thus quite attractive. At the L2 point the gravity due to the Earth and the Sun add up exactly to give an orbital period for a circular orbit of 1 terrestrial year. Around the L2 point, there are elliptical orbits in the ecliptic plane, and oscillatory vertical orbits perpendicular to the plane all of period approximately $1/2$ year. In addition, there are unstable modes in the ecliptic of similar time constant. These periods/time-constants are independent of distance from L2, for distances $\lsim 10^4\km$. At a distance $d$ from the L2 point the local gravity is therefore approximately $a \simeq 4\pi^2d/\tau^2$ where $\tau \simeq 150\hbox{--}200$ days. For $d$ in the range of $10^{(4\hbox{--}5)}$km, this gives $a\simeq 2\times 10^{-(5\hbox{--}6)}\meter\second^2$. This means that once BOSS is positioned to occult a star it takes at least ${\cal O}\left((1\hbox{--}3)\times10^3\right)$ seconds for BOSS to drift $10\meter$ off-center, with no corrective maneuvers. Typical orbital velocities around the L2 point are also relatively low: $v\simeq 2\pi d/\tau$, which is $(4\hbox{--}40)\meter\second^{-1}$ for $d=10^{(4\hbox{--}5)}\km$, making it relatively easy and inexpensive to reposition the BOSS and station-keep with respect to the telescope. For orbits out to approximately $10^5\km$, numerical simulations show similar stability and low orbital velocities. \subsection{Satellite Size} Suppose we can reliably position BOSS in space to a tolerance less than its linear dimensions. The satellite must then be large enough to guarantee that scheduled occultations occur despite the uncertainty, $\Delta\delta$, in the source location on the sky. One of the big uncertainties that would face the BOSS project is to what accuracy one can determine absolute star positions on the sky. Although Hipparcos determined the relative positions of stars in individual fields to about $1\mas$, the global fit to the positions of stars in the Hipparcos catalogue gives an accuracy of only about $\Delta\delta\ltsim0.1\as\simeq 5 \times10^{-7}\;$rad. Moreover, since the baseline for proper motions was only three years, as the epoch of the Hipparcos catalogue recedes into the past, the accuracy of the star positions decreases. This is an important effect for the stars which are the likely targets of searches for planets. However, the Hipparcos catalogue is likely to be supplemented by a SIM catalogue before the BOSS launch date. Since SIM can determine relative positions within a field to about $1\;\mu$as, we expect to have absolute angular positions for the target stars better than $\Delta\delta\ltsim0.01\as\simeq5\times10^{-8}\;$rad. We require that the BOSS be large enough to subtend a solid angle greater than the uncertainty in its desired position, $ r \Delta\delta$. If $r\leq 2\times 10^8\meter$ and $\Delta \delta = 0.01\as$, then we must require $x_\perp\geq 10\meter$. For a diameter of $70\meter$, BOSS at $1\times 10^8\meter$ distance would subtend about $0.14\as$. Thus if target stars have absolute positions determined to about $0.01\as$, then we can accurately aim to occult the star near the center of the satellite. Realtime examination of the diffraction image of the star would allow positioning well below $10\meter$. Even if we can know where the satellite {\it is\/} very accurately, can it be placed where it is needed when it is needed with the appropriate velocity? That question is difficult and is currently unanswered, but the answer is not obviously no. Relevant effects include solar radiation pressure, lunar gravity, eccentricity of the Earth's orbit, \etc., all of which need to be understood to address this question. The second important effect influencing the choice for BOSS's size is the satellite's nulling efficiency---the fraction of light from an occulted source that diffracts around the satellite decreases with the area of the satellite (see section \ref{sect:diffract}). The higher the contrasts between bright and dim sources that one wants to study, the larger the angular size that BOSS needs to be. On the other hand, the larger BOSS is, the further away on the sky the dim source must be from the bright source to avoid getting occulted itself. For high resolution imaging, there is less constraint since in the Fresnel limit the diffraction limit for the satellite is independent of the size of the satellite if $\lambda < x^2/r \simeq 500\unit{\mu m}$. We will use $\RBOSS=35\meter$ for BOSS in this paper, though further characterization of the performance BOSS as a function of size is essential and ongoing. The optimum answer will depend on the exact observing program and the nature of the objects to be occulted. \subsection{Satellite Mass} While there may be scientific pressures to make the occulter as large as possible, there are clear financial pressures to keep its mass as low as possible. In particular, the cost of launching a satellite increases dramatically with mass. Hence the maximum allowable satellite area $\pi \RBOSS^2$ depends on the minimum film thickness one can tolerate, the strength of any framework, etc\@. Since skin depths of good conductors in optical and IR bands are very much less than a micron, one can readily render the occulter as opaque as needed. The minimum film thickness is then determined by structural integrity not optical necessity. Appropriate films exist with surface density $\sigma\simeq10-20\gram\meter^{-2}$, so a single layer $70\meter$ diameter circle would have a mass of $40\hbox{--}80\kg$. The rigidity of the satellite structure could be achieved by supporting the film using inflatable, self-rigidizing struts. Because the struts rigidize after inflation, micrometeoroid hits would not affect the satellite structurally. Micrometeoroid impacts on the film might, however, affect the optical properties of BOSS\@. Small holes, with diameters less than the wavelength at which observations are planned ($\lambda\simeq 1\;\mu{m}$), would not severely impact the performance of the occulter until their areal density became quite large. This is because the photons encountering such holes would see them only as small changes in the effective transmission function of the satellite, and because diffraction would spread any passing light ray into a wide beam. Much larger holes would be more problematic, and could degrade performance significantly at even a very low level. Attitude and orbit corrections must be sufficiently gentle so as not to damage the framework or tear the film. Preliminary estimates show that struts and deployment system to support a $35\meter$ radius occulter able to withstand accelerations of 1-2 g's, would total another $100\hbox{--}200\kg$. Also, one would like the amplitude of oscillations of the satellite to be small so that sunlight is not reflected into the telescope. Detailed analysis of the modes of oscillation, damping times, maximum allowable amplitude of oscillation, as well as the permitted levels of film deterioration will also be required. \subsection{BOSS Geometry} \label{sect:geometry} Supporting a circular film is difficult. Putting a rigid, opaque ring around the circumference of BOSS destroys the nulling achieved by using a radially dependent transmission function unless the ring is extremely narrow. We instead suggest imprinting a circular transmission function on a transparent square film. The square will be supported by rigid, opaque struts along its diagonals. Our model configuration is thus made up of a $70\meter\times70\meter$ transparent square with a $35\meter$ radius, radially dependent, circular transmission function inscribed. It is supported by 20 inch ($0.5\meter$) diameter struts. The optical properties of this arrangement will be explored in section~\ref{sect:planet}. \section{Steerability} To change the satellite orbit to occult a particular object requires imparting an impulse: $\Delta {\vec p} = m\Delta {\vec v},$ resulting in a change of velocity \be \frac{\Delta\vec v_{\rm sat}}{v_{\rm sat}} = \frac{\Delta m_{\rm propellant}{\vec v}_{\rm ejection}}{m_{\rm sat}v_{\rm sat}} . \ee The solar radiation pressure will also play a significant role in determining the satellite's motion. This can either be viewed as a problem or an opportunity. The solar radiation pressure in the vicinity of the Earth (and L2) is approximately $P = 4\times10^{-6} \unit{N}\meter^{-2}$. For a satellite of effective surface density $0.05\kg\meter^{-2}$, this translates into an acceleration of $8\times10^{-5}\meter \second^{-2}$. This means that the satellite could cross a $30,000\km$ radius orbit in 1--2 days, or a $100,000\km$ radius orbit in 2--4 days using solar radiation pressure. Of course, the solar radiation pressure is pushing away from the sun, and L2 is unstable, so the use of solar sailing is somewhat limited. However, understanding solar radiation pressure {\it will\/} be essential, and utilizing it would be desirable. Whatever use BOSS makes of solar sailing, it will clearly be necessary to do some orbital adjustment using rockets. The number and size of such adjustments will probably be the limiting factor on the useful lifetime of the satellite. If $N$ is the number of desired major rocket-driven orbital corrections, then we must keep $(\Delta m_{\rm propellant}/m_{\rm sat})\leq N^{-1}$, where $\Delta m_{\rm propellant}$ is the typical mass of propellant expended per orbit reconfiguration. We therefore need \begin{equation} v_{\rm ejection} \geq N \Delta{\vec v_{\rm sat}} . \end{equation} As discussed above, satellite orbital velocities are in the range $4\hbox{--}40\meter\second^{-1}$ for the orbital radii of interest. If $\Delta{\vec v_{\rm sat}}\simeq v_{\rm sat}$, and $N \simeq 3\times10^3$ this implies $v_{\rm ejection}\simeq 12\hbox{--}120 \kmps$. $N={\cal O}(3\times10^3)$ is a reasonable number of corrections when it takes one or two days to reposition BOSS for a new occultation, since it implies a satellite lifetime of about $5$ years. With careful scheduling, and clever orbital dynamics, typical course corrections might not require $\Delta{\vec v_{\rm sat}}\simeq v_{\rm sat}$. Off-the shelf, low-cost ion engines are currently available with ejection velocities of $20 \kmps$, and more expensive systems with $30 \kmps$ performance have been developed. The issue of steerability therefore comes down to the sizes of the course corrections in which we are interested, the time scale over which we need to move the satellite to the new orbit, the effectiveness with which one can use the solar-sailing technique, and the state-of-the-art in high ejection velocity drives. The smaller the corrections and the longer we can wait, the less fuel we will need. Since planets are not transient sources, we can wait a relatively long time and use clever orbital dynamics to do much of the work for us. How many corrections we can make, therefore depends on exactly how we use the satellite. A reasonable program of observations seems possible, though careful scheduling of observations and clever design of orbital parameters will be essential. \section{Diffraction} \label{sect:diffract} Because we are interested in observing systems with very high contrast, we must include the effects of diffraction. The angular width of the satellite diffraction pattern in the region far from the satellite is \begin{equation} \label{diffraction} \Delta\theta_{\rm diff}\simeq \left\{ \begin{array}{ll} \frac\lambda{2R}, & \frac\lambda{2R} \ge \frac{2R}{\Dsat}\quad \hbox{(Fraunhoffer limit)} \\ \sqrt{\frac\lambda{\Dsat}}, & \frac\lambda{2R} \le \frac{2R}{\Dsat}\quad \hbox{(Fresnel limit)} \\ \end{array} \right. . \end{equation} Here $2R\simeq 70\meter$ is some typical size of the occulting patch. For $\lambda\simeq 1\unit{\mu m}$ and $\Dsat\simeq 1\times10^8\meter$, the distance to the satellite, the transition from Fresnel to Fraunhoffer behavior occurs for $R\simeq \sqrt{\lambda \Dsat}/2 \simeq 5\meter$. For $R=5\meter$ and $\lambda=1\unit{\mu m}$, $\theta_{\rm diff}\simeq 1\times 10^{-7} = 21\mas$. Inspection of (\ref{diffraction}), shows that once the size of the satellite has reached $R_{\rm crit}\simeq\sqrt{\lambda \Dsat}/2$ no further improvement in resolution can be obtained just by increasing the size. However, the fraction of the intensity of the occulted source which is not blocked decreases with the area of the occulter. For this reason, we choose $R\gg 5\meter$. The price is that we must use the Fresnel diffraction pattern of the sources as occulted by the satellite and as observed through the telescope. Here we briefly outline the necessary steps. The diffraction pattern for an arbitrary aperture ${\cal A}$ is given by \be U = - \frac{Bi}{\lambda r's'} \int\!\!\int\nolimits_{\cal A} \tau(S) e^{ik (r + s)}\,dS \ee where $B$ is a normalization constant related to the intensity of the source, $dS$ is a surface element in the aperture plane, $\tau(S)$ is the transmission function, $r$ ($s$) is the distance between the observation (source) point and a point in the aperture plane, and $r'$ ($s'$) is the distance between the observation (source) point and the origin of the aperture plane. Here we have assumed that the distances between the aperture plane and the source and observation planes are large and that all three planes are parallel. We may expand $r+s$ in the exponential in the limit of large separations. Note that we must keep terms of order $k\xi^2/2\RBOSS > 1$ where $\xi$ is a dimension in the aperture plane. The presence of this higher order term means that we are working in the Fresnel limit of diffraction. A direct application of Babinet's principle allows us to relate the diffraction pattern due to the occulting satellite $\Umask (\rho,\phi)$, to the complimentary problem of the diffraction pattern of a finite circular aperture. Thus we can write $\Umask$ as \be \Umask (\rho, \phi) = e^{-i\frac{k\tilde\rho^2}{2\Dsat}} + i\frac k{\Dsat} \int_0^{\RBOSS} dr\,r e^{i\frac{kr^2}{2\Dsat}} J_0 \left( \frac{k\tilde\rho r}{\Dsat} \right) \left[ 1 - \tau (r/\RBOSS) \right], \ee where $\tilde\rho^2 = \rho^2 + \rhosource^2 + 2 \rho\rhosource \cos (\phi-\phisource)$, $(\rhosource,\phisource)$ is the position of the source and $\tau$ is the transmission function of the satellite. Throughout this discussion have assumed a circular occulter and an occulted source on-axis. The above analysis has been performed for a square occulter and leads to comparable results. We want to look at this pattern with a telescope which also causes diffraction. The final pattern is given by an integral over the telescope \be U (w, \phi_0) = \frac{\sqrt\pi}{\lambda \Rtel } \int_0^{\Rtel} d\rho\,\rho \int_0^{2\pi}d\phi\, e^{-ik\Rtel w \cos (\phi-\phi_0)} \Umask (\rho, \phi). \label{Uimag} \ee Here $\Rtel$ is the radius of the telescope and $(w, \phi_0)$ is the angular position (in radians) of the observation point $(x_0, y_0)$ with respect to the center of the telescope. The prefactor normalizes the intensity in the observation such that \be \int_{0}^\infty dw\, w \int_0^{2\pi} d\phi_0\, \left| U (w, \phi_0) \right|^2 = 1. \ee We have treated the telescope as a circular aperture with a CCD at infinity which is used to image the diffraction pattern. The evaluation of $U$ is challenging. Straightforward numerical integration of the highly oscillatory integrand requires many hours to produce a single image. Using a scheme in which $\Umask$ is expanded in Tchebyshev polynomials we are able to evaluate all the integrals analytically and express (\ref{Uimag}) as a sum over Bessel functions which can be evaluated numerically. A complete image of an arbitrarily located point source can thus be obtained in only a few minutes. The transmission function allows us to apodise the diffraction pattern caused by the satellite thus making the planetary region darker. For the transmission function we follow the suggestion of Hyde~(1998). We write the transmission function as \be \tau_N (y) = \sum_{n=0}^N c_n y^n, \ee where $N$ is the order of the occulter and we define $\tau_0=0$. Here \be y = \frac{\left (\frac r{\RBOSS}\right)^2 - \epsilon}{1-\epsilon} \ee and $\epsilon$ is the fractional radius of the center of the occulter that is opaque ($\tau=0$). (All of the functions are made to satisfy $0\le \tau_N (y)\le1$ for $0\le y\le1$.) Throughout this work we will focus on the fourth order occulter \be \tau_4 (y) = 35y^4 - 84y^3 + 70y^6 - 20y^7, \ee and consider $\epsilon=0.15$. Note that this transmission function is obtained by optimizing the on-axis nulling of the occulter. Though this criterion is not optimal for planet detection, it provides a simple function that improves performance and can be used as a starting point for future investigations. We should also note that real materials do not have perfect transmission. In the near IR it is possible to get 97\% transmission in a film with an anti-reflective coating. Thus we will replace $\tau$ with $(1-\delta)\tau$ and consider $\delta = 0.03$. We will also use a constant transmission function of $1-\delta$ for the square in which our circular pattern is imprinted. Smaller values of $\delta$, if achieved, could lead to improvement in the performance of the BOSS. \section{The Satellite as a Source} Just like a planet, BOSS would shine by reflection, both coherent and diffuse, and by thermal emission. A serious concern is that BOSS might appear brighter than the sources it is occulting and thus be of little value. The concerns are less pronounced in the space-space configuration than they are in the space-ground configuration since by correctly orienting the satellite one can keep the sunlit side of the occulter away from the telescope. However the situation is complicated by using a radially dependent transmission function for the occulter since some sunlight is scattered in and through the portions of the occulter which are not completely opaque. Here we will discuss the constraints arising from all three sources: reflection, scattering, and thermal emission. The flux of photons arriving from a star of bolometric magnitude $m_b$ is approximately \be \label{eqn:phistar} \Phi_{\star} = 1.2 \times 10^{11} \left({T_\odot\over T_\star}\right) 10^{-0.4m_b} \pmsqs . \ee Since $m_b=-26.85$ for the Sun \be \Phi_{\odot} = 6.4\times 10^{21}\pmsqs . \ee We first consider the problem of reflection of sunlight off the satellite into the telescope. Consider a reflective patch on the satellite of area $A_{\rm patch}$ deliberately angled so as to reflect sunlight into the telescope. This patch will reflect photons at a rate \be \left({d n_\gamma \over dt}\right)_{\rm reflect} = 6.4 \times 10^{21} \frac{A_{\rm patch}}{\rm m^2}\unit{s^{-1}} . \ee {}From the position of the satellite however, the Sun subtends a much larger solid angle of the sky than the telescope, so only a fraction \be f_\Omega = {(\Rtel/\DBOSS)^2 \over (R_\odot/1.01\;{\rm AU})^2} = 7.5\times 10^{-11} \left({\Rtel\over 4\meter}\right)^2 \left({\DBOSS\over 1\times10^8\meter}\right)^{-2} \ee of what is reflected will enter the telescope. Here $\DBOSS$ is the distance from BOSS to the telescope and $\Rtel$ is the radius of the telescope. Thus the rate of photons collected by the telescope from the reflecting patch would be \be \Gamma_{\rm reflected} = 4.8\times 10^{11}\left ( \frac{A_{\rm patch}}{\rm m^2} \right) \left({\Rtel\over 4\meter}\right)^2 \left({\DBOSS\over 1\times10^8\meter}\right)^{-2}\unit{s^{-1}} . \ee Compare this to the rate of photons received in the telescope from a star, \be \Gamma_{\star} = 6 \times 10^{12-0.4m_b^\star} \left({T_\odot\over T_\star}\right) \left({\Rtel\over 4\meter}\right)^2 \unit{s^{-1}} . \ee We would like to keep the reflected light from being brighter than a 16 magnitude star, as this is the level at which it will begin to interfere with the performance of the occulter. We see that this requires \be A_{\rm patch}\leq 5 \unit{mm}^2 \left({T_\odot\over T_\star}\right) \left({\DBOSS\over 1\times10^8\meter}\right)^2 . \ee This seems a strenuous requirement and implies that careful attention must be given to the shape of the exposed surfaces of the occulter, especially the edges. However, it is reassuring that the Sun subtends only $6.8\times10^{-5}$ steradians, or $5.4\times10^{-6}$ of the sky, and the telescope mirror subtends only $2.2\times10^{-13}$ steradians. Random fluctuations of the occulter surface are therefore unlikely to reflect light into the telescope. So long as we are careful to appropriately orient the satellite and do not allow overly large fluctuations of the underlying surface we should be able to meet this specification. Not all of the sunlight incident on the satellite will be coherently reflected. Some will be absorbed and some will pass through the transmissive portions of the occulting body and be scattered diffusely. Let $\cal S$ be the fraction of photons that are scattered diffusely. Then, assuming isotropic scattering, the flux of photons at the telescope resulting from diffuse reflection from the satellite will be \be \Phi_{\rm diffuse} = \Phi_\odot {\cal S} \left({A_{\rm BOSS} \over 4\pi \DBOSS^2}\right) . \ee For this to be less than the flux of a star of bolometric magnitude $m_b$ (cf.\ equation \ref{eqn:phistar}) \be {\cal S} < 1.2\sci{-4} \left({\DBOSS \over 1\times 10^8\meter}\right)^2 \left({A_{\rm BOSS} \over (70\meter)^2}\right)^{-1} \left({T_\odot\over T_\star}\right) 10^{-0.4(m_b^\star - 16)}. \ee The diffuse scattering of light in transparent films, such as polyimides, which might be used for BOSS has been studied. For a $10\unit{\mu m}$ film ${\cal S} < 10^{-4}$ at $1\unit{\mu m}$ is not unusual (e.g.~Kowalczyk, \etal~1994). Intrinsic emission arises because BOSS is warmed by the sun. The integrated solar flux density at L2 is approximately \be f_\odot \simeq 1.38\times 10^6 \unit{erg}\unit{cm^{-2}}\unit{s^{-1}} . \ee The power absorbed by the occulter is \be P_{\rm abs} = \alpha f_\odot A_{\rm BOSS}\cos\theta, \ee where $\alpha$ is the fraction of incident solar radiation energy absorbed, and $A_{\rm BOSS}\cos\theta$ is the projected area presented to the sun. The power reemitted by BOSS is \be P_{\rm reemit} = e A'_{\rm BOSS} \sigma T_{\rm BOSS}^4. \ee Here $e$ is the emissivity and $\sigma$ is the Stefan-Boltzmann constant. Notice that $A'_{BOSS} \geq A_{BOSS}$, since the satellite can radiate from both sides, but absorbs sunlight from only one. If the satellite comes into thermal equilibrium with the solar radiation then \begin{eqnarray} T_{\rm BOSS} &=& \left( {\alpha f_\odot \over e \sigma} { A_{\rm BOSS}\over A'_{\rm BOSS}} \cos\theta\right)^{1/4}, \nonumber\\ &=& 394\unit{K} \left(\frac\alpha{e}{ A_{\rm BOSS}\over A'_{\rm BOSS}} \cos\theta\right)^{1/4}. \end{eqnarray} For emission from one side ($A'=A$) this gives $T_{\rm BOSS}=394\unit{K} \left({\alpha\over e} \cos\theta\right)^{1/4}$, while for equal two sided emission ($A'=2A$) it gives $T_{\rm BOSS}=332\unit{K} \left({\alpha\over e}\cos\theta\right)^{1/4}$. For many materials $\alpha/e\sim2$, but typically during observations $\theta$ will be greater than $60^o$, so $\alpha \cos\theta/e$ should not be large. The total photon flux at the telescope coming from the radiating satellite in thermal equilibrium in a waveband from $\lambda_1$ to $\lambda_2$ is \be \Phi_{\rm thermal} = \alpha\cos\theta {A_{\rm BOSS}\over 2\pi \DBOSS^2} {f_\odot\over 2.7 k T_{\rm BOSS}} g(\lambda_1,\lambda_2,T_{\rm BOSS}) , \label{bossflux} \ee where \be g(\lambda_1,\lambda_2,T) = {1\over 2 \zeta(3)}\ \int_{hc/\lambda_1 k T}^{hc/\lambda_2 k T} {x^2 dx \over e^x-1} \ee is the fraction of the total photons emitted in the relevant waveband. Comparing this to the flux from an occulted star in the same waveband we find \begin{eqnarray} {\Phi_{\rm thermal}(\lambda_1,\lambda_2,T_{\rm BOSS}) \over \Phi_{\rm occulted}(\lambda_1,\lambda_2,T_\star) } & = & 3\times 10^{5+0.4(m_b^\star-16)}\alpha\cos\theta \left({T_\star\over T_\odot}\right) \left({332\unit{K}\over T_{\rm BOSS}}\right) \left({A_{\rm BOSS}\over (70\meter)^2}\right) \nonumber \\ && \times \left({1\times10^8\meter\over \DBOSS}\right)^2 {g_{\rm thermal}(\lambda_1,\lambda_2,T_{\rm BOSS}) \over g_{\rm thermal}(\lambda_1,\lambda_2,T_\star) } . \end{eqnarray} For $T=332\unit{K}$, and $T_\star=T_\odot=5778\unit{K}$, ${\Phi_{\rm thermal}/\Phi_{\rm occulted}}$ varies from $1.4\times 10^{-8}$ for a $0.2\unit{\mu m}$ waveband centered at $1\unit{\mu m}$ to $2.8\times 10^{-3}$ at $1.5\unit{\mu m}$ to $1.7$ at $2.0\unit{\mu m}$. At $T=394\unit{K}$, these rise to $4.0\times 10^{-6}$ at $1\unit{\mu m}$, $0.1$ at $1.5\unit{\mu m}$ and $28$ at $2.0\unit{\mu m}$. Thus observations for wavelengths shortward of about $2.0\unit{\mu m}$ are unlikely to be seriously affected by the thermal emissions from BOSS. Allowing the satellite to be as bright as a magnitude 16 solar type star does not limit us to resolving pairs of magnitude 16 objects since only the shot noise in the flux from the satellite is relevant. The ratio of the square root of detected satellite photons to the number of photons from a star of bolometric magnitude $m_b$ is \be {{\sqrt {N_{\rm BOSS}}} \over N_\star} = 3.2\times 10^{-10} \left[ {50 \meter^2\over A_{\rm telescope}} {\second\over \Delta t}\right]^{1/2} {T_\star \over T_\odot} 10^{0.4m_b} . \ee For $T_\star=T_\odot$ and a $1\second$ exposure (the typical time it takes a satellite orbiting L2 to cross $0.1\mas$) a 23.7 bolometric magnitude target star would give $ {{\sqrt {N_{\rm BOSS}}} \over N_\star} \simeq 1$. It thus should be relatively easy to differentiate between stars of magnitude 20 or less at separations of better than $0.1\mas$. \section{Planet Searches} \label{sect:planet} Planets shine in two ways, in reflected light and in emitted light. In reflected light, the brightness of a Jupiter-like planet orbiting $1\AU$ from a star is approximately $10^{-7}$ times that of the star; falling off as $1/r^2$ as the orbital radius increases, but growing as the radius of planet squared (see figure~\ref{fig:planet-reflect}). In emitted light, a planet glows as approximately a black body characterized by its temperature (see figure~\ref{fig:planet-emit}), though molecular absorption can alter that dramatically in certain wavebands. A planet's temperature is a strong function of its age, the central star's type and proximity, atmospheric composition, and internal heat sources; however, typical brightness ratios are still $\ltsim 10^{-8}$ except for very young gas giants or planets very close to the central star (see e.g., Burrows \etal~1997). We will focus on a 1 micron wave band (0.9--1.1\unit{\mu m}) for our studies since the diffraction from an 8-m telescope is not too severe in this waveband. To study the benefits of employing BOSS in conjunction with a space telescope we have simulated the images produced by our solar system at $3\parsec$ and $10\parsec$ occulted by BOSS and viewed by a 8-m telescope. For our solar system we have included the Sun occulted on-axis and the four brightest planets; Venus, Earth, Jupiter, and Saturn. We have used the BOSS geometry described in section~\ref{sect:geometry} and have assumed that BOSS is located $1\times 10^8\meter$ from an 8-m space telescope at L2. All images have been calculated for the 1\unit{\mu m}\ waveband. Images of the occulted star and unocculted planets are calculated for 3 different wavelengths in this waveband as described in section~\ref{sect:diffract}. These images are then smoothed with the point spread function~(PSF) of the telescope. The PSF we employ has a diffraction limited core and a Gaussian halo with a $1\as$ FWHM that contains $10^{-3}$ of the light. The smoothed images provide the templates of the occulted star and unocculted planets. From these templates we can construct an image of a solar system. To construct a solar system image we first scale the planets' intensities by a fraction of the unocculted star's intensity. We then include a Gaussian random component equal to 5\% the intensity of each pixel to account for uncertainties in the PSF\@. After adding the planets to the occulted star we Poisson sample the resulting image to produce an image of the solar system. To identify planets, the template of the occulted star is subtracted off leaving behind the planets and the noise. A planet is detectable if it can be identified above this noise. \placetable{tab:solarsystem} Figures~\ref{fig:solar3} and~\ref{fig:solar10} show our solar system $3\parsec$ and $10\parsec$ away from us, respectively. The parameters describing out solar system are given in table~\ref{tab:solarsystem}. The bolometric magnitude of the Sun at $3\parsec$ is $m^\odot_b = 2.1$ and at $10\parsec$ is $m^\odot_b = 4.72$. The images are $2\as\times2\as$ CCD images with $0.01\as\times0.01\as$ pixels. The central $0.2\as$ square (roughly the size of BOSS) has been cut out. At $3\parsec$ Venus and the Earth are visible by-eye (see figure~\ref{fig:solar3}). Both Jupiter and Saturn would be easily observable but are outside the CCD shown here. At $10\parsec$ Venus and the Earth are occulted by BOSS but both Jupiter and Saturn are easily identified (see figure~\ref{fig:solar10}). A wider variety of images in color can be found at http://erebus.phys.cwru.edu/$\sim$boss/. In figures~\ref{fig:solar3} and~\ref{fig:solar10} our solar system is shown in reflected light. The relative intensity for Jupiter was calculated using its Bond albedo and assuming uniform scattering into $2\pi$ steradians. The intensity for Jupiter quoted in table~\ref{tab:solarsystem} is based on this. The intensity for the rest of the planets quoted in table~\ref{tab:solarsystem} are scaled from Jupiter using their size, distance, and albedo relative to Jupiter. We should also note that many improvements can be made in the analysis described above. A statistical image reconstruction technique could more reliably detect planets and may allow for identification of planets that the ``by-eye'' technique misses. Also, the reflected light from a planet would be polarized whereas starlight is not. Thus even further suppression of the starlight should be possible. Finally we comment that the $1\unit{\mu m}$ waveband is a fairly good choice for gas giant planets. Although $\rm CH_4$ (which is prevalent in gas giant atmospheres) has an absorption feature at $1\unit{\mu m}$, it is fairly weak and sufficient flux comes from the two edges of the waveband to compensate for this. Above about $1.1\unit{\mu m}$ strong $\rm CH_4$ absorption dominates the spectrum and drastically reduces the intensity of reflected light. Blackbody emission at $1\unit{\mu m}$ from planets similar to those in our solar is not important since the planets are too cool. A Jupiter-like planet would only be detectable at $1\unit{\mu m}$ if it were extremely hot ($T_{\rm planet} > 800\unit{K}$, see figure~\ref{fig:planet-emit}). Notice the sharp increase in intensity shown in figure~\ref{fig:planet-emit} at higher wavelengths. In particular at $2\unit{\mu m}$ planets down to a temperature of $T_{\rm planet} \approx 400\unit{K}$ could be observed in emission if a similar relative intensity limit can be achieved. However Jupiter has an effective blackbody temperature of $T_{\rm Jupiter} = 125\unit{K}$ and would still not be observed in emission at $2\unit{\mu m}$. In fact, about $5.5\unit{\mu m}$ is where Jupiter starts radiating brightly. Thus emission will not likely play an important role at $1\unit{\mu m}$ for this study. Larger wavelengths require larger telescopes to cut down on the size of the diffracted images. Even at $1.5\unit{\mu m}$, the ``by-eye'' identification technique employed here is not sufficient for finding faint planets since the light of both the planet and occulted star are spread out over larger regions. \section{Conclusion} Occultations provide a powerful technique for resolving both small angular scales and large separations in intensity. BOSS in a space-space configuration will offer distinct improvements over a lone space telescope allowing it to not only detect planets but also directly image them. BOSS blocks all but $4\sci{-5}$ of the light from a star in the $1\unit{\mu m}$ waveband. Furthermore, by doing the occultation well outside the telescope, degradation due to scattering inside the optics is minimized. The low effective gravity at L2 offers the possibility of obtaining very long exposures. In this work we have shown that a $70\meter\times 70\meter$ BOSS in conjunction with an 8-m space telescope at L2 would allow for the direct imaging of Venus, Earth, Jupiter, and Saturn at $3\parsec$. Earth and Venus would remain visible out to at least 5 pc. Both Jupiter and Saturn would be visible at $10\parsec$ and remain so out to about $20\parsec$. Solar systems with Jupiter-like planets closer to their star would be easily observable. Furthermore, photometric and spectrographic observations could be considered for these objects. Many improvements can be made over the techniques employed here to aid in finding dim planets. A sophisticated reconstruction technique would greatly improve the ``by-eye'' technique presented here. Also using polarized light from the planet versus the unpolarized light from the star would significantly increase the separation of planets from the star. In spite of the simple techniques used here BOSS has been shown to be a powerful tool for discovering and imaging planets. We have discussed only briefly high resolution observations using BOSS, though our early estimates lead us to anticipate a capability to resolve binary sources a the well sub-milliarcsecond level. Further discussions will appear in upcoming papers, but BOSS's potential for high resolution imaging, looks to be as exciting as its promise for planet discovery. \acknowledgements The authors would like particularly to thank C. Beichman for very extensive conversations on AO, planet finding, and IR astronomy, for ongoing encouragement and for arranging financial support from JPL; Art Chmielewski for financial and other support and, with Chuck Garner and Rick Helms, for educating them about inflatables; L. Close for telling them about state-of-the-art in AO; M. Dragovan for many useful comments and suggestions; F. Kustas for mission and systems analysis; L. Lichodziejewski for extensive design efforts; H. Morrison and P. Harding for educating them about PSF's, filters, and various nitty gritty astronomical details; and V. Slabinski for lengthy discussions on orbital dynamics and satellite control capabilities. They would also like to thank B. Madore for his interest and encouragement; P. McGregor for information on IR backgrounds; P. Taylor for assistance with questions on optics; W. Tobocman for special functions; and A. de Laix, L. Krauss, H. Mathur, M. Trodden and T. Vachaspati for ongoing useful comments. GDS would like to thank C. Alcock for very helpful discussions and encouragement during the initial phase of this research, and G. Marcy, B. Matisack, S. Drell and S. Tremaine for useful input. This work was supported by a CAREER grant to GDS from the National Science Foundation, a DOE grant to the theoretical particle and astrophysics group at CWRU, by a grant from NASA's Jet Propulsion Laboratory, and by funds from CWRU. Allan Fetters did some of the early detailed analysis of integration times for the space-ground configuration and prepared the BOSS home page. David Rear contributed heavily to early programming in the analysis of imaging capabilities.
\section{Introduction} High spatial and spectral resolution observations of the neutral hydrogen gas in nearby dwarf and Magellanic irregular galaxies have revealed remarkably intricate and complex structures (\cite{puche92}, hereafter P92; \cite{puche94}; \cite{west94}; \cite{staveley97}, \cite{kim98}). Numerous holes are visible in the \ion{H}{1}\ distributions of these galaxies, surrounded by shells of higher density. In Holmberg$\thinspace$II (Ho$\thinspace$II, DDO$\thinspace$50) for example, P92 identify 51 holes, and show that in many cases expansion of the gas surrounding the holes is directly detectable. Typical expansion velocities of 4 -- 10 km\thinspace s$^{-1}$\ have been measured, and holes as large as 1600 pc across are present. This hole--shell morphology for the \ion{H}{1}\ gas appears to be quite common, being present in essentially all of the nearby dwarfs studied to date with adequate spatial and velocity resolution. Previous studies of the \ion{H}{1}\ distributions in nearby spiral galaxies such as M31 and M33 (\cite{brinks86}; \cite{deul90}) also reveal the presence of \ion{H}{1}\ holes. The origin of these features has generally been attributed to stellar winds and supernova explosions (SNe) from young stellar associations and clusters. This previous work naturally suggested that the holes in Ho$\thinspace$II and other dwarfs were caused by a similar process. P92 put forth the hypothesis that the holes in Ho$\thinspace$II were due to the combined action of hot stellar winds from O and B stars, plus the SNe shocks from the same massive stars after they exploded at the end of their lifetimes. Based on their measured expansion rates and hole sizes, they determined that as much as $\sim$2 $\times$ 10$^{53}$ ergs of kinetic energy, the equivalent of $\approx$200 SNe, was required to produce the largest holes. The more typical holes require a few to several dozen SNe. Although the SNe hypothesis for explaining the \ion{H}{1}\ holes in Ho$\thinspace$II is an appealing one, it is not without problems. Several of the \ion{H}{1}\ holes are located at large distances from the center of the galaxy, in regions of both low optical surface brightness and low \ion{H}{1}\ column density. At least 18 holes are located beyond the Holmberg radius of the galaxy (the radius where the $B$-band surface brightness drops below 26.6 magnitudes/square arcsec), where it is unlikely that large numbers of massive stars could have formed. The \ion{H}{1}\ column densities at these radii are well below the canonical threshold density of 10$^{21}$ cm$^{-2}$ (\cite{kennicutt89}), a further indication that star formation is unlikely to occur at these locations in the galaxy. Therefore, it seemed appropriate to investigate further the SNe hypothesis of P92, to determine whether it was possible to corroborate the general picture or, alternatively, to rule it out. The supernova scenario described in P92 does in fact provide us with a direct observational test. Those authors found that many of the expanding holes in Ho$\thinspace$II require the kinetic energy input of $\sim$10 to as many as 200 SNe each. Such multiple supernova events should only occur in massive clusters or OB associations. Age estimates of the \ion{H}{1}\ holes found in Ho$\thinspace$II have been assigned based on their sizes and expansion velocities, and typical ages fall in the range of 10$^7$ to 10$^8$ years. Consequently --- if the ages are indeed accurate --- the upper main sequence stars (late-B, A and F) should still be present in the clusters which produced the SN precursors. During the brief time period since these clusters formed, they will not have dispersed significantly, and should still be observable as blue sources at the centers of the \ion{H}{1}\ holes. Since the surface brightness level associated with the underlying old population of stars in Ho$\thinspace$II is fairly low, the upper main sequence population in these clusters should readily stand out. If these young star clusters produced, for example, 50 OB stars which became SNe, then a Salpeter IMF would predict the existence of at least 300 upper main sequence stars which would still be present after 10$^8$ years. Although this population of stars would not be resolved from the ground (a single main sequence A0 star would have m$_V$ = 28 and 1 arc second corresponds to 15 parsecs at the distance of Ho$\thinspace$II), the total cluster brightness should be m$_V$ $\sim$ 22, which would be readily detectable in deep CCD images. In order to look for the young clusters which would provide evidence for the SN scenario, we decided to carry out a deep, multi-color imaging study of Ho$\thinspace$II. Our main goal was to obtain accurate optical magnitudes and colors of all objects appearing in and around the locations of the \ion{H}{1}\ holes found in the VLA maps. The information tabulated in P92 regarding the ages and energy requirements of the holes can be used to calculate the magnitudes and colors of the clusters that should be present in the holes if the SN scenario is correct. Direct comparison of these calculations and our imaging photometry allows us to carry out a sensitive test of the SN scenario. In the following section, details of the optical observations carried out for this study are discussed, along with the reduction steps executed on the data, and a complete description of the photometric measurements of the \ion{H}{1}\ hole regions and of individual sources located in or around the holes. Section~\ref{section:analysis} describes calculations of the sizes and brightnesses of the putative clusters made using the information tabulated in P92. The last two sections of the paper consist of a discussion of our findings and their implications for the scenario proposed in P92, followed by a summary and final remarks. \section{Observations} \subsection{Description and Preliminary Reductions} \label{section:observations} Observations of Ho$\thinspace$II were obtained in February 1994 and April 1995 with the 0.9-meter telescope at Kitt Peak National Observatory. The galaxy was imaged in three broad-band filters ($BV\!R$) during both observing runs, and in narrow-band H$\alpha$ during the February 1994 observing run. The detector used was a Tektronix 2048$\times$2048 CCD (T2KA), formatted to read out only the central 1280$\times$1280 pixels. Each pixel subtended 0.68\arcsec\ on the sky, resulting in a total field-of-view 14.5\arcmin\ on a side. Multiple exposures were obtained through each filter, and the telescope was dithered between exposures, to facilitate removal of particle events (cosmic rays) in the images. Total integration times were 2400 s in $B$, 1800 s in $V$, and 1200 s in $R$ for the 1994 data, and 2700 s in $B$, 1800 s in $V$, and 1800 s in $R$ for the 1995 data. Images were taken under clear observing conditions. Photometric standards (\cite{landolt83,landolt92}) were also observed for use as calibration sources. Observations of additional dwarf galaxies were obtained during these runs (Ho I, K 73, M81dwA, IC 2574, Leo A, Sex A, Gr8, DDO 147); the results for these objects will be presented in a subsequent paper (\cite{slaz99}). Images taken during the 1994 and 1995 observing runs were reduced separately. Preliminary reductions (overscan level subtraction, bias image subtraction, flat field division) were carried out following standard practices. Multiple images taken in sequence through a particular filter were aligned with the middle image in the sequence. Sky subtraction was accomplished by defining a series of rectangular regions surrounding the galaxy, measuring the mean flux level in those regions (after masking stars and cosmic rays), and subtracting that flux level from the image. Multiple images in a given filter were scaled to a common flux level (to preserve photometric integrity) and combined into a single image, using a pixel-rejection algorithm to eliminate cosmic rays. The combined $B$, $V$, and $R$ images were rotated to a north-up, east-left orientation, and their central coordinates were determined using an astrometry routine which measures the positions of Guide Star Catalog (\cite{lasker90}) objects appearing in the field. A total of six standard stars (\cite{landolt83,landolt92}) were used to calibrate the broad-band images of Ho II from the February 1994 run, while 21 were available for the April 1995 data. In all cases, the photometric zero-point constants were determined with an accuracy of $\sim$0.01 magnitude, i.e., the nights were photometric. In order to create the deepest possible $BV\!R$\ images with which to do photometry of the hole regions, the images from the two runs were convolved to a common resolution, then scaled, aligned, and combined to create a single, deep image in each filter. The fluxes were scaled to the April 1995 values, since the photometric calibration was deemed to be of slightly higher quality for that run. Photometry was carried out on faint point sources in the resultant images to quantify the detection limit. The 4-$\sigma$ limit on the brightness of any point source is $B$ $=$ 23. The resolution (PSF FWHM) of each of the final combined $BV\!R$\ images is 2.4\arcsec. A continuum-subtracted H$\alpha$ image was created using the images taken in February 1994 through the on-band ($\lambda_o$ = 6569 \AA, $\Delta\lambda$ = 89 \AA) and off-band ($\lambda_o$ = 6409 \AA, $\Delta\lambda$ = 88 \AA) filters. Both narrow-band images were aligned with the composite $BV\!R$\ frames and convolved to a common resolution. The off-band image was scaled to the on-band image by comparing the fluxes for several bright stars in the two images, after which it was subtracted from the on-band image to produce the final H$\alpha$ frame. Since the narrow-band images were acquired under non-photometric conditions, no flux calibration was attempted. The high-resolution (natural weight) \ion{H}{1}\ map of Ho$\thinspace$II from P92 was used for this study to provide information about the hole locations. To facilitate comparison between the optical and radio data, the format of the \ion{H}{1}\ map was modified to match that of the optical images. Specifically, the original \ion{H}{1}\ map (2\arcsec/pixel) was rescaled to the optical image scale (0.68\arcsec/pixel). The section of the radio map corresponding to the area covered by the optical frames --- i.e., having the same size and central coordinates as the optical images --- was then extracted and used to create a new image. Figure~\ref{fig:hi map} shows the high-resolution \ion{H}{1}\ map after it has been scaled and aligned with the optical image, and Figure~\ref{fig:combined BVR image} shows a composite $BV\!R$\ image, created by combining the $B$, $V$, and $R$-band data taken in April 1995. \subsection{Photometry} \label{section:photometry} \subsubsection{H$\thinspace$I Holes} To search for evidence of star clusters at the centers of the \ion{H}{1}\ holes, photometric measurements were carried out on the combined $B$, $V$, and $R$ images created by merging the data from the February 1994 and April 1995 observing runs. These combined frames provided us with the deepest images possible for our search. The photometry was performed using the following strategy. The hole positions given in Table 5 of P92 were used as an initial guess for locations of our synthetic apertures. These apertures were overlaid on the scaled \ion{H}{1}\ image, and the positions adjusted slightly when required to center the apertures within the \ion{H}{1}\ holes. This was necessary since the reported accuracy of the P92 positions was $\pm$0.1\arcmin. Next, we divided the holes into four categories according to size (again based on the data from P92), and assigned aperture sizes accordingly. The apertures adopted for the four size groups are listed in Table \ref{table:apsizes}. The decision to use variable aperture sizes for the photometry was motivated by several factors: (1) the central positions of the holes become increasingly uncertain for larger holes; (2) there is no {\it a priori} reason to expect the putative star clusters to be precisely in the hole centers; (3) the larger holes tend to be older so that diffusion would tend to increase the size of the clusters in these holes. The apertures used ensure that we did not miss the clusters. Because substantial galaxian background is present in many hole locations, we measured the flux in both a circular aperture corresponding to the hole location and in a concentric annular ring surrounding the aperture. This allows us to compare the fluxes and colors in the two regions. The area of each annular region was chosen to be approximately equal to that of the corresponding circular aperture to ensure comparable signal-to-noise for the two measurements. The test for the presence of a star cluster within the \ion{H}{1}\ holes then involves comparing the fluxes within each aperture/annulus pair to look for an excess of flux (expected to be blue in color) coming from the central circular aperture. In several instances, bright foreground stars were located inside the aperture or annular regions we had marked for photometry. In such cases we either adjusted the positions of the hole centers slightly, if possible, to avoid contamination from these objects, or excluded those holes from the sample. Photometry was successfully executed on 44 of the 51 hole regions identified in P92. The locations of the apertures and annuli are shown on the radio map (Figure~\ref{fig:hi map}) and on the composite $BV\!R$\ optical image (Figures~\ref{fig:combined BVR image}). Results from photometry of the aperture/annular regions are given in Table~\ref{table:photholes}. The equatorial coordinates of our aperture centers are given for each hole, along with the aperture (inner diameter) and annular (outer diameter) sizes in arc seconds, plus the $B$ band magnitude, $B$$-$$V$ and $B$$-$$R$ colors, and average $B$-band surface brightness of the integrated light within both the aperture and annulus. Throughout this paper we use 1950 coordinates, for consistency with P92. Note that the background-subtracted flux measured in the annulus of hole \#15, as well as for both the aperture and annulus of hole \#27, was slightly negative, resulting in indeterminate magnitudes and colors. These two holes contain no measurable light from either foreground objects or from Ho$\thinspace$II itself. \subsubsection{Point Sources in the HI Hole Regions} A number of the holes in Ho$\thinspace$II had faint objects which appeared within the region delimited by our circular apertures or in the surrounding annulus. Since these objects could be the remnant star clusters which produced the sequential SN explosions described in P92, we wished to measure accurate magnitudes and colors for them. Photometry was performed on 29 such objects using a 5\arcsec\ diameter aperture and a background annulus with an inner diameter of 8\arcsec\ and an outer diameter of 20\arcsec. Because we were measuring magnitudes and colors of individual sources rather than an entire hole region, we decided to use the $BV\!R$\ images from the April 1995 observing run alone for this step, rather than using the images merged from both observing runs. The April 1995 data were of better image quality compared to those from February 1994, and this improvement in resolution almost completely compensated for the added depth of the combined images with regard to point-source detection. The resolution (PSF FWHM) of the April 1995 broad-band images is $1.7-1.8$\arcsec, and the 4$\sigma$ limit on the brightness of a point source is $B = 23$. Results from the point source photometry are given in Table~\ref{table:photclusters}. The first column lists an object designation reflecting the hole number in which the object is located, the next two columns give the position of the object, and the $B$ magnitude and colors of the source are given in columns 4 -- 6. We note that the flux from {\it any} object within the apertures or annuli was measured, regardless of the object's brightness or location. Therefore, this list should not be taken as a list of possible candidate star clusters. In many cases the objects listed have magnitudes and colors consistent with their being faint foreground stars. \subsection{Results of Photometry} \label{section:photometry results} We summarize the results of our imaging and photometric analysis in Table~\ref{table:photsummary}. In this table, each of the 51 \ion{H}{1}\ holes is characterized with regard to the following key question: is there evidence for a star cluster at or near the center of the \ion{H}{1}\ hole? The categories into which each hole has been assigned are: (1) Empty Hole -- No source within the central aperture exceeding a flux level of 3$\sigma$ above the annular flux level. (2) Galaxian Background -- Source(s) present in the hole, but with the characteristics/appearance of general galaxian background light rather than the putative star clusters. (3) Possible Star Cluster -- Source(s) within the hole with the correct characteristics (color, brightness, appearance) of a genuine star cluster. (4) Possible Photoionization Region -- The hole is coincident with an \ion{H}{2}\ region and is most likely a cavity of photoionized gas rather than a wind/SNe-blown hole. (5) Faint Foreground Star -- Photometry reveals that the object in the hole has the brightness and colors consistent with it being a foreground star not associated with Ho II. (6) Contaminated/No Photometry -- A bright foreground star is present in the hole; no photometry was attempted. It should be stressed that the assignment of a given hole into one of the six categories is by no means unambiguous in all cases. For example, assigning objects to category 2 (galaxian background) as opposed to category 5 (faint foreground star) was at times subjective. Further, we emphasize that objects in category 3 (possible star cluster) are assigned that designation even if the color and brightness of the source is only {\it broadly} consistent with the expected values. An object classified as category 3 cannot be interpreted as having been shown to be a young star cluster within Ho$\thinspace$II. Rather, it is {\it consistent} with that hypothesis, but could just as likely be a foreground star (e.g., a halo white dwarf). Two holes (numbers 1 and 8) have point source photometry listed in Table~\ref{table:photclusters}, but the sources in question are located in the outer annuli. Hence, these two holes are classified as category 1. The main results from Table~\ref{table:photsummary} that we want to emphasize are: (1) nearly one third of the holes have no obvious optical source within our photometric apertures, and (2) only a minority of holes (6 of 44 for which photometry was obtained) have optical sources with the colors and brightnesses consistent with the expected star cluster. One hole (\#43) is located in a complex of bright \ion{H}{2}\ regions, and may represent a hole which is created by photoionization of the \ion{H}{1}\ gas by the O and B stars. The status of the remaining holes (category 2 and 5) are uncertain, owing to the foreground or galaxian light they contain. However, in most cases this light is inconsistent with the properties of the putative star clusters, as quantified in the following section. We note that the photometry listed in Tables~\ref{table:photholes} and \ref{table:photclusters} has not been corrected for reddening, either due to the Milky Way or internal to Ho$\thinspace$II. Burstein \& Heiles (1984) list a color excess of $E$($B$$-$$V$) $=$ 0.03 due to foreground reddening in the direction of Ho$\thinspace$II. With regard to intrinsic absorption in Ho$\thinspace$II, spectra of four \ion{H}{2}\ regions by Hunter \& Gallagher (1985) yield Balmer decrements (H$\alpha$/H$\beta$) consistent with zero or modest reddening. Hence, we believe that dust is not significantly affecting our photometry. This should be especially true in the outermost regions of Ho$\thinspace$II, where our photometry provides the most sensitive test of the SNe hypothesis. We will return to the photometry results in more detail in section~\ref{section:discussion}. \section{Modeling Analysis} \label{section:analysis} P92 catalogued the observed properties of the \ion{H}{1}\ holes in Ho$\thinspace$II, and from these, derived quantities such as the ages of the holes and the kinetic energies required to create them. Some of these observed and derived quantities (such as the radial expansion of the holes, and their energy requirements) were interpreted as evidence in favor of the stellar wind/multiple SNe scenario for the origin of the holes. In order to compare directly our observations with predictions that arise from the SN scenario, we have used the hole properties tabulated in P92 to derive the observable characteristics of the clusters which should exist if the SN hypothesis is correct. A number of the quantities tabulated in P92 are distance-dependent; for example, the kinetic energies they calculated are given by the equation (\cite{chev74}): \begin{equation} \label{equation:energy} E = 5.3 \times 10^{43}\ n_{HI}^{1.12}\ R^{3.12}\ V^{1.4}\ {\rm ergs} \end{equation} \noindent where $n_{HI}$ is the volume density of the surrounding medium in particles per cubic centimeter, $R$ is the radius of the hole in parsecs, and $V$ is the expansion velocity of the hole in km\thinspace s$^{-1}$. P92 adopted a distance of 3.2 Mpc for Ho$\thinspace$II. A more recent distance determination has been done by Hoessel et al.~(1998) using Cepheid variable stars in Ho$\thinspace$II. We have adopted their revised distance modulus for our calculations: $(m-M) = 27.42$, which yields a distance of 3.05 Mpc. Therefore we have used the data tabulated in P92 but have re-scaled it in accordance with this revised distance. Some of the relevant quantities that changed with the new distance have been included in Table~\ref{table:energy.out}. For most holes, the revised distance resulted in computed hole energies that are $\sim$14\%\ less than those tabulated by P92. Note, however, that a typographical error in P92 caused their published energy value for hole \#43 to be underestimated by more than an order of magnitude. Our table reflects the corrected energy for this hole. To derive the characteristics that should be observable if the remnant clusters are located inside the \ion{H}{1}\ holes, we proceeded as follows: using the kinetic energies of the holes, the number of supernova explosions required to create each hole was calculated. The energy imparted to the ISM by one supernova explosion was taken to be 10$^{51}$ ergs (cf.~\cite{mccray87}). If the number of supernova explosions required to create a hole was $<$1, that hole was omitted from the rest of the calculations. For each remaining hole, a Salpeter IMF (\cite{salpeter55}) was used to calculate the mass distribution for a model star cluster. This distribution was then scaled so that the number of stars of mass $\geq$ 7 $M_{\sun}$ was equal to the number of supernovae required to create that particular hole. (Stars with masses $\gtrsim$ 7 $M_{\sun}$ were assumed to be of sufficient mass to end their lives as Type II supernovae.) Next, stars which would have evolved off the main sequence over a time scale equal to the age of the hole were removed from the distribution. Finally, composite magnitudes and colors were calculated for each of the synthetic clusters, for comparison with our observations. The same calculations were repeated for a Miller-Scalo IMF (\cite{miller79}) and using a limit of $\geq$ 8 $M_{\sun}$ for the lowest mass stars which end their lives as Type II supernovae. Performing the calculation with these two sets of parameters provides us with a reasonable range of predicted brightnesses for the putative clusters, and reflects the current levels of uncertainty in both the slope of the IMF and lower mass limit for Type II supernova precursors. We note that in this simple calculation we make no effort to take into account the light that the post-main-sequence stars would contribute to the cluster total. For clusters in the age range considered here, red supergiants might make a significant contribution to the total light. By ignoring the evolved stars, our models are providing only {\it lower limits} to the total cluster brightness. This must be kept in mind when comparisons are made with the observations. The final results of the modeling analysis are given in Table~\ref{table:energy.out}. The hole energies and ages listed in columns 2 and 3 differ slightly from those tabulated by P92, for the reasons mentioned above. The predicted number of SNe for each hole is one-tenth of the number given in column 2 (i.e., the hole energy divided by 10$^{51}$). Only those holes with energies in excess of 10$^{51}$ ergs (i.e., those which require at least one SN) are listed in the table. The $B$ magnitude and $B$$-$$V$ color listed under the heading ``Model I'' are the composite values for the model clusters computed using the Salpeter IMF and a Type II SN mass limit of 7 $M_{\sun}$. The corresponding values listed under ``Model II'' are for models using a Miller-Scalo IMF and a mass limit of 8 $M_{\sun}$. In general, Model II predicts brighter clusters. For many of the holes, the predicted cluster brightnesses are well below our observational limits, and hence we are unable to say anything definitive regarding the presence or absence of a young stellar population. However, for holes requiring energies in excess of 10$^{52}$ ergs (or $>$ 10 SNe) the expected brightness of the putative star cluster is at or above the limits set by our data. These holes will receive careful scrutiny in the next section. \section{Discussion} \label{section:discussion} \subsection{Results} \label{section:results} The main results of our study are summarized in Table~\ref{table:comparison}, which combines data from Tables~\ref{table:photholes}, \ref{table:photclusters}, \ref{table:photsummary}, and \ref{table:energy.out} for specific holes. In Section~\ref{section:observations} we established that a point source with $B$ $=$ 23.0 would be detected at the 4$\sigma$ level in our images. We list in Table~\ref{table:comparison} all \ion{H}{1}\ holes that are predicted to contain star clusters as bright or brighter than $B$ $=$ 23.0 in one or both of the models presented in Section~\ref{section:analysis}. The entries are sorted by hole category, as taken from Table~\ref{table:photsummary}. A number of holes characterized in Table~\ref{table:photsummary} as category 1 (no objects visible in the hole) are seen to have limiting magnitudes significantly fainter than the expected brightness of the putative clusters. In other words, if the SNe hypothesis for the origin of the holes is correct, one should definitely be able to see the clusters in a number of cases. Specific holes for which this test fails include 10, 13, 47, 49, and 50. In all these cases, the observational limits are more than one magnitude fainter than the predicted values for the case of Model II (8 M$_\odot$ mass limit for SN). If the light from the evolved cluster stars is taken into account, the differences between the predicted brightnesses and the observed limits becomes even greater. It is interesting to note that these five holes are among the most energetic in Ho$\thinspace$II, each requiring between 32 and 65 supernovae to create them in the P92 scenario. For the specific \ion{H}{1}\ holes mentioned above, we can clearly rule out the presence of the predicted star clusters at the expected levels. Therefore, it seems extremely unlikely that these holes were created by multiple SNe as hypothesized by P92. This is perhaps no great surprise, since the holes in question are located in regions of extremely low surface brightness. In fact, most lie outside of the Holmberg radius of Ho$\thinspace$II. At these large distances from the galaxian center, the \ion{H}{1}\ gas column densities are quite low, well below the empirical limits suggested by Kennicutt (1989) necessary for star formation to occur. In general, little star formation should occur beyond the Holmberg radius, so that the existence of multiple massive star clusters in the outer regions of Ho II, as predicted by P92, would be at odds with what we know about star formation in other galaxies. The limits we can place on the presence of star clusters in the inner \ion{H}{1}\ holes are less interesting for two reasons. First, the galaxian background is much higher there, making it easier to hide the presence of any cluster light. Second, the inner holes tend to be smaller in size and hence require fewer SNe to create them. Consequently the brightnesses predicted for these holes tend to be fainter than our observational limits. Therefore, we cannot rule out the possibility that some of these inner \ion{H}{1}\ holes are produced by SNe. There is of course no compelling reason for believing that all the holes have the same origin; we can only adequately test the SNe hypothesis for the outer \ion{H}{1}\ holes. A few holes for which the observations may actually support the SNe hypothesis of P92 are numbers 21, 36, 44, and 48. These are listed in Table~\ref{table:comparison}. Holes 21 and 48 are the most likely candidates, although hole 21 is in a very crowded region with many point sources in the optical images. The sources in holes 36 and 44 are both significantly brighter than the model predictions. Further evidence against the SNe hypothesis comes from the H$\alpha$ image, which is shown in Figure 3 with the \ion{H}{1}\ hole apertures superposed. Given the predicted ages and energetics of the holes, one might expect that at least some of the holes would exhibit diffuse H$\alpha$ emission. No such emission is seen in any of the holes, with the possible exception of the few holes that are coincident with \ion{H}{2}\ regions (e.g., holes 16, 20, and 43). The locations of H$\alpha$ emission trace out the regions of high \ion{H}{1}\ column density seen in Figure 1, indicating that the current star formation is occurring at local density maxima in the neutral gas distributions. P92 interpreted this as SNe-induced star formation caused by the compression of swept-up gas. We note, however, that this interpretation does not appear to be consistent with most of the holes, since only the minority have H$\alpha$ emission nearby. ROSAT observations of Ho$\thinspace$II have failed to detect the presence of any diffuse X-ray gas inside the HI cavities (F.~Walter \& J.~Kerp, private communication). Such X-ray emission would be expected if the holes were filled with hot coronal gas from the SNe explosions. Furthermore, Stewart et al.~(1997) analyzed far-ultraviolet (FUV) images of Ho$\thinspace$II taken with the Ultraviolet Imaging Telescope and found no bright FUV knots located within the \ion{H}{1}\ hole locations identified by P92. They found instead that bright FUV emission, if present, was likely to occur outside the hole boundaries. \subsection{Alternative Explanations} \label{section:alternatives} The observational evidence presented here strongly suggests that at least some of the \ion{H}{1}\ holes in Ho$\thinspace$II are not caused by multiple SN explosions in the manner envisioned by P92. Here we discuss possible alternative explanations for the presence of these features. \subsubsection{Modified Hole Energetics} \label{section:modified energetics} One possibility is that the SNe hypothesis is correct, but that the numbers published by P92 for the energetics of the \ion{H}{1}\ holes are systematically overestimated. The energies and ages derived by P92 depend critically on the observed expansion velocities of the holes. These are difficult to measure precisely, and in some cases the evidence for the expansion is weak at best. If the measured expansion velocities given in P92 are systematically too high, then the holes might actually be substantially older, and the true required energies would be significantly reduced. This being the case, it might be possible to reduce the expected number of SNe, and hence the brightness of the remnant star clusters, below the observational limits found from our data. However, for the largest holes, a reduction in the total number of SNe of more than a factor of $\sim$5 is required. This would then imply an overestimation of the expansion velocities by a factor of more than three. Such a large error in the expansion velocities seems unlikely. Furthermore, the additional factors mentioned above regarding star formation beyond the Holmberg radius and the lack of H$\alpha$ and X-ray gas could still pose problems for this interpretation. Another possibility is that the energetics of the holes have been overestimated not because of errors in the the expansion velocities, but because of uncertainties in the energy calculation itself. Estimates for the amount of energy imparted to the ISM by stellar winds and supernovae are based on our knowledge of the efficiencies of these processes, which are not well-constrained. P92 calculated the energy associated with each of the \ion{H}{1}\ holes in Ho$\thinspace$II using an expression derived from a hydrodynamical model by Chevalier (1974), which describes the evolution of a single spherically-symmetric supernova remnant in a uniform medium. It seems possible that using the results from such a model to calculate the total energies associated with \ion{H}{1}\ holes in the non-uniform ISM of a galaxy like Ho$\thinspace$II could introduce uncertainties of at least a factor of a few. As explained in the previous paragraph, if the energies of the holes have been overestimated, then fewer SNe may be needed to create the \ion{H}{1}\ holes, which could push the putative cluster brightnesses below our observational limits. As before, however, this explanation still requires that some of the \ion{H}{1}\ holes be created by supernovae occurring well beyond the Holmberg radius, in regions of the galaxy where star formation does not appear to have occurred in the past. \subsubsection{A Non-Standard IMF} \label{section:non-standard IMF} Another way to retain the SNe hypothesis would be to invoke an unusual IMF for the putative clusters. This possibility was actually suggested by P92. A top-heavy IMF, rich in massive stars but poor in low-mass ones, could explain the observations. However, there is as yet no real evidence for significant variations in the IMF as measured in different environments in the Milky Way and other nearby galaxies (\cite{leitherer98}). Furthermore, since the stars which would be providing the expected cluster signature are themselves fairly massive, the IMF slope required to produce the requisite number of SNe would have to be quite severe, perhaps even inverted, in order to not produce a detectable population of B and A stars. Invoking such an unusual IMF is not a very compelling explanation. Since the SNe hypothesis has significant problems, even when allowances for the above variants are made, we are forced to consider the alternative that the outer \ion{H}{1}\ holes are not produced by the action of stellar winds and SNe. We again stress that our analysis does not exclude the possibility that the inner holes {\it are} produced by SNe/winds. \subsubsection{Gamma-Ray Bursts} \label{section:GRBs} One possible explanation that has gained attention recently is the suggestion that the holes are remnants of Gamma-Ray Burst events (GRBs). Recent work by Loeb \& Perna (1998) and Efremov et al.~(1998) proposes that some of the HI supershells and hole features seen in nearby galaxies (such as the dwarfs in our sample) are remnants of GRBs. These authors suggest that GRBs might be associated with the release of gravitational binding energy during, for example, the collapse of a single, massive star to a black hole. They argue that such an event could produce a blast wave with energy comparable to the multiple-SN events thought to be necessary to produce the \ion{H}{1}\ hole features. Since this proposal does not necessarily require that a large star cluster be left behind after the explosive event, the observational test applied here does not rule it out. Because only a single star can account for the GRB, one could hypothesize that the explosions which create the \ion{H}{1}\ holes do not occur in massive star clusters, but rather in smaller associations which we would have no hope of detecting in our data. However, like the SN scenario, the GRB scenario requires that massive stars be present at the centers of the \ion{H}{1}\ holes, in order to produce the expanding blast wave that creates the hole. As we have noted, many of the holes in Ho$\thinspace$II occur in extremely LSB regions which show no indication of recent massive star formation. In addition, the GRB hypothesis would still predict the presence of hot X-ray emitting gas within the \ion{H}{1}\ cavity. Hence, although the GRB hypothesis is an attractive alternative to the multiple-SNe scenario, and may well explain some of the \ion{H}{1}\ holes seen in Ho$\thinspace$II, we consider it unlikely to be correct for the outer holes, i.e., the same ones for which our current study has ruled out the SNe hypothesis. \subsubsection{Impacts from High-Velocity Clouds} \label{section:HVCs} If stellar energy sources are ruled out, the next two most likely explanations for the \ion{H}{1}\ holes are large-scale dynamical effects and ionization. The dynamical effects of a collision between an infalling neutral gas cloud and a galactic disk, resulting in an \ion{H}{1}\ hole, were modeled by Tenorio-Tagle (1980, 1981). Clear associations between \ion{H}{1}\ holes and high-velocity clouds are known in the Milky Way (Heiles 1985), M101 (van der Hulst \& Sancisi 1988; Kamphuis, Sancisi, \& van der Hulst 1991), NGC 628 (Kamphuis \& Briggs 1992), NGC 6946 (Kamphuis \& Sancisi 1993), and NGC 5668 (Schulman et al.~ 1996). The high-velocity clouds could be primordial material or the result of a galactic fountain. This mechanism is attractive for explaining the holes at large galactocentric radii in Ho$\thinspace$II, where there is very little starlight and very little likelihood for star formation. One of us (DJW) has re-examined the data cubes from P92 to search for candidate high-velocity clouds. The data cube with the greatest sensitivity, with a synthesized beam of 28 $\times$ 27 arc seconds and pixels of 10 arc seconds, was examined in the velocity range 64 to 244 km s$^{-1}$. The root-mean-square (RMS) signal in an apparently line-free channel was 1.88 mJy beam$^{-1}$. Candidate clouds were found by first making a statistical search for bright pixels, then searching for extended bright areas around those pixels. A candidate was required to have extended, superimposed signal in at least three adjacent channels. Only one candidate cloud met our requirements. It is located at 08$^h$ 16$^m$ 37$^s$ 70\arcdeg\ 42\arcmin\ 36\arcsec, with a central velocity of 223.5 km s$^{-1}$ and a total velocity width of 7.7 km s$^{-1}$. It is 17\arcmin\ from the center of Ho$\thinspace$II, giving a projected separation of about 15 kpc at the assumed distance of 3.05 Mpc. Pixels brighter than the background RMS are found in three channels. The peak intensity is 9.81 mJy beam$^{-1}$ (5.2 times the RMS) in the raw maps, and 19.63 mJy beam$^{-1}$ after correction for the primary beam. The maximum brightness temperature is 2.4 K. The cloud's integrated \ion{H}{1}\ line signal is 4.9 Jansky km s$^{-1}$, giving a total \ion{H}{1}\ mass of 1.2 $\times$ 10$^7 M_{\sun}$. Its total angular extent is 92\arcsec\ or 1.4 kpc. The properties of this candidate cloud are similar to those of the Milky Way high-velocity clouds described by Wakker \& van Woerden (1997). This cloud is not easily visible in the data cube, so no mention was made of it in P92. We wish to emphasize that this is a marginal detection at best, and must be confirmed by an independent observation before it can be considered anything but a candidate cloud. More than two million apparently empty pixels were searched for local peaks which might be candidate clouds. Only two pixels were found to be as bright as 5.2 times the RMS, and of these, only one had extended signal in three channels. In a sample of this size, two pixels as bright as 5.2 times the RMS are expected if the noise is Gaussian. There is a good possibility that we have simply found a noise peak rather than a real cloud. Even so, the candidate cloud can be used to set limits on the presence of high-velocity clouds within the observed velocity range. The cloud has a velocity of 65 km s$^{-1}$ relative to Ho$\thinspace$II, which has a systemic velocity of 158 km s$^{-1}$. Its kinetic energy in the frame of Ho$\thinspace$II is 5 $\times$ 10$^{46}$ Joules or 5 $\times$ 10$^{53}$ ergs, large enough to cause the largest holes in Ho$\thinspace$II. The possible detection of a single candidate cloud with enough energy to cause a hole does not prove or disprove the hypothesis that infall caused the holes. Infall could be episodic, or there could be clouds outside the observed velocity range. It does show the observational difficulty in identifying high-velocity clouds in galaxies at the distance of the M81 group --- even in observations as deep as those of P92, candidate clouds are nearly indistinguishable from 5-$\sigma$ noise peaks. \subsubsection{Large-Scale Turbulence} \label{section:turbulence} Holes might be unavoidable due to the nature of the interstellar medium. It is well known that molecular clouds are fractal --- see Beech (1987), Bazell \& D\'esert (1988), Falgarone (1989), Scalo (1990), Falgarone, Phillips, \& Walker (1991), and Elmegreen \& Falgarone (1996). It is becoming clear that \ion{H}{1}\ is fractal as well --- see Vogelaar \& Wakker (1994) and Westpfahl et al. (1999). The fractal dimension of the ISM is similar to that of structures seen in laboratory turbulence, which has led Elmegreen \& Efremov (1999) and others to conclude that interstellar clouds form by processes related to turbulence. The processes which cause fractal structure, including turbulence, usually produce an internal distribution of holes, characterized as lacunarity by Mandelbrot (1983). If \ion{H}{1}\ clouds are produced by processes related to turbulence, the holes may be a manifestation of the formation process. We note that the fractal nature of \ion{H}{1}\ distributions may change the energy required to form a hole via supernovae and stellar winds. A fractal structure with significant lacunarity may provide natural chimneys through which supernova ejecta can flow, thus significantly increasing the amount of energy injection needed to form an expanding shell. \subsubsection{Ionization} \label{section:ionization} Holes might also be formed by ionization. A source of ionizing photons in the outer regions of Ho$\thinspace$II might be the intergalactic UV field. The observed column densities in the HI gas at the locations of these outer holes is of order a few times 10$^{20}$ cm$^{-2}$, significantly above the densities at which the UV radiation field can keep a large fraction of the HI ionized for an HI disk of normal thickness. However, P92 argue convincingly that the scale height of the gas in Ho$\thinspace$II is significantly larger than in a typical spiral disk. If this is correct, the actual volume density of the gas in the outer regions would be significantly less than that for gas in a spiral disk with the same measured column density. In this case, it might be possible for lower density pockets of gas to approach the threshold volume density below which the ionization fraction of the gas remains high in steady state due to the ionizing UV photons. In other words, once a pocket of low density gas is created in the outer parts of a puffed-up disk as is envisioned for Ho$\thinspace$II, it could become ionized by the intergalactic UV radiation field and remain highly ionized for a long time. Eventually the holes would be destroyed by dynamical processes (e.g., rotational shear), but in the outer parts of the galaxy they might well maintain their integrity for a substantial length of time due to the slow rotation speeds in low-mass galaxies like Ho$\thinspace$II. \subsubsection{An Unresolved Question} \label{section:unresolved question} The actual origin of the \ion{H}{1}\ holes in the outer parts of Ho$\thinspace$II remains an open question. In the present study, we suggest strongly that at least some of the holes are not produced by the combined action of stellar winds and SNe explosions. In many ways the SNe hypothesis is a natural and sensible explanation for the holes, which was perhaps why it was almost universally accepted on face value. However, in several instances it clearly fails the direct observational test which we have applied. Additional work needs to be done on this problem in order for a clearer picture of the nature of these large-scale features to be developed. Our current lack of understanding leaves open a number of questions regarding the evolution of the ISM in irregular galaxies, and in particular the actual role of feedback from massive stars in shaping the ISM. \section{Summary \& Conclusions} \label{section:summary} We have carried out a deep, multi-color imaging study of Ho$\thinspace$II, a dwarf galaxy in the M81 group which has been shown to contain a large number of expanding holes in its neutral hydrogen distribution. The formation of the \ion{H}{1}\ holes in Ho$\thinspace$II and other galaxies like it has been attributed to multiple SNe occurring within wind-blown shells around young, massive star clusters. To search for evidence of the clusters, we have compared our optical images with the published \ion{H}{1}\ maps, and have measured accurate magnitudes and colors of all objects in and around the \ion{H}{1}\ holes. Photometry of 44 hole regions in Ho$\thinspace$II reveals that at least 16 holes contain no detectable point sources brighter than $B$ $=$ 23.0. Ten of these holes are located beyond the Holmberg radius. An additional 21 holes contain only red ($B$$-$$V$ $>$ 1.0) sources, which are most likely either faint foreground stars or diffuse galaxian background emission from Ho$\thinspace$II. Only 6 holes contain sources which could be interpreted as being young clusters of stars with the requisite brightness and color. Comparison of models which predict the brightness of the putative star clusters with the observational limits obtained from our imaging data appear to rule out the SN scenario as being the cause of at least several of the most substantial \ion{H}{1}\ holes. While convincing arguments cannot be made against the SNe scenario for the majority of the holes due to the lack of depth of our images coupled with severe crowding in the central portions of the galaxy, the fact that at least several holes appear to require an alternative explanation for their origin raises doubts about the SNe scenario in general. The lack of diffuse H$\alpha$ and X-ray emission from any of the holes further supports the possibility that the SNe scenario may be incorrect. Recent suggestions that the \ion{H}{1}\ holes in galaxies like Ho$\thinspace$II are caused by the events which create Gamma-ray bursts are also not favored by the current findings, although such scenarios are more difficult to rule out with the optical data since they require only a single massive star. A number of other alternative explanations for the existence of the \ion{H}{1}\ holes are explored, including errors in the hole energetics, non-standard IMFs, dynamical processes such as large-scale turbulence or impacts from high-velocity clouds, and ionization. None of these alternatives is clearly favored at this time, and the origin of the \ion{H}{1}\ holes remains an open question. There is no doubt that energy input from massive stars plays a major role in shaping the ISM in galaxies. The current study, however, suggests that one must interpret the observational evidence for such influence carefully. Although the scenario proposed by P92 appears sensible, it makes a direct observational prediction which is not verified by the current study. The precise role that winds from massive stars and SNe shocks play in sculpting the gaseous distribution in galaxies remains an open question, calling for continued careful work on both the observational and theoretical fronts. \acknowledgments We are grateful for the professional support of the staff of Kitt Peak National Observatory during our two observing trips. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. KLR, JJS and LAR acknowledge with gratitude financial support from Wesleyan University, the Keck Northeast Astronomy Consortium, the Research Corporation, and the National Science Foundation, all of whom provided partial support for this project. DJW gratefully acknowledges financial support from the New Mexico Space Grant Consortium, the Kahlmeyer Foundation, and the Research and Economic Development division of New Mexico Tech. We have benefited from discussions with many colleagues, including D. Puche, R. Larson, S. Van Dyk, F. Walter, and E. Brinks. JJS would like to recognize L. van Zee and NRAO-Socorro for their hospitality during the preparation of portions of this paper. We would also like to thank the anonymous referee for useful comments. \clearpage
\section*{Acknowledgements} I am indebted to Toni Riotto for many useful discussions. \newcommand\pl[3]{Phys. Lett. {\bf #1}, #2 (19#3)} \newcommand\np[3]{Nucl. Phys. {\bf #1}, #2 (19#3)} \newcommand\pr[3]{Phys. Rep. {\bf #1}, #2 (19#3)} \newcommand\prl[3]{Phys. Rev. Lett. {\bf #1}, #2 (19#3)} \newcommand\prd[3]{Phys. Rev. D{\bf #1}, #2 (19#3)} \newcommand\ptp[3]{Prog. Theor. Phys. {\bf #1}, #2 (19#3)}
\section{Introduction} Researchers have been analyzing the X-ray emission from supernova remnants for decades. Often the results were compared with analytical or computational models (for example, Hamilton and Sarazin 1984, Hughes and Helfand 1985, or the Sedov-Taylor solution combined with ionizational collisional equilibrium spectral codes) in order to determine the SNR's attributes or study SNR evolution. Much progress has been made with this approach. Although the preponderant share of attention has been devoted to young remnants and those in moderate density environments, some questions require models of older remnants and remnants in very diffuse media. For example, how many soft X-ray photons should be attributed to a single SNR during its entire lifetime; can SNRs produce the observed surface brightness and spectrum of the Galactic halo or external galaxies or must we consider other physical phenomena? What is the X-ray signature of a very old SNR and how does it compare with the observed spectrum of the Local Hot Bubble or of other quiescent X-ray regions? What do the results tell us? Such questions cannot be adequately addressed with models of young remnants alone because the more plentiful older remnants differ hydrodynamically and ionizationally. In the young remnants, those in the free expansion and energy conserving phases, an energetic shock front sweeps up and heats up the ambient gas. The temperature rises so quickly that the degree of ionization does not equilibrate with the gas temperature. Thus the plasma is ``underionized'' and in the process of ionizing. Eventually, the shock weakens substantially and the recently swept up gas is heated to much lesser temperatures. In contrast to the hotter gas in the center of the remnant, this gas quickly radiates its excess energy; the SNR is now in the radiative phase. The SNR quickly evolves to have two distinct components, a ``shell'' of warm/cool gas nearer to the shockfront and a bubble of hot gas nearer to the center of the remnant. The hot bubble was once extremely hot and highly ionized. During the remnant's expansion, this gas cooled faster than the atoms could recombine. Thus the atoms have become ``overionized'' and are in the process of recombining. These questions require detailed models that span the remnants' lifetimes. This paper was written to provide models for such inquiries. This paper reports on the X-ray appearance of thermally conductive supernova remnants evolving in cool, non-thermal pressure dominated, low density media. The remnants were simulated with a detailed hydrodynamic computer code coupled to a nonequilibrium ionization code. The paper provides detailed discussions of the remnants' physical states and atomic physics, high resolution 50 to 1150~eV spectra, spatial maps, estimates of the ROSAT PSPC $1/4$ and $3/4$~keV countrates, and \carthree, \nitfour, and \oxyfive\ column densities. The paper is organized as follows: Section 2 presents the simulation techniques and model parameters for three remnants. The primary hydrocode run (Remnant A) has an explosion kinetic energy ($E_o$) of $0.5 \times 10^{51}$~ergs, ambient density ($n_o$) of $10^{-2}$~atoms~cm$^{-3}$, ambient temperature ($T_o$) of $10^4$~K, and ambient non-thermal pressure ($P_{nt}$) of 1800~K~cm$^{-3}$. Remnant B has four times the ambient non-thermal pressure and Remnant C has four times the ambient non-thermal pressure and twice the explosion energy. Section 3 treats Remnant A as a case study, thoroughly discussing its physical structure, evolution, ionization states, spectra, total soft X-ray luminosity, spatial appearance in soft X-ray emission, and \oxyfive, \nitfour, and \carthree\ content. Sections 4 and 5 more succinctly discuss Remnants B and C. Section 6 shows how the modeled emission would be interpreted if it were observed and examined using popular observing instruments and analysis techniques. These analyses include the determination of the color temperature, electron density, thermal pressure, and O VII - O VIII diagnostics. Provisos are discussed in Section 7. Section 8 applies the simulations to the Galactic halo and the Local Bubble. Section 9 summarizes the major results. Moreover, this paper, Paper I, and Paper III compose a larger study. Paper I provides a very detailed examination of the high stage ions (\oxyfive, \nitfour, and \carthree) in Remnant A - like SNRs. Paper III combines the results of Papers I and II with the predicted number of high latitude supernova progenitors then compares the X-ray and high stage ion quantities and signatures with the observations. \section{Modeling Method} The supernova remnants are simulated using a Lagrangian mesh style hydrocode, which employs non-equilibrium ionization and recombination, thermal conduction, and non-thermal pressure. The hydrocode assumes spherical symmetry and that the electrons and ions have equal kinetic temperatures. Paper I described the hydrocode in detail. The Raymond and Smith (1977, 1993) code was used to calculate the non-equilibrium spectra for the given ionizational and hydrodynamic states. Between Paper I and this paper, the implementation of the Raymond and Smith code was improved. The gas phase abundances of the ambient medium are taken as Grevesse \& Anders (1989) solar abundances. The simulations do not include additional metals from the progenitor. Section 7 and Paper I discuss provisos on the model. Remnant A has an explosion energy ($E_o$) of $0.5 \times 10^{51}$~ergs, while the ambient medium has a temperature ($T_o$) of $10^4$~K, ambient density of atoms ($n_o$) of $10^{-2}$~cm$^{-3}$, and ambient nonthermal pressure ($P_{nt}$) of $1.8 \times 10^3$~Kcm$^{-3}$. Paper I also explains the choice of these conditions. Remnants B and C were made with less conservative choices of the explosion energy and ambient nonthermal pressure. Remnant B has $E_o = 0.5 \times 10^{51}$~ergs, $T_o = 10^4$~K, $n_o = 10^{-2}$~cm$^{-3}$, and $P_{nt} = 7.2 \times 10^3$~Kcm$^{-3}$ and Remnant C has $E_o = 1.0 \times 10^{51}$~ergs, $T_o = 10^4$~K, $n_o = 10^{-2}$~cm$^{-3}$, and $P_{nt} = 7.2 \times 10^3$~Kcm$^{-3}$. \section{Remnant A: $n_o = 0.01$~cm$^{-3}$, $P_{nt} = 1.8 \times 10^3$, and $E_o = 0.5 \times 10^{51}$~ergs} \subsection{Remnant A -- Physical Structure} The temperature, density, and pressure of the simulated SNR at $10^4$~yrs of age are shown as solid lines in Figures~\ref{temperature}a, \ref{density}a, and \ref{pressure}a, while the later times are shown in other linestyles. The explosion creates an extremely hot, rarefied cavity enveloped by a hot, dense, rapidly expanding rim of swept up material. Thermal conduction has reduces the gradients in temperature and density. The blast wave or energy conserving phase of SNR evolution begins when the expanding shock front has swept up much more mass than was ejected in the supernova explosion. This takes place within roughly $5 \times 10^4$~yrs. During this phase, the remnant continues to expand rapidly. The thermal energy of the hot bubble is conserved and so as the remnant expands, the interior temperature declines. Between $10^5$ and $2.5 \times 10^5$~yrs, the dense rim begins to radiate copiously and thus cool quickly. By $5 \times 10^5$~yrs, it has cooled to the ambient temperature, creating a spatially thick, cool shell (see Figures~\ref{temperature}a and \ref{density}a). The shell is still bounded by a shock front which sweeps up and thermalizes the gas it encounters. Now, however, the shock is weaker and the post shock temperature is less than $10^5$~K. The SNR is now in its snow plow or momentum conserving phase. In time, the shell widens because the shock front expands much faster than the hot bubble. The hot bubble (also called the interior) continually erodes as the hot gas nearest its periphery cools and adds to the cool shell. In addition, the ambient pressure restrains, and, at about $2 \times 10^6$ yrs, halts the bubble's expansion. Subsequently, the hot bubble cools and diminishes very slowly. See Figures~\ref{temperature}b, c, and d, and \ref{density}b, c, and d. Between $1.6 \times 10^7$ and $1.7 \times 10^7$~yrs it disappears entirely. A more expanded description and additional illustrations are provided in Paper I. \subsection{Remnant A -- Overview to the X-Ray Emission} X-ray emission rates are governed by multiple effects. The important emission processes (collisionally excited line, bremsstrahlung, two-photon, and recombination emission) depend on $n^2$ and kinetic temperature. Collisionally excited line, two-photon, and recombination emission in the $1/4$ and $3/4$~keV bands additionally require that the metal atoms be highly ionized. Typically, most of the soft X-ray power is due to line emission from highly ionized, hot gas. In collisional equilibrium\footnote{The assumption of collisional equilibrium is a commonly used tool for discussing X-ray production, even in cases such as this, in which the ions are not in collisional equilibrium.}, the ions which emit strongly in the $\sim100$ to $\sim300$ eV energy range (useful for comparing with the ROSAT $1/4$~keV band countrate, the Wisconsin All Sky Survey B and C bands countrates, or the lower part of ASCA's range) are most abundant if the gas is $\sim10^6$~K. The ions that emit strongly in the $\sim400$ to $\sim1000$ eV energy range (useful for comparing with the ROSAT $3/4$~keV band countrate or the Wisconsin All Sky Survey M band countrate) are most abundant if the gas is $\sim5 \times 10^6$~K. Before the cool shell forms, the hot, comparatively dense gas behind the shock front produces the vast majority of the SNR's X-ray emission. After the cool shell forms, the region behind the shock is no longer hot or highly ionized and so is no longer emissive. Thus, most of the emission derives from the remnant's ``hot'', highly ionized interior. As a result of this shift, the spectrum and total flux level change markedly. \subsection{Remnant A -- Ionization States} The metal atoms in the hot bubble are seldom in the collisional ionizational equilibrium with the gas temperature. This challenge to the discussion of the plasma's ionization state can be resolved by defining an ionization temperature, $T_i$. If the out-of-equilibrium gas has the same distribution of ionization states as collisional ionization equilibrium gas having a kinetic temperature of $T_1$, then the ionization temperature can be defined as the equivalent equilibrium temperature, in this case $T_1$. Although the populations of ionization states in the SNR gas parcels do not exactly match equilibrium populations and although different elements may indicate different $T_1$'s, it is still useful to compare ratios of prevalent ions with the ratios found in collisional equilibrium in order to find approximate ionization temperatures, $T_i$'s. The ratios of Si$^{+8}$/Si$^{+9}$, Si$^{+7}$/Si$^{+8}$, O$^{+6}$/O$^{+7}$, and O$^{+5}$/O$^{+6}$ ions were tested as possible indicators of the ionization temperature, with the Si$^{+8}$/Si$^{+9}$ ratio chosen for presentation in Figure~\ref{fig:iont}. That figure presents the kinetic temperatures (the temperatures corresponding to the random velocities of the atoms) and ionization temperatures of the plasma for the five epochs prior to and just after the onset of shell formation as well as for a selection of later epochs. At $10^4$~years, the SNR is most dramatically underionized\footnote{In an underionized plasma, the ionization temperature is less than the kinetic temperature.}. The kinetic temperature, $T_k$, of the SNR is $\sim3 \times 10^7$~K, but the ionization temperature is only $\sim1 \times 10^6$~K. The ionization temperature of the gas in the interior is rising slowly because of the low collision rate in this low density, high kinetic temperature gas. The kinetic temperature in the interior is dropping rapidly because the SNR's expansion spreads the interior's thermal energy across a rapidly increasing volume. Around $10^5$~yrs, the kinetic and ionization temperatures converge, largely due to the drop in the kinetic temperature. By $2.5 \times 10^5$~yrs, $T_k$ in the interior is less than $T_i$; thus the plasma has become overionized\footnote{In an overionized plasma, the ionization temperature is greater than the kinetic temperature.}. The gas remains overionized for the rest of the SNR's life because cooling (via emission, thermal conduction to more rapidly cooling gas, and expansion) proceeds at a faster rate than recombination (see Figure~\ref{fig:iont}b). \subsection{Remnant A -- Spectra} The nonequilibrium X-ray emission from the simulated remnant was calculated with the Raymond \& Smith spectral code (1977, 1993), using the non-equilibrium ionic abundances calculated in the hydrodynamic simulation. The spectra, shown in logarithmic form in Figure~\ref{fig:spectralog} includes line emission, a bremsstrahlung continuum, recombination edges, and two-photon continua. The line emission from 50 to $\sim300$~eV is mainly due to highly ionized N, O, Ne, Mg, Si, S, and Fe. The line emission from $\sim300$~eV to $\sim600$~eV is mainly due to highly ionized C, O, Ar, and Ca. The lines between $\sim600$ and $\sim900$~eV are mainly due to highly ionized N, O, and Fe. The lines between $\sim900$ and 1150~eV are emitted mainly by highly ionized Ne, Fe, and Ni. The O VII triplet ($\sim570$~eV) and the O VIII Lyman $\alpha$ line (653~eV) merit special attention. They have been used as plasma diagnostics in the past (Vedder, {\it{et al.}} 1986, Canizares (1990)) and may be used in analyzing future high resolution data. Furthermore, the spectra for $10^4$, $2.5 \times 10^4$, $5 \times 10^4$, and $10^5$~yrs shows a rounded continuum running from $\sim$50~eV to $\sim$570~eV due to a combination of two-photon continua emitted by O, C, and N. The spectra for later epochs exhibit recombination edges. For example, the feature at 740~eV in the $10^6$~yr spectra is due to O$^{+7}$ recombining to O$^{+6}$. At $10^4$~yrs, the emitting plasma is exceptionally hot, but not yet highly ionized. Some of the lines above about 700~eV are still weak and the two photon continua below about 570~eV is very strong. Between $10^4$ and $2.5 \times 10^4$~yrs, the ionization temperature rises by about $50\%$ (to about $1.5 \times 10^6$~K) and the kinetic temperature drops by a factor of three (to about $10^7$~K). More lines populate the high energy end, and the ratios of various emission lines shift. In the ensuing 75,000~years, the kinetic temperature of the material behind the shockfront decreases substantially, slightly softening the bremsstrahlung continuum. The ionization temperature begins to exceed the kinetic temperature and tiny recombination edges begin to appear. For example, the O$^{+7}$ to O$^{+6}$ recombination edge at 740~eV marginally appears in the $10^5$~yr spectrum. The higher energy emission lines weaken. Between $1.0$ and $2.5 \times 10^5$~yrs, the gas within several parsecs of the shock front ceases to be X-ray luminous. With only the SNR's tenuous interior contributing, the total X-ray luminosity wanes and the spectrum alters. Lower energy emission lines dominate the spectrum. Recombination edges become more prominent, while line and two-photon emission diminish. The remnant's physical characteristics continue to evolve until an age of about one million years. Then for several million years afterwards, the kinetic and ionization temperatures remain nearly constant. Similarly, the spectra radically evolve between $2.5 \times 10^4$ and $10^6$~yrs, then remain fairly unchanged. Compared with the spectra from earlier epochs, the spectra for 1, 5, and $10 \times 10^6$~yrs (Figure~\ref{fig:spectralog}) have reduced strengths in the higher photon energy emission lines, increased prominence in the recombination edges and steepened slopes in the recombination continua. At $1.5 \times 10^7$~yrs, the kinetic temperature (only $\sim10^5$~K and less) begins to rapidly drop and the spectral characteristics change again. Now, the extremely steep bremsstrahlung component is much dimmer than those from the other processes. An ensemble of two-photon emission curves defines the shape of the continuum between 250 and 900~eV. A series of recombination edges punctuate the spectrum at 54, 126 and 387 eV and a relatively sparse collection of emission lines rises above the continua. (Note that for the low temperatures of this epoch, the Raymond and Smith spectral code terminated some predictions below 920 eV.) \subsection{Remnant A -- Total Luminosity and Spatial Appearance} The SNR's luminosity and spatial appearance are most easily presented in terms of broad band countrates. For this case, the ROSAT PSPC $1/4$ and $3/4$~eV bands are used. The ROSAT PSPC $1/4$~keV band is composed of the R1 band ($\sim110$ to 284 eV) and the R2 band ($\sim$140 to 284 eV), while the ROSAT PSPC $3/4$~keV band is composed of the R4 band ($\sim$440 to $\sim$1010 eV) and the R5 band ($\sim560$ to $\sim1210$eV). Figure~\ref{bandfunctions} depicts the band response functions. The R2 band is harder than the R1 band, although both bands have the same upper boundary at 284 eV. The R2 band also has some sensitivity above 500~eV. The simulated spectra discussed in the previous subsection were convolved with the ROSAT PSPC response matrix, effective area, window transmission and gas transmission coefficients (Briel {\it{et al.}} 1996) to determine the following results. During the first few epochs, the hot, relatively dense gas just behind the shock produces most of the soft X-rays, causing the remnant to appear luminous (Figure~\ref{fig:totalxray}) and edge brightened (Figures~\ref{fig:R1flux} and \ref{fig:R4flux}) Between $5 \times 10^4$ and $10^5$~yrs, the spectra softens. Consequently, the SNR dims and loses its edge brightened appearance in the $3/4$~keV band. Between $10^5$ and $2.5 \times 10^5$~yrs, the densest part of the SNR cools to less than $10^5$~K; the ``cool shell'' forms and the shock front weakens substantially. The region behind the shock now dims in the $1/4$~keV band. With only the SNR interior providing X-rays, the SNR ceases to appear edge brightened or bright. During the course of a couple million years, the total $1/4$~keV luminosity diminishes by a factor of $\sim 100$. It remains near this level until around $1.5 \times 10^7$~yrs when the ancient, relatively cool, relatively dense SNR begins a phase of rapid decline. Integrating the $1/4$~keV luminosity with respect to time yields $5.8 \times 10^{59}$ counts~cm$^{2}$ Integrating the $3/4$~keV luminosity with respect to time yields $1.2 \times 10^{58}$ counts~cm$^{2}$~yr~s$^{-1}$. \subsection{Remnant A -- The \carthree, \nitfour, and \oxyfive\ Content} Paper I provided a detailed study of the \carthree, \nitfour, and \oxyfive\ ions in Remnant A. This subsection summarizes the important results. While the SNR is young, (for example, at $10^4$~yrs) its atoms are ionizing up through the \carthree, \nitfour, and \oxyfive\ states. As a result, very large column densities of UV ions exist in the young remnant, however their bulk velocities and thermal broadening can be enormous. By $2.5 \times 10^4$~yrs, the interior gas has ionized beyond these states and the \carthree, \nitfour, and \oxyfive\ exist only in the recently heated, underionized gas just behind the shock front. As the shell forms, the shock front becomes too weak to ionize the swept up ambient gas up to the \carthree, \nitfour, and \oxyfive\ states. Henceforth, these ions only derive from cooling, recombining gas. Highly ionized oxygen can recombine to O$^{+4}$ before the nitrogen recombines to N$^{+3}$ or the carbon recombines to C$^{+2}$ and the gas cools before all of the carbon has recombined to the C$^{+2}$ level. As a result, the \carthree\ extends out to greater radii in the cooling gas than does the \nitfour\ or \oxyfive, and at very late times it resides in cooled gas. In Remnant A, some \carthree\ remains a million years after the SNR cools. The column densities are only weakly dependent on the age of the SNR, particularly after $5 \times 10^4$~yrs. The time integrals of the number of high-stage ions contained by Remnant A are $7.8 \times 10^{69}$ \oxyfive seconds, $7.2 \times 10^{68}$ \nitfour seconds, and $1.6 \times 10^{69}$ \carthree seconds. Dividing by the time integral of the area covered by the \carthree, \nitfour, or \oxyfive\ ions gives the time and impact parameter averaged column densities for a sightline which intersects the high stage ions\footnote{A subtilty is that the \carthree\ covers $15\%$ more area than the \oxyfive.}: $5.2 \times 10^{13}$ \oxyfive cm$^{-2}$, $4.7 \times 10^{12}$ \nitfour cm$^{-2}$, and $9.8 \times 10^{12}$ \carthree cm$^{-2}$. \section{Remnant B: $n_o = 0.01$~cm$^{-3}$, $P_{nt} = 7.2 \times 10^3$, and $E_o = 0.5 \times 10^{51}$~ergs} Remnant B has four times the ambient nonthermal pressure as Remnant A and half of the explosion energy of Remnant C. By comparing the structures, evolutions, spectra, and luminosities of these remnants we can see the effects of varying these parameters. \subsection{Remnant B -- Structure} Figures~\ref{temperature5uG}, \ref{density5uG}, and \ref{pressure5uG} depict Remnant B's kinetic temperature, density of atoms, and thermal and total pressure as a function of radius for a variety of ages. In the original simulation the shockfront traveled to the edge of the grid, reflected, and traveled back towards the hot bubble, colliding with the bubble between 1.1 and $1.2 \times 10^7$~yrs which is just before the end of the SNR's life. Thus the post $1.1 \times 10^7$~yrs results have been replaced with those from an additional, simulation performed with half the spatial resolution. The SNR has completely disappeared by $1.3 \times 10^7$~yrs. Before their cool shells form, this remnant and Remnant A are similar in size, temperature, density, and thermal and total pressure. The early evolution has been little affected by quadrupling the ambient nonthermal pressure. Sometime after their cool shells begin to form (which occurs only slightly earlier in Remnant B than in Remnant A), however, each bubble's thermal pressure comes to approximate the total ambient pressure. Thus, after the shell forms, Remnant B's bubble evolves to be of higher thermal pressure, hotter, and denser than Remnant A. Consequently Remnant B radiates away its energy on a shorter timescale and so contains a warm or hot bubble for only $\sim 1.2 \times 10^7$~yrs, whereas Remnant A's bubble was warm or hot for $\sim 1.6 \times 10^7$~yrs, \subsection{Remnant B -- Spectra, Total Luminosity, and Spatial Appearance} During Remnant B's energy conserving phase, its spectra (Figure~\ref{fig:spectralogB}) is nearly identical to that of Remnant A. Small differences begin to appear around 250,000~yrs and are obvious after about a million years. Many of the emission lines below 400~eV grow to be much stronger (even an order of magnitude stronger) than the lines in Remnant A and until Remnant B is on the brink of death, its recombination spectra are not as strongly sloped as those of Remnant A at the same ages. Figure~\ref{luminosityB} presents Remnant B's luminosity as a function of time For the first $2.5 \times 10^5$~yrs the luminosities in the ROSAT PSPC $1/4$ and $3/4$~keV bands nearly trace those of Remnant A. Afterward, both remnants dim, but because of its larger temperature, density, and thermal pressure, Remnant B dims less. Integrating the $1/4$~keV luminosity with respect to time, yields $1.4 \times 10^{60}$ counts~cm$^{2}$. The integrated $3/4$~keV luminosity luminosity is $1.1 \times 10^{58}$ counts~cm$^{2}$. Thus, during its lifetime, Remnant B produces over twice as many $1/4$~keV photons as Remnant A, and, unlike Remnant A, produces most of its $1/4$~keV photons after the cool shell forms. Its production of $3/4$~keV photons is no higher than that of Remnant A. Remnant B's spatial appearance can be discussed in terms of its surface brightness as a function of impact parameter. Like Remnant A, Remnant B initially appears edge brightened and later evolves to appear centrally filled. (see Figure~\ref{fig:halo29R1flux} and \ref{fig:halo29R4flux}). During its youth, Remnant B has a similar luminosity and spatial appearance as Remnant A. During its centrally filled phase, however, the ancient Remnant B's $1/4$~keV flux is a couple hundred $\times 10^{-6}$ counts s$^{-1}$ arcmin$^{-2}$, which is about ten times greater than that of Remnant A and bright enough to be observable to an analysis like that done by Snowden {\it{et al.}} (1998). \subsection{Remnant B -- The \carthree, \nitfour, and \oxyfive\ Content} The time integrals of the number of high-stage ions contained in Remnant B are $6.0 \times 10^{69}$ \oxyfive seconds, $4.6 \times 10^{68}$ \nitfour seconds, and $1.1 \times 10^{69}$ \carthree seconds. After the cool shell forms, the number of ions contained by the remnant is a slowly varying function of time and impact parameter. As a result, the age and impact parameter averaged column densities (found by dividing by the time integrals of the number of ions by the time integrals of the areas\footnote{The average area covered by \carthree\ ions is greater than that covered by \nitfour\ or \oxyfive\ ions.}) is a reasonable estimate of the column density expected for a sightline which traverses the high-stage ions in and around the hot bubble. These values are: $7.8 \times 10^{13}$ \oxyfive cm$^{-2}$, $5.7 \times 10^{12}$ \nitfour cm$^{-2}$, and $1.2 \times 10^{13}$ \carthree cm$^{-2}$. \section{Remnant C: $n_o = 0.01$~cm$^{-3}$, $P_{nt} = 7.2 \times 10^3$, and $E_o = 1.0 \times 10^{51}$~ergs} \subsection{Remnant C -- Structure} The kinetic temperature, density of atoms, and thermal and total pressure are plotted in Figures~\ref{fig:temperatureC}, \ref{fig:densityC}, and \ref{fig:pressureC}. Remnant C is hotter, larger, more diffuse, generally more pressurized, and longer lived. than its lower explosion energy cohort, Remnant B. \subsection{Remnant C -- Spectra, Total Luminosity, and Spatial Appearance} Figure~\ref{fig:spectralogC} depicts Remnant C's spectra. For the first 5 million years, the spectral features nearly trace those of Remnant B, except that Remnant C is brighter. Correspondingly, Remnant C's luminosity as a function of time (Figure~\ref{fig:luminosityC}) follows the pattern set by Remnant B, but with a greater magnitude. The time integrated ROSAT PSPC $1/4$~keV luminosity is $3.5 \times 10^{60}$ counts cm$^{2}$, which is 2.3 times that of Remnant B. Over $80\%$ of this emission is produced after the cool shell forms. The time integrated ROSAT PSPC $3/4$~keV luminosity is $2.8 \times 10^{58}$ counts cm$^{2}$, which is 2.5 times that of Remnant B. Only about $25\%$ of the $3/4$~keV emission is produced after the cool shell forms. The spatial appearance (Figures~\ref{fig:halo31R1flux} and \ref{fig:halo31R4flux}) is edge brightened during the remnant's energy conserving phase and centrally filled during the post shell formation phase. The progression is very similar to that of Remnant B. \subsection{Remnant C -- The \carthree, \nitfour, and \oxyfive\ Content} The time integrals of the number of high-stage ions contained in Remnant B are $1.2 \times 10^{70}$ \oxyfive\ seconds, $8.9 \times 10^{68}$ \nitfour\ seconds, and $2.4 \times 10^{69}$ \carthree\ seconds. Dividing these time integrals by the time integrated area covered by each of the respective ions yields the age and impact parameter averaged column densities. They are $8.7 \times 10^{13}$ \oxyfive\ cm$^{-2}$, $6.0 \times 10^{12}$ \nitfour\ cm$^{-2}$, and $1.5 \times 10^{13}$ \carthree\ cm$^{-2}$. \section{The Observer's Perception of the Simulated Data: Color Temperature, Electron Density, Thermal Pressure, and Oxygen line diagnostics} Oft-used observational analysis procedures include determining an object's spectrum's color temperature, electron density, and thermal pressure. In addition, with very high spectral resolution data, the ratios of the O~VII emission line fluxes and the ratio of the O~VII to O~VIII line fluxes have been used as diagnostics of the plasma's temperature and ionization history. In this section, the simulated data is subjected to the identical treatment in order to determine how the ``observed quantities'' map to the original physical properties of the emitting gas and to provide observational signatures of a SNR evolving in a diffuse, pressure dominated region. Remnant A is used as the test case. Considering that Remnants B and C are hotter, they may have higher color temperatures. \subsection{Color Temperature} One wonders how hot one of the simulated SNRs would appear to an astronomer using only ROSAT data. In order to answer this question, a temperature yardstick must be created. (As is commonly done in analyses of observational data, the yardstick is calibrated as if the observed spectrum derives from ions which are in collisional equilibrium.) While the assumption may be far from true, the {\underline{apparent}} temperatures are still useful as descriptors of the measured spectra. Following the example of Snowden {\it{et al.}} (1998) a temperature measure (called the color temperature, $T_c$) was crafted by calculating the ROSAT PSPC R1/R2 and R2/R4 band ratios for equilibrium spectra simulated with the Raymond \& Smith code (1977, 1993), Grevesse and Anders (1989) abundances, and various assumed temperatures. The R1/R2 ratio is steep and single valued (making a good yardstick) between $T_{cR1R2} = 10^5$ and $10^6$~K. The R2/R4 ratio is single valued below $T_{cR2R4} = 2.5 \times 10^6$~K. See Figure~\ref{fig:colort} for the color temperature versus impact parameter. Interestingly, even when both color temperatures are within their single-valued ranges, the R1/R2 color temperature is generally less than the R2/R4 color temperature. Some evolution of the color temperatures does occur, but the most striking characteristic of the SNR is how small that variation is. At $10^4$~yrs, the spectrum from a pointing toward the center of the remnant has a R1/R2 color temperature of about $10^6$~K and a R2/R4 color temperature of about $2.0 \times 10^6$~K, while the spectrum from a pointing toward the edge of the remnant has slightly lower color temperatures. Between $10^4$ and $2.5 \times 10^4$~yrs, $T_{cR1R2}$ as a function of fractional impact parameter increases slightly while $T_{cR1R2}$ as a function of fractional impact parameter remains approximately constant. From $2.5 \times 10^4$ to $10^5$~yrs, $T_{cR1R2}$ and $T_{cR2R4}$ drop slowly. Over the remainder of Remnant A's lifetime, the the R1/R2 color temperature drops to several hundred thousand degrees, the R2/R4 color temperature hovers just above a million degrees, and the profiles develop and upward curve. This exercise is telling us that the observationally determined color temperature may not directly reveal either the kinetic or ionization temperatures, yet it may be possible to use simulations to interpret the observationally determined color temperature. A note of clarification: As x-ray photons transit interstellar material, the lower energy photons are preferentially absorbed, making the spectrum appear 'harder', as if it derives from hotter plasma. Thus, in order to determine the true spectra, the observed spectra must be 'de-absorbed', as is done in Snowden, {\it{et al.}} (1998)'s determination of the halo color temperature. In contrast, de-absorption is hardly an issue in analyzing the Local Bubble spectrum because that light transits a negligible absorption column density before reaching the solar system. \subsection{Electron Density and Thermal Pressure} In this subsection, the electron density ($n_e$) and thermal pressure ($P_{th}$) are calculated using an observational-type analysis of the simulated emission and compared with the $n_e$ and $P_{th}$ taken directly from the simulation. Using low spectral resolution detectors, such as those on ROSAT, the observationally determined variables are often inconsistent with the true values, suggesting that further progress in connecting the broad band data with the source phenomena may be possible via mappings such as the following. The standard analysis finds $n_e$ from the emission measure ($n_e^2(l) dl$), where $l$ is the depth of the luminous region and is generally taken as equal to its width or a simple function of the width). The emission measure is calculated from the observed countrate as follows. The color temperature (found from the observations) is taken as an approximation of the kinetic temperature (this is known to be a false approximation), the gas is assumed to be in ionizational equilibrium (also generally false), using these temperatures as inputs and using a given emission measure, the high energy resolution spectrum is then calculated with a spectral code (such as the Raymond and Smith code) and folded through the detector response functions. That act yields a countrate per emission measure which is then compared with the observed countrate in order to determine the emission measure. A hypothetical pointing toward the center of the $10^4$~year old simulated remnant (Remnant A) has an R2 countrate of $1.1 \times 10^{-3}$ counts s$^{-1}$ arcmin$^{-2}$, an R1/R2 color temperature of $1.2 \times 10^6$~K, and a depth of 60~pc. Assuming that the color temperature approximates the kinetic and ionization temperatures and using Grevesse and Anders (1989) abundances, spectral calculations made with the Raymond and Smith (1977, 1993) spectral code yield an R2 countrate per emission measure of $2.3 \times 10^{-20}$ counts s$^{-1}$ arcmin$^{-2}$ cm$^{5}$. The electron density calculated from the unrounded values is 0.016~cm $^{-3}$. Assuming full ionization of the hydrogen and helium and using $T_c$ as $T_k$ yields $P_{th}/k = 3.8 \times 10^4$~K cm$^{-3}$. How do these values compare with the emission measure, temperature, and thermal pressure found directly from the hydrocode? The observational-type analysis implies an $\int n_e^2(l) dl$ that is roughly 3 times higher than the $\int n_e^2(l) dl$ calculated directly from the hydrocode; the color temperature is two orders of magnitude below the kinetic temperature, and the estimated thermal pressure is one seventh of that from the hydrocode. By $10^5$~yrs, the values have come into better agreement. By $10^6$~yrs, the $\int n_e^2(l) dl$ calculated by this method is an order of magnitude less than the $\int n_e^2(l) dl$ calculated directly from the hydrocode, the color temperature is about $3/2$ the kinetic temperature, and the expected thermal pressure is less than the thermal pressure calculated directly from the hydrocode.\\ \subsection{O VII and O VIII Emission Lines} We are entering an era in which very high resolution X-ray spectrometers will record the signals of individual lines and complexes. The spectra should be complex and difficult to fit with simple models. One approach to the complexity is to extract physical information from specific emission lines. The O~VIII Lyman $\alpha$ line at 653 eV and the O~VII ``triplet'' around 570 eV are good candidates for examination because these lines provide useful measures of the temperature and ionization state, are strong, are not closely clustered with other strong lines, and because model estimates depend on relatively well known atomic constants. The energy range spanned by the O VII ``triplet'' actually contains several emission lines. There are forbidden (1s1s $-$ 1s2s ($^3S_1$)), intercombination (1s1s $-$ 1s2p ($^3P_1$) and 1s1s $-$ 1s2p ($^1P_1$)), resonance (1s1s $-$ 1s2p ($^1P_1$)), and dielectronic recombination satellite (1s2nl$^2 -$ 1s2pnl) lines, as well as a line due to innershell excitation while the oxygen is in its \oxyfive\ state. The Raymond and Smith code combines the strengths of the two intercombination lines and reports them as a single entry at 569 eV. Currently the code does not consider innershell excitation of oxygen in its \oxyfive\ state. Thus, the spectra presented here include four lines, the forbidden (561 eV), merged intercombination (569 eV), satellite (571 eV), and resonance (574 eV). The forbidden, intercombination, and resonance transitions can result from collisions of O$^{+6}$ or recombinations from O$^{+7}$ to O$^{+6}$. In contrast, the dielectronic recombination satellite transition can only result from recombinations from O$^{+7}$ to O$^{+6}$. In collisional equilibrium, collisional excitations of O$^{+6}$, inner shell ionizations of \oxyfive, and recombinations of O$^{+7}$ each play a role. In an underionized plasma, the O$^{+6}$ excited states are populated largely via collisional excitation, but also via inner shell ionizations of \oxyfive. In this case, the resonance line is strongest. In an overionized plasma, the O VII excited states are populated largely via recombinations from O$^{+7}$ to O$^{+6}$, as well as collisional excitations. The $^3S_1$ and $^3P_1$ states have larger statistical weights than the $^1P_1$ state, increasing the probability that the electrons's route to the ground state will be via the forbidden or intercombination transitions. Figure~\ref{fig:ovii} presents the line luminosities of the four O VII lines and the O VIII Lyman alpha (653 eV) line. There is a tradition of plotting the ratio of the O VII forbidden to resonance line luminosities against the the ratio of the O VIII to O VII lines for underionized plasmas, (Vedder {\it{et al.}} (1986), Canizares (1990), and Sanders {\it{et al.}} (1997)) The ratio of the O VII lines is thought to provide an indicator of the ionization state of the plasma (though the temperature also plays a role). The ratio of the O VIII to O VII luminosities is thought to provide an indicator of the temperature, with the provisos that the ionization time plays a role in underionized plasma, under and overionized plasmas produce different O VIII to O VII ratios, and above $T {\sim} 2 \times 10^6$~K, the choice of equilibrium, isobaric or isochoric cooling plays a role for cooling plasma. One subtlety is that depending on the spectral resolution, the intercombination and satellite lines may or may not be resolvable from the resonance line. Figure~\ref{fig:oviioviii} is a variation on the standard figure, having the ratio of the O~VII forbidden to resonance plus satellite line luminosities plotted against the ratio of the O~VIII to O~VII resonance plus satellite line luminosities. The satellite line is much weaker than the resonance line; excluding it does not alter the character of the curve. The figure also includes the ratios for a plasma in ionizational equilibrium. Underionized plasmas correspond to the region below this curve and overionized plasmas correspond to the region above the curve. Aside from the first datapoint ($10^4$ yrs), the OVII - OVIII signature of the young SNR accurately describes it as an underionized plasma having a low ionization parameter ($nt$, where t is the age) and a fairly high thermal temperature. A quasi ionizational equilibrium is reached when the kinetic temperature drops down to the ionization temperature at $\sim1 \times 10^5$~yrs (epoch 4). Afterwards, the gas is recombining, and predictably, the OVII forbidden to resonance line ratio rises. It peaks at $10^6$ yrs (epoch 7) before it and the O~VIII to O~VII ratio begin their slow descents. \section{Caveats} Paper I discussed a long list of assumptions and approximations pertaining to the modeling and interpretation of the \carthree, \nitfour, and \oxyfive\ calculations. Additional considerations, which more strongly affect the X-ray emission are presented here along with the most pertinent caveats from Paper I. In the hydrocode, the electron and ion kinetic temperatures are set equal to each other, although during the earliest part of the evolution, the electrons should be cooler than the ions. To find an upper limit on the timescale for the electrons to come into equilibrium with the ions, they will be assumed to equilibrate solely via Coulomb collisions and in the absence of thermal conduction. In this case, the equilibrium timescale is $t_{eq} = 5000 {\rm{yr}} E_{51}^{3/14} n_o^{-4/7}$, where $E_{51}$ is the explosion energy in units of $10^{51}$~ergs (Cox and Anderson 1982). For Remnants A and B, this formula gives an equilibration timescale of 60,000~yrs and for Remnant C, it gives only 10,000~yrs more. Plasma instabilities and electron and ion thermal conduction will bring the particles into equilibrium faster. Compared with the lifetimes of the SNRs, and especially when compared with the lifetimes of their energy conserving phases, this timescale is short. While in effect, the disequilibrium can change the bremsstrahlung power and spectral shape, ionization and recombination rates, and line emission. Some of these changes may have longterm ramifications. One must access a spectral model in order to calculate the radiative cooling rates and spectral features. In this case, the Raymond and Smith (1977, 1993) spectral model was used. Applying others, such as those of Landini and Monsignori Fossi (1990), Mewe {\it{et al}} (1985), Mewe {\it{et al}} (1986), Kaastra (1992), or Masai (1984) could lead to some differences in the detailed spectra, though the differences in predicted broad band count rates should be small. Note that the Raymond and Smith model was primarily designed for calculations of collisional ionizational equilibrium, low energy resolution spectra. Better non collisional ionizational equilibrium emission calculations may require inner-shell ionization and other processes not currently written into the Raymond and Smith code. The spectral code community is strenuously working to improve the modeling and update the atomic constants. Furthermore, because the hydrodynamic and spectral codes do not include cosmic ray acceleration, the predicted spectra do not include X-ray synchrotron emission like that observed toward a few very young remnants (Koyama {\it{et al.}} 1995, Allen {\it{et al.}} 1997). \section{Selected Applications} \subsection{Application to the nearby lower Galactic halo} Significant numbers of X-ray photons are produced above the HI layer of the Galactic Disk. Snowden, {\it{et al.}}'s 1998 analysis found the de-absorbed ROSAT $1/4$~keV surface brightness of the gas above the HI layer (after subtracting the extragalactic flux) to be 400~counts~sec$^{-1}$~arcmin$^{-2}$ for the south and 1150~counts~sec$^{-1}$~arcmin$^{-2}$ for the north. The average northern surface brightness is much more that of the south because the north contains anomalous regions such as Loop I. The high latitude sky is also rich in \oxyfive, \nitfour, and \carthree\ (Shelton 1998, Savage, {\it{et al.}} 1997). Determining the source of the X-ray emission and high stage ions challenges researchers; many physical scenarios are plausible and multiple mechanisms may be at work. One type of contributor is the population of isolated SNRs originating above the Galactic disk. This subsection reports on the X-ray contribution from those SNRs, for the case in which the explosion energy is $10^{51}$~ergs and the halo environment is tepid ($\sim10^4$~K) with a nonthermal pressure (magnetic and cosmic ray combined) of 7200~K cm$^{-3}$. The supernova remnant simulations use $n = 0.01$~cm$^{-3}$, but preliminary estimates show that during their lifetimes, SNRs in somewhat denser media produce similar numbers of soft X-ray photons. Hence the lifetime integrated results for this simulation (Remnant C) may be used to approximate those of SNRs at a range of heights above the Galactic disk. In order to estimate the contributed X-ray surface brightness, the simulation results must be combined with the statistical distribution of isolated supernova progenitors residing at least 160~pc above or below the Galactic disk (to avoid overlap with the region occupied by the Local Bubble and to be above most of the Galactic Disk's HI layer). The progenitor rates are not precisely known and so here I will take two rates as upper and lower estimates. The first set of isolated progenitor rates is that of Ferri\`{e}re (1995). The second is a combination of McKee and Williams (1997) massive star progenitors with Ferri\`ere's Galactocentric radial distribution and her Type 1a rates (See Paper III). With these rates, a population of Remnant C - like SNRs produces an average ROSAT $1/4$~keV surface brightness of 230 to 390~counts~sec$^{-1}$~arcmin$^{-2}$. Given the uncertainties, this compares well with Snowden {\it{et al.'s}} 1998 observationally based estimate of the Southern Galactic halo's surface brightness. The image produced by a population of Remnant C - like SNRs is consistent with the observations of the Galaxy's southern halo. The population of Remnant C - like SNRs would appear as a couple of bright (1000's of counts~sec$^{-1}$~arcmin$^{-2}$), limb brightened regions and dozens of dim (100's of counts~sec$^{-1}$~arcmin$^{-2}$), centrally filled regions scattered across the high latitude sky. The remnants are sufficiently plentiful and longlived as to cover roughly half of the high latitude sky (including overlap of remnants). This compares well with the observations of the southern halo, which can be described as a mottled ``background'' overlayed with scattered bright features. Paper III (Shelton 1999) provides greater detail and additional calculations (such as the SNRs' area coverage, and hot gas volume filling). Paper I shows the a population of Remnant A-like SNRs produces as much \oxyfive, \nitfour, and \carthree\ as is observed via absorption measurements toward stars within the first kiloparsec of the plane. This paper shows that Remnant B and C-like SNRs produce similar or larger quantities of these ions, thus extending the conclusions of Paper I to include these types of remnants. An idea which escaped Paper I is the notion of buoyancy. The hot diffuse gas in the SNRs should rise. As it does, it moves to more tenuous and less pressurized surroundings. Thus, the remnant bubbles will further expand and their gas densities decrease. With lower gas densities, the cooling and recombinations will slow. The remnants will live longer, causing the time integrals of the numbers of \oxyfive, \nitfour, and \carthree\ per SNR to increase. The \carthree\ will be especially effected. Not only should this phenomena increase the estimated number of high stage ions produced by the population of high latitude SNRs, but it should also increase the theoretical estimates of the scale heights and ratio of \carthree\ to \oxyfive\ atoms. \subsection{Application to the Local Hot Bubble} The Local Hot Bubble (LHB), also called the Local Bubble, is a large (diameter $\sim 50$ to 100~pc), diffuse ($n \sim 0.05$ cm$^{-3}$), presumably hot ($T_c = 10^6$~K), X-ray emissive region containing the Sun (Snowden, {\it{et al}} 1990, Warwick, {\it{et al.}} 1993, Cox and Reynolds 1987). The LHB is situated within a cavity called the Local Cavity. In some directions the Local Cavity extends far beyond the X-ray emitting region (Welsh {\it{et al.}} 1994, Snowden {\it{et al.}} 1998). The Local Hot Bubble also contains several parsec-scale clouds and complexes of clouds, including the Local Cloud complex which, as expected from its name, surrounds the sun (Lallement 1998). Although its existance has long been suspected, understanding the Local Hot Bubble's origin has proven to be a very challenging enterprise\footnote{The reader will find a fascinating set of readings on this topic in ``The Local Bubble and Beyond'', 1998, edited by Breitschwerdt, Freyberg, and Tr\"umper}. The observations impose a host of difficult constraints on LHB modeling. Snowden {\it{et al}} (1998) determined that the local region's $1/4$~keV surface brightness is $\sim250$ to $\sim820 \times 10^{-6}$ counts s$^{-1}$ arcmin$^{-2}$ and the R1$/$R2 color temperature is $\sim10^6$~K. The interior of the Local Bubble appears to produce little $3/4$~keV emission (Snowden, McCammon, and Verter 1993) and no evidence has been found for an X-ray bright edge in either the $1/4$ or $3/4$~keV band. Furthermore, high spectral resolution observations by the Diffuse X-ray Spectrometer (Sanders {\it{et al.}} 1998) force constraints on any detailed spectral model. The average \oxyfive\ column density on a sightline from the earth though the Local Cloud complex and the Local Hot Bubble, is $\sim1.6 \times 10^{13}$ \oxyfive\ cm$^{-2}$, with a velocity centroid near 0~km~s$^{-1}$ (Shelton \& Cox 1994). In order to explain the high column density of approximately stationary \oxyfive, either the edge of the hot, highly ionized region is nearly stationary, most of the \oxyfive\ is located within the bubble's stationary interior, or the slow-moving Local Cloud is producing the observed \oxyfive\ and either the Local Hot Bubble's \oxyfive\ component is moving faster than {\it{Copernicus}}'s velocity range ($\sim \pm 100$~km~s$^{-1}$) or is negligibly small. Simultaneously satisfying all of constraints is difficult. The Remnant C simulations may be pointing the way to a promising portion of parameter space. When Remnant C is several million years old, it has an X-ray emitting bubble of about the appropriate size, emits copious $1/4$~keV X-rays, is not limb brightened in X-rays, is fairly dim in $3/4$~keV X-rays, and contains plentiful quantities of stationary \oxyfive\ (the \oxyfive\ column density for a sightline looking out from the center is $\sim 2.7 \times 10^{13}$ cm$^{-2}$). Like in the Breitschwerdt and Schmutzler (1994) model, the gas has cooled from a much hotter temperature and contains overionized atoms. Is the proposed physical scenario reasonable -- is the ambient medium relatively diffuse and nonthermal pressure dominated? Yes. The Local Bubble is situated inside the Local Cavity, a low density region. If the Local Cavity predates the Local Bubble, then the Local Bubble has been evolving in a rarefied medium. Furthermore, if the Local Cavity gas is tepid and the total pressure is typical of the total pressure in the Galactic midplane, then the pressure would be largely nonthermal. Thus, the proposed nature of the environment is reasonable. Some adjustments and further spectral testing of the model are needed, of course. For example, the local region is expected to have a higher total pressure than Remnant C and so the author is currently working on higher pressure models. \section{Summary} It is difficult to do justice while condensing the life histories of the three SNRs into a mere paragraph or two. With that said, here is an attempt. The gross features of their life histories are similar. During their energy conserving phases, each remnant is hot, highly pressured and rapidly expanding. The relatively dense gas swept up by the shockfront contributes nearly all of the X-ray photons, causing the remnants to appear edge brightened. The remnants are also very luminous, with surface brightnesses of thousands of $1/4$~keV counts s$^{-1}$ arcmin$^{-2}$. The ionization timescales significantly lag the dynamic timescales, initially causing the gas to be drasticly underionized. Between 100,000 and 250,000 years, the remnants enter their radiative phases. The shockfront of each remnant slows to the extent that it is no longer able to dramaticly heat the gas it encounters and a cool shell develops between the shockfront and the hot bubble. Even without a very hot outer edge, the hot bubbles continue to emit X-rays, but with lesser luminosities and without the strong edge brightening of the young remnants. Remnants B and C are an order of magnitude brighter than Remnant A. Compared with Remnant A, their larger ambient nonthermal pressures better compress the SNR bubbles, elevating the bubbles' temperatures and densities, and hence X-ray luminosities. Remnants B and C have surface brightnesses of hundreds of $1/4$~keV counts s$^{-1}$ arcmin$^{-2}$ during their old evolutionary phases. (Remnants B and C produce more $1/4$~keV photons during their old evolutionary phases than during their youth.) The recombination timescale lags the cooling timescale, causing the atoms to become overionized around the time that the cool shell forms and to remain overionized until the hot bubbles disappear, some 12 million or more years later. The simulation parameters, lifetimes, sizes, $1/4$~keV, $3/4$~keV output, and the numbers of \oxyfive, \nitfour, and \carthree\ atoms are compiled into Table 1. An effort was made to understand how such remnants would appear if observed with modern or future facilities and analyzed with common techniques which assume that the ions are in collisional equilibrium with respect to the gas temperature. The ``color temperature'' was found from the ratios of the ROSAT PSPC R1, R2, and R4 band countrates, the electron density and thermal pressures were calculated from the color temperatures and surface brightnesses, and the O VII and O VIII ratios were compared with those of collisional equilibrium plasma. Remnant A was used as an example. When seen through the lens of these techniques, the young remnant appears, denser, and less pressurized than it is and the old remnant appears hotter, more rarefied, and less pressurized than it is. The O VII and O VIII diagnostics accurately identify the plasma in the young remnant as being underionized and the plasma and the old remnant as containing overionized, recombining gas. The results were combined with the Galaxy's progenitor statistics in order to compare with the observed $1/4$~keV surface brightness of the Galaxy's southern halo, and analyzed for clues as to the origin of the Local Hot Bubble. In the case of the Galactic halo, the number of type O and B runaway stars and type Ia progenitors which explode a few hundred parsecs from the disk is remarkably large, as is the number of soft X-ray photons emitted during Remnant C's lifetime. Combining these values shows that a population of Remnant C-like remnants could explain roughly 200 to 400 counts s$^{-1}$ arcmin$^{-2}$ of the observed 400 counts s$^{-1}$ arcmin$^{-2}$ in the ROSAT PSPC $1/4$~keV band. These SNRs can also explain the observed spatial emission pattern which consists of a few bright regions, a dim mottled background, and about half of the high latitude southern sky having nothing but the local and extragalactic fluxes. In the case of the Local Hot Bubble, a one million year old Remnant C comes within sight of explaining the Local Hot Bubble's size, $1/4$~keV surface brightness, lack of limb brightening, and column density of $\sim0$~km~s$^{-1}$ bulk velocity \oxyfive\ ions. \vspace{1cm} \noindent{\bf{Acknowledgements}} While at the University of Wisconsin, Department of Physics, the author received invaluable assistance from Don Cox and thanks him for sharing his hydrocode expertise, astrophysics intuition, and good cheer. While at the NASA/Goddard Space Flight Center, Laboratory for High Energy Astrophysics, the author received irreplaceable assistance from Rob Petre, Steve Snowden, and Kip Kuntz, and thanks them for sharing their X-ray analysis expertise, bibliographic memories, and moral support. This work was supported under NASA grant NAG5-3155 and by an award from the Wisconsin Space Grant Consortium while the author was at the University of Wisconsin -- Madison, and by a grant from the National Research Council while the author was at the NASA/Goddard Space Flight Center, Laboratory for High Energy Astrophysics. \clearpage \begin{table} \begin{center} \begin{tabular}{|c|c|c|c|} \hline \hline & Remnant A & Remnant B & Remnant C \\ \hline \hline \ ambient $n$ & 0.01 cm$^{-3}$ & 0.01 cm$^{-3}$ & 0.01 cm$^{-3}$ \\ \hline \ ambient $T$ & $1.0 \times 10^4$~K & $1.0 \times 10^4$~K & $1.0 \times 10^4$~K \\ \hline \ ambient $B$ & 2.5 $\mu$G & 5 $\mu$G & 5 $\mu$G \\ \hline \ explosion energy & $0.5 \times 10^{51}$ ergs & $0.5 \times 10^{51}$ ergs & $1.0 \times 10^{51}$ ergs \\ \hline \ lifetime & 16 million yrs & 12 million yrs & 14 million yrs \\ \hline \ maximum radius & 140 pc & 110 pc & 140 pc \\ \hline \ $1/4$~keV output & $5.8 \times 10^{59}$ counts cm$^{2}$ & $1.4 \times 10^{60}$ counts cm$^{2}$ & $3.5 \times 10^{60}$ counts cm$^{2}$ \\ \hline \ $3/4$~keV output & $1.2 \times 10^{58}$ counts cm$^{2}$ & $1.1 \times 10^{58}$ counts cm$^{2}$ & $2.8 \times 10^{58}$ counts cm$^{2}$ \\ \hline \ \oxyfive & $7.8 \times 10^{69}$ \oxyfive\ sec & $6.0 \times 10^{69}$ \oxyfive\ sec & $1.2 \times 10^{70}$ \oxyfive\ sec \\ \hline \ \nitfour & $7.2 \times 10^{68}$ \nitfour\ sec & $4.6 \times 10^{68}$ \nitfour\ sec & $8.9 \times 10^{68}$ \nitfour\ sec \\ \hline \ \carthree & $1.6 \times 10^{69}$ \carthree\ sec & $1.1 \times 10^{69}$ \carthree\ sec & $2.4 \times 10^{69}$ \carthree\ sec \\ \hline \hline \end{tabular} \end{center} \caption[] {Simulation parameters, lifetimes, maximum radius of the hot bubbles, and the time integrated numbers of soft X-ray counts and high stage ions for Remnants A, B, and C.} \label{table1} \end{table} \clearpage
\section{INTRODUCTION} Clusters of galaxies remain a valuable venue for study. Morphology and evolution of galaxies differ in clusters from the field (e.g., Hubble \& Humason 1931, Oemler 1974, Dressler et al. 1997). As the nearest large cluster, the Virgo cluster offers the opportunity for studying details of environmental effects on galaxy kinematics. Virgo studies began over two centuries ago, when Messier (1784) noted the concentration of nebulae ($``$sans etoile") in Virgo. Yet 150 years passed before Shapley and Ames (1932 and references therein) initiated a large-scale study of the properties of galaxies in the cluster region. The most extensive study of galaxies in the Virgo cluster is due to Binggeli, Sandage and Tammann (1985, 1987, BST; called VCC for Virgo Cluster Catalogue), who cataloged 2100 galaxies in the Virgo region, and identified 1277 as members. Virgo galaxies have also been studied with H$\alpha$ images (Kennicutt \& Kent 1983; Koopmann \& Kenney 1998, 1999), integrated HI profiles (Giovanelli \& Haynes 1983; Helou, Hoffman, \& Salpeter 1984; Warmels 1988; Cayatte et al. 1990), CO (Kenney \& Young 1989), radio continuum (Kotanyi 1980), and x-ray (Fabbiano, Kim, \& Trinchieri 1992). Surveys of Virgo galaxy properties and comparisons with non-cluster galaxies provide evidence that within the core ($\approx$6$^{\rm o}$) of the cluster, galaxies are subjected to environmental effects, such as galaxy-cluster and galaxy-galaxy tidal interactions, mergers, gas accretion, collisions, and gas stripping by the intracluster medium (ICM). Many processes operate in clusters (Oemler 1992; Valluri 1993; Moore et al. 1996; Kenney \& Koopmann 1999; Barton et al. 1999), although we do not yet know which are dominant, nor the precise evolutionary effects on the disturbed galaxies. The Virgo cluster is not yet in overall dynamical equilibrium, although some galaxy populations within it are more relaxed than others. There are significant sub-structures in the cluster, further evidence of a lack of dynamical equilibrium (BST 1985, 1987, Huchra 1885). While the ellipticals and lenticulars have a Gaussian distribution of line-of-sight (l-o-s) systemic velocities indicating that they may be in the process of establishing a relaxed population, the spirals have a complex, multi-peaked distribution (Huchra 1985) suggesting that they may be clumped into discrete subunits (Sandage, Tammann, \& Binggeli (1985). Many of them may be presently arriving into the cluster on elongated orbits (Huchra 1985). Tully \& Shaya (1984) employed a flow model for the local supercluster, offering evidence that many spirals in the general region of the Virgo cluster (including the Milky Way) are on orbits which will carry them into the Virgo cluster, some within a fraction of a Gyr. Internal galaxy kinematics are sensitive to tidal interactions, yet there are few previous measures of gravitational interactions within a single cluster population. Hence, some years ago we embarked on a program to obtain CCD optical long-slit spectra for disk galaxies in the Virgo cluster. From these high spectral resolution emission line spectra, we determine rotation curves and study the ionized gas kinematics for S0$-$Sm galaxies. Our data are well-suited to study irregularities in velocity fields, which trace disturbances in the galaxies' gravitational potentials. Non-axisymmetric and non-bisymmetric features in the velocity fields are likely due to tidal events, whether high or low velocity encounters/collisions, mergers, or accretion events. Tidal interactions produce a non-bisymmetric disturbance when (m/M)(r/R)$^3$ between the perturbee and the perturber are large; (m/M) is the ratio of the masses, and (r/R) is the ratio of the radius of the perturber to the distances between the two galaxies. Thus galaxy-cluster tidal effects and near galaxy-galaxy tidal encounters between galaxies with nearly similar masses are more likely to produce non-bisymmetric disturbances. We think that these types of disturbances are responsible for most kinematic irregularities we discuss in this paper. Since m=2 structure is common in galaxies, the m=2 mode must be fairly long-lived, and cannot unambiguously tell us about recent interactions. In contrast, most other non-axisymmetric kinematic disturbances smooth out in several rotation periods. Hence disturbed kinematics identify galaxies in which gravitational disturbances have taken place within the last billion years. \section{OBSERVATIONS} In this paper, we present major axis velocity data for 89 Virgo galaxies (plus one galaxy with only a minor axis spectrum), rotation curves for 81, and minor axis velocities for 42. Galaxy properties and a journal of observations are compiled in Table 1. For an additional 14 galaxies, most of early type located in the central dense Markarian chain, emission is undetected or possibly detected only in the nucleus (Table 1 notes). The observed galaxies (mostly spirals) are fairly well distributed across the cluster and its southern extension (Fig. 1), and include Hubble types E7/S0 to Spec brighter than about 14.5 (Fig. 2). The sample contains 72 NGC galaxies, 13 UGC (Nilson 1973; UGC), and 5 blue compact objects (BST 1985). Most galaxies in Table 1 are identified as members of the Virgo cluster. Seventy are classified certain members and 11 are possible members (BST). Nine others have more ambiguous designations. NGC 4165 (V$_{\rm o}$=1882 km s$^{-1}$) and UGC 7181 (276 km s$^{-1}$) were called background by BST. Although their low rotation velocities imply faint absolute magnitudes, they plot within the scatter of Virgo galaxies on a Tully-Fisher plot. Seven others (NGC 4064, 4651, 4710, 4713, 4808, 4866, and UGC 7249) are located just outside the BST fields. Their velocities range from 637 to 1114 km s$^{-1}$, except NGC 4866 with V$_{\rm o}$=1994 km s$^{-1}$, and suggest that these are in the outskirts of the cluster. Contamination from non-cluster galaxies is expected to be minimal. There are 42 spirals classified Sa through Scd in the VCC, with B(t) $\le$ 13, i $\le$ 40$^{\rm o}$. We have observed 39 (93\%) of these. Including galaxies classified ``possible" members, with the same Hubble type and inclination limits, we have observed 44 of 56 (79\%). Corresponding numbers (members and possible members) for B(t)$\le$14 are 60 of 91 observed (66\%) and for B(t)$\le$15 are 64 of 124 observed (52\%). Fractions are higher if we limit the sample to ``members". Of the remaining galaxies in our sample, 24 are classified other than Sa-Scd, 5 are outside the VCC area in VCC, and a few have i $\le$ 40$^{\rm o}$. About two-thirds of our sample galaxies have no previously published optical velocities extending beyond the nucleus. Twenty-three have ionized gas rotation velocities from Sperandio et al. (1995), and a few more from Sofue et al. (1998). Resolved HI kinematic data exist for fewer than 25 galaxies observed here (Section 3). Spectra centered near H$\alpha$ were obtained with the Palomar 200-inch and 60-inch telescopes, and Kitt Peak 4-m telescope (Tables 1, 2) for 104 galaxies in the Virgo region. Spectral resolution was $\approx$2 or 3 pixels, and hence near 1\AA~(Table 2). Position angles of the major axes, determined from available images, are generally within a degree or two of those in the UGC. For a few galaxies with uncertain position angles, spectra were taken in several position angles. Slit widths of 1.3$''$ to 2$''$ were chosen to match the instrumental resolution and seeing. A few large galaxies were observed in two frames. Major axis spectra are of high quality, but minor axis spectra taken during intervals of poor weather or bright moon are of lower quality. For each telescope plus CCD detector combination, normal processing procedures were followed. In spectra with a strong galaxy continuum, a mean continuum (of the near H$\alpha$ region) was removed before measuring. Details of the velocity measuring procedure are given elsewhere (Rubin, Hunter, \& Ford 1991). Nightsky OH lines are used for the two-dimensional wavelength calibration and velocity zero-point. At successive distances from the nucleus along the major axis, velocities of both H$\alpha$ and [NII] (occasionally [SII]) are found from the centroid of each emission line. This procedure produces a good estimate of the l-o-s gas velocity at a given position, except near the central few arcseconds of those galaxies with steep velocity gradients. Several galaxies of kinematic interest have previously been discussed in detail: NGC 4550 for its counterrotating stellar disks (Rubin, Graham, \& Kenney 1992); the highly disturbed spiral NGC~4438 and its apparent companion NGC 4435, which may have experienced a high-velocity collision (Kenney et al. 1995); the peculiar merger remnant NGC 4424 (Kenney et al. 1996); the spiral NGC~4522, which is experiencing ICM-ISM stripping (Kenney \& Koopmann, 1999); and galaxies with kinematically distinct circumnuclear disks (Rubin, Kenney, \& Young 1997). Major and minor axis velocities are included here; velocities in other position angles are contained in the references. Heliocentric systemic velocity as a function of galaxy right ascension is shown in Fig. 3. The mean heliocentric cluster velocity, 1150$\pm$51 km s$^{-1}$ (Huchra 1988) or 1050$\pm$35 km s$^{-1}$ (Binggeli, Popescu, \& Tammann 1993), is poorly defined by our predominantly spiral sample, a recurrent problem with Virgo studies. The high systemic velocity dispersion is apparent (Fig. 3) throughout much of the cluster. We adopt M87 as the cluster center, based on ROSAT x-ray observations (Bohringer et al. 1994) \section{ROTATION VELOCITIES AND CENTRAL VELOCITIES} Major axis velocities as a function of radius are shown in Fig. 4. Mean velocities, for all measures within a small radial bin, are shown along with their 1$\sigma$ errors. For 10 galaxies observed in 1987 (U7181, N4165, U7259, N4237, N4383, N4420, U7590, U7676, N4584, U7932), the position angle of the slit is known, but the algorithm for determining the direction (e.g., NW or SE) of one end of the slit has been lost. For these, no orientations are included in Fig. 4. Note that the outermost point for NGC 4519 comes from a companion galaxy, 2.2$'$ NW along the major axis (see images, Sandage \& Bedke 1994; Koopmann, Kenney, \& Young 1999). Minor axis velocities with respect to the galaxy systemic velocity are shown in Fig. 5. For NGC 4651, the observed position angle is displaced 9$^{\rm o}$ from the adopted minor axis. For UGC 7623 we have only a minor axis spectrum. We discuss the forms and amplitudes of the rotation curves in Sections 5 and 6 below. Representative nuclear spectra (Fig. 6) illustrate the range in emission line properties, for galaxies ranging from types Amorphous to Sb to dE?, and blue magnitudes from 10.8 to 18. NGC 4548 is a Liner which shows the less common intensity ratio H$\alpha$/[NII] $\le$ 1; NGC 4639 is a broad lined Sy 1, while NGC 4388 is a Sy 1.9. NGC 4694 exhibits a typical emission line spectrum, with H$\alpha$ intensity greater than [NII]; and VCC 1941 has H$\alpha$/[NII] $\ge$ 10. Its strong [SII] lines, relative to virtually absent [NII] lines, identify it as a very low luminosity object (Rubin, Ford, \& Whitmore 1984), consistent with its absolute magnitude M = $-$13 at the cluster distance. The four NGC objects are classified as members; V1941, with no previously known velocity, was called a possible member (BST). Its velocity here, V$_{\rm o}$=1213 km s$^{-1}$, supports its membership. Resolved HI observations (as opposed to integrated HI profiles) exist for only a small fraction of Virgo galaxies. Guhathakurta et al. (1988), Cayatte et al. (1990), and Warmels (1988) have published HI velocity fields for about 30 Virgo galaxies in common with our sample, but most are at low resolution, and rotation curves are presented only for a subset. An early comparison (Rubin et al. 1989) of 8 Virgo galaxies (then in common) showed agreement between optical and HI velocities only for the 3 galaxies observed at 15$''$ resolution. Four Virgo galaxies have recent HI observations: N4254, (Phookun, Vogel, \& Mundy, 1993), NGC 4321 (Knapen et al. 1993), N4532 (Hoffman et al, 1999) , and N4654 (Phookun, \& Mundy, 1995). Only NGC 4532 was not on the Guhathakurta et al. list. It is disappointing that {\it stellar} rotation velocities exist for only 9 of these galaxies: NGC 4374 (Davies \& Birkinshaw 1988), N4435 (Simien \& Prugniel 1997), N4450 and N4569 (Filmore, Boroson, \& Dressler 1986), 4459 (Peterson 1978), N4477 (Jarvis et al. 1988), N4550 (Rubin, Graham, \& Kenney 1992, Rix et al. 1992), N4579 (Palacios et al. 1997), and N4698 (Corsini et al. 1997), and velocity dispersions for 6. Only for NGC 4435, N4550, and N4698 do measured stellar velocities extend as far (or farther) then our gas velocities. For the others, gas velocities extend beyond stellar measures by factors ranging from 2 to over 10. We cannot draw conclusions from this limited material. Values of V$_{\rm o}$, the heliocentric systemic velocity for 90 galaxies, are listed in Table 1. For galaxies with only nuclear emission velocities, the measured (centroid) nuclear velocity is accepted as V$_{\rm o}$. For galaxies with emission beyond the nucleus, the choice of systemic velocity takes account of the velocity at the strongest continuum, and the outer symmetry of the resulting curve. The uncertainty of a single emission line velocity is generally a few km s$^{-1}$, but larger near the nucleus of a galaxy with a steep nuclear velocity gradient, or where extinction complicates the choice of center. Uncertainty in V$_{\rm o}$ is generally of order 10 km s$^{-1}$, increasing to 15 km s$^{-1}$ for galaxies with complex velocity patterns. For 74 galaxies with previously published HI systemic velocities (de Vaucouleurs et al. 1991; RC3), the agreement is excellent, with mean value (V$_{\rm o}-$V$_{21}$) = $-$2.0 km s$^{-1}$, and mean of the absolute values of (V$_{\rm o}-$V$_{21}$) = 14 km s$^{-1}$. Five additional galaxies have RC3 velocities, but we follow Roberts et al. (1991) and do not accept as detections NGC 4293, 4374, 4506, and 4550. For example, the 21-cm velocity attributed in RC3 to NGC 4374 is an old measure, unsubstantiated by new, more sensitive observations (Huchtmeier \& Richter 1989); it is 181 km s$^{-1}$ lower than our emission line velocities. The fifth galaxy, NGC 4526, is undetected in HI (Kumar \& Thonnard 1983), but detected in CO (Sage \& Wrobel 1989), with a broad velocity range. The agreement of our velocities with RC3 {\it optical} velocities is sometimes poor, (e.g., V$_{\rm o}-$V$_{old}$ $\ge$ 300 km s$^{-1}$ in 2 cases), due to large uncertainties in older values. Rotation velocities, (V$_{\rm obs}-$V$_{\rm o}$)/sin $i$, as a function of nuclear distances are shown (Fig. 7) for 77 galaxies; the galaxies are arranged according to angular extent of measured emission. We assume circular, planar orbits. To transform from angular to linear distances (upper axis, Fig. 7), we adopt m$-$M=31.0, so the cluster distance is 15.8 Mpc (Jacoby et al. 1992; Freedman et al. 1994) and 1$''$=77 pc, but this choice is irrelevant to our discussion. Major axis rotation velocities for all galaxies, and velocities along minor axes and other observed position angles, are available in ASCII form from the AJ data bank. The first page of data is shown in Table 3. For a few galaxies with uncertain inclinations, we set i=xx in Table 1, and only (V$_{\rm obs}-$V$_{\rm o}$) is tabulated in the file. As our paper was nearing completion, the spectacular HST long slit (STIS) spectrum of NGC 4374 (M84, E1) was obtained by Bower et al. (1998). We compare with some amusement our measured ground based velocities with the HST measures (Fig. 8). Although slit widths differ by a factor of 10 (2.0$''$ vs. 0.2$''$), and exposures correspondingly (600s vs. $\approx$5000s), except for the all-important sub-arcsecond velocity peak, the agreement is satisfactory. This comparison supports the maxim that a factor of ten improvement leads to discovery. \section{EXTENT OF MEASURED EMISSION} Success in detecting extended emission depends on the instrument, the weather, the integration time, and the galaxy. For our spectra, the combination of relatively high spectral resolution and long exposures results in a relatively high S/N, especially for the nuclear and HII regions (Fig. 6). Hence, we believe that the measured radial extent depends principally on the morphological type and the environmental history of each galaxy. The extent of measured emission as a function of Hubble type is illustrated in Fig. 9. The distance of the farthest measured point along the major axis, r$_{f}$, relative to the isophotal radius R$_{25}$ (Table 1), ranges from a low of 0.02 for NGC 4374 (E1) to 1.40 for UGC 7590 (Sbc). As expected, the median value increases, from r$_{f}$/R$_{25} \approx$ 0.3 for S0 galaxies, to $\approx$ 0.7 for Sb and Sc. Note, however, that the assigned Hubble type of some galaxies in Virgo is misleading, since many galaxies classified Sa are actually disk systems with small bulges and low SFRs (Koopmann \& Kenney 1998). This may increase the scatter in Fig. 9. Galaxies with no detected emission, or only possible nuclear emission (Table 1 notes) are included in the plot. Rotation curve forms are discussed in Section 5 below. To illustrate the combined effects of Hubble type and distance from the cluster core (Fig. 10), we show the extent of detected emission as a function of projected galaxy distance from M87. For E and S0 galaxies, emission extents vary from 0 to almost 0.4R$_{25}$, with little dependence upon distance from M87. In contrast, Sa spirals show a large variation with core distance. At 1$^{\rm o}- 3^{\rm o}$ from M87, emission extent varies from 0 to 0.7R$_{25}$ (mean=0.25R$_{25}$). In the 3$^{\rm o}- 6^{\rm o}$ zone, emission extent increases to 0.2$-$1.0R$_{25}$ (mean=0.55R$_{25}$). For types Sb-Sd, the emission extents are lower within 2$^{\rm o}$ of M87, and higher, between 3$^{\rm o}$and 6$^{\rm o}$ from M87. Thus the well known property of spirals near the Virgo core to have stripped outer neutral gas disks (Haynes, Giovanelli, \& Chincarini 1984; Warmels 1988; Cayatte et al. 1990) is seen also in the ionized gas in many Virgo spirals (Section 5 below; Koopmann \& Kenney 1998, 1999). \section{THE ROTATION CURVES: FORMS} From these relatively high accuracy rotation curves for Virgo spirals, there are three properties which we can study. The simplest procedure is to classify the form of each rotation curve as Regular or Disturbed, and to examine its relation to other galaxy parameters. This is the primary goal of this paper. A more difficult parameter to evaluate is Vmax; we do this in the following section. The most difficult parameter to determine is M, the absolute magnitude of each galaxy, via Vmax and Tully-Fisher relation calibrators; we do not attempt this here. As expected for galaxies in regions of high density, many Virgo cluster galaxies exhibit complex velocity curves. Abnormalities consist generally of: asymmetrical rotation velocities on the two sides of the major axis, falling outer velocities, inner velocity peculiarities, dips in the rotation velocities at intermediate radii, or velocities near zero for the near nuclear gas. Although forms are classified by a comparison with the synthetic rotation curves (SRC) from our earlier studies (Thonnard \& Rubin 1981; Rubin et al. 1985), Disturbed forms are easily identified from knowledge of what normal rotation curves looks like. We review the formation of SRC in Section 6. Examples of velocity curves are shown in Fig. 11. NGC 4419 has a normal rotation pattern for a high luminosity Sa. In contrast, the starburst galaxy NGC 4383 (Amorphous) has an unusually shallow nuclear velocity rise. The H$\alpha$+[NII] image of NGC~4383 shows filaments of ionized gas, suggesting significant non-circular gas motions due to the starburst (Koopmann et al. 1999). The asymmetrical outer major axis velocities in NGC 4567 (Sc) are probably tidally induced, rising steeply only on the NE where its close companion, NGC 4568, is superposed. NGC 4498 has a relatively normal velocity pattern on the NW but not on the SE; it has an irregular morphology with the outer galaxy off-center and warped (Koopmann et al. 1999), suggesting a strong gravitational interaction. For 81 galaxies, we have classified the major axis rotation curve form as Regular (38) or Disturbed (43: Table 4, Fig. 12). We attempt a quasi-quantitative measure of the irregularities, and call a rotation curve Disturbed if we detect a distortion of order 15\% or larger in velocity, extending over 1 kpc or more, in comparison with the SRC. Minor axis velocities help with difficult decisions. If only the outermost velocity is discrepant, the rotation curve is not classified falling or disturbed, for outermost points are notoriously difficult to measure. Rotation curve forms are not assigned for 5 BCD galaxies, and 3 others with few measured points. With this classification, we make the following conclusions. 1. {\it One-half of the (mostly spiral) Virgo cluster galaxies we observed have rotation patterns called Regular, while about 50\% exhibit more complex internal kinematics, which we label Disturbed}. Abnormalities include asymmetrical rotation velocities on the two sides of the major axis, falling outer velocities, inner velocity peculiarities, dips in the rotation velocities at intermediate radii, or velocities all near zero at small radii. For example, the Sc galaxies NGC~4567 and NGC~4568 (sometimes called the Siamese Twins), an apparent pair with similar l-o-s velocities, both show signs of kinematic disturbance. Although NGC~4567 is significantly disturbed, the larger NGC~4568 shows peculiar velocities only in the outermost data points, and thus is classified as Regular by our criteria. NGC~4647 (Sc) and NGC~4649 (E) form an apparent pair with similar velocities. While NGC~4647 exhibits an asymmetric H$\alpha$ morphology suggesting a tidal interaction (Koopmann \& Kenney 1999), its kinematics appear nearly normal. Enough time may not have elapsed for the kinematic disturbance to be maximally manifest. The largest kinematic disturbances often appear after closest approach (Toomre \& Toomre 1972, Moore et al. 1997); galaxies with mild disturbances may be at early stages of their interactions. At the opposite end of the interaction time scale are old encounters. NGC~4550 and NGC~4698, with counterrotating or misaligned stellar components, have fairly normal rotation curves. Each merger probably occured a sufficiently long time ago that the velocity fields have become regular. Complex rotation patterns can arise from galaxy-galaxy interactions. Self-consistent N-body models which explore the first pass of two disk galaxies (Barton, et al. 1999) produce rotation curves with bumps and dips in the mid regions, and outer portions which rise or fall differently on the two sides of the major axis, much like those we observe. But kinematic disturbances from galaxy-cluster interactions might also produce similar effects. Models of disk galaxies falling for the first time into the cluster mean tidal field show (Valluri 1993) show that m=1 (warp) and m=2 (bar and spiral arms) perturbations result. Although we do not attempt to distinguish galaxy-galaxy from galaxy-cluster effects, it is noteworthy that eight of our sample galaxies are predicted to fall into Virgo (Tully and Shaya 1984) with time scales ranging from 0.1 to 2.2 Gyr (median 0.5 Gyr), and seven of these have rotation forms classified as Disturbed. Of these eight, 2 are called cluster members, 4 possible members, and 2 are outside the VCC boundaries (BST). Gravitationally disturbed galaxies will return to regular, axisymmetric or bisymmetric kinematics in a few rotation periods, $\simeq$1 Gyr. Consequently, significant kinematic disturbances must occur at least once per 1 Gyr for galaxies in Virgo. Most of these galaxies have likely had several significant tidal encounters during their lifetime. This conclusion is broadly consistent with the results of simulations which show frequent galaxy-galaxy tidal encounters in clusters (Moore et al. 1996, 1997). While there are few statistics on the fraction of {\it field} spirals with disturbed rotation curves, the Virgo set appears unlike those observed in (nominally) field spirals. The galaxies we studied earlier (Rubin et al. 1980, 1982, 1985) were chosen to have no near neighbors, and to have relatively normal, unbarred morphology. A reexamination of their rotation curves reveals that most (74\%) of them, 11 of 14 Sa, 14 of 22 Sb, and 16 of 21 Sc, are {\it very} regular. The remaining curves have minimal disturbances. Only two or three show the large scale disturbances we see in some of the Virgo galaxies. Hence we feel secure in attributing the kinematic distortions to the cluster environment. Other statistics of kinematically disturbed spirals are fragmentary. Based only on {\it morphology} in the visual, Zaritsky and Rix, (1997) and infrared, Odewahn (1996), the fraction of disturbed field spirals has been estimated from 15\% to 50\%. From studies of non-symmetrical integrated HI profiles, Haynes et al. (1998) estimated the fraction of disturbed galaxies to be high, near 50\%. However, HI profiles arise from a integration of the velocity and the HI distributions; resolved HI observations offer more direct information. A study by Swaters et al. (1999) based on kinematically resolved HI observations of lopsided galaxies discusses data for only two such galaxies, but also estimates the disturbed fraction to be at least 50\%. The fraction of galaxies which are classified as disturbed in a given study depends on the criteria for disturbance and the range of galactocentric radii considered. Since these vary among the studies cited, we cannot directly compare our disturbed fraction with others. 2. {\it There is no strong correlation of rotation curve complexity with Hubble type, with galaxy luminosity, or with HI deficiency.} Hubble types are fairly equally distributed among galaxies with Regular and Disturbed rotation forms, (Fig 9, except Sb). Both samples contain mostly Sc's, and equivalent numbers of S0$-$Sa. Luminosities are also similar among both samples. The median apparent magnitudes, B$_{t}^{\rm o}$=11.90 (Disturbed) and B$_{t}^{\rm o}$=11.93 (Regular), are similar, and correspond to M$_{B}$=$-$19.1 at the adopted distance. There is no strong correlation of HI deficiency parameter (Giovanelli \& Haynes 1985) with rotation curve forms (Fig. 13). A comparison of this parameter for Regular and Disturbed rotation galaxies clearly shows that Disturbed galaxies are not preferentially HI-deficient. This indicates that the mechanism which causes HI deficiency, presumably ICM-ISM interactions, is different from the mechanism which causes kinematic disturbances, presumably tidal interactions. In contrast, there is a small (and marginally significant) excess of Disturbed rotation galaxies with normal HI emission (i.e., low HI deficiency). This small excess might arise if is easier to measure disturbed gas kinematics in galaxies with extended gas disks (see 5. below). 3. {\it There is no clear correlation of rotation curve classification with local galaxy density}, as tabulated by Tully (1988). There are both Regular (N4647 and 4808) and Disturbed (N4536) galaxies in the regions of lowest (0.21-0.46 gal/Mpc$^{3}$) local density, as well as both Regular (N4456) and Disturbed (N4477) galaxies in regions of the highest local density (4.06 gal/Mpc$^{3}$). NGC 4647 is a good illustration of the difficulties of sorting out local density and rotation curve form. With a Regular rotation curve which shows little or no sign of an interaction, in a region called low density, it has a close companion on the sky, NGC 4649 (S0, which we have not observed). But the Tully density parameter is not especially sensitive to pairs. Furthermore, the high relative velocities in many galaxy-galaxy interactions makes it likely that any external culprit responsible has long since moved away or merged, so that present local density is only weakly related to earlier density when the interaction occurred. An examination of the VCC suggests that low velocity interactions within gravitationally bound pairs or groups is not the major cause of the kinematic disturbances. Of the 11 sample galaxies which are likely to be in bound pairs (e.g., companions within 10$'$ = 46 kpc at the adopted distance, velocity differences of less than 300 km $^{-1}$, and apparent brightness differences of less than 3 magnitudes), 4 are classified as Disturbed, 7 as Regular. These disturbed galaxies in pairs likely owe their disturbances to low-velocity tidal interactions with their companions. While there are undoubtedly more physically bound pairs and groups in Virgo than these 11 (Ferguson 1992), relaxing the identification criteria reduces the probability of finding true bound systems. A comparison of the nearest neighbors on the sky and in velocity shows no significant difference in the probability of belonging to a group or similar-mass pair for the Disturbed and Regular galaxies. This suggests that low velocity galaxy-galaxy interactions are not the only cause of kinematic distortions. High velocity tidal interactions (Moore et al. 1997) or galaxy-cluster interactions (Valluri 1993) are probably responsible for a significant fraction of the disturbances. 4. {\it The distribution of galaxy systemic velocities is different for galaxies with Regular rotation curves and galaxies with Disturbed rotation curves}. We show in Fig. 14 the distribution of heliocentric radial velocities for each class (Table 4). While galaxies with Regular rotation forms have a relatively flat distribution of observed velocities ranging from $-$300 km s$^{-1}$ to +2600 km s$^{-1}$, galaxies with Disturbed rotation curves have a nearly Gaussian velocity distribution which peaks near the cluster mean velocity. A maximum likelihood fit (Pryor \& Meylan 1993) gives $<$V$>$ = 1172$\pm$100 km s$^{-1}$, and $\sigma_{\rm v}$ = 654$\pm$71 km s$^{-1}$. These two different distributions are reminiscent of the distributions which result when galaxies in Virgo are sorted by Hubble type and systemic velocity (Huchra 1985, BST 1987). Early type galaxies show a Gaussian distribution peaked at the cluster velocity ($<$V$>$ = 1200$\pm$46 km s$^{-1}$; $\sigma_{\rm v}$ = 581 (+35/-30) km s$^{-1}$; n=164; Huchra 1985). In contrast, late type galaxies have a flatter velocity distribution ($<$V$>$ = 1144$\pm$64 km s$^{-1}$) and consequently higher velocity dispersion ($\sigma_{\rm v}$ = 871 (+49/-42) km s$^{-1}$; n=163; Huchra 1985). In Fig. 15 we superpose the spirals with Disturbed kinematics on the distribution of ellipticals in Virgo (Huchra 1985); the agreement is remarkable. While the velocity distribution of the Disturbed spirals matches that of the Virgo ellipticals, their spatial distributions are quite different, with the ellipticals $\sim$30-50\% more centrally concentrated than the Disturbed spirals. Our new results on the velocity distributions offer additional evidence that galaxies with Disturbed kinematics have orbits with large radial (within the cluster) components. In Fig. 16, the fraction of galaxies with observed velocities farthest from the cluster mean (smallest squares) is highest within 3$^{\rm o}$ of M87; here radial orbital motions are aligned close to our l-o-s. With increasing distance from M87, radial orbits make larger angles with the l-o-s, so observed velocities are closer to the cluster mean velocity. The dispersion about the mean cluster velocity (adopted as 1100 km/s) decreases from 675 km s$^{-1}$ for Disturbed galaxies observed within 1$^{\rm o}-$2.9$^{\rm o}$ of M87 to 256 km s$^{-1}$ for Disturbed galaxies observed within 5$^{\rm o}-$6.9$^{\rm o}$ of M87. Orbits with large radial components will bring galaxies close to the cluster center, where tidal interactions with the cluster and other galaxies are strong and/or more common. The undisturbed spirals exhibit a significantly less pronounced effect, suggesting that their observed velocity distribution arises from orbits with relatively smaller radial components. Regardless of the orbital intricacies which produce the observations, we have been able to identify two populations of Virgo spirals by the characteristics of their rotation patterns. Spirals in the Disturbed population have recently ($\le$1 Gyr) undergone a close encounter with another galaxy or with the cluster tidal field, with consequent kinematic disturbance. In contrast, the spiral population with normal rotation properties may have suffered no major encounters within the past $\approx$10$^{9}$ years. One interesting result of this division of spirals is that five of the six spirals in the sample with negative heliocentric systemic velocities have rotation forms classified Regular. This is not evidence that they are foreground objects, but rather that they have not had a recent tidal interaction. 5. {\it Spirals with Disturbed rotation curves are not located preferentially close to M87} (Fig. 16). As the distance from M87 increases from 1$^{\rm o}$ to 7$^{\rm o}$, in steps of 2$^{\rm o}$, the fraction of galaxies with Disturbed kinematics, relative to Regular, increases from 0.38 to 0.54 to 0.88. If we include the Sa galaxies without emission in the Disturbed category, the fractions change only to 0.44, 0.54, and 0.88. Although the numbers are small, the relative increase in Disturbed forms, $\Delta$M87$\le7^{\rm o}$, seems fairly secure. This observation follows as a natural consequence of radial orbits (see 4. above). Galaxies on radial orbits spend most of their time near apocenter. Hence we expect to find Disturbed galaxies far from the cluster core, even though the disturbances occur closer to the cluster core. We have attempted to examine if there is a bias in our rotation curve classification. The few spirals near the inner core would have truncated gas disks. Are we more likely to classify an extended rotation curve as Disturbed than we are to classify a truncated curve as Disturbed? Of the 16 galaxies within 2$^{\rm o}$ of M87, 8 are classified Regular, 7 as Disturbed, and one unclassified. The range of measured gas extents, r$_{f}$/R$_{25}$, are essentially the same for both types. Of 4 S0s with small inner gas disks, 3 are classified Regular. Among the galaxies with very extended gas disks, r$_{f}$/R$_{25} \ge$ 0.75, 6 are classified Regular, 11 Disturbed. We conclude that any bias in the classification is small, and that the increasing fraction of Disturbed galaxies at the core periphery is real. \section{THE ROTATION CURVES: AMPLITUDES} Vmax, the maximum amplitude of rotation in the plane of the galaxy, is an important parameter for obtaining estimates of absolute magnitudes and distance estimates via the Tully-Fisher relation (Yasuda, Fukugita, \& Okamura 1997; Federspiel, Tammann, \& Sandage 1998). We list in the final column of Table 1, values of Vmax we have determined. Each value comes from the maximum of the rotation curve, or for galaxies with less extended emission, from extrapolating velocities to the R$_{25}$ isophotal radius (Fig. 12). In evaluating Vmax, we have been guided by the forms of the synthetic rotation curves (SRC). These curves were produced some years ago (Thonnard \& Rubin 1981; Rubin et al. 1985) from our rotation curves of normal (mostly) field spirals, with systemic velocities ranging from about 1000 km s$^{-1}$ to 8000 km s$^{-1}$, well distributed over the sky. Rotation curves for galaxies of Hubble type Sa (11), or Sb (23), or Sc (20) were used to produce for that Hubble type a sequence of SRC as a function of absolute magnitude. To determine absolute magnitudes, each galaxy was placed at the distance inferred from its velocity, for an adopted Hubble constant. Note however, that in making use of the SRC here, we use only their forms; the absolute magnitude scale is not relevant to the discussion. To review the procedure to form the SRC, we show in Fig. 17 plots from the 1980 Carnegie yearbook (Thonnard \& Rubin 1981). The upper panels show the variation of rotational velocity with absolute magnitude for 20 Sc galaxies, each plot for a different value of r/R$_{25}$. Each fitted line is a TF-like relation, but evaluated at a fixed fraction of the galaxy radius, smaller than the R$_{25}$ radius. The bottom plot shows the superposition of 5 such fits. To form the SRC for B$_{t}^{o}$ = $-$19 (for example), the velocity at successive values of 0.1 in r/R$_{25}$ is read from the intersections, and a smooth curve drawn. This process is repeated for successive values of B$_{t}^{o}$. We have no doubt that this is a valid and valuable procedure. But we must point out that the (mostly) field sample of 60 contained 3 Virgo cluster galaxies, and these three, NGC 4321 (Sc), 4419 (Sa), and 4698 (Sa), all classified Regular, are in the present sample. As described in those early papers, these three galaxies were placed at the mean Virgo distance. We identify the points for NGC 4321 in Fig. 17. A check on the validity of this procedure comes from a comparison (Fig. 18) of our Vmax(optical) values with Vmax(21-cm) (Federspiel et al. 1998). These authors correct published values of W$_{20}$ for inclination, turbulent- and z-motions. Our optical values include only inclination corrections and extrapolations discussed above; our adopted inclinations differ only trivially from those of Federspiel et al. The agreement of the two sets is excellent; the perpendicular displacement of the points from the line of slope 1 gives $\sigma$(log line width) = 0.037. Compared with the 21-cm line width, galaxies with Disturbed rotation curves show no larger scatter than galaxies with Regular rotation curves. This is evidence that the 21-cm line width and the extrapolated optical rotation velocity measure the same quantity, and the quality of the optical Vmax values are not compromised by irregularities in the rotation curves. The points which lie above the line (Fig. 18) have optical velocities (times 2) higher than the 21-cm profile widths. The few points which lie highest above the line have observed values of HI deficiency $\ge$0.5. The optical extrapolation procedure may compensate for the HI truncation, and offer an accurate value of the disk Vmax. While it is not the aim of this paper to present a Tully-Fisher analysis nor to determine a distance for the Virgo cluster, we show several TF plots (Fig. 19). Plots (a) and (c) show the correlation of blue apparent magnitude with log Vmax; galaxies are identified by their rotation classification. Galaxies with uncertain values of Vmax are excluded. Galaxies both Regular and Disturbed are well mixed, and exhibit similar scatter. When the galaxies are coded by Hubble type, (Fig. 19b), the characteristic separation of early types toward higher rotation velocity is apparent. When Hubble types S0$-$Sb are excluded (Fig. 19c), the scatter is reduced. Although Hubble type dependence disappears for H magnitudes, there are only 18 Virgo galaxies in our sample (Fig. 19d) with H magnitudes (Aaronson et al. 1982). A few Virgo galaxies have exceptionally small rotation velocities. Notable are NGC 4064, 4424, 4506, 4584, 4694; all have emission detected only over a small fraction ($\approx0.2R_{25}$) of their radii, with velocities close to zero. Their location on Tully-Fisher plots (Figs. 17, 19) shows that their optical velocities are anomalously low for their optical luminosities, if the l-o-s velocities arise from circular orbits in a circular disk, and if the galaxies are at the distance of the Virgo cluster. While we have no direct measures of their distances, their small gas extents and peculiar morphologies suggest that they are indeed in the cluster. They are difficult to classify. Several are called Amorphous or pec, a few are barred, several have no obvious nucleus. N4424 is an apparent merger remnant (Kenney et al. 1996). Perhaps they are en route to becoming spheroidals. A few others, UGC 7171, 7249, and 7784, with low Vmax but with more normally extended gas, define the low-luminosity end of the Tully-Fisher plot (Fig. 19). We may have detected galaxies in which the gas is not rotationally supported. Such galaxies have been produced in N-body cluster simulations (Moore, Katz, \& Lake 1995). Following several strong encounters, the galaxy loses angular momentum, and ends up as a prolate figure with gas at the center. Hence rotation along the most elongated axis may be small; rotation along the apparent minor axis can be larger. Both NGC 4064 and 4424 (Kenney et al. 1996) show approximately zero velocities along their major axes, and small but velocity gradients along their minor axes; we have no minor axis spectra for the others in this set. Such galaxies deserve more study. \section{CONCLUSIONS} The primary purpose of this paper is to present optical kinematic information for a significant number of galaxies in the Virgo cluster, and to identify kinematic disturbances in these galaxies. We present evidence that there are two populations of spirals in the cluster, distinguished by their internal kinematics. Those with Regular rotation properties exhibit a distribution of systemic velocities which is flat, V$_{\rm o}$ = $-$300 km s$^{-1}$ to V$_{\rm o}$ = +2500 km s$^{-1}$. Those with Disturbed rotation exhibit a Gaussian distribution of systemic velocities which peaks near the cluster systemic velocity, V$_{\rm o}$ = +1172$\pm$100 km s$^{-1}$ and $\sigma_{\rm v}$ = 654$\pm$71 km s$^{-1}$ (Fig. 14,15). Disturbed galaxies are not preferentially near M87, but are prominent near the 6$^{\rm o}$ core periphery. The Disturbed galaxies probably have orbits with larger radial (relative to the cluster center) components than the undisturbed galaxies, and some of them may be falling into the core for the first time. Identifying this population of spirals with disturbed kinematics is the major result of this study. Other conclusions which follow have been discussed in Sections 5 and 6. We cannot distinguish between galaxy-galaxy or galaxy-cluster gravitational interactions as the sources which produced these disturbances, although the required perturbations should not be bisymmetric, in order to match the observations. Perhaps there are multiple causes. In any cluster, galaxies will be exposed to both local (galaxy-galaxy) and global (galaxy-cluster) interactions. The effects may be more pronounced in a non-virialized cluster like Virgo, where tidal effects from substructure are likely important, and where infalling galaxies will experience these effects for the first time. Regardless of the cause, the processes which produce Disturbed rotation curves must also play a role in altering the galaxies' morphologies, and perhaps also in driving the Virgo cluster toward dynamical equilibrium. In their systemic velocity distribution, the galaxies with Disturbed kinematics resemble the nearly-virialized elliptical population more than they resemble the non-virialized population of spirals with Regular kinematics. Ellipticals and the Disturbed spirals have similar distributions of l-o-s orbital velocities, yet the Disturbed spirals are on average located further from the cluster center. This raises an interesting question about galaxy morphology and environment. Do elliptical galaxies owe their morphology to their orbits within the cluster? Galaxies we identify as kinematically Disturbed are only that subset presently within the required time window following the event. Interacting galaxies observed at earlier or later interaction stages will have more modest kinematic distortions. This may explain the lack of simple correlation between galaxies with kinematic disturbances and apparent pairs, and why other galaxies which may indeed be experiencing tidal interactions display only minimal kinematic irregularities. Thirty years ago, Peebles (1970; see also Lynden-Bell 1967) made a 300-body calculation, matched to velocity and position parameters of galaxies in the Coma cluster. The results confirmed that a gravitationally bound cluster could collapse in an expanding universe, an idea that was then not yet universally accepted. Shectman (1982) extended the calculation to study a more detailed infall model for the formation of the Coma cluster. His conclusion, that a similar calculation ``... should be possible for the Virgo Cluster" is now almost 20 years old. Perhaps its time will soon come. With the availability of larger telescopes, observational interests are moving from nearby to very distant clusters. Data from the Virgo cluster should offer a baseline for study of distant clusters, as we attempt to learn when and how galaxies and clusters formed. But the data have relevance to other questions. They are input for the Tully-Fisher correlation and for our infall to Virgo. They aid in understanding the role that interactions play in the formation and evolution of galaxies within a cluster and for the cluster as a whole, and in learning how evolutionary processes differ for field and cluster spirals. Unfortunately, we lack data concerning {\it stellar} orbits within most galaxies. Stellar kinematics can provide valuable evidence for disturbances in gas-poor galaxies, for mergers and accretion events, and also answer questions concerning the similarity of stellar and gas motions. Needed most are three dimensional orbital motions; then we will make real progress in sorting out the kinematic evolution of the cluster. \acknowledgments We thank the Directors of Kitt Peak National Observatory and Palomar Observatories for telescope time, and numerous telescope operators for cheerful assistance. VR thanks Dr. W. Kent Ford, Jr., who participated in early observations of these galaxies, Dr. Paul Schechter for introducing her to the Palomar double spectrograph, Drs. Neta Bahcall, Jay Gallagher, John Graham, Gus Oemler, Richard Larson, and Francois Schweizer for helpful conversations and references, Dr. Allan Sandage for valuable comments on an early draft of this paper, and especially Sandy Keiser for recovering spectra from old 9-track tapes. We thank Drs. J. van Gorkom and D. Burstein for confirming that resolved 21-cm HI observations and H magnitudes are as few as we found. Thoughtful comments from a referee prompted us to enlarge the scope and improve this paper. J.D.P.K. received support from NSF grant AST-9322779, and NASA support to V. R. and J.D.P.K. comes from HST program GO-5375. We acknowledge the use of the NASA Extragalactic Database NED operated by IPAC and the Observatoire de Lyon Hypercat. \clearpage
\section{Introduction} \setcounter{equation}{0} The inflationary paradigm \cite{L90} offers the attractive possibility of resolving many of the puzzles of standard hot big bang cosmology. The crucial ingredient of most successful inflationary scenarios is a period of ``slow-roll'' evolution of a scalar field $\varphi $ (the ``inflaton''), during which the potential energy $V(\varphi )$ stored in $\varphi $ dominates its kinetic energy $\dot{\varphi}^{2}/2$ and drives a quasi-exponential expansion of the Universe. At present there exists no preferred concrete inflationary scenario based on a convincing realistic particle physics model. In particular, though string theory provides one with several very weakly coupled scalar fields (the moduli), which could be natural inflaton candidates, their non-perturbative potentials $V(\varphi )$ do not seem fit to sustain slow-roll inflation because, for large values of $\varphi $, they tend either to grow, or to tend to zero, too fast. It is therefore important to explore novel possibilities for implementing an inflationary evolution of the early Universe. The aim of this work is to point out that, even in absence of any potential energy term, a general class of non-standard (i.e. non-quadratic) \textit{kinetic-energy} terms $\mathcal{L}(\dot{\varphi})$, for a scalar field $\varphi $, can drive an inflationary evolution of the same type as the usually considered potential driven inflation. By ``usual type of inflation'' we mean here an accelerated \textit{expansion} (in the Einstein conformal frame) during which the curvature scale starts around a Planckian value and then \textit{% decreases} monotonously. By contrast, the pre-big bang scenario \cite{PBB} uses a \textit{standard} (quadratic) kinetic-energy term $\dot{\varphi}^{2}/2$ to drive an accelerated \textit{contraction} (in the Einstein frame) during which the curvature scale \textit{increases}. Though we shall motivate below the consideration of non-standard kinetic terms by appealing to the existence, in string theory, of higher-order corrections to the effective action for $\varphi $, we do not claim that the structure needed for implementing our ``kinetically driven'' inflation arises inevitably in string theory. The aim of this work is more modest. We draw attention to a new mechanism for implementing inflation. A large class of the (toy) models we shall consider satisfy two of the most crucial requirements of inflationary scenarios: (i) the scalar perturbations are ``well-behaved'' during the inflationary stage, and (ii) there exist natural mechanisms for exiting inflation in a ``graceful'' manner. We leave to future work a more detailed analysis of the observational consequences of our kinetically driven inflation (spectrum of scalar and tensor perturbations, reheating, \ldots). We hope that, by enlarging the basic ``tool-kit'' of inflationary cosmology, our mechanism could help to locate which sector of string-theory (if any) has inflated a strongly curved initial state into our presently observed large and weakly curved Universe. \section{General Model} \label{sec:general-model} \setcounter{equation}{0} We consider a single scalar field $\varphi $ interacting with gravity through non-standard kinetic terms, \begin{equation} \label{eq:einstein-action} S=\int d^{4}x\,\sqrt{g}\left[ -\frac{R}{6\kappa ^{2}}+p(\varphi ,\nabla \varphi )\right] . \end{equation} Here, $\kappa ^{2}\equiv 8\pi G/3$ and we use the signature $(+---)$. For simplicity, we only consider Lagrangians $p$ which are local functions of $% \nabla _{\mu }\varphi $ and therefore, depend only on the scalar \begin{equation} X\equiv \frac{1}{2}(\nabla \varphi )^{2}. \end{equation} As we bar here the consideration of potential terms, we must impose that the function $p(\varphi,X)$ vanish when $X\rightarrow 0$. Near $X=0$, a generic kinetic Lagrangian $p(\varphi,X)$ is expected to admit an expansion of the form \begin{equation} \label{eq:p-expansion} p(\varphi ,X)=K(\varphi )X+L(\varphi )X^{2}+\cdots. \end{equation} One of the possible particle theory motivations for taking this kind of Lagrangian seriously could be the following. Let us consider only gravity and some moduli field $\varphi $ (which could be the dilaton, or some other moduli) in string theory. It is well-known that $\alpha^{\prime}$ corrections (due to the massive modes of the string) generate a series of higher-derivative terms in the low-energy effective action $S_{\rm eff}$, while string-loop corrections generate a non-trivial moduli dependence of the coefficients of the various kinetic terms. This leads to a structure of the type (in the string frame $\hat{g}_{\mu\nu}$) \begin{eqnarray} \label{eq:string-effective-action} S_{\rm eff} &=&\frac{1}{6\kappa^{2}}\int d^{4}x\sqrt{\hat{g}}\Big\{ -B_{g}(\varphi)\hat{R}-B_{\varphi}^{(0)}(\hat{\nabla}\varphi)^{2} \\ &&+\alpha^{\prime}\left[c_{1}^{(1)}B_{\varphi}^{(1)}(\varphi)(\hat{ \nabla}\varphi)^{4}+\cdots\right] +O(\alpha'\,^{2})\Big\}, \nonumber \end{eqnarray} where the ellipsis stands for other four-derivative terms (like $(\Box \varphi)^{2},\hat{R}_{\mu\nu\rho\sigma}^{2}, \ldots )$. In the case where $\varphi $ is the dilaton the coupling functions are of the form \begin{eqnarray*} B_{g}(\varphi ) &=&e^{-\varphi }+c_{g,0}+c_{g,1}e^{\varphi }+\cdots , \\ B_{\varphi }^{(0)}(\varphi ) &=&e^{-\varphi }+c_{\varphi ,0}+c_{\varphi ,1}e^{\varphi }+\cdots , \\ B_{\varphi }^{(1)}(\varphi ) &=&e^{-\varphi }+\cdots , \end{eqnarray*} where the ellipsis contain higher contributions in $g_{\rm string}^{2}= e^\varphi$, including non-perturbative ones. Transforming Eq. (\ref{eq:string-effective-action}) to the Einstein frame $g_{\mu\nu }=B_{g}(\varphi)\hat{g}_{\mu\nu}$, and neglecting the other possible four derivative terms in Eq. (\ref{eq:string-effective-action}), leads to the effective action (\ref{eq:einstein-action}) with $\varphi$-kinetic terms of the form (\ref{eq:p-expansion}) where \begin{eqnarray} K(\varphi ) &=&3\frac{B_{g}^{\prime }{}^{2}(\varphi )}{B_{g}^{2}(\varphi )}-2 \frac{B_{\varphi }^{(0)}(\varphi )}{B_{g}(\varphi )}, \\ L(\varphi ) &=&c_{1}^{(1)}\frac{2\alpha ^{\prime }}{3\kappa ^{2}}B_{\varphi }^{(1)}(\varphi ). \end{eqnarray} In the case where $\varphi $ is the dilaton (so that $g_s=e^{\varphi /2}$ is the string coupling), we have, in the weak-coupling limit $g_s^{2}=e^{\varphi}\ll 1$, $B_{g}(\varphi )\simeq B_{\varphi}^{(0)} (\varphi)\simeq B_{\varphi}^{(1)}(\varphi)\simeq e^{-\varphi}$ so that $K(\varphi)\simeq 1$ and $L(\varphi)\propto e^{-\varphi}$. However, when $g_s$ becomes of the order of unity it is not a priori excluded that $\ K(\varphi)$ and $L(\varphi)$ could become more complicated functions of $\varphi$. We shall give below some examples of such possible complicated behaviours. Returning now to a general Lagrangian $p(\varphi ,X)$ the ``matter'' energy-momentum tensor reads \begin{equation} \label{eq:energy-momentum-tensor} T_{\mu \nu }\equiv \frac{2}{\sqrt{g}}\frac{\delta S_{\varphi }}{\delta g^{\mu \nu }}=\frac{\partial p(\varphi ,X)}{\partial X}\nabla _{\mu }\varphi \nabla _{\nu }\varphi -p(\varphi ,X)g_{\mu \nu }. \end{equation} Equation (\ref{eq:energy-momentum-tensor}) shows that, if $\nabla _{\mu }\varphi $ is time-like (i.e. $X>0$), our scalar field action is ``equivalent'' to a perfect fluid ($T_{\mu \nu }=(\varepsilon +p)u_{\mu }u_{\nu }-p\,g_{\mu \nu }$) with pressure \begin{equation} \label{eq:pressure} p=p(\varphi ,X), \end{equation} energy density \begin{equation} \label{eq:energy-density} \varepsilon =\varepsilon (\varphi ,X)\equiv 2X\frac{\partial p(\varphi ,X)} {\partial X}-p(\varphi ,X), \end{equation} and four-velocity \begin{equation} \label{eq:four-velocity} u_{\mu }=\sigma \frac{\nabla _{\mu }\varphi }{\sqrt{2X}}, \end{equation} where $\sigma $ denotes the sign of $\dot{\varphi}=\nabla _{0}\varphi $. As usual, in inflationary cosmology, we consider a flat background Friedmann model $ds^2=dt^2-a^2(t)d\mathbf{x}^2$ and an homogeneous background scalar field $X=\frac{1}{2}\dot{\varphi}^2$. A convenient minimal set of independent evolution equations for $a(t)$ and $\varphi(t)$ is (with $H\equiv% \dot{a}/a$) \begin{eqnarray} \label{eq:Friedmann} H^2 &=&\kappa^2 \varepsilon \, , \\ \label{eq:energy-conservation} \dot{\varepsilon}&=&-3H(\varepsilon+p). \end{eqnarray} It will be also useful to refer to other (redundant) forms of the evolution equations: \begin{eqnarray} \frac{\ddot{a}}{a}&=&-\frac{1}{2}\kappa^2(\varepsilon+3p) \, , \\ \label{eq:canonical} \frac{1}{a^3}\frac{d}{dt}(a^3\pi)&=&\dot{\pi}+3H\pi=\frac{\partial p} {\partial \varphi}, \end{eqnarray} where $\pi\equiv\partial p/\partial\dot{\varphi}=\dot{\varphi}\,\partial p/\partial X$ denotes the momentum conjugate to $\varphi$. [Note that equation (\ref{eq:energy-density}) reads $\varepsilon=\dot{\varphi}\partial p/\partial \dot{\varphi}-p$, which is the usual energy associated to the Lagrangian $p(\varphi,\dot{\varphi})$.] In this work we shall only consider solutions of Eqs. (\ref{eq:Friedmann})-(% \ref{eq:canonical}) which describe an expanding Universe (in Einstein frame), that is $H>0.$ This reduces the evolution equations (\ref {eq:Friedmann}), (\ref{eq:energy-conservation}) to the master equation \begin{equation} \label{eq:master} \dot{\varepsilon}=-3\sqrt{\varepsilon }(\varepsilon +p). \label{master} \end{equation} Here, and in the following, we use units such that $\kappa ^{2}\equiv 8\pi G/3=1$. Note that, from Eq. (\ref{eq:Friedmann}), $\varepsilon $ was constrained to be positive. \section{Kinetically Driven Inflation - Basic Idea} \label{sec:k-inflation} \setcounter{equation}{0} As a warm up, let us first consider the case where the Lagrangian $p$ depends only on $X=\frac{1}{2}(\nabla \varphi )^{2}$, and not on $\varphi $: $p=p(X)$. From Eq. (\ref{eq:energy-density}), the energy-density depends also only on $X$: $\varepsilon (X)=2X\partial p/\partial X-p(X)$. In hydrodynamical language it means that we have here an ``isentropic'' fluid, with a general equation of state relating $p$ to $\varepsilon $: $% p=f(\varepsilon )$. The master evolution equation can then be qualitatively solved by looking at the graph of the equation of state $p=f(\varepsilon )$. The shape of this graph depends very much on the shape of the function $p=p(X)$, or better, on the shape of the function $p=p(\psi)$ where $\psi \equiv \dot{\varphi}$, so that $X=\frac{1}{2}\psi^{2}$. Indeed, the relation \begin{equation} \varepsilon =\psi \partial p/\partial \psi -p \label{eq:varepsilon-psi} \end{equation} shows that $\varepsilon $ can be read geometrically off the graph $p=p(\psi)$. More precisely $-\varepsilon$ is the ``intercept''of the tangent to the curve $p=p(\psi)$ at the point $\psi$, i.e. its intersection with the vertical axis $\psi=0$. The minus sign means that $\varepsilon $ is positive (negative) if the tangent intersects the vertical axis below (above) the $p=0$ horizontal axis. Therefore, if the function $p(\psi )=\frac{1}{2}K\psi ^{2}+\frac{1}{4}L\psi ^{4}+\cdots $ is, for instance, always \textit{convex}, $\partial ^{2}p/\partial \psi ^{2}>0$, as will be the case if all the coefficients $K,L,\ldots $ appearing in the expansion (\ref{eq:p-expansion}) are positive, $p$ and $\varepsilon $ will be always positive. In such a case, Eq. (\ref{eq:master}) shows that $\varepsilon $ will monotonically decrease towards zero, and the evolution will be driven to the ``attracting solution'' \begin{equation} \label{eq:vacuum-attractor} \varepsilon =H^{2}\approx \frac{1}{9t^{2}}\quad ,\quad a\approx a_{0}\,t^{1/3}. \end{equation} This attractor corresponds to the asymptotic equation of state $p\approx \varepsilon $ valid near $\varepsilon =0$ where the usual kinetic term $% p\simeq \frac{1}{2}K\psi ^{2}$ dominates. On the other hand, if the function $p(\psi )$ is non-convex and has some oscillatory behaviour as $\psi $ increases (i.e., if we consider the general case where the expansion coefficients $K,L,\ldots $ in Eq. (\ref{eq:p-expansion}) may take negative values) the graph $p=f(\varepsilon )$ can be more complicated and can allow for exponential-type inflationary behaviour. Let us first note that, because of Eq. (\ref{eq:varepsilon-psi}), the extrema of the function $p=p(\psi )$ (or $p=p(X)$) correspond to values where $p=-\varepsilon $, i.e. to \textit{% fixed points} of the master evolution equation (\ref{eq:master}). For a general function $p=p(X)$ the graph of the (multiform) equation of state might resemble Fig.\ref{fig:state}. \begin{figure}[!t] \begin{center} \epsfig{file=state.eps} \end{center} \caption{Graph of the equation of state linking $p$ to $\protect\varepsilon$ for an hypothetical general kinetic Lagrangian $p(\dot{\protect\varphi})$. The evolution for expanding, flat cosmologies proceeds along the indicated arrows. The shaded region ($\protect\varepsilon<0$) is excluded. Except for the origin and the point above it on the vertical axis, the attractors of the evolution are inflationary fixed points with $p=-\protect\varepsilon$.} \label{fig:state} \end{figure} From the master Eq. (\ref{eq:master}) \ it follows that $\varepsilon $ will decrease above the line $p=-\varepsilon $, and increase below it. Fig. \ref {fig:state} then shows that all the intersection points with the $% p=-\varepsilon $ line are \textit{attractors} of the (future) evolution. The arrows in Fig. \ref{fig:state} indicate the evolutive flow, which reverses (along the graph) at the extrema of $\varepsilon $. The region where $% \varepsilon <0$ is excluded because it cannot be reached by flat cosmologies (see Eq. (\ref{eq:Friedmann})). The fixed points lying at the $% p=-\varepsilon $ line correspond to an exponential inflation \begin{equation} \label{eq:inflationary-attractor} H_{\rm att}^2=\varepsilon_{\rm fixed}\quad ,\quad a_{\rm att}(t)=a_{0}\exp\left(\sqrt{\varepsilon_{\rm fixed}}t\right) \end{equation} Apart from these inflationary attractors, there are two other attractors (in the case depicted in Fig. \ref{fig:state}): (i) the origin, where the evolution is driven toward the solution (\ref{eq:vacuum-attractor}) (corresponding to the ``hard'' equation of state $p=\varepsilon $), and (ii) the point above the origin on the vertical axis. As we will see later the latter point lies in the region of absolute instability and has therefore no physical significance. In this work we shall focus on the inflationary attractors (\ref{eq:inflationary-attractor}) because they exhibit the novel possibility of getting, for a large set of initial conditions, quasi-exponential (or power law) inflation out of a purely kinetic Lagrangian. Note that the condition for the existence of these inflationary attractors can also be seen in Eq. (\ref{eq:canonical}). In absence of a $\varphi $ dependence, Eq. (\ref{eq:canonical}) says that $a^{3}\pi $ is constant so that $\pi $ is attracted toward zero. As the momentum $\pi=\partial p/\partial \dot{\varphi}=% \dot{\varphi}\,\partial p/\partial X$, we see that non trivial ($\dot{\varphi% }\neq 0$) attractors can exist if the kinetic terms are non-standard so that $\partial p/\partial X$ can vanish for non zero values of $\dot{\varphi}$ . The extremal values of $p$ correspond to the inflationary attractors discussed above. The labels ``s'' and ``u'' in Fig. \ref{fig:state} (which stand for ``stable'' and ``unstable'') indicate whether, for our present isentropic equation of state $p=f(\varepsilon)$, the squared speed of sound $c_s^2=dp/d\varepsilon$ is positive or negative, respectively. This issue will be further discussed below. Note that the ``price'' to pay for having inflationary attractors as in Fig. \ref{fig:state} is the existence of regions, in phase space $(\varphi ,% \dot{\varphi})$, with negative energy density. We shall assume in this work that this is not physically forbidden. All the cosmological evolutions that we shall consider below stay always in the positive energy regions, and the existence, elsewhere in phase space, of negative $\varepsilon $ regions does not necessarily cause some instabilities along our evolutionary tracks. The simple case of a kinetic Lagrangian depending only on $\dot{\varphi}$ considered above is the analog, for kinetically driven inflation (or ``\textit{k}-inflation'' for short), of a de Sitter model with constant energy density. It is clear that both models should have similar problems. Namely, there is no natural graceful exit, no smooth transition to a Friedmann Universe and the cosmological perturbations are ``ill-defined''. To avoid these problems we should allow the coefficients in the expansion (\ref{eq:p-expansion}) of the Lagrangian $p$ to depend on the scalar field $\varphi$. \section{``Slow-Roll'' \textit{k}-Inflation} \label{sec:slow-roll-inflation} \setcounter{equation}{0} The simplest way to realize successfully the idea of \textit{k}-inflation is to consider the analog of ``slow-roll'' potential driven inflation, in which the potential $V(\varphi )$ in the Lagrangian $\mathcal{L}(\varphi ,\dot{\varphi}% )=\frac{1}{2}\dot{\varphi}^{2}-V(\varphi )$ dominates the kinetic term $\dot{% \varphi}^{2}/2$ and evolves slowly. For the concept of \textit{k}-inflation to have a relevance to a large class of models, we need to consider a general kinetic Lagrangian $p(\varphi ,X)$. The idea is therefore to find the conditions under which the influence of the non-trivial $\varphi $ dependence of $p(\varphi ,X)$ will represent only a relatively small perturbation of the attraction toward exponential inflation discussed in Section \ref{sec:k-inflation}. To do that in a concrete manner it is convenient to focus henceforth on the simplest kinetic Lagrangian, containing only $\dot{\varphi}^{2}$ and $\dot{\varphi}^{4}$ terms, namely \begin{equation} \label{eq:simplest-p-expansion} p(\varphi ,X)=K(\varphi )X+L(\varphi )X^{2}=\frac{1}{2}K(\varphi )(\nabla \varphi )^{2}+\frac{1}{4}L(\varphi )(\nabla \varphi )^{4}. \end{equation} Let us first motivate the possibility of rather arbitrary functions $% K(\varphi ),L(\varphi )$ by considering again the low-energy effective action of string theory (\ref{eq:string-effective-action}). As we mentioned above, in the weak coupling limit $K(\varphi )\simeq 1$ and $L(\varphi )\propto e^{-\varphi }.$ However, when $g_{s}$ becomes of order unity it is not a priori excluded that $K(\varphi )$ could change sign. For instance we could consider a simple model (of the type considered in Refs. \cite{DP94}, \cite{DV96}), where the coupling functions $B(\varphi)$ in the action (\ref{eq:string-effective-action}) are the same, that is $B_{g}(\varphi)=B_{\varphi}^{(0)}(\varphi)=B_{\varphi}^{(1)}(\varphi)=B(\varphi) $. In this model \begin{equation} K(\varphi )=3\left(\frac{B'(\varphi)}{B(\varphi)}\right)^2-2 \end{equation} might become negative when $g_{s}$ reaches values of order unity. In fact a model, incorporating string loop corrections, considered in Ref. \cite{FMS99} has \begin{equation} \label{eq:FMS-K} K(\varphi)=1-3k\,e^{\varphi}\frac{6+k\,e^{\varphi}}{(3+k\,e^{\varphi})^2}, \end{equation} where $k=3\delta^{\rm GS}/8\pi ^{2}$ is a positive parameter. The R.H.S. of Eq. (\ref{eq:FMS-K}) becomes negative when $e^{\varphi}>3(\sqrt{6}-2)/2k$. Motivated by these examples, we shall assume, as simplest toy model exhibiting interesting dynamics, a Lagrangian of the form (\ref{eq:simplest-p-expansion}% ) with a function $K(\varphi)$ which is positive in some range of values of $\varphi $ (``weak-coupling-domain'') and becomes negative in some other range (``strong-coupling-domain''). On the other hand, to ensure the positivity of $\varepsilon $ for large field gradients $X$, we shall assume that the function $L(\varphi )$ remains always positive. The equation of state in the model (\ref{eq:simplest-p-expansion}) is parametrically given by \begin{eqnarray} p&=&K(\varphi)X+L(\varphi)X^2, \\ \varepsilon&=&K(\varphi)X+3L(\varphi)X^2. \end{eqnarray} We represent in Fig. \ref{fig:model-state} the change in the form of the equation of state $p=f(\varepsilon,\varphi)$ as $\varphi$ varies from the weak-coupling region ($K>0$) to the strong-coupling one ($K<0$). Note that for large values of $X$, the equation of state asymptotes the one of radiation ($p=\varepsilon/3$), while it is tangent to the hard equation of state ($p=\varepsilon$; with $\mathrm{sgn}(\varepsilon)=\mathrm{sgn}(K)$) when $X\to0 $. In the strong coupling domain there appears (in the adiabatic approximation where $\varphi$ is treated as constant) an inflationary fixed point where $p_{\rm fixed}=-\varepsilon_{\rm fixed}$. \begin{figure}[!b] \begin{center} \epsfig{file=model-state} \end{center} \caption{Change of form of the equation of state $p=f(\protect\varepsilon, \protect\varphi)$ in the model (\ref{eq:simplest-p-expansion}) as $\protect\varphi$ varies from the weak-coupling region ($K(\protect\varphi)>0$) to the strong-coupling one ($K(\protect\varphi)<0$).} \label{fig:model-state} \end{figure} Let us investigate under what conditions on the functions $K(\varphi )$ and $L(\varphi)$ one can indeed approximately solve the master evolution equation (\ref{eq:master}) by considering that the variation of $\varphi$ brings only a small perturbation onto the simple $\varphi $-independent evolution studied above. We shall simplify this study by redefining the scalar field and working with the new field variable $\varphi^{\rm new}=\int d\varphi^{\rm old}L^{1/4}(\varphi^{\rm old})$ (which is well defined because we assume $L(\varphi)>0$). With this definition $L_{\rm new}(\varphi^{\rm new})=1$ and $K_{\rm new}(\varphi^{\rm new})=K_{\rm old}(\varphi^{\rm old})/L_{\rm old}^ {1/2}(\varphi^{\rm old})$. In other words, we can assume, without loss of generality, that $L(\varphi)=1$. With this simplification the zeroth-order ``slow-roll'', or ``adiabatic'', solution to Eq. (\ref{eq:master}), i.e. the instantaneous attractive fixed point of Eq. (\ref{eq:master}) (solution of $0=\varepsilon+p=2X\partial p/\partial X)$ corresponds to \begin{eqnarray} \label{eq:X-slow-roll} X_{\rm sr} &=&\frac{1}{2}\overline{K}(\varphi_{\rm sr}), \\ \dot{\varphi}_{\rm sr} &=&\sigma\sqrt{\overline{K}(\varphi_{\rm sr})}, \\ \varepsilon_{\rm sr} &=&\frac{1}{4}\overline{K}^{2}(\varphi_{\rm sr}), \\ \label{eq:H-slow-roll} H_{\rm sr} &=&\frac{1}{2}\overline{K}(\varphi_{\rm sr}). \end{eqnarray} Here $\overline{K}(\varphi)\equiv -K(\varphi)$ is positive (in the slow-roll domain) and $\sigma $ denotes the sign of $\dot{\varphi}$. The time evolution of the slow-roll \textit{k}-inflation is given, from Eqs. (\ref {eq:X-slow-roll})-(\ref{eq:H-slow-roll}), by simple quadratures (the subscript ``in'' denotes initial values) \begin{eqnarray} t-t_{\rm in} &=&\sigma \int_{\varphi_{\rm in}}^{\varphi}\frac{d\varphi} {\sqrt{\overline{K}(\varphi)}}, \\ N &\equiv &\ln \left( \frac{a(t)}{a_{\rm in}}\right)=\frac{\sigma}{2} \int_{\varphi_{\rm in}}^{\varphi}\sqrt{\overline{K}(\varphi)}d\varphi . \end{eqnarray} [The notation $N$ is introduced to denote the number of $e$-folds of inflation.] The post-slow-roll approximation, $X=X_{\rm sr}+\delta {X}$, is then obtained by rewriting the master equation (\ref{eq:master}) as \begin{equation} \frac{\partial p}{\partial X}=-\overline{K}+2X=2\delta X=-\frac{\dot{\varepsilon}} {6X\sqrt{\varepsilon}}, \end{equation} and replacing the slow-roll approximations (\ref{eq:X-slow-roll})-(\ref {eq:H-slow-roll}) in the R.H.S. This yields \begin{equation} \label{eq:slow-roll-parameter} \frac{\delta X}{X_{\rm sr}}\simeq -\frac{1}{12}\frac{\dot{\varepsilon}} {\varepsilon ^{3/2}}\simeq -\frac{\sigma}{3}\frac{\overline{K}^{\prime}} {\overline{K}^{3/2}}\simeq +\frac{2\sigma }{3}\left(\frac{1}{\sqrt{\overline{K}}} \right)', \end{equation} where the prime denotes a derivative with respect to $\varphi $. The criterion for the validity of our previous slow-roll solution (\ref{eq:X-slow-roll})-(\ref{eq:H-slow-roll}) is \begin{equation} \label{eq:slow-roll-condition} \frac{\delta X}{X}\ll 1, \end{equation} i.e. $(\overline{K}^{-1/2})^{\prime}\ll 3/2$ (when keeping $L(\varphi)\neq 1$ it would read $L^{-1/4}\partial (L^{1/4}\overline{K}^{-1/2})$ $/\partial\varphi \ll 3/2$). This condition is as easily satisfied as the usual slow-roll condition for potential driven inflation. Examples of functions $\overline{K}(\varphi)$ that satisfy this condition are: (i) any power law or exponential (or super-exponential) growth as $\varphi\rightarrow\infty$, (ii) any levelling off of $\overline{K}(\varphi)$ ($\overline{K}(\varphi)\rightarrow {\rm limit}$, with $\overline{K}^{\prime}(\varphi)\rightarrow 0$) as $\varphi\rightarrow\infty $, or (iii) a sufficiently fast pole-like growth of $\overline{K}(\varphi)\propto (\varphi_*-\varphi)^{-\alpha }$, with $\alpha>2$, as $\varphi\rightarrow\varphi_*$. Note that during slow-roll \textit{k}-inflation the following useful relation \begin{equation} \label{eq:e-plus-p} \frac{\varepsilon +p}{\varepsilon }\simeq 4\frac{\delta X}{X_{\rm sr}}, \end{equation} is satisfied, that is, the fractional compensation of the energy density by the negative pressure is proportional to the small parameter $\delta X/X_{sr}.$ Therefore, in those models where $\varepsilon+p> 0,$ or equivalently, the energy density decreases in the course of expansion, $\delta X$ is positive. It is also obvious from (\ref{eq:e-plus-p}), that inflation ends when $\delta X/X_{\rm sr}$ becomes of the order of unity, i.e. when the slow-roll condition (\ref{eq:slow-roll-condition}) is violated. For any function $\overline{K}(\varphi)$ (and more generally for any $p(\varphi ,X)$) satisfying the slow-roll criterion we can visualize our \textit{k}-inflationary behaviour as being one point (i.e. one value of $X=X_{\rm sr}+\delta X$) on an adiabatically varying equation of state graph of the type of Figs. {\ref{fig:state}} or \ref{fig:model-state}. [The adiabatic variation we mention corresponding to the fact that each graph corresponds to some specific, instantaneous value of $\varphi$ , which is itself evolving]. Eq. (\ref{eq:slow-roll-parameter}) (or its generalization to a generic $p(\varphi,X)$) then tells us that the point $X=X_{\rm sr}+\delta X$ is always \textit{displaced} away from the intersections of the ($\varphi $-instantaneous) graph with the $p=-\varepsilon$ line. Overall, the qualitative behaviour of the solutions we are focussing on is the following: Initially, we start with some representative point in the $(\varepsilon ,p)$ plane lying on an equation of state graph corresponding to some initial value of $\varphi $, deep into the strong coupling domain. We assume that, for strong coupling, the slow-roll criterion is very well satisfied. In a first evolution stage, we can neglect the $\varphi$-dependence of the equation of state because there is a fast attraction taking just a few $e$-folds of the representative point toward the nearest inflationary attractor (this stage is described by the arrows in Fig. \ref {fig:state}). After this initial stage, we can consider that our representative point follows the (post-) slow-roll motion $X=X_{\rm sr}+\delta X$, corresponding to a representative point near but away from the $p=-\varepsilon $ line (such a point is indicated in Fig. \ref{fig:model-state}). As the evolution continues, the slow-roll condition is less and less well satisfied and the representative point straggles more and more away from the $p=-\varepsilon $ line. At some point in the evolution the slow-roll criterion (\ref{eq:slow-roll-condition}) becomes violated ($\delta X/X\sim 1$) and one naturally exits the inflationary stage. We shall come back later to this exit mechanism. The qualitative picture of the evolution just represented (based on a succession of graphs in the $(\varepsilon,p)$ plane) can also be globally visualized by a phase-space picture in the $(\varphi,\dot{\varphi})$ plane (see Fig. \ref{fig:phase}). ``Slow-roll'' inflation on this graph corresponds to the portion of the separatrix (attractor) given by \begin{equation}\label{eq:separatrix} \dot{\varphi}=\sigma \sqrt{\stackrel{\_}{K}}\left( 1+\frac{\delta X} {X_{\rm sr}}+...\right)^{1/2}, \end{equation} where $\delta X/X\ll 1.$ When $\delta X/X$ becomes of the order of one a graceful exit from inflation takes place, see Section \ref{sec:exit-mechanisms}. The phase diagram Fig. \ref{fig:phase} is very similar to the one of potential driven slow-roll inflation \cite{BGZK85}. We see that the set of initial configurations of the scalar field which lead to inflation has nonzero measure. Therefore the problem of initial conditions here is very similar to what one has in the case of chaotic inflation \cite{L90}. We expect that in analogy to the other models of inflation self-reproduction of the Universe \cite{LLM94} can take place in our \textit{k-}inflationary model. However this question needs a special investigation and we leave it to future work. \begin{figure} \begin{center} \epsfig{file=phase.eps} \end{center} \caption{\label{fig:phase}Schematic phase diagram of ``slow-roll'' $k$-inflation. Trajectories approach the attractor but do not reach the line $\epsilon+p=0$ where the speed of sound vanishes. Around the point where the slow-roll condition is violated, the solutions leave the inflationary stage and approach then smoothly the vacuum $\dot{\varphi}=0$.} \end{figure} Let us also mention that one can easily build a model in which one starts initially in the weak coupling regime, with the field $\varphi$ evolving towards the strong coupling regime. The function $K$ can then be arranged in such a way that the Universe leaves the weak coupling regime to enter an inflationary stage and finally exits inflation. \section{``Power Law'' \textit{k}-Inflation} \label{sec:power-law-kinflation} \setcounter{equation}{0} It is well known that, in the usual potential driven inflationary scenario, if the potential depends exponentially on the scalar field, there exists an attractor solution which describes a power law inflating Universe (see for instance \cite{LM85}). There is no graceful exit from inflation if the potential is exponential everywhere. Therefore to solve the graceful exit problem one should assume that the exponential potential is a valid approximation of a more realistic and complicated potential only within some limited range of values of the scalar field $\varphi$. As we show in this section, one can get an analogous power law \textit{k}-inflation within the class of models which we consider in this paper. Let us again consider the model with the Lagrangian (\ref{eq:simplest-p-expansion}). For the purposes of this section it is convenient to make a new field redefinition (valid only in the region where $K<0$) and rewrite the Lagrangian (\ref{eq:simplest-p-expansion}) in terms of the new field variable $\varphi^{\rm new}=\int\sqrt{L(\varphi^{\rm old})/ |K(\varphi^{\rm old})|}\,d\varphi^{\rm old}$. This yields \begin{equation} \label{eq:new-lagrangian} p=f\left(\varphi\right)\left(-X+X^{2}\right), \end{equation} where $f\left(\varphi\right)\equiv f(\varphi^{\rm new})=K^{2}(\varphi ^{\rm old})/L(\varphi^{\rm old})$ and $X\equiv X^{\rm new}= (L/|K|)X^{\rm old}$. Working with this Lagrangian one can try to find out whether there is a function $f\left(\varphi\right)$ for which the master equation (\ref{eq:master}) has an \textit{exact} solution which describes power law inflation. In the case of power law inflation \begin{equation} \label{eq:gamma} \varepsilon +p=\gamma \varepsilon, \end{equation} where $\gamma$ is a constant. Substituting $\varepsilon $ and $p$ into the last equation we find immediately that if a solution exists, then \begin{equation} \label{eq:power-law-X} X=X_{0}=\frac{2-\gamma }{4-3\gamma }=const. \end{equation} Expressing $\varepsilon$ in terms of $p$ from (\ref{eq:gamma}) and substituting (\ref{eq:new-lagrangian}) with $X$ given by (\ref{eq:power-law-X}) into the master Eq. (\ref{eq:master}), we get a simple equation for $f$, which is solved by \begin{equation} \label{eq:power-law-f} f\left( \varphi \right) =\frac{4}{9}\frac{\left( 4-3\gamma \right)}{\gamma ^{2}}\frac{1}{\left(\varphi -\varphi_*\right) ^{2}}. \end{equation} Therefore, a model with Lagrangian (\ref{eq:new-lagrangian}), with $f\left(\varphi\right)$ given by Eq. (\ref{eq:power-law-f}), has an attractor solution which describes power law expansion, \begin{equation} \label{eq:power-law-a} a\left(t\right) \varpropto t^{\frac{2}{3\gamma}}. \end{equation} If $0<\gamma <2/3$ then this solution describes the usual power law inflation. If one takes a negative value of $\gamma$ then one gets pole-like super-inflation in \textit{Einstein frame}. However, this pole-like inflation has a ``graceful exit problem'' which is very similar to the one of the pre-big bang scenario \cite{PBB}. We were unable to find a simple solution to this problem and we doubt that such a solution can be meaningfully discussed within the effective field Lagrangian formalism considered here. Therefore we think that pole-like inflation does not help toward bringing a solution of the main cosmological problems. In distinction from pole-like inflation, in the model of power law inflation the graceful exit problem can be easily solved if in some range of $\varphi$ the function $f$ is modified in an obvious way. A natural generalization of the Lagrangian (\ref{eq:new-lagrangian}), \begin{equation} \label{eq:general-lagrangian} p=f\left( \varphi \right) g\left(X\right), \end{equation} where $g$ is a rather arbitrary function of $X$, opens the possibility to realize power law inflation in a wide class of theories. Actually, taking the function $f$ to be \begin{equation} \label{eq:generalized-f} f\left( \varphi\right) =-\frac{4}{9}\frac{\left(1-\gamma\right)}{\gamma ^{2}}\frac{2X_{0}}{g\left( X_{0}\right)}\frac{1}{\left(\varphi-\varphi _*\right)^2}, \end{equation} where $X_{0}$ is a solution of the equation \begin{equation} \label{eq:solX} 2X\frac{\partial\ln g}{\partial X}=\frac{\gamma}{\gamma-1}, \end{equation} one can easily verify that power law inflation (\ref{eq:power-law-a}) is a solution of the corresponding theory. \section{Stability} \setcounter{equation}{0} A detailed study of the spectrum of quantum perturbations in our slow-roll \textit{k}-inflation scenario will be done in a forthcoming publication. We shall only discuss here a necessary condition for the \textit{stability} of our models with respect to arbitrary, high-frequency scalar perturbations. The equation for the canonical ``quantization variable'' $v$ describing the collective metric and scalar field perturbations in the case of the action (\ref{eq:einstein-action}) can be written down in the standard way \cite{MFB} and takes the form \cite{GM99} \begin{equation} \label{eq:perturbations} v''-c_s^2\nabla ^2 v-\frac{z''}{z}v=0 , \end{equation} where the only relevant piece of information for our stability analysis is the appearance of the ``speed of sound'' \begin{equation} \label{eq:sound-velocities} c_{s}^{2}=\frac{p_{,X}}{\varepsilon_{,X}}=\frac{p_{,X}}{p_{,X}+2Xp_{,XX}} \end{equation} in front of the Laplacian. Here a comma denotes a partial derivative with respect to $X.$ It is clear that if $c_{s}^{2}$ is negative then the model is absolutely unstable. The increment of instability is inversely proportional to the wavelength of the perturbations, and therefore the background models for which $c_{s}^{2}<0$ are violently unstable and do not have any physical significance. This stability requirement ($c_{s}^{2}=p_{,X}/\varepsilon _{,X}>0$) is non trivial within our scenario, because, for instance, in slow-roll \textit{k}-inflation in the zeroth-order slow-roll approximation the inflationary attractors are defined by $p_{,X}=0$, and therefore $c_{s}^{2}=0$. However, as discussed in Section \ref{sec:slow-roll-inflation}, in the post-slow-roll approximation, with $X=X_{\rm sr}+\delta X$, $p_{,X}$ does not vanish. To first order in $\delta X$ we can write $p_{,X}\simeq p_{,XX}^{\rm sr}\delta X$. Using the second equation (\ref{eq:sound-velocities}) we get \begin{equation} c_{s}^{2}\simeq \frac{\delta X}{2X_{\rm sr}}. \end{equation} Therefore stability requires that $\delta X>0$, i.e. that on the equation of state graphs of Figs. \ref{fig:state} and \ref{fig:model-state}, the $% \varphi $-gradients of $p(\varphi,X)$ be such that they displace the real, non-adiabatic, slow-roll attractor \textit{beyond} the $p=-\varepsilon $ line (``beyond'' meaning here ``further away'' as one runs along the $% p=f(\varepsilon ,\varphi )$ graph following the natural $X$ parametrization). These stable stretches of the $% (\varepsilon ,p)$ graphs are labelled $s$ in Fig. \ref{fig:state}. They are also the stretches where the slope $(dp/d\varepsilon)$ is positive (as is clear from Eq. (\ref{eq:sound-velocities}) which says that the velocity of sound is given by the usual formula $c_{s}^{2}=dp/d\varepsilon$, when keeping $\varphi$ fixed). Let us now discuss the simple model (\ref{eq:simplest-p-expansion}). For this model we have computed $\delta X/X_{\rm sr}$, and we can therefore assess under what conditions \textit{k}-inflation will be stable under scalar perturbations. We see from Eq. (\ref{eq:slow-roll-parameter}) that the conditions can be expressed in two (equivalent) forms: (i) the energy density $\varepsilon=\sqrt{\overline{K}}$ must \textit{decrease} during the slow-roll of $\varphi$, or (ii) $-\sigma\overline{K}_{,\varphi}=\sigma K_{,\varphi}$ must be positive. The satisfaction of this condition is very natural within the intuitive picture we have in mind: namely, starting at some high ($\sim$ Planckian) energy density, i.e. a large \textit{negative} value of $K(\varphi)=-\varepsilon^2$, and then letting $\varphi$ evolve toward the weak-field coupling domain where $K(\varphi)$ vanishes before becoming positive. During the slow-roll phase (with $K(\varphi)<0$), it is natural (and even necessary if $K(\varphi)$ is monotonic) to have a decreasing $\overline{K}=-K$. One consequence applicable to a general model $p(\varphi,X)$ is that slow-roll implies a small value $|c_{s}^{2}|\ll 1$. It is interesting to ask for which (non slow-roll) models one can have both a continued \textit{k}-inflation and a constant speed of sound of order unity. Let us consider for that purpose power law inflation. In the case where the Lagrangian takes the form (\ref{eq:new-lagrangian}) the speed of sound during the inflationary stage is \begin{equation} \label{eq:power-law-cssq} c_{s}^{2}=\frac{\gamma}{8-3\gamma}. \end{equation} If we restrict ourselves to the inflationary range $0<\gamma<2/3$ this speed can not exceed $c_{s}^{2}=1/9.$ The smaller values of $\gamma $ correspond to very fast (nearly exponential) expansion and small speed of sound in complete agreement with our analysis of slow-roll \textit{k}-inflation. Pole-like inflation is violently unstable in this model. However, if we consider more general Lagrangians (\ref{eq:general-lagrangian}) we can avoid these restrictions. Actually in this case the speed of sound during the inflationary stage is given by the expression \begin{equation} c_{s}^{2}=\frac{g_{,X}\left( X_{0}\right)}{g_{,X}\left(X_{0}\right) +2X_{0}\,g_{,XX}\left(X_{0}\right)}, \end{equation} where $X_{0}$ is the solution of equation (\ref{eq:solX}). The necessary conditions for power law inflation given in Section \ref{sec:power-law-kinflation} imply that $g_{,X}\left( X_{0}\right) \not= 0$ and do not involve any restrictions on the second derivative $g_{,XX}\left( X_{0}\right)$. Therefore, for a power law inflationary stage with any a priori given value of the parameter $\gamma$, one can always find a corresponding function $g\left(X\right)$ to arrange any required speed of sound. Note that it follows from here that one can also easily build a theory with pole-like inflation which is stable with respect to scalar perturbations. \section{Exit Mechanisms} \label{sec:exit-mechanisms} \setcounter{equation}{0} In the simple model of Section \ref{sec:slow-roll-inflation} (and in the normalization $L(\varphi)=1$) the total number of inflationary $e$-folds is (considering for definiteness that $\varphi$ decreases during slow-roll) \begin{equation} N_{\rm inf}=\frac{1}{2}\int\limits_{\varphi_{\rm end}}^{\varphi_{\rm in}} \sqrt{\overline{K}(\varphi)}\,d\varphi. \end{equation} Here $\varphi_{\rm in}$ is the initial value of $\varphi$, and $\varphi_{\rm end}$ the end of slow-roll, i.e. the value of $\varphi$ where $\delta X/X$ becomes of order unity, i.e. (from Eq. (\ref{eq:slow-roll-parameter})), such that \begin{equation} \label{eq:slow-roll-end} \overline{K}^{-3/2}(\varphi_{\rm end})\overline{K}^{\prime}(\varphi_{\rm end})\sim 1. \end{equation} We shall not investigate here the problem of the choice of initial conditions within our model. We shall assume that some large parameter is present (or at least possible) in the problem and allows $N_{\rm inf}$ to be larger that 60 or so. [One simple possibility would be a function $K(\varphi )$ of the type of Eq. (\ref{eq:FMS-K}) which levels off to a negative constant when $\varphi $ increases. All the couples $(\varphi,X)$ where $X\sim 1$ (in Planck units) and $\varphi $ is arbitrary large lead to an energy density of order $1$. A random initial condition with $\varepsilon (\varphi_{\rm in},X_{\rm in})\sim 1$ could have an arbitrary large $\varphi$.] Assuming this we note not only that our mechanism contains a natural exit from inflation (because of the evolution of $K$ and its final change of sign), but that this exit is generically expected to take place within a small number of $e$-folds. Indeed, condition (\ref{eq:slow-roll-end}) signalling the end of slow-roll \textit{k}-inflation can be rewritten (using Eqs. (\ref {eq:X-slow-roll})-(\ref{eq:H-slow-roll})) as $[\dot{K}/(KH)]_{\rm end}\sim 1$, which means that $K$ changes by 100\% in a Hubble time, around the time of exit of \textit{k}-inflation. One therefore expects that one can approximately match slow-roll \textit{k}-inflation with a post-inflationary phase where $K(\varphi )$ has become positive. We think that, in most cases, this way of exiting \textit{k}-inflation provides a naturally graceful exit. Indeed, it is clear from Fig. \ref{fig:model-state} that, after the transition to the $K>0$ (``weak-coupling'') branch, the cosmological evolution will quickly be attracted toward an approximate $p\approx\varepsilon$ equation of state. The corresponding expansion was discussed in Eq. (\ref{eq:vacuum-attractor}) and corresponds to a very fast decrease of the energy density: $\varepsilon_{\varphi}\propto a^{-6}$. As this decrease is much faster than the decay of the energy density in radiation ($\varepsilon_{\rm rad}\propto a^{-4}$) [and in non relativistic matter ($\varepsilon_{\rm matter}\propto a^{-3}$) if any is present], even small traces of the latter forms of energy present at the end of \textit{k}-inflation will ultimately dominate the expansion. The situation is very similar to what has been recently discussed in Ref. \cite{PV98}. As we have in mind that, in our model, the scalar $\varphi$ could be the dilaton (or a moduli), i.e. a field which modifies the coupling constants of all the other matter fields, we expect that the nearly uniform time variation of $\varphi$ during \textit{k}-inflation will generate quantum particles at a uniform spacetime rate. During \textit{k}-inflation the produced particles are constantly diluted by the fast expansion and are not expected to cause a strong back reaction, but the particles produced in the last $e$-fold of \textit{k}-inflation should be sufficiently numerous to dominate soon the expansion. Even without this assumption (that the couplings of $\varphi$ are efficient in producing particles), the mere effect of the variable gravitational coupling (at the end of inflation) is sufficient to create any scalar particle with energy density (at birth) $\sim 10^{-2}H_{\rm end}^4$ \cite{Ford} \cite{DV96}. To discuss more precisely the evolution of the inflaton after it exits from \textit{k}-inflation, i.e. when $K(\varphi)$ has become positive, and the higher-derivative term $L\,X^2$ has become negligible, it is convenient to introduce the canonical scalar field $\phi=\int d\varphi \sqrt{K(\varphi)}$. In terms of $\phi$ the equation of motion reads $\ddot{\phi}+3H\dot{\phi}% =a^{-3}d(a^3\dot{\phi})/dt\approx 0$, with $H=\sqrt{\varepsilon_\phi+% \varepsilon_{\rm rad}}$. Hence $a^3\dot{\phi}\simeq c=\mathrm{const}$ and $\phi$ evolves according to $\phi(t)\approx c\int dt\,a^{-3}$. In a first phase after \textit{k}-inflation $\varepsilon_\phi$ probably dominates over $\varepsilon_{\rm rad}$ and the evolution follows the $p=\varepsilon$ attractor solution Eq. (\ref{eq:vacuum-attractor}). During this initial phase $a(t)\propto t^{1/3}$ so that $\phi(t)$ drifts logarithmically: $\phi(t)\propto \int_{t_{\rm end}}^t dt/t=\ln(t/t_{\rm end})$. Later, $\varepsilon_{\rm rad}\propto a^{-4}$ will take over $\varepsilon_\phi\propto a^{-6}$ and the evolution will become radiation dominated. In this second phase, $a(t)\propto t^{1/2}$ and $\phi(t)\simeq c\int^t dt a^{-3}\propto \int^t dt\,t^{-3/2}$ stops drifting logarithmically to converge toward some final value $\phi_f$: $\phi(t)\approx \phi_f-c^{\prime}t^{-1/2}$. Note that, in this generic exit mechanism, the final value $\varphi_f$ of the original field (corresponding to the final value $\phi_f$ of the canonical field) is arbitrary. Therefore, if $\varphi$ is the dilaton or a moduli our \textit{k}-inflationary mechanism does not, by itself, provide a mechanism for fixing $\varphi$ to a particular value. To do that one must appeal either to the presence, at low-energies, of a $\varphi$-dependent potential energy term $V(\varphi)$ (which may have been negligible at high energies), or to a non-trivial structure of couplings to matter \cite{DP94}. We wish, however, to point out that some variants of our general model can also provide another way of fixing the end location of $\varphi$ very near a particular value. Indeed, if the kinetic function $K(\varphi)$ (of the original $\varphi$ variable) happens to have a pole singularity $% K(\varphi)\propto (\varphi-\varphi_{p})^{-\alpha}$ with $\alpha\geq 2$ in the positive-$K$ domain the corresponding canonical field $\phi=\int d\varphi \sqrt{K(\varphi)}$ diverges when $\varphi\to\varphi_{p}$. Therefore, if this pole singularity is in the way of the evolution of $\varphi$ after the exit from \textit{k}-inflation (e.g., if $\varphi_{p}<\varphi_{\rm end} <\varphi_{\rm in}$ in the case where $\dot{\varphi}<0$), the typically large logarithmic drift of the canonical field $\phi$ during the first phase after exit, $\Delta \phi\simeq c\int_{t_{\rm end}}^{t_{\rm rad}}dt\,a^{-3}\propto \ln(t_{\rm rad}/t_{\rm end})$ (where $t_{\rm rad}$ denotes the beginning of radiation domination), will mean that the original field $\varphi$ will end up very near $\varphi_p$. Evidently, this mechanism assumes that, except for $K(\varphi)$, all the functions describing the coupling of $\varphi$ to matter (like the one giving the $\varphi$-dependence of the gauge couplings) are regular (or, at least, less singular) at $\varphi=\varphi_p$. Then, in the notation of Ref. \cite{DP94}, the observable coupling strength of $\varphi$ to matter $\alpha_A(\varphi)=d\ln m_A(\phi)/d\phi=(K(\varphi))^{-1/2}d\ln m_A(\varphi)/d\varphi$ is driven near zero by our mechanism. \section{Conclusions} \setcounter{equation}{0} We have pointed out that a general class of higher-order scalar kinetic terms $p(\varphi,\nabla \varphi)=\frac{1}{2}K(\varphi)(\nabla\varphi )^{2}+\frac{1}{4}L(\varphi)(\nabla\varphi)^{4}+\cdots$ can drive an inflationary evolution starting from rather arbitrary initial conditions. Under not very restrictive conditions on the $\varphi $-dependence of the kinetic term $p(\varphi,\nabla\varphi)$ the early cosmological evolution will be attracted toward a slow-roll kinetically driven inflationary stage. In a large class of models this slow-roll behaviour has the following attractive features: (i) it drives the evolution from an initial high-curvature phase down to lower curvatures while dilating space in a quasi-exponential or power law manner, (ii) it is stable under (high-frequency) scalar perturbations, and (iii) it contains a natural exit mechanism because of the $\varphi$-dependence of the kinetic terms. We have briefly discussed the exit of kinetically driven inflation and found that it seems to be naturally ``graceful'' in lending itself to a smooth transition toward a stage dominated by the radiation produced (either through the $\varphi$-dependence, or through purely gravitational effects) at the end of slow-roll. We have also pointed out that the presence of pole-like singularities in the $\varphi$-dependence of the kinetic terms can have some useful consequences: (i) a $\varphi$-dependence of the form $\sim (\varphi-\varphi_*)^{-2}\tilde{p}(\nabla\varphi)$ in the initial field domain can ensure a nearly constant speed of sound of order unity, while (ii) a pole-like singularity in the coefficient $(\nabla\varphi)^2$ in the final field domain can help to fix the value of $\varphi$ to a specific value. We leave to future work the investigation of general important issues: the choice of (a measure on the) initial conditions, and the computation of the perturbation spectra generated by this new type of inflation. We are aware that the compatibility of the latter issues with observations will probably necessitate the presence of some small (or large) dimensionless parameters in $p(\varphi,\nabla\varphi)$. In this work, we have used the structure of the effective action in string theory as a partial motivation for considering higher-order kinetic terms $p(\varphi,\nabla\varphi)$. Having found that such terms can generically drive an inflationary behaviour, we hope that our mechanism might be useful in suggesting new ways in which the dilaton and moduli fields of string theory might be compatible with inflation. \section*{Acknowledgements} The work of V.M. and C.A.P. was supported in part by the \textit{Sonderforschungsbereich 375-95 der Deutschen Forschungsgemeinschaft}. T.D.'s work was partly supported by the NASA grant NAS8-39225 to Gravity Probe B. C.A.P. thanks the \textit{Institut des Hautes \'Etudes Scientifiques} for its kind hospitality.
\section{Introduction} It is generally believed that the hierarchical structure today seen in the universe is generated by gravitational instability of cold dark matter (CDM) whose density is about 10 times higher than that of baryonic matter, and galaxies form in virialized dark matter haloes by cooling of baryonic gas and subsequent star formation. The characteristic mass range of galaxies is about $10^{8-12}M_\odot$, which can be understood as a mass range in which baryonic gas can cool sufficiently within a dark matter halo to form a galaxy (e.g., Blumenthal et al. 1984 for a review). However, spheroidal components reside only in relatively massive galaxies with $M = 10^{10-12} M_\odot$ and less massive galaxies are mostly irregular types or late-type spiral galaxies (Burstein et al. 1997; Hunter 1997). [There is a population called as dwarf spheroidal galaxies, but they are generally considered to be different from giant elliptical galaxies and bulges. They are more similar to dwarf irregulars (Bender et al. 1992; Mao \& Mo 1998).] Elliptical galaxies and bulges appear to form a very uniform, old stellar population with very little scatter in metallicity and star formation history, and they are generally considered to have formed by intensive starbursts at high redshifts (e.g., Renzini 1999). The trigger mechanism for starbursts is unknown, and it is often assumed to be mergers of disk galaxies when galaxy formation is modeled in the context of hierarchical structure formation in the CDM universe (Kauffman, White, \& Guiderdoni 1993; Baugh, Cole, \& Frenk 1996). However, there is no evidence for spheroidal galaxy formation through mergers, and the merger hypothesis does not give a clear explanation for the fact that all spheroidal galaxies are very old and only in massive galaxies. These trends appear to run opposite to the expectation of structure formation in the CDM universe, in which smaller objects form earlier and then massive objects later. This is currently considered as a major challenge for galaxy formation in the CDM universe (e.g., Renzini 1999), and it is interesting to seek for a mechanism which triggers starbursts only in massive and high-redshift objects. In this letter we show that a possible strong dependence of star formation activity on magnetic field strengths, which is suggested by observations of magnetic fields in local galaxies, gives a candidate for such a mechanism. \section{Galactic Magnetic Fields and Star Formation Activity} Interstellar magnetic fields in galaxies are strong enough to significantly affect interstellar gas dynamics on both global scales characteristic of galactic structure and small scales characteristic of star formation (see, e.g., Vall\'ee 1997; Zweibel \& Heiles 1997 for a review). Fitt and Alexander (1993; hereafter FA) derived strengths of volume-averaged magnetic fields for 146 spiral galaxies, which is a complete sample including various types from Sa to irregular galaxies, and they found a strange result: the dispersion in strength is surprisingly narrow within a factor of less than 2 with an average strength of 3 $\mu$G, in spite of a wide range of absolute radio luminosity of galaxies (almost 4 orders of magnitude). Why is magnetic field strength so uniform at $\sim 3 \mu$G for various types of galaxies with such a wide range of luminosity? Clearly there should be a physical reason for this strange fact. Vall\'ee (1994) investigated the relation between the magnetic field strength and star formation activity by using 48 galaxies out of the FA sample, and found that there is a correlation between magnetic field strength ($B$) and star formation efficiency (SFE) as B $\propto ({\rm SFE})^i$ with $i=0.13 \pm 0.04$. [The same relation was found also for star formation rate (SFR) and $B$.] From this result, one may conclude that star formation activity does not strongly affect the strength of magnetic fields. However, it is also likely that star formation rate {\it is} affected by magnetic fields. If we take the result of Vall\'ee in this context, the observed correlation suggests that star formation activity depends quite sensitively on $B$ as SFE $\propto B^q$ with $q = 1/i \sim 7.7 \pm 2.6$. Here we point out that, if this is the case, the surprisingly narrow dispersion in galactic magnetic fields is naturally explained by an observational selection effect. If the strength of magnetic field in an object is significantly smaller than a few $\mu$G, the star formation activity becomes much lower than in typical spiral galaxies, and such object cannot produce enough stars to be observed as a galaxy. On the other hand, if magnetic fields are stronger than a few $\mu$G, such objects would experience strong starburst and all interstellar gas will be converted into stars. If the star formation time scale is shorter than the dynamical time scale for disk formation, they will become present-day spheroidal systems. Magnetic fields in elliptical galaxies are difficult to measure since there are no relativistic electrons to illuminate the magnetic fields by synchrotron radiation. Then the reason why the observed dispersion of magnetic fields is quite small can be understood by an observational selection effect; we cannot observe or measure magnetic field strength when the strength is significantly different from the typically observed value of $\sim 3 \mu$G. Therefore the observed correlation between $B$ and SFE, as well as the surprisingly narrow dispersion of measured values of magnetic field strength, well suggests a strong dependence of star formation activity on magnetic field strength. This idea is further supported by an amorphous spiral galaxy M82, which has abnormally large magnetic field of $\sim 10 \mu$G in the sample of FA. M82 is actually known as an archetypal starburst galaxy, and the strong field strength suggested by FA is caused by the very strong field ($\sim 50 \mu$G) in the nuclear region of this galaxy, where intense star formation is underway (Klein, Wielebinski, \& Morsi 1988; Vall\'ee 1995a). Unfortunately our knowledge for physics of star formation is poor and it is not clear from a theoretical point of view why SFE depends so strongly on $B$, but it is not unreasonable. In fact, at least one effect is known by which magnetic fields help star formation: the magnetic braking (e.g., McKee et al. 1993 for a review). In order for a protostellar gas cloud to collapse into stars, a significant amount of initial angular momentum must be transported outward. Magnetic fields play an important role for this angular momentum loss by Alfv\'en waves launched into ambient medium. Strongly magnetized objects with large scale high gas density would be followed by magnetic braking of individual regions with angular momentum losses, ending in outgoing Alfv\'en waves, gas collapse, and star formation on a large scale. It should be noted that the above hypothesis does not contradict with the well-known Schmidt law of star formation. Vall\'ee (1995b, 1997) found a relation between magnetic field strengths and interstellar gas density on a scale of a whole galaxy as $B \propto n^k$ with $k = 0.17 \pm 0.03$. By using this relation and $B$-SFR relation, we find SFR $\propto n^{k/i}$ with $k/i = 1.3 \pm 0.1$. This relation agrees well (Vall\'ee 1997) with some direct estimates of the relation between SFR and $n$ (e.g., Kennicutt 1989; Shore \& Ferrini 1995; Niklas \& Beck 1997). At the same time, these relations suggest that the thermal energy of interstellar matter is not in equipartition with the magnetic field energy or cosmic ray energy (Kamaya 1996). It is possible that what is physically affected by magnetic fields is gas density, and then the gas density determine SFE by the Schmidt law, producing the observed $B$-SFE relation. Anyway, the correlation between $B$ and SFE has actually been observed, and it is natural to consider this relation holds also in high redshift objects. [The chance probability to observe such a correlation from an uncorrelated parent population is only 12\% (Vall\'ee 1994).] Therefore the following analysis is valid unless some unknown processes violate this empirical relation in high redshift objects. \section{Magnetic Fields in Hierarchical Structure Formation} As discussed above, it is suggested that spheroidal components of galaxies formed by starbursts induced by strong magnetic fields. In the following we consider the magnetic field generation and formation of spheroidal systems in the framework of the standard cosmological structure formation in the CDM universe, and show that spheroidal systems form only in relatively massive objects at high redshifts. Recently it is discussed that galactic magnetic fields are generated at the stage of collapse of protogalactic clouds (Kulsrud et al. 1997). Although there are some other scenarios as the origin of galactic magnetic fields, such as earlier field generation on cosmological scales or dynamo amplification after galaxies form (see, e.g., Zweibel \& Heiles 1997), we assume here that magnetic fields are generated during collapse of protogalactic clouds. When a dark matter halo decouples from the expansion of the universe and collapses into a virialized object, the gravitational energy of baryonic gas is converted into turbulent motions and thermal energy by generated shocks. It can be shown that a very weak magnetic field ($\sim 10^{-21}$ G), which has been generated by thermoelectric current before collapse (i.e., the battery mechanism), is amplified up to a strength nearly in equipartition with turbulent energy (Kulsrud \& Anderson 1992; Kulsrud et al. 1997). The time scale for equipartition is given by $\sim r/\upsilon$, where $r$ is the length scale of the system and $\upsilon$ the turbulent velocity. (The essence of this result can be understood by a dimensional analysis. The equation of magnetic field generation is $d{\bf B}/dt = {\rm rot}({\bf v \times B})$, where ${\bf v}$ is the velocity field of fluid, and when we estimate rot by the inverse of the system scale, it is clear that the field evolution time scale is given by $\sim r/\upsilon$.) If we take the radius and three-dimensional velocity dispersion of a virialized halo as $r$ and $\upsilon$, it is straightforward to see that this time scale is given by the dynamical time scale of the halo. Then we can estimate the strength of magnetic fields by the turbulent energy density of baryonic gas in a collapsed object, which is determined by using the well-known spherical collapse model (e.g., Peebles 1980). For a dark halo with mass $M_h$ and formation redshift $z$, the turbulent energy density becomes $\epsilon_B \sim (3 \Omega_B M_h \upsilon^2 / 8 \pi \Omega_0 r^3) = 6.4 \times 10^{-13} h^{8/3} M_{12}^{2/3} (1+z)^4 \Omega_B \rm \ erg \ cm^{-3}$, where $h$ is the Hubble constant normalized at 100 km/s/Mpc, $M_{12} = M_h / (10^{12} M_\odot)$, and $\Omega_B$ and $\Omega_0$ are the baryon and matter density in units of the critical density in the universe. In the second expression above, we have assumed the Einstein-de Sitter universe ($\Omega_0 = 1$), but extension to other cosmological models is easy. According to the theory of Kulsrud and Anderson (1992), we assume that one-sixth of the turbulent energy is converted into magnetic field energy and then the equipartition magnetic field becomes $B \sim 1.6 h^{4/3} M_{12}^{1/3} (1+z)^2 \Omega_B^{1/2} \ \mu$G. The observational properties of disk galaxies including our Galaxy are well understood if they are considered to have formed at $z \sim$ 0--1 (Mao \& Mo 1998), and hence this theory of magnetic field generation gives a roughly correct strength for our Galaxy with $M_h \sim 10^{12} M_\odot$. \section{Emergence of the Hubble Sequence} Then spheroidal galaxies are expected to form at high redshifts in massive dark matter halos, because magnetic field becomes stronger with increasing mass and redshift as $B \propto M_h^{1/3} (1 + z)^2$. In order to discuss more quantitatively, we equate the previously defined star formation efficiency (in \S 2) to $\nu = (t_{SF})^{-1}$, where $t_{\rm SF}$ is the time scale on which interstellar gas is converted into stars. The star formation rate $\dot M_*$ in a galaxy is given by $\dot M_* = M_{\rm gas}/t_{\rm SF}$, where $M_{\rm gas}$ is the mass of interstellar gas in the galaxy. As mentioned earlier, the observed relation between $B$ and $\nu$ is $\nu \propto B^q$ with $q \sim 7.7 \pm 2.6$. We set the normalization of this relation by SFE at our Galaxy. The star formation rate in our Galaxy is about a few $M_\odot$/yr and mass of interstellar gas in the disk is about $6 \times 10^9 M_\odot$ at present, suggesting that $t_{\rm SF} \sim 2$ Gyr at $B \sim 3 \mu$G (e.g., Binney \& Tremaine 1987). We also normalize the strength of $B$ to 3 $\mu$G for a halo whose baryonic mass is $6 \times 10^{10} M_\odot$ and whose virialization occurred at $z=1$, supposing our Galaxy. Then we can calculate the ratio of baryonic dynamical time to $t_{\rm SF}$, $\Gamma \equiv t_{\rm dyn} / t_{\rm SF}$, which can be considered as a criterion for spheroidal formation. In calculation of $t_{\rm dyn}$, we use a baryon density $\lambda^{-3}$ times higher than the value at virialization, considering the contraction of baryons due to cooling and dissipation, where $\lambda \sim 0.05$ is typical dimensionless angular momentum of dark haloes (Warren et al. 1992). If $\Gamma \gg 1$, spheroidal systems are expected to form. In Figure \ref{fig:hubble}, we have plotted the contour of $\Gamma$ = 1, 10, and 100 by thin solid lines in a plane of baryon mass of a collapsed object ($M_{\rm baryon}$) and its formation redshift, assuming $q=5$. (Observed mass of galaxies is considered to be mainly baryonic mass.) For reference, we have also plotted the contour of magnetic field strengths in Fig. \ref{fig:mag}. We have used a flat, $\Lambda$-dominated cosmological model, where $\Lambda$ is the cosmological constant, with the standard cosmological parameters of $(h, \Omega_0, \Omega_\Lambda, \sigma_8) = (0.7, 0.3, 0.7, 1)$. This universe is well consistent with various observations such as ages of globular clusters, high redshift type Ia supernovae (Perlmutter et al. 1998), cosmic microwave background (Bunn \& White 1997), and abundance of clusters of galaxies (Kitayama \& Suto 1997). In the following we discuss the emergence of the Hubble sequence using Fig. \ref{fig:hubble}. Objects lying above the thin solid lines are expected to form spheroidals, but we have to check how such objects are typical in the context of cosmological structure formation in the CDM universe. For this purpose we have plotted mass scale of $n$-$\sigma$ fluctuation as defined in Blumenthal et al. (1984), by dashed lines. The lines are defined by $n \sigma(M_h, z) = \delta_c$, where $\sigma(M_h, z)$ is the root-mean-square of density fluctuation predicted by linear theory in the CDM universe at scale of $M_h$ and at redshift $z$ (Peacock \& Dodds 1994; Sugiyama 1995), and $\delta_c \sim 1.69$ is a critical density contrast at which an object virializes (Peebles 1980). The three dashed lines correspond to $n$ = 1, 2, and 3, and these lines represent typical mass scales of collapsed objects as a function of $z$. (For larger $n$, the cosmological abundance of objects is statistically suppressed as $\propto e^{-n^2/2}$.) In the region where the dashed lines are above the solid lines, spheroidal galaxies can form as cosmologically typical objects. Figure \ref{fig:hubble} shows that objects with $M_{\rm baryon} \gtilde 10^{9-10} M_\odot$ with $\sim 2$--$3 \sigma$ fluctuation will form spheroidal systems because $\Gamma \gg 1$, at high redshifts of $z \sim$ 3--10 with $B \sim 50 \mu$G (see also Fig. \ref{fig:mag}). It is interesting to note that this magnetic field strength is roughly the same with that in the starbursting nucleus of M82, as mentioned in \S 2. Spheroidal formation with $M_{\rm baryon} \ltilde 10^{9-10} M_\odot$ is inhibited because of low $\Gamma$ for cosmologically typical objects. On the other hand, if $M_{\rm baryon}$ is larger than $\sim 10^{12} M_\odot$, formation of galaxies is inhibited by too long cooling time ($t_{\rm cool}$) compared to the Hubble time at each redshift ($t_H$), as shown by dotted lines which are contours of $t_{\rm cool}/t_H$ (Blumenthal et al. 1984). Therefore, our model can explain the observed mass range ($10^{10-12} M_\odot$) and old stellar populations seen in spheroidal galaxies. The region for spheroidal galaxy formation is schematically indicated in Figs. \ref{fig:hubble} and \ref{fig:mag} by the thick solid line. After spheroidal galaxy formation, gas accretion onto some of them is possible at lower redshifts. Since such gas accreting recently is located below the thin solid lines in Figure \ref{fig:hubble}, it results in disk formation of spiral galaxies at $z \ltilde 1$. On the other hand, if recent gas accretion is negligible, such object will be seen as present-day elliptical galaxies. Since solid lines (constant $\Gamma$) and the dashed lines ($n$-$\sigma$ lines) are approximately parallel in the region of spheroidal formation, our scenario for the origin of the Hubble sequence predicts that elliptical galaxies should lie on a constant $n$-$\sigma$ density fluctuation with $n \sim$ 2--3, and with decreasing $n$, the type of galaxies becomes later in the Hubble sequence, i.e., from early to late spiral, and then into irregular galaxies at $\sim 1 \sigma$ fluctuation. In fact, this trend is exactly what has been observed in galaxies (Blumenthal et al. 1984; Burstein et al. 1997), giving a further support for our scenario. Most properties or relations observed in present-day galaxies, such as the Tully-Fisher or Faber-Jackson relation, existence of the fundamental plane for galaxies and their distribution on it, can be explained by formation of early type galaxies from higher $n$-$\sigma$ fluctuations than later Hubble types (Burstein et al. 1997). The density-morphology relation, which is the correlation between galaxy types and number density, is also explained; higher $n$-$\sigma$ fluctuations occur preferentially in denser regions destined to become rich clusters, and hence one expects to find more ellipticals there, as is observed (Blumenthal et al. 1984). Higher $n$-$\sigma$ objects are expected to show stronger spatial clustering than lower ones (e.g., Mo \& White 1996), and it is consistent with the stronger clustering observed for elliptical galaxies than late type galaxies (Loveday et al. 1995). Various observations suggest that giant galaxies formed at higher redshifts of $z \gtilde 3$, and were then followed by a sequence of less and less massive galaxies forming at lower and lower redshifts, leading down to the formation of dwarfs at recent ($z \ltilde 0.5$) (Fukugita, Hogan \& Peebles 1996; Sawicki, Lin, \& Yee 1997). The proposed scenario gives an explanation for this trend which is sometimes termed as ``downsizing'', otherwise it seems opposite to the expectation in the CDM universe. \section{Discussion \& Conclusions} For very low mass objects, star formation is strongly suppressed by the absence of physical triggers when magnetic fields are weak, and this may give an explanation for the fact that the faint end of luminosity function of galaxies is much flatter than expected from the mass function of dark haloes (Kauffmann et al. 1993; Baugh et al. 1996). Because of the strong dependence of star formation activity on magnetic fields, stellar luminosity of galaxies would quite rapidly decrease with decreasing mass of galaxies. Then it is expected that, in the faint end of galaxy luminosity function, the mass of galaxies does not so change compared to the change in luminosity. Since the number density of objects is determined by mass of objects, the faint end is expected to be relatively flat, as observed. Irregular galaxies or late-type galaxies are expected to show a wide range of star formation activity within a narrow range of galaxy masses depending on the magnetic field strength in them, and in fact such a trend has been observed (Hunter 1997). It is well known that galaxy interactions or mergers induce intensive starbursts, but the mechanism which triggers starbursts in galaxy interactions is still poorly known. Mergers are followed by cloud collisions leading to high gas density, strong turbulences and strong magnetic fields, and ending in gas collapse and star formation in clouds. Therefore the hypothesis presented in this letter, i.e., strong dependence of star formation activity on magnetic field strengths, may also be important in the merger-induced starbursts. Some fraction of elliptical galaxies may have formed by such a process, but we have argued that the observed trends of ellipticals, i.e., being massive and old, originate mainly from properties of gravitationally bound objects in the standard theory of cosmological structure formation. We have presented a new idea for the origin of the Hubble sequence of galaxies. The key of this scenario is the strong dependence of star formation activity on average magnetic fields in galaxies, which is just a speculation from a theoretical point of view, but motivated by several observational facts about magnetic fields in nearby galaxies. This only one speculation provides a simple explanation for the surprisingly narrow dispersion in the magnetic field strengths observed in spiral galaxies, and also for most properties of galaxies seen along the Hubble sequence. To verify this ``magnetic galaxy formation'' scenario as the origin of the Hubble sequence, it is indispensable to confirm observationally that starbursts are actually triggered by stronger magnetic fields. Comprehensive measurements of galactic magnetic fields in larger samples of galaxies are necessary, especially for starburst galaxies at local as well as at high redshifts. The author would like to thank T. Kitayama, M. Shimizu, and an anonymous referee for useful comments and discussions.
\section{Introduction}\label{section1} Harmonic analysis on the compact quantum group $SU_q(2)$ has been studied by several authors, see for example \cite{VS}, \cite{NM}, \cite{K2}, \cite{Koel}. One possible approach is studying the restriction of the Haar functional and of the action of the quantum Casimir to ``functions'' in $SU_q(2)$ of a given, fixed bi-$T$-type, where $T$ is the standard compact torus in $SU(2)$. This reduces the harmonic analysis to the spectral analysis of an explicit second order $q$-difference operator considered as unbounded linear operator on an explicit Hilbert space. The Hilbert space is the $L^2$-space corresponding to the orthogonality measure of the little $q$-Jacobi polynomials, while the second order $q$-difference operator is diagonalized by the little $q$-Jacobi polynomials, so this leads to an interpretation of (the orthogonality relations of) the little $q$-Jacobi polynomials on $SU_q(2)$. This approach was employed by Kakehi, Masuda and Ueno \cite{KMU} and Kakehi \cite{K} for the development of harmonic analysis on the non-compact quantum group $SU_q(1,1)$, see also Vaksman and Korogodsky \cite{KV}. The restriction of the Casimir to functions on $SU_q(1,1)$ of a given, fixed bi-$T$-type leads to the same second order $q$-difference operator as for $SU_q(2)$, while the restriction of a Haar functional on $SU_q(1,1)$ to the given bi-$T$-type leads to a $L^2$-space with respect to a discrete measure with unbounded support of the form $[0,\infty(z))_q=\{zq^k\}_{k\in\ensuremath{\mathbb{Z}}}$ ($z\not=0$), where $0<q<1$. The spectral analysis of the second order $q$-difference operator considered as unbounded linear operator on this Hilbert space was developed in \cite{KMU}, \cite{K}. It leads to the Plancherel formula and the inversion formula for the little $q$-Jacobi function transform. It is well known that harmonic analysis on $SU_q(2)$ can be generalized by replacing the role of the torus $T$ in $SU_q(2)$ by a ``conjugate'' $T_t$ ($t\in \ensuremath{\mathbb{R}}$), which is defined in terms of twisted primitive elements in the quantized universal enveloping algebra $U_q\bigl(su(2)\bigr)$, see \cite{K2}. In particular, considering the harmonic analysis on $SU_q(2)$ with respect to left $T$-types and right $T_t$-types leads to an interpretation of the big $q$-Jacobi polynomials on $SU_q(2)$, see \cite{NM}. Furthermore, the restriction of the Haar functional to functions of a given left $T$-type and right $T_t$-type can be computed directly and identified with the orthogonality measure of the big $q$-Jacobi polynomials, see \cite{KV} for the spherical case. In this paper we develop the functional analytic aspects related to the harmonic analyis on $SU_q(1,1)$ with respect to left $T$-types and right $T_t$-types. Concretely, we give a detailed description of the spectral properties of the second order $q$-difference operator $L$ which is diagonalized by the big $q$-Jacobi polynomials, considered as an unbounded linear operator on a one-parameter family of Hilbert spaces. The Hilbert spaces are $L^2$-spaces corresponding to explicit discrete measures with unbounded support $[-1,\infty(z))_q=\{-q^k\}_{k\in\ensuremath{\mathbb{Z}}_+}\cup \{zq^k\}_{k\in\ensuremath{\mathbb{Z}}}$ ($z>0$). They can be interpreted for specific parameter values as the restriction of a Haar functional on $SU_q(1,1)$ to functions of a given left $T$-type and right $T_t$-type. We construct a one-parameter family of dense domains for which $L$ is self-adjoint. With respect to one of these domains, we explicitly compute the resolution of the identity of $L$. This leads to the big $q$-Jacobi function transform, its Plancherel formula and its inversion formula. Furthermore, we show that the corresponding dual orthogonality relations lead to a one-parameter family of non-extremal orthogonality measures for the continuous dual $q^{-1}$-Hahn polynomials, and explicit sets of functions complementing these polynomials to orthogonal bases of the corresponding Hilbert spaces. The little (respectively big) $q$-Jacobi function transform is a $q$-analogue of the Jacobi function transform on the interval $[0,\infty)$ (respectively $[-1,\infty)$), see e.g. Braaksma and Meulenbeld \cite{BM}, Koornwinder \cite{K1} and references given there. In the classical setting, these two transforms are related by a dilation of the geometric parameter in which the end-point $0$ of the interval $[0,\infty)$ is mapped onto the end-point $-1$ of the interval $[-1,\infty)$. The key point in our study of the big $q$-Jacobi function transform is to show that the limiting point $0$ of the $q$-interval $[-1,\infty(z))_q$ does not play a special role in the spectral analysis, while the role of the end-point $-1$ is similar to the role of the limiting point $0$ for the little $q$-Jacobi function transform, see \cite{KMU}, \cite{K}. We will report elsewhere in detail about the connection between the big $q$-Jacobi function transform and harmonic analysis on $SU_q(1,1)$. Furthermore, we will report elsewhere on the Plancherel formula and the inversion formula for the Askey-Wilson function transform, which is associated to harmonic analysis on $SU_q(1,1)$ with respect to left $T_s$-types and right $T_t$-types ($s,t\in\ensuremath{\mathbb{R}}$). The organization of the paper is as follows. In section 2 we introduce the second order $q$-difference operator $L$ and the Hilbert spaces on which we consider $L$ as an unbounded linear operator. Furthermore, we derive dense domains for which $L$ is self-adjoint. The domains are given explicitly in terms of continuity and differentiability conditions of the functions at the origin. In section 3 we derive criteria when it is possible to extend an arbitrary eigenfunction of $L$ on the positive real axis to a global eigenfunction in such a way that the solution is ``continuously differentiable'' at the origin. We furthermore give, in terms of basic hypergeometric series, two explicit eigenfunctions which are continuously differentiable at the origin. One of them is the spherical function for the big $q$-Jacobi function transform. In section 4 we introduce the asymptotically free eigenfunction of $L$ on the positive real axis and we give the corresponding $c$-function expansion of the spherical function on the positive real axis. In section 5 we extend for generic parameters the asymptotically free solution to a global eigenfunction of $L$ in such a way, that the global eigenfunction is continuously differentiable at the origin. This leads to the $c$-function expansion of the spherical function on the whole support of the measure. In section 6 we define the Green function of $L$ in terms of the spherical function and the extended asymptotically free eigenfunction. In section 7 we use the Green function to compute the continuous contribution to the resolution of the identity of $L$. Furthermore, we derive the Plancherel formula and the inversion formula for the continuous part of the big $q$-Jacobi function transform. In section 8 we derive the discrete contribution of the resolution of the identity of $L$ and we derive the orthogonality relations and the quadratic norm evaluations of the corresponding spherical functions, which are square integrable in these cases. We furthermore give a precise description of the spectrum of $L$. In section 9 we state and prove the main results of the paper. We derive the Plancherel formula and the inversion formula for the big $q$-Jacobi function transform. Furthermore, we show that the dual orthogonality relations lead to a one-parameter family of non-extremal orthogonality measures for the continuous dual $q^{-1}$-Hahn polynomials, as well as explicit sets of functions which complement the polynomials to orthogonal bases of the corresponding Hilbert spaces. Finally we derive in section 10 a functional analytic proof of the orthogonality relations and the quadratic norm evaluations for the big $q$-Jacobi polynomials. Proofs will only be sketched in this section, we mainly emphasize the differences between the compact setting, which corresponds to $SU_q(2)$ and the big $q$-Jacobi polynomials, and the non-compact setting, which corresponds to $SU_q(1,1)$ and the big $q$-Jacobi functions. {\it Notations:} We assume throughout the paper that $0<q<1$ is fixed. We follow the notation of Gasper and Rahman \cite{GR} concerning $q$-shifted factorials and basic hypergeometric series. We write \[\theta(x_1,\ldots, x_r)=\theta(x_1)\ldots\theta(x_r) \] for products of (renormalized) Jacobi theta-products $\theta(x)=\bigl(x,q/x;q\bigr)_{\infty}$. We write $\ensuremath{\mathbb{Z}}_+=\{0,1,2,\ldots,\}$ and $\ensuremath{\mathbb{N}}=\{1,2,\ldots\}$. {\it Acknowledgements:} The research for this paper started when both authors were affiliated to the University of Amsterdam, the first author as post-doc supported by the Netherlands Organization for Scientific Research (NWO) under project number 610.06.100. Part of the research was done while the second author was a post-doc at the Centre de Math{\'e}matiques de Jussieu, Universit{\'e} Paris VI Pierre et Marie Curie, Paris, France, supported by the EC TMR network ``Algebraic Lie Representations'', Grant no. ERB FMRX-CT97-0100. The second author is also supported by a NWO-TALENT stipendium of the Netherlands Organization for Scientific Research (NWO). We thank Tom H. Koornwinder for sharing with us his private notes on the little $q$-Jacobi function transform and for his interest and valuable comments on the subject. \section{The second order $q$-difference operator}\label{sdomain} In this section we consider the second order $q$-difference operator $L$ associated with the big $q$-Jacobi polynomials as an unbounded operator on a one-parameter family of Hilbert space. We determine suitable domains of definition for which $L$ is self-adjoint. The second order $q$-difference operator $L$ depends on three parameters $(a,b,c)$. The corresponding one-parameter family of Hilbert spaces thus depends on four parameters $(a,b,c,z)$, where the parameter $z$ labels the Hilbert spaces. In sections \ref{sdomain}--\ref{PID} we assume that the four parameters $(a,b,c,z)$ satisfy $z>0$ and $(a,b,c)\in V$, where \begin{equation}\label{con1} V=\{(a,b,c) \,\, | \,\, a,b,c>0,\quad ab,ac,bc<1\}, \end{equation} unless explicitly stated otherwise. Observe that two of the three parameters $a,b,c$ take their values in the open interval $(0,1)$. We define the $q$-interval $I$ by \[ I=[-1,\infty(z))_q=\{ -q^k \, | \, k\in\ensuremath{\mathbb{Z}}_+\}\cup \{zq^k \, | \, k\in\ensuremath{\mathbb{Z}}\}, \] which is regarded as a discrete $q$-analogue of the interval $[-1,\infty)$. We use the notation $(-1,\infty(z))_q$ for the $q$-interval $I\setminus \{-1\}=[-q,\infty(z))_q$. Let ${\mathcal{F}}(I)$ be the linear space of complex-valued functions $f: I\to\ensuremath{\mathbb{C}}$, and define a linear operator $L\in {\hbox{End}}_{\ensuremath{\mathbb{C}}}\bigl({\mathcal{F}}(I)\bigr)$ by \begin{equation}\label{equiv2} L=A(\cdot)\bigl(T_q-\hbox{Id}\bigr)+B(\cdot)\bigl(T_{q^{-1}}-\hbox{Id}\bigr), \end{equation} with the $q$-shift operators defined by $\bigl(T_{q^{\pm 1}}f\bigr)(x)=f(q^{\pm 1}x)$, and with \begin{equation}\label{ABbig} A(x)=a^2\left(1+\frac{1}{abx}\right)\left(1+\frac{1}{acx}\right),\quad B(x)=\left(1+\frac{q}{bcx}\right)\left(1+\frac{1}{x}\right). \end{equation} Here $\bigl(Lf\bigr)(-1)$ is by definition given by \begin{equation}\label{special} \bigl(Lf\bigr)(-1)=A(-1)\bigl(f(-q)-f(-1)\bigr),\quad f\in {\mathcal{F}}(I), \end{equation} which is formally compatible with the fact that $B(-1)=0$. In the next lemma we rewrite the operator $L$ as a second order operator in the $q$-difference operator $D_q$, which is defined by \[\bigl(D_qf\bigr)(x)=\frac{f(x)-f(qx)}{(1-q)x}. \] \begin{lem}\label{self-adjointform} Let $f\in {\mathcal{F}}(I)$, then \begin{equation}\label{equiv1} \bigl(Lf\bigr)(x)= \begin{cases} {\displaystyle{p(x)\bigl(D_q(r(\cdot)D_qf)\bigr)(q^{-1}x)}},\qquad &x\in (-1,\infty(z))_q,\\ {\displaystyle{\frac{-qp(x)r(x)}{(1-q)x}}}\bigl(D_qf\bigr)(x), \qquad &x=-1, \end{cases} \end{equation} where \begin{equation}\label{pr} \begin{split} p(x)&= \frac{\bigl(-abx,-acx;q\bigr)_{\infty}} {\bigl(-bcx, -qx;q\bigr)_{\infty}},\\ r(x)&= \frac{(1-q)^2}{qbc} \frac{\bigl(-bcx, -qx;q\bigr)_{\infty}} {\bigl(-qabx,-qacx;q\bigr)_{\infty}}. \end{split} \end{equation} \end{lem} \begin{proof} Observe that $p$ and $r$ are well defined as functions on the $q$-interval $I$ since $(a,b,c)\in V$. The formula for $x\in (-1,\infty(z))_q$ follows by observing that $p(\cdot)T_{q^{-1}}D_qr(\cdot)D_q$ is of the form \eqref{equiv2} with $A$ and $B$ given by \[ A(x)=\frac{qp(x)r(x)}{(1-q)^2x^2}, \quad B(x)=\frac{q^2p(x)r(q^{-1}x)}{(1-q)^2x^2}. \] By inserting the explicit functions $p$ and $r$, we see that the corresponding $A$ and $B$ coincide with the ones given by \eqref{ABbig}. The formula for $Lf$ in the point $x=-1$ follows now immediately from the definition \eqref{special} of $\bigl(Lf\bigr)(-1)$. \end{proof} We have that $p(x)>0$ for all $x\in I$ since $(a,b,c)\in V$. We can thus define the Hilbert space \begin{equation} {\mathcal{H}}=\{\,\, f\in {\mathcal{F}}(I) \,\,\,\,\, | \,\,\,\,\, \|f\|^2=\langle f,f\rangle <\infty \,\,\}, \end{equation} where \begin{equation} \langle f,g\rangle = \int_{-1}^{\infty(z)}f(x){\overline{g(x)}} \frac{d_qx}{p(x)}, \end{equation} with the Jackson $q$-integral for $\alpha,\gamma\in \ensuremath{\mathbb{C}}^*$ defined by \begin{equation}\label{Jackson} \begin{split} \int_{\alpha}^{\beta}f(x)d_qx&=\int_{0}^{\beta}f(x)d_qx- \int_{0}^{\alpha}f(x)d_qx,\quad \beta=\gamma,\infty(\gamma)\\ \int_{0}^{\gamma}f(x)d_qx&=(1-q)\sum_{n=0}^{\infty}f(\gamma q^n)\gamma q^n,\\ \int_{0}^{\infty(\gamma)}f(x)d_qx&= (1-q)\sum_{n=-\infty}^{\infty}f(\gamma q^n)\gamma q^n, \end{split} \end{equation} for functions $f$ such that the sums converge absolutely. If $\alpha=q^k\gamma$ for certain $k\in \ensuremath{\mathbb{Z}}_+$, then the $q$-integral from $\alpha$ to $\gamma$ can be defined for arbitrary functions $f$, since the $q$-integral reduces then to a finite sum. We regard $L$ as an unbounded operator on the Hilbert space ${\mathcal{H}}$. In the remainder of this section, we define a one-parameter family of dense subspaces ${\mathcal{D}}\subset{\mathcal{H}}$ such that $L$, with domain of definition ${\mathcal{D}}$, is self-adjoint. In order to determine the subspaces, it is convenient to consider the symmetry of $L$ with respect to a truncated version of the inner product $\langle .,.\rangle$, which is defined for $k\in \ensuremath{\mathbb{Z}}_+$ and $l,m\in \ensuremath{\mathbb{Z}}$ with $l<m$ by \begin{equation*} \begin{split} \langle f,g \rangle_{k;l,m}&=\left(\int_{-1}^{-q^{k+1}}+ \int_{zq^{m+1}}^{zq^l}\right)f(x){\overline{g(x)}}\frac{d_qx}{p(x)}\\ &= \sum_{n=0}^kf(-q^n){\overline{g(-q^n)}}\frac{(1-q)q^n}{p(-q^n)}+ \sum_{n=l}^mf(zq^n){\overline{g(zq^n)}}\frac{(1-q)zq^n}{p(zq^n)}. \end{split} \end{equation*} Observe that for all $f,g\in {\mathcal{H}}$, we have \begin{equation}\label{lim} \lim_{\stackrel{{\scriptstyle{k,m\to\infty}}}{l\to -\infty}} \langle f,g\rangle_{k;l,m}=\langle f,g\rangle. \end{equation} We have now the following lemma. \begin{lem}\label{stoktermen} Let $f,g\in {\mathcal{F}}(I)$, $k\in \ensuremath{\mathbb{Z}}_+$ and $l,m\in \ensuremath{\mathbb{Z}}$ with $l<m$, then \begin{equation*} \langle Lf,g\rangle_{k;l,m}-\langle f,Lg\rangle_{k;l,m}= W(f,{\overline{g}})(zq^{l-1})-W(f,{\overline{g}})(zq^m) +W(f,{\overline{g}})(-q^k), \end{equation*} with ${\overline{f}}(x)={\overline{f(x)}}$ and with the Wronskian $W(f,g)\in {\mathcal{F}}(I)$ defined by \begin{equation}\label{Wronskian} \begin{split} W(f,g)(x)&= \frac{qr(x)}{(1-q)x}\bigl(f(x)g(qx)-f(qx)g(x)\bigr)\\ &= qr(x)\bigl((D_qf)(x)g(x)-f(x)(D_qg)(x)\bigr) \end{split} \end{equation} for all $x\in I$. \end{lem} \begin{proof} Since $p$ and $r$ (respectively $A$ and $B$) are real valued on the $q$-interval $I$, we may restrict the proof to real valued functions $f$ and $g$. Then we derive by a direct computation using Lemma \ref{self-adjointform} that \begin{equation*} \begin{split} \Bigl\{\bigl(Lf\bigr)(x)g(x)- f(x)\bigl(Lg\bigr)(x)\Bigr\}&\frac{(1-q)x}{p(x)}\\ =&\begin{cases} W(f,g)(q^{-1}x)-W(f,g)(x),\,\,\, &x\in (-1,\infty(z))_q,\\ -W(f,g)(-1), &x=-1. \end{cases} \end{split} \end{equation*} The lemma is now an easy consequence of this formula since the finite sums become telescoping. \end{proof} In the remainder of this section, we use some standard terminology on unbounded linear operators, see \cite[Chapter XII]{DS} and \cite[Chapter 13]{R}. Let ${\mathcal{D}}\subset {\mathcal{H}}$ be a dense linear space satisfying $L\bigl({\mathcal{D}}\bigr)\subset {\mathcal{H}}$. Then ${\mathcal{D}}$ may be considered as a domain of definition for the unbounded operator $L$ on ${\mathcal{H}}$. Lemma \ref{stoktermen} and \eqref{lim} then show that $(L,{\mathcal{D}})$ is a densely defined symmetric operator if and only if for all $f,g\in {\mathcal{D}}$, the three limits \[ \lim_{l\to -\infty}W(f,{\overline{g}})(zq^{l-1}), \quad \lim_{m\to\infty}W(f,{\overline{g}})(zq^m),\quad \lim_{k\to\infty}W(f,{\overline{g}})(-q^k) \] exist, and \[ \lim_{l\to-\infty}W(f,{\overline{g}})(zq^{l-1}) -\lim_{m\to\infty}W(f,{\overline{g}})(zq^m)+\lim_{k\to\infty} W(f,{\overline{g}})(-q^k)=0. \] The following lemma determines the behaviour of the Wronskian in the limit to infinity. \begin{lem}\label{Winfty} Let $f,g\in {\mathcal{H}}$, then $\lim_{m\to \infty}W(f,g)(zq^{-m})=0$. \end{lem} \begin{proof} Using the formula \begin{equation}\label{qshifttransform} \bigl(xq^{1-m};q\bigr)_{\infty}=(-x)^mq^{-m(m-1)/2}\bigl(1/x;q\bigr)_m \bigl(qx;q\bigr)_{\infty},\qquad x\in\ensuremath{\mathbb{C}}^*, \end{equation} it follows that the behaviour of the weights of $\langle .,.\rangle$ at infinity is given by \begin{equation}\label{pinfty} \frac{(1-q)zq^{-m}}{p(zq^{-m})}=K a^{-2m}\left(1+{\mathcal{O}}(q^m)\right),\qquad m\to\infty, \end{equation} where $K$ is the positive constant \begin{equation}\label{Kconstant} K=(1-q)z\frac{\theta(-bcz,-qz)} {\theta(-abz,-acz)}. \end{equation} This implies that $\lim_{m\to\infty}a^{-m}f(zq^{-m})=0$ for all $f\in {\mathcal{H}}$. The proof follows now from the definition of the Wronskian \eqref{Wronskian} and the fact that \begin{equation}\label{rinfty} \frac{qr(zq^{-m})}{(1-q)zq^{-m}}=Ka^{2-2m}\left(1+{\mathcal{O}}(q^m)\right), \qquad m\to\infty. \end{equation} \end{proof} It follows from Lemma \ref{Winfty} that the domains ${\mathcal{D}}\subset {\mathcal{H}}$ for which $(L,{\mathcal{D}})$ is symmetric are determined by vanishing properties of the Wronskian at the origin. Before we define suitable domains of definition for $L$ explicitly, we first introduce some convenient notations. For $f\in {\mathcal{F}}(I)$, we define \begin{equation} \begin{split} f(0^+)&=\lim_{k\to\infty}f(zq^k),\quad f(0^-)=\lim_{k\to\infty}f(-q^k),\\ f'(0^+)&=\lim_{k\to\infty}\bigl(D_qf\bigr)(zq^k),\quad f'(0^-)=\lim_{k\to\infty}\bigl(D_qf\bigr)(-q^k) \end{split} \end{equation} provided that the limits exist. {}From now on we tacitly assume that the limits exist whenever we write $f(0^+)$, $f(0^-)$ etc. Let $\ensuremath{\mathbb{T}}=\{ \alpha\in \ensuremath{\mathbb{C}} \, | \, |\alpha|=1 \}$ be the unit circle in the complex plane. \begin{Def}\label{Defdomain} Let $\alpha\in \ensuremath{\mathbb{T}}$. We write ${\mathcal{D}}_{\alpha}\subset {\mathcal{H}}$ for the subspace of functions $f\in {\mathcal{H}}$ satisfying $Lf\in {\mathcal{H}}$, $f(0^+)=\alpha f(0^-)$ and $f'(0^+)=\alpha f'(0^-)$. \end{Def} Observe that ${\mathcal{D}}_{\alpha}$ contains the functions with finite support, hence ${\mathcal{D}}_{\alpha}\subset {\mathcal{H}}$ is dense. \begin{lem}\label{suitable} Let $\alpha\in\ensuremath{\mathbb{T}}$. {\bf (i)} There exists a function $f\in {\mathcal{D}}_{\alpha}$ such that $f(0^+)\not=0$ and $\bigl(D_qf)(zq^k)=0$, $\bigl(D_qf\bigr)(-q^l)=0$ for $k,l\in\ensuremath{\mathbb{Z}}_+$. {\bf (ii)} There exists a function $g\in {\mathcal{D}}_{\alpha}$ such that $g(0^+)=0$ and $g'(0^+)\not=0$. {\bf (iii)} ${\mathcal{D}}_{\alpha}\not={\mathcal{D}}_{\beta}$ for $\alpha,\beta\in \ensuremath{\mathbb{T}}$ with $\alpha\not=\beta$. \end{lem} \begin{proof} Part {\bf (iii)} follows immediately from {\bf (i)} and Definition \ref{Defdomain}. For the proof of {\bf (i)}, we observe first that the weights of $\langle .,.\rangle$ around zero behave like \begin{equation}\label{asymtozero} \frac{(1-q)q^k}{p(-q^k)}={\mathcal{O}}(q^k), \quad \frac{(1-q)zq^k}{p(zq^k)} = {\mathcal{O}}(q^k),\qquad k\to\infty. \end{equation} Define now the function $f\in {\mathcal{F}}(I)$ by $f(x)=1$ if $-1\leq x<0$, $f(x)=\alpha$ if $0<x\leq z$ and $f(x)=0$ if $x>z$. Then $f\in {\mathcal{H}}$ follows from \eqref{asymtozero}, and $Lf\in {\mathcal{H}}$ since $(Lf)(x)=0$ if $x\not\in \{z,q^{-1}z\}$. Furthermore, $f(0^+)=\alpha=\alpha f(0^-)$ and $f'(0^+)=0=\alpha f'(0^-)$, hence $f\in {\mathcal{D}}_{\alpha}$. By construction we have $f(0^+)\not=0$ and $\bigl(D_qf)(zq^k)=0$, $\bigl(D_qf\bigr)(-q^l)=0$ for $k,l\in\ensuremath{\mathbb{Z}}_+$. For the proof of {\bf (ii)} we define the function $g\in {\mathcal{F}}(I)$ by $g(x)=x$ if $-1\leq x<0$, $g(x)=\alpha x$ if $0<x\leq z$, and $g(x)=0$ if $x>z$. Then $g\in {\mathcal{H}}$. Furthermore, $g(0^+)=0=\alpha g(0^-)$ and $g'(0^+)=\alpha=\alpha g'(0^-)$. So it remains to show that $Lg\in {\mathcal{H}}$. Using the explicit expression \eqref{equiv2} for the $q$-difference operator $L$, we see that $\bigl(Lg\bigr)(-q^k)={\mathcal{O}}(1)$, $\bigl(Lg\bigr)(zq^k)={\mathcal{O}}(1)$ as $k\to\infty$. Combined with \eqref{asymtozero}, it follows that $Lg\in {\mathcal{H}}$. \end{proof} \begin{lem}\label{symmetryoperator} Let $\alpha\in \ensuremath{\mathbb{T}}$, then $(L,{\mathcal{D}}_{\alpha})$ is a symmetric operator. \end{lem} \begin{proof} We have to show that \[ \langle Lf,g\rangle-\langle f,Lg\rangle=0, \quad \forall f,g\in {\mathcal{D}}_{\alpha}. \] By Lemma \ref{stoktermen}, Lemma \ref{Winfty}, \eqref{lim} and the fact that $h, Lh\in {\mathcal{H}}$ for $h\in {\mathcal{D}}_{\alpha}$, it suffices to show that $W(f,{\overline{g}})(0^+)=W(f,{\overline{g}})(0^-)$ for all $f,g\in {\mathcal{D}}_{\alpha}$. Since $r(0^+)=(1-q)^2/qbc=r(0^-)$, we have for all $f,g\in {\mathcal{D}}_{\alpha}$, \begin{equation*} \begin{split} W(f,{\overline{g}})(0^+)=& \frac{(1-q)^2}{bc} \bigl(f'(0^+){\overline{g(0^+)}}-f(0^+){\overline{g'(0^+)}}\bigr)\\ =&\frac{(1-q)^2}{bc}\alpha{\overline{\alpha}} \bigl(f'(0^-){\overline{g(0^-)}}-f(0^-){\overline{g'(0^-)}}\bigr)= W(f,{\overline{g}})(0^-), \end{split} \end{equation*} as desired. \end{proof} For $\alpha\in \ensuremath{\mathbb{T}}$, we write $(L^*, {\mathcal{D}}_{\alpha}^*)$ for the adjoint of the operator $(L,{\mathcal{D}}_{\alpha})$. Since $(L,{\mathcal{D}}_{\alpha})$ is a symmetric operator, we have $(L,{\mathcal{D}}_{\alpha})\subset (L^*,{\mathcal{D}}_{\alpha}^*)$. Recall that $L$ was initially defined as a linear operator on the linear space ${\mathcal{F}}(I)$ of complex-valued functions on $I$. In particular, $L$ is well defined as a linear map $L:{\mathcal{D}}_{\alpha}^*\rightarrow {\mathcal{F}}(I)$ by restriction. We claim that \begin{equation}\label{connect*} L^*=L|_{{\mathcal{D}}_{\alpha}^*}. \end{equation} To prove the claim, we observe that \begin{equation*} \langle Lf,g\rangle=\langle f, Lg\rangle \end{equation*} for functions $f,g\in {\mathcal{F}}(I)$ such that $f$ has finite support (compare with the proof of Lemma \ref{stoktermen}). For $g\in {\mathcal{D}}_{\alpha}^*$ this implies that $\langle f, Lg\rangle=\langle f, L^*g\rangle$ for functions $f$ with finite support. Applying this formula with a non-zero function $f$ having support in one point $x\in I$, we arrive at $\bigl(Lg\bigr)(x)=\bigl(L^*g\bigr)(x)$. The claim \eqref{connect*} follows, since $x\in I$ can be chosen arbitrarily. \begin{prop}\label{s} Let $\alpha\in\ensuremath{\mathbb{T}}$, then $(L,{\mathcal{D}}_{\alpha})$ is self-adjoint, i.e. ${\mathcal{D}}_{\alpha}={\mathcal{D}}_{\alpha}^*$. \end{prop} \begin{proof} Since $(L,{\mathcal{D}}_{\alpha})$ is symmetric, it suffices to prove the inclusion ${\mathcal{D}}_{\alpha}^*\subset {\mathcal{D}}_{\alpha}$. Let $g\in {\mathcal{D}}_{\alpha}^*$. Then by \eqref{connect*}, $Lg=L^*g\in {\mathcal{H}}$. For any $f\in {\mathcal{D}}_{\alpha}$ we have, using \eqref{connect*}, Lemma \ref{stoktermen}, Lemma \ref{Winfty} and \eqref{lim}, that \[0=\langle Lf,g\rangle-\langle f, L^*g\rangle =W\bigl(f,{\overline{g}}\bigr)(0^-) -W\bigl(f,{\overline{g}}\bigr)(0^+). \] Now using Lemma \ref{suitable}, one derives from the existence of the limits \begin{equation} \begin{split} W\bigl(f,{\overline{g}}\bigr)(0^+)&=\frac{(1-q)^2}{bc}\lim_{k\to\infty} \bigl(\bigl(D_qf\bigr)(zq^k){\overline{g(zq^k)}}- f(zq^k){\overline{\bigl(D_qg\bigr)(zq^k)}}\bigr)\nonumber\\ W\bigl(f,{\overline{g}}\bigr)(0^-)&=\frac{(1-q)^2}{bc}\lim_{k\to\infty} \bigl(\bigl(D_qf\bigr)(-q^k){\overline{g(-q^k)}}- f(-q^k){\overline{\bigl(D_qg\bigr)(-q^k)}}\bigr)\nonumber \end{split} \end{equation} for all $f\in {\mathcal{D}}_{\alpha}$ that $g(0^+)=\alpha g(0^-)$ and $g'(0^+)=\alpha g'(0^-)$. It follows that $g\in {\mathcal{D}}_{\alpha}$, as desired. \end{proof} Observe that the operator $L\in \hbox{End}\bigl({\mathcal{F}}(I)\bigr)$ preserves the subspace of functions $f\in {\mathcal{F}}(I)$ with support in $I_-=[-1,0]_q=\{-q^k\}_{k\in\ensuremath{\mathbb{Z}}_+}$, as well as the subspace of functions $f\in {\mathcal{F}}(I)$ with support in $I_+=[0,\infty(z))_q=\{zq^l\}_{l\in\ensuremath{\mathbb{Z}}}$. We denote $L_-$ (respectively $L_+$) for the restriction of $L$ to functions $f\in {\mathcal{F}}(I)$ with support in $I_-$ (respectively $I_+$) and we write ${\mathcal{H}}={\mathcal{H}}^-\oplus {\mathcal{H}}^+$ for the orthogonal direct sum decomposition where ${\mathcal{H}}^{\pm}$ are the functions $f\in {\mathcal{H}}$ with support in $I_\pm$. We end this section by indicating the relation of this splitting of $L$ with the possible choices of domains for $L$. The following results will not be used in the remainder of the paper, so we omit detailed proofs. We start with the symmetric operator $(L_+,{\mathcal{D}}_{fin}^+)$ on ${\mathcal{H}}_+$, where ${\mathcal{D}}_{fin}^+\subset {\mathcal{H}}_+$ is the subspace of functions with finite support. Its minimal closure is given by $(L,{\mathcal{D}}_0^+)$, where \[{\mathcal{D}}_0^{+}=\{f \in {\mathcal{H}}^{+} \, | \, L_{+}f\in {\mathcal{H}}^{+},\,\,\, f(0^{+})=f'(0^{+})=0 \}, \] and the deficiency indices of $(L,{\mathcal{D}}_0^+)$ are $(1,1)$. The spectral properties of the self-adjoint extensions of $(L_{+},{\mathcal{D}}_0^{+})$ are highly sensitive with respect to the choice of self-adjoint extension. Furthermore, the spectral analysis of $L_+$ with respect to a fixed choice of self-adjoint extension seems to result in a rather implicit description of the spectrum and of the resolution of the identity. In order to obtain better spectral properties, we ``blow up'' the Hilbert space ${\mathcal{H}}_+$ to ${\mathcal{H}}={\mathcal{H}}_-\oplus {\mathcal{H}}_+$ and we regard $L$ now as an unbounded operator on ${\mathcal{H}}$ with domain ${\mathcal{D}}_{fin}$ given by the functions in ${\mathcal{H}}$ with finite support. Its minimal closure is given by $(L,{\mathcal{D}}_0)$, where \[ {\mathcal{D}}_0=\{f\in {\mathcal{H}} \, | \, Lf\in {\mathcal{H}}, \,\,\,\, f(0^{+})=f(0^-)=0=f'(0^{-})=f'(0^+) \}. \] The closed, symmetric operator $(L,{\mathcal{D}}_0)$ has deficiency indices $(2,2)$. By the increase of deficiency indices we have a larger choice of self-adjoint extensions for $(L,{\mathcal{D}}_0)$ than for $(L_+,{\mathcal{D}}_0^+)$. In particular, the self-adjoint extensions $(L,{\mathcal{D}}_{\alpha})$ ($\alpha\in\ensuremath{\mathbb{T}}$) are self-adjoint extensions for which $(L_+,{\mathcal{D}}_{\alpha}\cap {\mathcal{H}}_+)$ are {\it not} self-adjoint. It turns out that the spectral analysis of $(L,{\mathcal{D}}_{\alpha})$ is essentially independent of $\alpha\in \ensuremath{\mathbb{T}}$, so we restrict attention in this paper to the unbounded self-adjoint operator $L$ on ${\mathcal{H}}$ with domain of definition ${\mathcal{D}}={\mathcal{D}}_1\subset {\mathcal{H}}$. In particular, any function $f\in {\mathcal{D}}$ is {\it continuously differentiable} at the origin, i.e. $f$ satisfies $f(0^+)=f(0^-)$ and $f'(0^+)=f'(0^-)$. In the next sections we show that $(L,{\mathcal{D}})$ is a self-adjoint extension of $(L,{\mathcal{D}}_0)$ for which the spectral analysis can be derived in a very explicit manner. The reason is that for a given generic eigenvalue, one has two linear independent eigenfunctions of $L$ which are explicitly given in terms of basic hypergeometric series and which are continuously differentiable at the origin. The corresponding Wronskian, as well as their Wronskian with the asymptotically free solutions of the corresponding eigenvalue equation, can be computed explicitly. This allows us to derive the spectral properties in a very explicit manner. Let us finally make some remarks on the ``blowing up'' procedure of the Hilbert space ${\mathcal{H}}_+$ as described in the previous paragraphs. It is a well known principle in harmonic analysis that the spectral properties of the unbounded self-adjoint operator under consideration is essentially determined by the behaviour of the eigenfunctions at infinity (along the support of the measure). The fact that ${\mathcal{H}}_-$ is a $L^2$-space of functions supported on a compact space implies that the situation at infinity does not alter in the enlarged setting. Hence, the spectral properties of the extension of $L_+$ to the unbounded operator $L$ on ${\mathcal{H}}$ should be essentially determined by spectral properties of $L_+$. We will justify this principle in this paper by deriving most of the spectral properties of $L$ from the spectral properties of $L_+$. The blowing up of the Hilbert space ${\mathcal{H}}_+$ is canonical in the sense that the extra piece ${\mathcal{H}}_-$ added to the Hilbert space ${\mathcal{H}}_+$ is essentially determined by the following three properties: \begin{enumerate} \item[\bf{(i)}] The support of the measure of ${\mathcal{H}}^-$ is a {\it bounded} $q$-interval, \item[\bf{(ii)}] The operator $(L,{\mathcal{D}}_{fin}^-)$ on ${\mathcal{H}}^-$ is symmetric, where ${\mathcal{D}}_{fin}^-$ consists of functions in ${\mathcal{H}}^-$ with finite support, \item[\bf{(iii)}] The weight function is continuous at the origin. \end{enumerate} Indeed, the condition that the support of the measure of ${\mathcal{H}}^-$ is a bounded $q$-interval implies that the support is of the form $[y,0]_q=\{yq^k\}_{k\in\ensuremath{\mathbb{Z}}_+}$ with $B(y)=0$. This leads to the possibilities $y=-1$ or $y=-bc/q$. Choosing $y=-1$, we derive from condition {\bf (ii)} that the weight function corresponding to ${\mathcal{H}}^-$ is given by $1/p(\cdot)$ up to a positive constant, while condition {\bf (iii)} implies that the constant is one. Since $p(x;a,b,c)=p(bcx/q;a,q/b,q/c)$ and a similar property holds for $L$, see Remark \ref{symmetryrem}, we may take $y=-1$ without loss of generality. \section{Eigenfunctions of $L$}\label{eigenf} In sections \ref{eigenf}--\ref{PID} we consider $L$ with domain of definition ${\mathcal{D}}={\mathcal{D}}_1$. In particular, any function $f$ in the domain of definition ${\mathcal{D}}$ is continuously differentiable at the origin. We set $f(0)$ (respectively $f'(0)$) for the common limits $f(0^\pm)$ (respectively $f'(0^\pm)$). Furthermore, we need to restrict sometimes the choice of parameters $(a,b,c)\in V$ to a dense subdomain $V_z^{gen}$, which is defined by \begin{equation}\label{con3} V_z^{gen}=\{(a,b,c)\in V \, | \, a^2,b^2,c^2,ab,ac,bc,a/b,a/c,a^2b^2c^2z^2 \not\in \{q^k\}_{k\in\ensuremath{\mathbb{Z}}}\}. \end{equation} In this section we consider eigenfunctions of the linear operators $L$ and $L_{\pm}$, and study their behaviour at the origin. We write $(-1,0]_q=I_-\setminus \{-1\}=[-q,0]_q$, and define for $\mu\in\ensuremath{\mathbb{C}}$ the linear spaces \begin{equation}\label{subspaces} \begin{split} V_{\mu}^{\pm}&=\{\, f: I_{\pm} \to\ensuremath{\mathbb{C}} \, | \, L_{\pm}f=\mu f \,\, \hbox{on} \,\, I_{\pm} \,\},\\ {\hat{V}}_{\mu}^{-}&=\{\, f: I_{-} \to\ensuremath{\mathbb{C}} \, | \, L_{-}f=\mu f \,\, \hbox{on} \,\, (-1,0]_q \,\},\\ V_{\mu}&=\{ f\in {\mathcal{F}}(I) \, | \, Lf=\mu f \, \hbox{ on }\, (-1,\infty(z))_q,\, f\hbox{ cont. differentiable in } 0 \}. \end{split} \end{equation} Observe that $V_{\mu}^-\subset {\hat{V}}_{\mu}^{-}$. \begin{lem}\label{generaleigenfunction} Let $\mu\in\ensuremath{\mathbb{C}}$.\\ {\bf (i)} $\hbox{dim}\bigl(V_{\mu}^+\bigr)=2$, $\hbox{dim}\bigl(V_{\mu}^-\bigr)=1$ and $\hbox{dim}\bigl({\hat{V}}_{\mu}^{-}\bigr)=2$. \\ {\bf (ii)} If $f_1,f_2\in V_{\mu}^+$ \textup{(}respectively $f_1,f_2\in {\hat{V}}_{\mu}^{-}$\textup{)}, then $W(f_1,f_2)$ is constant on $I_{+}$ \textup{(}respectively constant on $I_-$\textup{)}.\\ {\bf (iii)} If $f_1, f_2\in V_{\mu}$, then $W(f_1,f_2)\in {\mathcal{F}}(I)$ is constant on $I$. \end{lem} \begin{proof} {\bf (i)} We start with computing $\hbox{dim}\bigl(V_{\mu}^-\bigl)$. The coefficient $A(x)$ in the expression \eqref{equiv2} of $L$ is non-zero for $x\in I_-$ since $(a,b,c)\in V$. It follows that solutions of $L_-f=\mu f$ on $I_-$ are in one to one correspondence with solutions $(a_k)_{k\in\ensuremath{\mathbb{Z}}_+}$ of a recurrence relation of the form \begin{equation*} \mu a_k= \alpha_ka_{k-1}+\beta_ka_k+\gamma_ka_{k+1},\qquad \gamma_k\not=0,\,\,k\in\ensuremath{\mathbb{Z}}_+ \end{equation*} with $a_{-1}=0$ by definition. The correspondence is obtained by associating the sequence $a_k=f(-q^k)$ ($k\in\ensuremath{\mathbb{Z}}_+$) to $f\in V_{\mu}^-$. The coefficients of the recurrence relation are then given by $\alpha_k=-B(-q^k)$, $\beta_k=-A(-q^k)-B(-q^k)$ and $\gamma_k=A(-q^k)$. Any solution of such a recurrence relation is uniquely determined by $a_0=f(-1)\in\ensuremath{\mathbb{C}}$, hence $\hbox{dim}\bigl(V_{\mu}^-\bigl)=1$. Then $\hbox{dim}\bigl({\hat{V}}_{\mu}^{-}\bigl)=2$ follows from the fact that functions $f\in {\hat{V}}_{\mu}^-$ are in one to one correspondence with solutions $(a_k)_{k\in\ensuremath{\mathbb{Z}}_+}$ of the recurrence relation \begin{equation*} \mu a_k=\alpha_ka_{k-1}+\beta_ka_k+\gamma_ka_{k+1},\qquad \gamma_k\not=0,\,\, k\in \ensuremath{\mathbb{N}}. \end{equation*} Finally, in order to show that $\hbox{dim}\bigl(V_{\mu}^-\bigl)=2$, observe that solutions of $L_+f=\mu f$ on $I_+$ are in one to one correspondence with solutions $(a_k)_{k\in\ensuremath{\mathbb{Z}}}$ of a double infinite recurrence relation of the form \[ \mu a_k=\alpha_ka_{k-1}+\beta_ka_k+\gamma_ka_{k+1},\qquad \alpha_k\not=0,\, \gamma_k\not=0, \,\,k\in\ensuremath{\mathbb{Z}},\] since $A(x)$ and $B(x)$ \eqref{ABbig} in the expression \eqref{equiv2} of $L$ are non-zero for $x\in I_+$. This is a two-dimensional space.\\ {\bf (ii)} Let $f_1, f_2\in {\mathcal{F}}(I)$. Using Lemma \ref{self-adjointform}, the product rule for the $q$-derivative \begin{equation}\label{productrule} \bigl(D_q(fg)\bigr)(x)=\bigl(D_qf\bigr)(x)g(x)+f(qx)\bigl(D_qg\bigr)(x), \end{equation} and the second equality of \eqref{Wronskian}, we have for all $x\in (-1,\infty(z))_q$, \begin{equation}\label{consta} \begin{split} q^{-1}p(x)&\bigl(D_qW(f_1,f_2)\bigr)(q^{-1}x)\\ &=\bigl(Lf_1\bigr)(x)f_2(x)+p(x)r(q^{-1}x)\bigl(D_qf_1\bigr)(q^{-1}x) \bigl(D_qf_2\bigr)(q^{-1}x)\\ &-\bigl(Lf_2\bigr)(x)f_1(x)-p(x)r(q^{-1}x)\bigl(D_qf_2\bigr)(q^{-1}x) \bigl(D_qf_1\bigr)(q^{-1}x)\\ &=\bigl(Lf_1\bigr)(x)f_2(x)-\bigl(Lf_2\bigr)(x)f_1(x). \end{split} \end{equation} If $f_1,f_2\in {\hat{V}}_{\mu}^-$, then it follows from \eqref{consta} that $\bigl(D_qW(f_1,f_2)\bigr)(x)=0$ for all $x\in (-1,0]_q$ since $p(x)\not=0$ $(x\in I_-)$, hence $W(f_1,f_2)$ is constant on $I_-$. A similar argument shows that $W(f_1,f_2)$ is constant on $I_+$ if $f_1,f_2\in V_{\mu}^+$.\\ {\bf (iii)} Let $f_1,f_2\in V_{\mu}$. The continuously differentiability of the $f_j$'s at the origin yields that $W(f_1,f_2)(0^+)=W(f_1,f_2)(0^-)$. Indeed, both sides are equal to $qr(0)\bigl(f_1'(0)f_2(0)- f_1(0)f_2'(0)\bigr)$ (compare with the proof of Lemma \ref{symmetryoperator}). The result follows now from {\bf (ii)}. \end{proof} Observe that $f|_{I_-}\in {\hat{V}}_{\mu}^-$ and $f|_{I_+}\in V_{\mu}^+$ for $f\in V_{\mu}$. \begin{prop}\label{generaleigenfunctionprop} We have ${\hbox{dim}}(V_{\mu})\leq 2$. Furthermore, the following three statements are equivalent:\\ {\bf (i)} The linear map $f\mapsto f|_{I_-}:\,\, V_{\mu}\to {\hat{V}}_{\mu}^-$ is a bijection.\\ {\bf (ii)} The linear map $f\mapsto f|_{I_+}: V_{\mu}\to V_{\mu}^+$ is a bijection.\\ {\bf (iii)} $\hbox{dim}(V_{\mu})=2$. \end{prop} \begin{proof} Suppose that $\hbox{dim}(V_{\mu})\geq 2$, then we claim that the map $f\mapsto f|_{I_+}:\,\, V_{\mu}\to V_{\mu}^+$ is injective. Suppose that the map is not injective. Then there exist two linearly independent functions $f_1,f_2\in V_{\mu}$ such that $f_1|_{I_+}, f_2|_{I_+}\in V_{\mu}^+$ are linearly dependent. The Wronskian $W(f_1,f_2)$ is not identically zero as function on $I$ by the linear independence of $f_1$ and $f_2$. Since $W(f_1,f_2)$ is constant on $I$ by Lemma \ref{generaleigenfunction}{\bf (iii)}, it follows that $W(f_1,f_2)$ is not identically zero on $I_+$, which contradicts the linear dependence of $f_1|_{I_+}$ and $f_2|_{I_+}$. So if $\hbox{dim}(V_{\mu})\geq 2$, then it follows from Lemma \ref{generaleigenfunction}{\bf (i)} and from the injectivity of the map $f\mapsto f|_{I_+}:\,\, V_{\mu}\to V_{\mu}^+$ that $\hbox{dim}(V_{\mu})=2$. This proves that $\hbox{dim}(V_{\mu})\leq 2$ for all $\mu\in \ensuremath{\mathbb{C}}$, and it proves the implication ${\bf (iii)}\Rightarrow {\bf (ii)}$. The implication ${\bf (iii)}\Rightarrow {\bf (i)}$ is proved in a similar manner, while the implications ${\bf (i)}\Rightarrow {\bf (iii)}$ and ${\bf (ii)}\Rightarrow {\bf (iii)}$ are immediate consequences of Lemma \ref{generaleigenfunction}{\bf (i)}. \end{proof} \begin{cor}\label{extensionpossible} If $\hbox{dim}(V_{\mu})=2$, then any $g\in {\hat{V}}_{\mu}^-$ \textup{(}respectively $g\in V_{\mu}^+$\textup{)} extends uniquely to a function $g\in V_{\mu}$. \end{cor} In the remainder of this section we introduce two explicit functions which live in $V_{\mu}$. We define for $\gamma\in \ensuremath{\mathbb{C}}^*$, \begin{equation}\label{eigenfunction} \mu(\gamma)=-1-a^2+a(\gamma+\gamma^{-1}), \end{equation} and we write $X_n=\bigl(bcx^2/q\bigr)^n\bigl(-q/bcx,-1/x;q\bigr)_ng_{\gamma}(xq^{-n})$ where $g_{\gamma}$ is a function satisfying $\bigl(Lg_{\gamma}\bigr)(xq^{-n})= \mu(\gamma)g_{\gamma}(xq^{-n})$. Then the $X_n$ satisfy the three term recurrence relation \cite[(2.1)]{GIM} with parameters $(A,B,C,D,z)$ in \cite[(2.1)]{GIM} given by $A=-q/bcx$, $B=-1/x$, $C=-q/acx$, $D=-q/abx$ and \[z=\frac{abcx^2}{q}(\gamma+\gamma^{-1})= \frac{q}{ABCD\lambda_{\pm}}+\lambda_{\pm},\qquad \lambda_{\pm}=\frac{abcx^2}{q}\gamma^{\pm 1}. \] In \cite{GIM} it is was shown that the three term recurrence relation \cite[(2.1)]{GIM} is the one satisfied by the associated continuous dual $q$-Hahn polynomials. Furthermore, several explicit solutions of the recurrence relation \cite[(2.1)]{GIM} were derived explicitly in \cite[section 2]{GIM}. In particular, the explicit solution \cite[(2.13)]{GIM} of the recurrence relation \cite[(2.1)]{GIM} implies that \begin{equation}\label{phiB} \phi_{\gamma}(x)=\phi_{\gamma}(x;a,b,c)= {}_3\phi_2\left( \begin{array}{c} a\gamma, a/\gamma, -1/x \\ ab, ac \end{array} ; q,-bcx \right),\quad |bcx|<1 \end{equation} satisfies $\bigl(L\phi_{\gamma}\bigr)(x)=\mu(\gamma)\phi_{\gamma}(x)$ for $x\in (-1,\infty(z))_q$ with $|x|<q/bc$. Observe that $\phi_{\gamma}$ is well defined since $(a,b,c)\in V$. The explicit solution \cite[(2.13)]{GIM} of the recurrerence relation \cite[(2.1)]{GIM} implies that \begin{equation}\label{psiB} \psi_{\gamma}(x;a,b,c) ={}_3\phi_2\left( \begin{array}{c} a\gamma, a/\gamma, -q/bcx \\ qa/b, qa/c \end{array} ; q,-qx \right), \quad |qx|<1 \end{equation} satisfies $\bigl(L\psi_{\gamma}\bigr)(x)=\mu(\gamma)\psi_{\gamma}(x)$ for $x\in (-1,\infty(z))_q$ with $|x|<1$ when $(a,b,c)\in V_z^{gen}$. Since $bc<1$, we have that $\phi_{\gamma}$ and $\psi_{\gamma}$ are well defined on $I_-$ and that they are solutions of $Lg=\mu(\gamma)g$ on $(-1,0]_q$. Since $\phi_{\gamma}$ and $\psi_{\gamma}$ are solutions of $\bigl(Lg\bigr)(x)=\mu(\gamma)g(x)$ for $x\in I_+$ with $x<1$, it follows that $\phi_{\gamma}$ and $\psi_{\gamma}$ uniquely extend to functions on $I$ satisfying the eigenvalue equation $Lg=\mu(\gamma)g$ on $(-1,\infty(z))_q$, cf. the proof of Lemma \ref{generaleigenfunction}{\bf (i)}. For given $x\in I$, $\phi_{\gamma}(x)$ and $\psi_{\gamma}(x)$ depend analytically on $\gamma\in\ensuremath{\mathbb{C}}^*$ and are invariant under $\gamma\leftrightarrow \gamma^{-1}$. The extension of $\phi_{\gamma}$ to a function on $I$ can also be obtained by the transformation formula \cite[(3.2.10)]{GR}, which yields \begin{equation}\label{ancontB} \phi_{\gamma}(x)= \frac{\bigl(a\gamma, bc,-abcx/\gamma;q\bigr)_{\infty}} {\bigl(ab, ac, -bcx;q\bigr)_{\infty}} {}_3\phi_2\left( \begin{array}{c} b/\gamma, c/\gamma, -bcx \\ bc, -abcx/\gamma \end{array} ; q,a\gamma \right). \end{equation} This gives a single valued analytic continuation for $\phi_{\gamma}(x)$ for $\gamma\in\ensuremath{\mathbb{C}}^*$ with $|\gamma|<a^{-1}$ to $x\in \ensuremath{\mathbb{C}}\setminus (-\infty, -1/bc]$. Observe that $(-\infty,-1/bc]\cap I=\emptyset$ since $(a,b,c)\in V$. If $a<1$, then the analytic continuation for $|\gamma|\geq a^{-1}$ can be obtained by replacing $\gamma$ by $\gamma^{-1}$ in the right hand side of \eqref{ancontB}. If $a\geq 1$, then $b<1$, and we can use \cite[(3.2.7)]{GR} to obtain \begin{equation}\label{ancontB2} \phi_{\gamma}(x)= \frac{\bigl(b\gamma, -abcx/\gamma;q\bigr)_{\infty}} {\bigl(ab, -bcx;q\bigr)_{\infty}} {}_3\phi_2\left( \begin{array}{c} a/\gamma, c/\gamma, -acx \\ ac, -abcx/\gamma \end{array} ; q,b\gamma \right), \end{equation} which yields then the explicit expression for the analytic continuation of $\phi_{\gamma}$ ($\gamma\in \ensuremath{\mathbb{C}}^*$) in a similar manner as for $a<1$. A similar remark holds for the solution $\psi_{\gamma}$. \begin{rem}\label{symmetryrem} If $f(\cdot)=f(\cdot;a,b,c)$ satisfies the eigenvalue equation $Lf=\mu(\gamma)f$, then $g(x)=f(bcx/q;a,q/b,q/c)$ satisfies the same eigenvalue equation. It is easy to check using the explicit expressions \eqref{phiB} and \eqref{psiB} for $\phi_{\gamma}$ and $\psi_{\gamma}$ that the solutions $\phi_{\gamma}$ and $\psi_{\gamma}$ are interchanged by this symmetry, i.e. \begin{equation}\label{symmetryphi} \psi_{\gamma}(x;a,b,c)= \phi_{\gamma}(bcx/q;a,q/b,q/c). \end{equation} \end{rem} {}From \eqref{phiB} we see that $\phi_{\gamma}(-1)=1$ and by direct computation we have \begin{equation}\label{dphi} \bigl(D_q\phi_{\gamma}\bigr)(x;a,b,c)=\frac{bc\mu(\gamma)} {(1-q)(1-ab)(1-ac)}\phi_{\gamma}(x;qa,b,c),\qquad x\in I. \end{equation} It follows that \begin{equation}\label{valuephi} \bigl(D_q\phi_{\gamma}\bigr)(-1)= \frac{bc\mu(\gamma)}{(1-q)(1-ab)(1-ac)}, \end{equation} so $\bigl(L\phi_{\gamma}\bigr)(-1)=\mu(\gamma)\phi_{\gamma}(-1)$ by \eqref{special}. We conclude that $\phi_{\gamma}|_{I_-}\in V_{\mu(\gamma)}^-$ for all $\gamma\in \ensuremath{\mathbb{C}}^*$. For $\psi_{\gamma}$ we have the formula \begin{equation}\label{dpsi} \bigl(D_q\psi_{\gamma}\bigr)(x;a,b,c)= \frac{q\mu(\gamma)}{(1-q)(1-qa/b)(1-qa/c)} \psi_{\gamma}(x;qa,b,c). \end{equation} We stress already the fact that in general we have $\bigl(L\psi_{\gamma}\bigr)(-1)\not= \mu(\gamma)\psi_{\gamma}(-1)$, so that $\psi_{\gamma}|_{I_-}\in \hat{V}_{\mu(\gamma)}\setminus V_{\mu(\gamma)}$, see Corollary \ref{psifoutind}{\bf (iii)}. \begin{lem}\label{oknul} Let $\gamma\in\ensuremath{\mathbb{C}}^*$. The function $\phi_{\gamma}\in {\mathcal{F}}(I)$ is continuously differentiable at the origin, i.e. $\phi_{\gamma}\in V_{\mu(\gamma)}$. In fact, we have \begin{equation}\label{nul} \begin{split} \phi_{\gamma}(0)&={}_2\phi_2\left( \begin{array}{c} a\gamma, a/\gamma \\ ab, ac \end{array} ; q, bc \right),\\ \phi_{\gamma}'(0)&=\frac{bc\mu(\gamma)}{(1-q)(1-ab)(1-ac)} {}_2\phi_2\left( \begin{array}{c} qa\gamma, qa/\gamma \\ qab, qac \end{array} ; q, bc \right). \end{split} \end{equation} For $(a,b,c)\in V_z^{gen}$ we have $\psi_{\gamma}\in V_{\mu(\gamma)}$ and \[ \psi_{\gamma}(0;a,b,c)=\phi_{\gamma}(0;a,q/b,q/c),\qquad \psi_{\gamma}'(0;a,b,c)=\frac{bc}{q}\phi_{\gamma}'(0;a,q/b,q/c), \] where we have extended the definition of $\phi_{\gamma}(0)$ and $\phi_{\gamma}'(0)$ to generic parameters $(a,b,c)\in \ensuremath{\mathbb{C}}^{\times 3}$ by analytic continuation of the right hand sides of \eqref{nul}. \end{lem} \begin{proof} The proof for $\phi_{\gamma}$ is a direct consequence of \eqref{dphi} and the explicit expression \eqref{phiB} for $\phi_{\gamma}$. The proof for $\psi_{\gamma}$ follows then from \eqref{psiB} and \eqref{dpsi}. \end{proof} In section \ref{section5} we will evaluate the Wronskian $W\bigl(\psi_{\gamma},\phi_{\gamma}\bigr)$ explicitly, see Proposition \ref{Wpsiphiprop}. In particular, this will give explicit criteria on the spectral parameter $\gamma$ for which we have $\hbox{dim}(V_{\mu(\gamma)})=2$, i.e. for which Corollary \ref{extensionpossible} is applicable. But first we need to study yet another solution of the eigenvalue equation $Lg=\mu(\gamma)g$, the so called asymptotic solution. \section{The asymptotic solution}\label{A} In this section we determine the asymptotic solution $\Phi_{\gamma}$ of the eigenvalue equation $L_+g=\mu(\gamma)g$ on $I_+$. We furthermore determine the $c$-function expansion of $\phi_{\gamma}$ on $I_+$, i.e. we write $\phi_{\gamma}|_{I_+}$ explicitly as linear combination of the asymptotic solutions $\Phi_{\gamma}$ and $\Phi_{\gamma^{-1}}$. We define singular sets by \begin{equation}\label{Ssing} S_{sing}^{+}=\{\pm q^{-\frac{1}{2}k}\}_{k\in\ensuremath{\mathbb{N}}},\qquad S_{sing}=\{\pm q^{\frac{1}{2}k}\}_{k\in \ensuremath{\mathbb{Z}}}, \end{equation} and we write $S_{reg}^{+}$, $S_{reg}$ for the complements of these singular sets in $\ensuremath{\mathbb{C}}^*$. The $c$-function expansion of $\phi_{\gamma}$ on $I_+$ which we derive in this section, holds for $\gamma\in S_{reg}$. We consider for $y>0$ the minimal solution \cite[(2.26)]{GIM} of the three term recurrence relation \cite[(2.1)]{GIM} with corresponding parameters $(A,B,C,D,\lambda_-)$ given by \[ A=-q/bcy,\quad B=-q/aby,\quad C=-q/acy,\quad D=-1/y,\quad \lambda_-=\frac{abcy^2}{q}\gamma. \] This leads for $\gamma\in S_{reg}^+$ with $|\gamma|<a^{-1}$ to the following explicit solution $\Phi^y_{\gamma}(x)=\Phi^y_{\gamma}(x;a,b,c)$ of $Lf=\mu(\gamma)f$ on $[0,\infty(y))_q$, \begin{equation}\label{PhiB} \begin{split} \Phi_{\gamma}^y(x)=& \frac{\bigl(-q\gamma/ax,-q^2\gamma/abcx,a\gamma;q\bigr)_{\infty}} {\bigl(-q/abx,-q/acx,q\gamma^2;q\bigr)_{\infty}}\\ &.\bigl(a\gamma\bigr)^{-k}{}_3\phi_2\left( \begin{array}{c} q\gamma/a, -q/abx, -q/acx \\ -q\gamma/ax, -q^2\gamma/abcx \end{array} ; q,a\gamma \right), \quad x=yq^k. \end{split} \end{equation} For given $x\in [0,\infty(y))_q$, we have a single valued analytic extension of $\Phi^y_{\gamma}(x)$ to $\gamma\in S_{reg}^+$. Indeed, for $x\in [0,\infty(y))_q$ with $x>bc/q$ we can apply \cite[(3.2.7)]{GR} to arrive at \begin{equation}\label{PhiBalt} \begin{split} \Phi^y_{\gamma}(x)=& \frac{\bigl(-q/bcx,-q\gamma/ax;q\bigr)_{\infty}} {\bigl(-q/abx,-q/acx;q\bigr)_{\infty}}\\ &.\bigl(a\gamma\bigr)^{-k}{}_3\phi_2\left( \begin{array}{c} q\gamma/a, b\gamma, c\gamma \\ -q\gamma/ax, q\gamma^2 \end{array} ; q,-q/bcx\right), \quad x=yq^k, \end{split} \end{equation} which gives the analytic extension in this case. For $x\in [0,\infty(y))_q$ with $y\leq bc/q$ we can use the eigenvalue equation $L\Phi^y_{\gamma}=\mu(\gamma)\Phi^y_{\gamma}$ on $[0,\infty(y))_q$ to derive the analytic extension for $\Phi^y_{\gamma}(x)$ from the analytic extension of $\Phi^y_{\gamma}(u)$ with $u>bc/q$. It follows from \eqref{PhiBalt} that the asymptotics to infinity of $\Phi^y_{\gamma}$ is given by \begin{equation}\label{asymptoticsPhiB} \Phi_{\gamma}^y(yq^{-m})=\bigl(a\gamma\bigr)^m\bigl(1+{\mathcal{O}}(q^m)\bigr), \qquad m\to\infty. \end{equation} \begin{Def}\label{asymptoticsolutiondef} Let $\gamma\in S_{reg}^+$. We call $\Phi_{\gamma}(\cdot;a,b,c)=\Phi^z_{\gamma}(\cdot;a,b,c)$ the asymptotic solution of $L_+f=\mu(\gamma)f$ on $I_+=[0,\infty(z))_q$. \end{Def} Definition \ref{asymptoticsolutiondef} is justified by the following lemma. \begin{lem}\label{l2} Let $\mu\in \ensuremath{\mathbb{C}}^*$ and $K=\{zq^{-k}\}_{k\in\ensuremath{\mathbb{Z}}_+}\subset I_+$. Set \[M_{\mu}=\{ h: I_+\to\ensuremath{\mathbb{C}} \, | \, L_+h=\mu h \,\, \hbox{ on } I_+ \,\, \hbox{ and }\,\, \|h_K\|^2<\infty\}, \] where $h_K\in {\mathcal{F}}(I)$ is defined to be equal to zero on $I\setminus K$ and to be equal to $h$ on $K$. Then $M_{\mu}=\hbox{span}\{\Phi_{\gamma}\}$ for all $\mu\in \ensuremath{\mathbb{C}}\setminus \ensuremath{\mathbb{R}}$, where $\gamma\in \ensuremath{\mathbb{C}}^*$ is the unique non-zero complex number such that $\mu=\mu(\gamma)$ and $|\gamma|<1$. \end{lem} \begin{proof} First of all, observe that if $\gamma\in \ensuremath{\mathbb{C}}^*$ and $|\gamma|<1$, then $\gamma\in S_{reg}^+$, hence $\Phi_{\gamma}$ is well defined. Let $\mu\in\ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$ and let $\gamma\in\ensuremath{\mathbb{C}}^*$, $|\gamma|<1$ be such that $\mu=\mu(\gamma)$. Using the asymptotics \eqref{pinfty} for the weights of the inner product $\langle .,. \rangle$, it follows from \eqref{asymptoticsPhiB} that $\Phi_{\gamma}\in M_{\mu(\gamma)}$ because $|\gamma|<1$. It remains to show that $\hbox{dim}(M_{\mu})=1$. For the proof we use some well known results from the theory of the classical moment problem, see for instance \cite{Ak} or \cite{Simon}. If $f:I_+\to\ensuremath{\mathbb{C}}$ satisfies $L_+f=\mu f$ on $I_+$, then by setting \[ a_k=f(zq^{-k}){\sqrt{\frac{(1-q)zq^{-k}}{p(zq^{-k})}}}, \qquad k\in \ensuremath{\mathbb{Z}}, \] we see that $\bigl(a_k\bigr)_{k\in\ensuremath{\mathbb{Z}}}$ satisfies the recurrence relation \begin{equation}\label{recassdualqHahn} \alpha_{k-1}a_{k-1}+\beta_ka_k+\alpha_ka_{k+1}=\mu a_k \end{equation} with \[ \alpha_k=a\sqrt{\left(1+\frac{q^{k+1}}{abz}\right) \left(1+\frac{q^{k+1}}{acz}\right) \left(1+\frac{q^{k+1}}{bcz}\right)\left(1+\frac{q^k}{z}\right)} \] and $\beta_k=-A(zq^{-k})-B(zq^{-k})$. It follows from \cite[Corollary 4.5]{Simon} and from the fact that the sequence $\bigl(\alpha_k\bigr)_{k\in\ensuremath{\mathbb{Z}}_+}$ is bounded that the Hamburger moment problem corresponding to the recurrence relation \eqref{recassdualqHahn} for $k\in\ensuremath{\mathbb{Z}}_+$ is determined. By \cite[Theorem 3]{Simon} this implies that $\hbox{dim}(M_{\mu})=1$, as desired. \end{proof} It follows from Lemma \ref{l2} that \begin{equation}\label{symmetryPhi} \Phi_{\gamma}^{bcz/q}(bcx/q;a,q/b,q/c)=\Phi_{\gamma}(x;a,b,c),\qquad x\in I_+ \end{equation} for parameters $(a,b,c)\in V$ such that $(a,q/b,q/c)\in V$. Indeed, both sides of \eqref{symmetryPhi}, considered as function of $x\in I_+$, are solutions of $L_+f=\mu(\gamma)f$ on $I_+$ by Remark \ref{symmetryrem}, and they have the same asymptotics to infinity by \eqref{asymptoticsPhiB}. Formula \eqref{symmetryPhi} is also obvious from the explicit expression \eqref{PhiB} for $\Phi^y_{\gamma}$. \begin{lem}\label{dep} For $\gamma\in S_{reg}$ and $x\in I_+$, we have \[ W\bigl(\Phi_{\gamma},\Phi_{\gamma^{-1}}\bigr)(x)=aK \bigl(\gamma-\gamma^{-1}\bigr)\not=0, \] where $K$ is the positive constant defined by \eqref{Kconstant}. In particular, $\{\Phi_{\gamma}, \Phi_{\gamma^{-1}}\}$ is a basis of $V_{\mu(\gamma)}^+$ when $\gamma\in S_{reg}$. \end{lem} \begin{proof} The explicit expression for the Wronskian follows by computing the limit $\lim_{m\to\infty}W\bigl(\Phi_{\gamma},\Phi_{\gamma^{-1}}\bigr)(zq^{-m})$ using the first expression of \eqref{Wronskian} and the formulas \eqref{rinfty} and \eqref{asymptoticsPhiB}. Since the Wronskian is non-zero, it follows that $\Phi_{\gamma}$ and $\Phi_{\gamma^{-1}}$ are linear independent, hence they form a basis of $V_{\mu(\gamma)}^+$ by Lemma \ref{generaleigenfunction}{\bf (i)}. \end{proof} It follows from Lemma \ref{dep} that $\phi_{\gamma}|_{I_+}$ and $\psi_{\gamma}|_{I_+}$ can be written uniquely as linear combination of $\Phi_{\gamma}$ and $\Phi_{\gamma^{-1}}$ for $\gamma\in S_{reg}$. The corresponding coefficients can be expressed in terms of the $c$-function $c(\gamma)=c(\gamma;a,b,c;z)$, which is defined by \begin{equation}\label{cB} c(\gamma)=\frac{1}{\bigl(ab, ac;q\bigr)_{\infty}\theta(-bcz)} \frac{\bigl(a/\gamma, b/\gamma, c/\gamma;q\bigr)_{\infty} \theta(-q/abcz\gamma)}{\bigl(1/\gamma^2;q\bigr)_{\infty}}. \end{equation} \begin{prop}\label{conncoef} Let $\gamma\in S_{reg}$. Then we have \[\phi_{\gamma}(x)=c(\gamma)\Phi_{\gamma}(x)+ c(\gamma^{-1})\Phi_{\gamma^{-1}}(x), \qquad x\in I_+. \] The same formula holds for $\psi_{\gamma}(x)$ when $(a,b,c)\in V_z^{gen}$, with $c(\gamma)$ replaced by \begin{equation}\label{tildecB} \begin{split} \tilde{c}(\gamma;a,b,c;z)&=c(\gamma;a,q/b,q/c;bcz/q)\\ &=\frac{1}{\bigl(qa/b, qa/c;q\bigr)_{\infty}\theta(-1/z)} \frac{\bigl(a/\gamma, q/b\gamma, q/c\gamma;q\bigr)_{\infty} \theta(-1/az\gamma)}{\bigl(1/\gamma^2;q\bigr)_{\infty}}. \end{split} \end{equation} \end{prop} \begin{proof} We first prove the connection coefficient formula for $\phi_{\gamma}$. Observe that $c(\gamma^{\pm 1})$ is well defined for $\gamma\in S_{reg}$ since $(a,b,c)\in V$. We fix $x=zq^k$ with $k\in \ensuremath{\mathbb{Z}}$ such that $q/acx<1$. Furthermore, we assume that $a<1$ and we fix $\gamma\in \ensuremath{\mathbb{C}}\setminus (-\infty,0]$ such that $\gamma\in S_{reg}$ and $a<|\gamma|<1/a$. By the assumptions on $x$ and $\gamma$, we may apply the three term recurrence relation \cite[(3.3.3)]{GR} with $a\to -bcx$, $b\to b/\gamma$, $c\to c/\gamma$, $d\to -abcx/\gamma$ and $e\to bc$. We arrive at \begin{equation*} \begin{split} &{}_3\phi_2\left( \begin{array}{c} -bcx, b/\gamma, c/\gamma \\ bc, -abcx/\gamma \end{array} ; q,a\gamma \right)\\ &=\frac{\bigl(c\gamma, b\gamma,-q/bx\gamma,-q\gamma/abcx;q\bigr)_{\infty}} {\bigl(bc,-q/abx,-q/bcx, \gamma^2;q\bigr)_{\infty}} {}_3\phi_2\left( \begin{array}{c} c/\gamma, a/\gamma, q/b\gamma \\ -q/bx\gamma, q/\gamma^2 \end{array} ; q,-q/acx \right)\\ &\quad-\frac{\bigl(-q\gamma/abcx,-q\gamma/ax,b/\gamma,c/\gamma,a/\gamma ;q\bigr)_{\infty}\theta(-abcx\gamma/q)} {\bigl(-abcx/q\gamma,bc,-q/acx,-q/abx,-q/bcx;q\bigr)_{\infty} \theta(\gamma^2)}\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad \qquad.{}_3\phi_2\left( \begin{array}{c} q\gamma/a, -q/acx, -q/abx \\ -q^2\gamma/abcx, -q\gamma/ax \end{array} ; q,a\gamma \right). \end{split} \end{equation*} The ${}_3\phi_2$ on the left hand side is the ${}_3\phi_2$ in the expression \eqref{ancontB} for $\phi_{\gamma}$ and the second ${}_3\phi_2$ on the right hand side is the ${}_3\phi_2$ in the expression \eqref{PhiB} for $\Phi_{\gamma}$. Again by the assumptions on $x$ and $\gamma$, we may rewrite the first ${}_3\phi_2$ on the right hand side using \cite[(3.2.10)]{GR} with $a\to c/\gamma$, $b\to a/\gamma$, $c\to q/b\gamma$, $d\to q/\gamma^2$ and $e\to -q/bx\gamma$. This gives \begin{equation*} \begin{split} &{}_3\phi_2\left( \begin{array}{c} c/\gamma, a/\gamma, q/b\gamma \\ -q/bx\gamma, q/\gamma^2 \end{array} ; q,-q/acx \right)\\ &\qquad\qquad= \frac{\bigl(a/\gamma,-q^2/abcx\gamma,-q/ax\gamma;q\bigr)_{\infty}} {\bigl(q/\gamma^2,-q/bx\gamma,-q/acx;q\bigr)_{\infty}} {}_3\phi_2\left( \begin{array}{c} q/a\gamma, -q/abx, -q/acx \\ -q^2/abcx\gamma, -q/ax\gamma \end{array} ; q,a/\gamma \right). \end{split} \end{equation*} The ${}_3\phi_2$ on the right hand side is the ${}_3\phi_2$ in the expression \eqref{PhiB} for $\Phi_{\gamma^{-1}}$. So substituting this formula in the three term recurrence relation, and simplifying the formulas using in particular the functional relation \begin{equation}\label{functheta} \theta(q^kx)= \begin{cases} q^{-k(k-1)/2}(-x)^{-k}\theta(x),\qquad &k\in \ensuremath{\mathbb{Z}}_+,\\ q^{k(k+1)/2}(-x)^{-k}\theta(x), &k\in -\ensuremath{\mathbb{Z}}_+ \end{cases} \end{equation} for the Jacobi theta function, we arrive at the desired result for restricted choices of $x$, $a$ and $\gamma$. The extension to all $x\in I_+$ is made using the fact that the left hand side and the right hand side of the $c$-function expansion are solutions of the eigenvalue equation $L_+f=\mu(\gamma)f$ on $I_+$. Finally, the restrictions on $a$ and $\gamma$ can be removed by analytic continuation. The proof of the connection coefficient formula for $\psi_{\gamma}$ follows by analytic continuation from the connection coefficient formula for $\phi_{\gamma}$ using \eqref{symmetryphi} and \eqref{symmetryPhi}. \end{proof} For $(a,b,c)\in V_z^{gen}$ and $\gamma\in S_{reg}^+$ with $|\gamma|<a^{-1}$ we define $\Phi^-_{\gamma}(x)$ for $x\in (\infty(-1),0]_q$ by \begin{equation}\label{PhiB-} \begin{split} \Phi^-_{\gamma}(x)=& \frac{\bigl(-q^2\gamma/abcx,-q\gamma/ax,a\gamma;q\bigr)_{\infty}} {\bigl(-q/abx,-q/acx,q\gamma^2;q\bigr)_{\infty}}\\ &.\bigl(a\gamma\bigr)^{-k}{}_3\phi_2\left( \begin{array}{c} q\gamma/a, -q/abx, -q/acx \\ -q^2\gamma/abcx, -q\gamma/ax \end{array} ; q,a\gamma \right), \qquad x=-q^k. \end{split} \end{equation} Observe that $\Phi^-_{\gamma}$ is obtained by taking $y=-1$ in the definition of the eigenfunction $\Phi^{y}_{\gamma}$, see \eqref{PhiB}. \begin{lem}\label{collaps} Let $(a,b,c)\in V_z^{gen}$ and $\gamma\in S_{reg}^+$ with $|\gamma|<a^{-1}$. Then \[\Phi^-_{\gamma}(x)=\frac{\bigl(q\gamma/a,q\gamma/b, q\gamma/c;q\bigr)_{\infty}}{\bigl(q/ab,q/ac,q\gamma^2;q\bigr)_{\infty}} \phi_{\gamma}(x),\qquad x\in I_-. \] \end{lem} \begin{proof} Using the minimal solution \cite[(2.26)]{GIM} of the three term recurrence relation \cite[(2.1)]{GIM}, it follows that $\Phi^-_{\gamma}$ is a solution of $Lf=\mu(\gamma)f$ on $(\infty(-1),0]_q$, where $L$ is the second order $q$-difference operator defined by \eqref{equiv2}. In particular, we have $A(-1)(\Phi^-_{\gamma}(-q)-\Phi^-_{\gamma}(-1))= \mu(\gamma)\Phi^-_{\gamma}(-1)$ since $B(-1)=0$. It follows that $\Phi^-_{\gamma}|_{I_-}\in V_{\mu(\gamma)}^-$. We have $\hbox{dim}(V_{\mu(\gamma)}^-)=1$ by Lemma \ref{generaleigenfunction}{\bf (i)} and $0\not=\phi_{\gamma}|_{I_-}\in V_{\mu(\gamma)}^-$. It follows that $\Phi^-_{\gamma}|_{I_-}=C_{\gamma}\phi_{\gamma}|_{I_-}$ for a unique constant $C_{\gamma}$. The explicit expression for $C_{\gamma}$ can be found using $\phi_{\gamma}(-1)=1$ and by evaluating $\Phi^-_{\gamma}(-1)$ using the $q$-Gauss sum \cite[(1.5.1)]{GR}. \end{proof} Observe that the Wronskian $W(\phi_{\gamma}, \Phi^-_{\gamma})$ is identically zero on $I_-$ by Lemma \ref{collaps}. On the other hand, the Wronskian $W(\phi_{\gamma},\Phi_{\gamma})$ on $I_+$ is non-zero for generic $\gamma$ by Lemma \ref{dep} and Proposition \ref{conncoef}. Suppose now that for a given (generic) $\gamma\in\ensuremath{\mathbb{C}}^*$ with $|\gamma|<a^{-1}$ we have an extension of $\Phi_{\gamma}\in V_{\mu(\gamma)}^+$ to a function $\Phi_{\gamma}\in V_{\mu(\gamma)}$. Then the Wronskian $W(\phi_{\gamma},\Phi_{\gamma})$ is constant on $I$ by Lemma \ref{generaleigenfunction}{\bf (iii)} and Lemma \ref{oknul}, so $\Phi_{\gamma}|_{I_-}$ is {\it not} a multiple of $\Phi^-_{\gamma}$ on $I_-$. This shows that the extension of $\Phi_{\gamma}\in V_{\mu(\gamma)}^+$ to a function $\Phi_{\gamma}\in V_{\mu(\gamma)}$ can not be obtained by taking the explicit expression for $\Phi_{\gamma}$ and extending its definition in an obvious manner to the whole $q$-interval $I$. We derive in the next section the extension of $\Phi_{\gamma}\in V_{\mu(\gamma)}^+$ to a function $\Phi_{\gamma}\in V_{\mu(\gamma)}$ for $(a,b,c)\in V_z^{gen}$ by writing $\Phi_{\gamma}$ explicitly as a linear combination of the two eigenfunctions $\phi_{\gamma}\in V_{\mu(\gamma)}$ and $\psi_{\gamma}\in V_{\mu(\gamma)}$. \section{The extension of the asymptotic solution}\label{section5} In this section we first evaluate the Wronskian $W\bigl(\psi_{\gamma},\phi_{\gamma}\bigr)$ explicitly in product form. {}From this evaluation we derive explicit conditions on the spectral parameter $\gamma$ for which we have $\hbox{dim}(V_{\mu(\gamma)})=2$. For these values of the spectral parameter, there exists a unique extension of $\Phi_{\gamma}$ which lives in $V_{\mu(\gamma)}$ by Corollary \ref{extensionpossible}. We will give this extension explicitly as a linear combination of $\phi_{\gamma}$ and $\psi_{\gamma}$. Observe that if we compute the Wronskian by taking the limit to $0$, \[W(\psi_{\gamma},\phi_{\gamma})(0)= \frac{(1-q)^2}{bc}\bigl(\psi_{\gamma}'(0)\phi_{\gamma}(0)- \psi_{\gamma}(0)\phi_{\gamma}'(0)\bigr), \] and by applying Lemma \ref{oknul}, we get an explicit expression as a linear combination of products of ${}_2\phi_2$'s, while evaluation of $W(\psi_{\gamma},\phi_{\gamma})(x)$ at $x=-1$ using \eqref{valuephi} and $\phi_{\gamma}(-1)=1$ gives an expression of the Wronskian as a linear combination of two ${}_3\phi_2$'s. {}From both these expressions, the (non)vanishing properties of the Wronskian $W(\psi_{\gamma},\phi_{\gamma})$ are hard to derive. In the next proposition we evaluate the Wronskian by substitution of the $c$-function expansions for $\phi_{\gamma}$ and $\psi_{\gamma}$. Recall that the Wronskian $W\bigl(\psi_{\gamma},\phi_{\gamma}\bigr) $ is constant on $I$ by Lemma \ref{generaleigenfunction}{\bf (iii)} and Lemma \ref{oknul}. \begin{prop}\label{Wpsiphiprop} Let $\gamma\in\ensuremath{\mathbb{C}}^*$ and $(a,b,c)\in V_z^{gen}$, then \begin{equation}\label{Wpsiphi} W\bigl(\psi_{\gamma},\phi_{\gamma}\bigr)= (1-q)\frac{\bigl(a\gamma,a/\gamma;q\bigr)_{\infty}\theta(bc)} {\bigl(ab,ac,qa/b,qa/c;q\bigr)_{\infty}}. \end{equation} \end{prop} \begin{proof} We first assume that $\gamma\in S_{reg}$. Using Proposition \ref{conncoef} we then have \begin{equation}\label{st1} W\bigl(\psi_{\gamma},\phi_{\gamma}\bigr)= \tilde{c}(\gamma)c(\gamma^{-1}) W\bigl(\Phi_{\gamma}, \Phi_{\gamma^{-1}}\bigr)(z) +\tilde{c}(\gamma^{-1})c(\gamma) W\bigl(\Phi_{\gamma^{-1}}, \Phi_{\gamma}\bigr)(z), \end{equation} where $c(\cdot)$ and $\tilde{c}(\cdot)$ are given by \eqref{cB} and \eqref{tildecB}, respectively. By Lemma \ref{dep}, the right hand side of \eqref{st1} is equal to \begin{equation}\label{st2} W\bigl(\psi_{\gamma},\phi_{\gamma}\bigr)=-aK \left(c(\gamma){\tilde{c}}(\gamma^{-1})- c(\gamma^{-1}){\tilde{c}}(\gamma)\right)\bigl(\gamma-\gamma^{-1}\bigr), \end{equation} where $K$ is the positive constant defined by \eqref{Kconstant}. It follows by direct computation that \begin{equation*} \begin{split} c(\gamma){\tilde{c}}(\gamma^{-1})- c(\gamma^{-1}){\tilde{c}}(\gamma)= &\frac{1}{\bigl(ab,ac,qa/b,qa/c;q\bigr)_{\infty}\theta(-q/bcz,-1/z)}\\ .\frac{\bigl(a\gamma,a/\gamma;q\bigr)_{\infty}} {\bigl(\gamma^2, \gamma^{-2};q\bigr)_{\infty}} &\left\{\theta\bigl(q\gamma/b, c/\gamma, -q/abcz\gamma, -\gamma/az\bigr)\right.\\ &\left.\quad-\theta\bigl(q/b\gamma, c\gamma,-q\gamma/abcz,-1/az\gamma\bigr) \right\}. \end{split} \end{equation*} Now we can apply the $\theta$-product identity \begin{equation}\label{thetaidentity} \theta(x\lambda, x/\lambda,\mu\nu,\mu/\nu)-\theta(x\nu,x/\nu,\lambda\mu,\mu/\lambda)= \frac{\mu}{\lambda}\theta(x\mu,x/\mu,\lambda\nu,\lambda/\nu), \end{equation} see \cite[Exercise 2.16]{GR}, with parameter values \[x=c\sqrt{\frac{q}{bc}}, \quad \lambda=\gamma\sqrt{\frac{q}{bc}}, \quad \mu=-\frac{1}{az}{\sqrt{\frac{q}{bc}}}, \quad \nu=\frac{1}{\gamma}\sqrt{\frac{q}{bc}}, \] to obtain \begin{equation}\label{st3} \begin{split} c(\gamma){\tilde{c}}(\gamma^{-1})- c(\gamma^{-1}){\tilde{c}}(\gamma)= &\frac{-1}{\bigl(ab,ac,qa/b,qa/c;q\bigr)_{\infty}\theta(-q/bcz,-1/z)}\\ &.\frac{1}{az\gamma}\frac{\bigl(a\gamma,a/\gamma;q\bigr)_{\infty}} {\bigl(\gamma^2, \gamma^{-2};q\bigr)_{\infty}} \theta(-q/abz,-acz,q/bc,\gamma^2). \end{split} \end{equation} The proposition now follows for $\gamma\in S_{reg}$ by substitution of \eqref{st3} in \eqref{st2} and using $\theta(x)=\theta(q/x)$. By continuity in $\gamma$, it follows that \eqref{Wpsiphi} holds for all $\gamma\in \ensuremath{\mathbb{C}}^*$. \end{proof} In the following corollary we give a characterization of the eigenfunction $\phi_{\gamma}$ of $L$ for generic values of the spectral parameter $\gamma$. In section \ref{PID} we show that the eigenfunction $\phi_{\gamma}$ plays the role of the spherical function for the big $q$-Jacobi function transform. \begin{cor}\label{psifoutind} Let $\gamma\in \ensuremath{\mathbb{C}}^*$ and $(a,b,c)\in V_z^{gen}$.\\ {\bf (i)} $\{\phi_{\gamma}, \psi_{\gamma}\}$ is a linear basis of $V_{\mu(\gamma)}$ if \begin{equation}\label{con4} \gamma\not\in \{aq^n,a^{-1}q^{-n}\}_{n\in \ensuremath{\mathbb{Z}}_+}. \end{equation} {\bf (ii)} Assume that \eqref{con4} is satisfied. Let $f\in {\mathcal{F}}(I)$ be a function satisfying {\bf --} $Lf=\mu(\gamma)f$ on $I$; {\bf --} $f$ is continuously differentiable at the origin; {\bf --} $f(-1)=1$.\\ Then $f=\phi_{\gamma}$.\\ {\bf (iii)} If \eqref{con4} is satisfied, then $\bigl(L\psi_{\gamma}\bigr)(-1)\not=\mu(\gamma)\psi_{\gamma}(-1)$. \end{cor} \begin{proof} We fix $\gamma\in\ensuremath{\mathbb{C}}^*$ satisfying \eqref{con4}. Then the Wronskian $W(\psi_{\gamma},\phi_{\gamma})$ is non-zero since $(a,b,c)\in V_z^{gen}$, see Proposition \ref{Wpsiphiprop}. Hence $\{\phi_{\gamma}, \psi_{\gamma}\}$ is a linear basis of $V_{\mu(\gamma)}$ by Proposition \ref{generaleigenfunctionprop} and Lemma \ref{oknul}. For the proof of {\bf (ii)}, we observe first that $\phi_{\gamma}$ satisfies the three properties as stated in {\bf (ii)}, see section \ref{eigenf}. If $f$ is another function satisfying the same three properties, then $f|_{I_-}=\phi_{\gamma}|_{I_-}$ since $f(-1)=\phi_{\gamma}(-1)=1$ and $\hbox{dim}\bigl(V_{\mu(\gamma)}^-\bigr)=1$ by Lemma \ref{generaleigenfunction}. Hence $f=\phi_{\gamma}$ on $I$ by Corollary \ref{extensionpossible}. Finally, for the proof of {\bf (iii)}, observe that $\psi_{\gamma}|_{I_-}$ and $\phi_{\gamma}|_{I_-}$ are linearly independent by part {\bf (i)} of the corollary and by Proposition \ref{generaleigenfunctionprop}. Since $\phi_{\gamma}|_{I_-}\in V_{\mu(\gamma)}^-$ by part {\bf (ii)} of the corollary and $\hbox{dim}\bigl(V_{\mu(\gamma)}^-\bigr)=1$ by Lemma \ref{generaleigenfunction}{\bf (i)}, it follows that $\psi_{\gamma}|_{I_-}\in \hat{V}_{\mu(\gamma)}^-\setminus V_{\mu(\gamma)}^-$. So $\bigl(L\psi_{\gamma}\bigr)(-1)\not=\mu(\gamma)\psi_{\gamma}(-1)$, as desired. \end{proof} Set $\gamma_n=aq^n$ ($n\in \ensuremath{\mathbb{Z}}_+$). We have the following description of the excluded set $\{\gamma_n^{\pm 1}\}_{n\in\ensuremath{\mathbb{Z}}_+}$ in Corollary \ref{psifoutind}. \begin{prop}\label{polcase} Let $S_{pol}\subset \ensuremath{\mathbb{C}}^*$ be the set of spectral parameters $\gamma\in \ensuremath{\mathbb{C}}^*$ for which the eigenvalue equation $\bigl(Lf\bigr)(x)= \mu(\gamma)f(x)$ on $I$ has a solution $f\not=0$ which is polynomial in $x$. Then $S_{pol}=\{\gamma_n^{\pm 1}\}_{n\in\ensuremath{\mathbb{Z}}_+}$. The non-zero polynomial eigenfunction corresponding to the eigenvalue $\mu(\gamma_n^{\pm 1})$ is the big $q$-Jacobi polynomial of degree $n$ and is explicitly given by \begin{equation}\label{connpol} \begin{split} \phi_{\gamma_n}(x)&=\frac{\bigl(qa/b,qa/c;q\bigr)_n} {\bigl(ab,ac;q\bigr)_n}\left(\frac{bc}{q}\right)^n\psi_{\gamma_n}(x)\\ &=\frac{\bigl(qa/c;q\bigr)_n}{\bigl(ac;q\bigr)_n} \left(\frac{-c}{aq^{(n+1)/2}}\right)^n {}_3\phi_2\left( \begin{array}{c} q^{-n}, -abx, q^{n}a^2 \\ ab, qa/c \end{array} ; q,q \right). \end{split} \end{equation} \end{prop} \begin{proof} This is well known, see for instance \cite{AA} and \cite[section 7.3]{GR}. The connection between $\phi_{\gamma_n}$ and $\psi_{\gamma_n}$ with the ${}_3\phi_2$ in the last equality of \eqref{connpol} follows from \cite[(3.2.5)]{GR}. \end{proof} \begin{rem} Proposition \ref{Wpsiphi} and Proposition \ref{polcase} are essential ingredients for a functional-analytic derivation of the orthogonality relations and the quadratic norm evaluations for the big $q$-Jacobi polynomials, see section \ref{polsection}. \end{rem} We write $S_{pol}^{\pm}=\{\gamma_n^{\pm 1}\}_{n\in\ensuremath{\mathbb{Z}}_+}$, where $\gamma_n=aq^n$. Observe that $S_{pol}^+\subset S_{pol}\subset S_{reg}\subset S_{reg}^+$ when $(a,b,c)\in V_z^{gen}$. \begin{prop}\label{extensionPhi} Let $\gamma\in S_{reg}^+\setminus S_{pol}^+$ and $(a,b,c)\in V_z^{gen}$. Then \begin{equation}\label{connpsiphi} \Phi_{\gamma}(x)=K(\gamma)\phi_{\gamma}(x)+ \tilde{K}(\gamma)\psi_{\gamma}(x),\qquad x\in I_+, \end{equation} with $K(\gamma)=K(\gamma;a,b,c;z)$ given by \[ K(\gamma)=\frac{\bigl(ab,ac,q\gamma/b,q\gamma/c;q\bigr)_{\infty}} {\bigl(q\gamma^2,a/\gamma;q\bigr)_{\infty}} \frac{\theta(-bcz,-az/\gamma)}{\theta(bc,-abz,-acz)}, \] and $\tilde{K}(\gamma)=\tilde{K}(\gamma;a,b,c;z)= K(\gamma;a,q/b,q/c;bcz/q)$. \end{prop} \begin{proof} We first prove \eqref{connpsiphi} for $\gamma\in S_{reg}\setminus S_{pol}$. By Corollary \ref{psifoutind}{\bf (i)} and Proposition \ref{generaleigenfunctionprop}, there exist unique $K(\gamma), \tilde{K}(\gamma)\in \ensuremath{\mathbb{C}}$ such that \eqref{connpsiphi} holds for all $x\in I_+$. These coefficients can be expressed in terms of Wronskians by \begin{equation}\label{exprW} K(\gamma)= \frac{W(\psi_{\gamma},\Phi_{\gamma})(z)}{W(\psi_{\gamma},\phi_{\gamma})}, \qquad \tilde{K}(\gamma)= \frac{W(\Phi_{\gamma},\phi_{\gamma})(z)}{W(\psi_{\gamma},\phi_{\gamma})}. \end{equation} By Lemma \ref{dep} and Proposition \ref{conncoef} we have \begin{equation}\label{laat} \begin{split} W(\psi_{\gamma},\Phi_{\gamma})(z)&= aK\tilde{c}(\gamma^{-1})\bigl(\gamma^{-1}-\gamma\bigr),\\ W(\Phi_{\gamma},\phi_{\gamma})(z)&= aKc(\gamma^{-1})\bigl(\gamma-\gamma^{-1}\bigr), \end{split} \end{equation} where $K$ is the positive constant defined by \eqref{Kconstant}. Formula \eqref{connpsiphi} now follows by substituting \eqref{laat}, the explicit formula \eqref{Wpsiphi} for the Wronskian $W(\psi_{\gamma},\phi_{\gamma})$, and the explicit expressions \eqref{cB} and \eqref{tildecB} for $c(\gamma)$ and $\tilde{c}(\gamma)$ in \eqref{exprW}, and using the theta function identities $\theta(x)=\theta(q/x)$ and \eqref{functheta}. It follows now by continuity in $\gamma$ that \eqref{connpsiphi} is valid for $\gamma\in S_{reg}^+\setminus S_{pol}^+$. \end{proof} \begin{rem} Fix $\gamma\in S_{reg}\setminus S_{pol}$ and suppose that $(a,b,c)\in V_z^{gen}$. The proof of Proposition \ref{Wpsiphiprop} implies the matrix equation \begin{equation}\label{matrixequation} \begin{pmatrix} c(\gamma) & c(\gamma^{-1})\\ \tilde{c}(\gamma) & \tilde{c}(\gamma^{-1}) \end{pmatrix} =\begin{pmatrix} K(\gamma) & \tilde{K}(\gamma)\\ K(\gamma^{-1}) & \tilde{K}(\gamma^{-1}) \end{pmatrix}^{-1}. \end{equation} Indeed, \eqref{matrixequation} follows easily from the explicit formula \eqref{st3} for the determinant of the matrix on the left hand side of \eqref{matrixequation}. Let $x\in I_+$, then the matrix equation \eqref{matrixequation} implies that the two connection coefficient formulas \[ \phi_{\gamma}(x)=c(\gamma)\Phi_{\gamma}(x)+ c(\gamma^{-1})\Phi_{\gamma^{-1}}(x),\quad \psi_{\gamma}(x)=\tilde{c}(\gamma)\Phi_{\gamma}(x)+ \tilde{c}(\gamma^{-1})\Phi_{\gamma^{-1}}(x) \] are equivalent to the two connection coefficient formulas $\Phi_{\gamma^{\pm 1}}(x)=K(\gamma^{\pm 1})\phi_{\gamma}(x)+ \tilde{K}(\gamma^{\pm 1})\psi_{\gamma}(x)$. \end{rem} \begin{cor}\label{imp} Let $\gamma\in S_{reg}^+\setminus S_{pol}$ and $(a,b,c)\in V_z^{gen}$. The unique extension of $\Phi_{\gamma}\in V_{\mu(\gamma)}^+$ to a function $\Phi_{\gamma}\in V_{\mu(\gamma)}$ is given by \begin{equation}\label{extensionPhi!} \Phi_{\gamma}(x)=K(\gamma)\phi_{\gamma}(x)+\tilde{K}(\gamma)\psi_{\gamma}(x), \qquad x\in I, \end{equation} with $K(\gamma)$, $\tilde{K}(\gamma)$ as defined in Proposition \ref{Wpsiphiprop}. \end{cor} \begin{proof} Immediate from Corollary \ref{extensionpossible}, Corollary \ref{psifoutind}{\bf (i)} and Proposition \ref{extensionPhi}. \end{proof} In section \ref{A} we have seen that $\Phi_{\gamma}(x)$ depends analytically on $\gamma\in S_{reg}^+$ for all $x\in I_+$. We have the following analogous result for the extension \eqref{extensionPhi!} of $\Phi_{\gamma}\in V_{\mu(\gamma)}^+$. \begin{lem}\label{reg} Let $\gamma\in S_{pol}$ and $(a,b,c)\in V_z^{gen}$. There exists a unique extension of $\Phi_{\gamma}\in V_{\mu(\gamma)}^+$ to a function $\Phi_{\gamma}\in V_{\mu(\gamma)}$ such that $\tilde{\gamma}\mapsto \Phi_{\tilde{\gamma}}(x)$ is analytic at $\tilde{\gamma}=\gamma$ for all $x\in I$. If $\gamma=\gamma_n^{-1}\in S_{pol}^-$, then the extension of $\Phi_{\gamma}$ is given by \eqref{extensionPhi!}. If $\gamma=\gamma_n\in S_{pol}^+$, then the extension of $\Phi_{\gamma_n}$ is given by \begin{equation}\label{anext} \Phi_{\gamma_n}(x)= \left(\underset{\gamma=\gamma_n}{\hbox{Res}}K(\gamma)\right) \frac{\partial\phi_{\gamma}}{\partial\gamma}(x)|_{\gamma=\gamma_n} +M_n\phi_{\gamma_n}(x)+ \left(\underset{\gamma=\gamma_n}{\hbox{Res}} \tilde{K}(\gamma)\right) \frac{\partial\psi_{\gamma}}{\partial\gamma}(x)|_{\gamma=\gamma_n} \end{equation} for $x\in I$, with $M_n$ given by the existing limit $M_n=\lim_{\gamma\to\gamma_n}M_n(\gamma)$, where \begin{equation}\label{Mngamma} M_n(\gamma)=\frac{\bigl(qa/b,qa/c;q\bigr)_n}{\bigl(ab,ac;q\bigr)_n} \left(\frac{bc}{q}\right)^nK(\gamma)+\tilde{K}(\gamma). \end{equation} \end{lem} \begin{proof} The proof for $\gamma\in S_{pol}^-$ is trivial since $K(\tilde{\gamma})$, $\tilde{K}(\tilde{\gamma})$, $\phi_{\tilde{\gamma}}(x)$ and $\psi_{\tilde{\gamma}}(x)$ are regular at $\tilde{\gamma}=\gamma$. For $\gamma_n\in S_{pol}^+$ ($n\in\ensuremath{\mathbb{Z}}_+$), observe that $K(\gamma)$ and $\tilde{K}(\gamma)$ have simple poles at $\gamma=\gamma_n$ and that $\phi_{\gamma}(x)$ and $\psi_{\gamma}(x)$ are regular at $\gamma=\gamma_n$. It follows from \eqref{extensionPhi!} and the first equality of \eqref{connpol} that the singularity of $\Phi_{\gamma}(x)$ at $\gamma=\gamma_n$ is removable for $x\in I$ if the (at most simple) singularity of $M_n(\gamma)$ at $\gamma=\gamma_n$ is removable. This can be checked by direct computions using the theta function identities $\theta(x)=\theta(q/x)$ and \eqref{functheta}. It follows that $\lim_{\gamma\to\gamma_n}M_n(\gamma)$ exists and that \eqref{anext} holds. The extension \eqref{anext} of $\Phi_{\gamma_n}$ lies in $V_{\mu(\gamma_n)}$ because $\phi_{\gamma_n}$ and the derivatives of $\phi_{\gamma}$ and $\psi_{\gamma}$ at $\gamma=\gamma_n$ are continuously differentiable at the origin, cf. Lemma \ref{oknul}. \end{proof} For future reference, we collect here the main results concerning the asymptotic solution of the eigenvalue equation $Lf=\mu(\gamma)f$. \begin{thm}\label{samenvatting} Let $x\in I$ and $(a,b,c)\in V_z^{gen}$. {\bf (i)} $\Phi_{\gamma}\in V_{\mu(\gamma)}$ for all $\gamma\in S_{reg}^+$. {\bf (ii)} $\Phi_{\gamma}(x)$ is analytic at $\gamma\in S_{reg}^+$. {\bf (iii)} $\Phi_{\gamma}\in {\mathcal{D}}$ for $\gamma\in \ensuremath{\mathbb{C}}^*$ with $|\gamma|<1$. {\bf (iv)} $\phi_{\gamma}(x)=c(\gamma)\Phi_{\gamma}(x)+ c(\gamma^{-1})\Phi_{\gamma^{-1}}(x)$ for all $\gamma\in S_{reg}$. {\bf (v)} Let $\gamma\in S_{reg}^+$. The Wronskian $W(\gamma)=W\bigl(\Phi_{\gamma},\phi_{\gamma}\bigr)$ is constant on $I$, and $W(\gamma)=aKc(\gamma^{-1})\bigl(\gamma-\gamma^{-1}\bigr)$, where $K$ is the positive constant defined by \eqref{Kconstant}. \end{thm} \begin{proof} {\bf (i)} and {\bf (ii)} follow from Corollary \ref{imp} and Lemma \ref{reg}.\\ {\bf (iii)} Let $\gamma\in \ensuremath{\mathbb{C}}^*$ with $|\gamma|<1$, then $\gamma\in S_{reg}^+$, hence $\Phi_{\gamma}$ is well defined. Then $\Phi_{\gamma}\in {\mathcal{D}}$ follows from part {\bf (i)} of the theorem and from Lemma \ref{l2}.\\ {\bf (iv)} We first prove the connection coefficient formula for $\gamma\in S_{reg}\setminus S_{pol}$. Then the connection coefficient formula is valid for all $x\in I_+$, see Proposition \ref{conncoef}. Since $\phi_{\gamma},\Phi_{\gamma^{\pm 1}}\in V_{\mu(\gamma)}$, it follows by the uniqueness property of extensions of eigenfunctions, see Corollary \ref{extensionpossible} and Corollary \ref{psifoutind}{\bf (i)}, that the connection coefficient formula is valid for all $x\in I$. The connection coefficient formula holds then for all $\gamma\in S_{reg}$ by continuity.\\ {\bf (v)} This follows from Lemma \ref{generaleigenfunction}{\bf (iii)}, part {\bf (iv)} of the theorem, and Lemma \ref{dep}. \end{proof} \begin{rem} Observe that the statements of Theorem \ref{samenvatting}{\bf (i)} and {\bf (iii)} are not in contradiction with the self-adjointness of $(L,{\mathcal{D}})$, see Proposition \ref{s}. Indeed, let $\gamma\in \ensuremath{\mathbb{C}}\setminus \ensuremath{\mathbb{R}}$ with $|\gamma|<1$, then $\gamma\in S_{reg}^+$ and $\mu(\gamma)\in \ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$. By Theorem \ref{samenvatting}{\bf (i)} and {\bf (iii)} we have that $\Phi_{\gamma}\in {\mathcal{D}}$ and $\bigl(L\Phi_{\gamma}\bigr)(x)=\mu(\gamma)\Phi_{\gamma}(x)$ for all $x\in (-1,\infty(z))_q$, but the eigenvalue equation does not hold in the end-point $x=-1$. In fact, the self-adjointness of $(L,{\mathcal{D}})$ forces that $\bigl(L\Phi_{\gamma}\bigr)(-1)\not= \mu(\gamma)\Phi_{\gamma}(-1)$. Another proof of this inequality can be given using the non-vanishing of the Wronskian $W(\phi_{\gamma},\Phi_{\gamma})$, cf. the proof of Corollary \ref{psifoutind}{\bf (iii)}. \end{rem} \section{The Green function} In this section we define the Green function of the self-adjoint operator $(L,{\mathcal{D}})$. With the proper extensions of the asymptotic expansion $\Phi_{\gamma}$ now at hand, see Theorem \ref{samenvatting}, the construction of the Green function is a straightforward extension of the construction given by Kakehi \cite{K} and Kakehi, Masuda and Ueno \cite{KMU} for the little $q$-Jacobi function transform. We use some standard terminology and results for unbounded self-adjoint operators and their spectral measures for which we refer to Dunford and Schwartz \cite[Chapter XII]{DS} and Rudin \cite[Chapter 13]{R}. Let $(a,b,c)\in V_z^{gen}$ and $\mu\in\ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$. Let $\gamma\in\ensuremath{\mathbb{C}}^*$ be the unique non-zero complex number such that $\mu=\mu(\gamma)$ and $|\gamma|<1$. Observe that $\gamma\in\ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$, hence $\gamma\in S_{reg}$ and $W(\gamma)\not=0$, see Theorem \ref{samenvatting}{\bf (v)}. We define the Green kernel $K_{\gamma}(x,y)$ for $x,y\in I$ by \begin{equation}\label{K} K_{\gamma}(x,y)= \begin{cases} W(\gamma)^{-1}\Phi_{\gamma}(x)\phi_{\gamma}(y), \qquad &y\leq x,\\ W(\gamma)^{-1}\phi_{\gamma}(x)\Phi_{\gamma}(y), \qquad &y\geq x. \end{cases} \end{equation} Observe that $K_{\gamma}(x,\cdot), K_{\gamma}(\cdot,x)\in {\mathcal{D}}$ for all $x\in I$ in view of Lemma \ref{oknul} and Theorem \ref{samenvatting}{\bf (i)} and {\bf (iii)}, so we have a well defined linear map ${\mathcal{H}}\to {\mathcal{F}}(I)$ mapping $f\in {\mathcal{H}}$ to \begin{equation}\label{Greenfunction} G_f(x,\gamma)=\langle f, {\overline{K_{\gamma}(x,\cdot)}}\rangle, \qquad x\in I. \end{equation} Written out explicitly, we arrive at the formula \begin{equation}\label{expl} G_f(x,\gamma)=W(\gamma)^{-1} \left(\Phi_{\gamma}(x)\int_{-1}^xf(y)\phi_{\gamma}(y)\frac{d_qy}{p(y)} +\phi_{\gamma}(x)\int_{x}^{\infty(z)}f(y)\Phi_{\gamma}(y) \frac{d_qy}{p(y)}\right). \end{equation} By the self-adjointness of $(L,{\mathcal{D}})$ we have that the resolvent $\bigl(L-\mu.\hbox{Id}\bigr)^{-1}$ is a one to one, continuous map from ${\mathcal{H}}$ onto ${\mathcal{D}}$ for all $\mu\in\ensuremath{\mathbb{C}}\setminus \ensuremath{\mathbb{R}}$. \begin{prop}\label{G} Let $(a,b,c)\in V_z^{gen}$, $f\in {\mathcal{H}}$ and $\mu=\mu(\gamma)\in \ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$ with $\gamma\in \ensuremath{\mathbb{C}}^*$ and $|\gamma|<1$. Then $G_f(\cdot,\gamma)=\bigl(L-\mu.\hbox{Id}\bigr)^{-1}f$. In particular, $G_f(\cdot,\gamma)\in {\mathcal{D}}$. \end{prop} \begin{proof} We first prove that \begin{equation}\label{check} \bigl(\bigl(L-\mu(\gamma)\bigr)G_f(\cdot,\gamma)\bigr)(x)=f(x), \qquad \forall f\in {\mathcal{H}}. \end{equation} For the proof we need to consider the two cases $x\in (-1,\infty(z))_q$ and $x=-1$ seperately.\\ {\it Case 1:} $x\in (-1,\infty(z))_q$. Observe that the product rule \eqref{productrule} for the $q$-derivative $D_q$ and Lemma \ref{self-adjointform} imply the following product rule for $L$: \begin{equation*} \begin{split} \bigl(L(fg)\bigr)(x)&=\bigl(Lf\bigr)(x)g(x)+f(x)\bigl(Lg\bigr)(x) +qp(x)r(x)\bigl(D_qf\bigr)(x)\bigl(D_qg\bigr)(x)\\ &\qquad\qquad\qquad +p(x)r(q^{-1}x)\bigl(D_qf\bigr)(q^{-1}x)\bigl(D_qg\bigr)(q^{-1}x). \end{split} \end{equation*} Combined with the easily verified formulas \begin{equation*} D_q\left(x\mapsto \int_{-1}^xf(y)d_qy\right)=f(x),\qquad D_q\left(x\mapsto\int_{x}^{\infty(z)}f(y)d_qy\right)=-f(x) \end{equation*} and the definition \eqref{expl} of $G_f(x,\gamma)$, we obtain \begin{equation*} \begin{split} &W(\gamma)\left(\bigl(L- \mu(\gamma)\bigr)G_f(\cdot,\gamma)\right)(x)=\\ &\quad\qquad\Phi_{\gamma}(x)p(x)\left(D_q(p^{-1} rf\phi_{\gamma})\right)(q^{-1}x) -\phi_{\gamma}(x)p(x)\left(D_q(p^{-1}rf\Phi_{\gamma})\right)(q^{-1}x)\\ &\quad\qquad+p(x)\left(p^{-1}rf\bigl(D_q\Phi_{\gamma}\bigr) \phi_{\gamma}\right)(q^{-1}x) -p(x)\left(p^{-1}rf\bigl(D_q\phi_{\gamma}\bigr) \Phi_{\gamma}\right)(q^{-1}x)\\ &\quad\qquad+q\left(rf\bigl(D_q\Phi_{\gamma}\bigr)\phi_{\gamma}\right)(x) -q\left(rf\bigl(D_q\phi_{\gamma}\bigr)\Phi_{\gamma}\right)(x) \end{split} \end{equation*} for $f\in {\mathcal{H}}$. By writing out the $D_q$-terms in this formula, we see that the right hand side reduces to $W(\Phi_{\gamma},\phi_{\gamma})(x)f(x)$, which in turn is equal to $W(\gamma)f(x)$ by Theorem \ref{samenvatting}{\bf (v)}. This completes the proof of \eqref{check} for $x\in (-1,\infty(z))_q$.\\ {\it Case 2:} $x=-1$. Observe that the Green function in the end-point $-1$ reduces to \[G_f(-1,\gamma)=\int_{-1}^{\infty(z)}f(y)\Phi_{\gamma}(y)\frac{d_qy}{p(y)} \] since $\phi_{\gamma}(-1)=1$. Using the product rule \eqref{productrule}, we then have that \begin{equation*} \begin{split} W(\gamma)&\bigl(D_qG_f(\cdot,\gamma)\bigr)(-1)= \Phi_{\gamma}(-q)\phi_{\gamma}(-1)\frac{f(-1)}{p(-1)}\\ &\qquad\qquad-\phi_{\gamma}(-q)\Phi_{\gamma}(-1)\frac{f(-1)}{p(-1)} +\bigl(D_q\phi_{\gamma}\bigr)(-1)G_f(-1,\gamma). \end{split} \end{equation*} Combined with the expression of $L$ in the end-point $x=-1$ in terms of the $q$-derivative and the functions $p(\cdot)$ and $r(\cdot)$, see Lemma \ref{self-adjointform}, and using the fact that $(L\phi_{\gamma})(-1)=\mu(\gamma)\phi_{\gamma}(-1)$, we obtain \[ W(\gamma)\left(\bigl(L-\mu(\gamma)\bigr)G_f(\cdot,\gamma)\right)(-1)= W(\Phi_{\gamma},\phi_{\gamma})(-1)f(-1). \] Hence \eqref{check} is valid for the end-point $x=-1$ since $W(\Phi_{\gamma},\phi_{\gamma})(-1)=W(\gamma)$ by Theorem \ref{samenvatting}{\bf (v)}. It remains to show how \eqref{check} leads to a proof of the proposition. Let ${\mathcal{D}}_{fin}$ be the set of functions $f:I\to\ensuremath{\mathbb{C}}$ with finite support. Then ${\mathcal{D}}_{fin}\subset {\mathcal{D}}\subset {\mathcal{H}}$ as dense subspaces. Let $f\in {\mathcal{D}}_{fin}$, then $G_f(\cdot,\gamma)$ is continuously differentiable at the origin since $\phi_{\gamma}, \Phi_{\gamma}\in V_{\mu(\gamma)}$. Furthermore, $G_f(x,\gamma)$ is a constant multiple of $\Phi_{\gamma}(x)$ for $x\gg 0$, hence $G_f(\cdot,\gamma)\in {\mathcal{H}}$. By \eqref{check} it follows that $LG_f(\cdot,\gamma)\in {\mathcal{H}}$, hence $G_f(\cdot,\gamma)\in {\mathcal{D}}$. Combined with \eqref{check}, this proves that \begin{equation}\label{check2} \bigl((L-\mu(\gamma).\hbox{Id})^{-1}f\bigr)(x)=G_f(x,\gamma)= \langle f,{\overline{K_{\gamma}(x,\cdot)}}\rangle,\qquad \forall f\in {\mathcal{D}}_{fin} \end{equation} for all $x\in I$. By continuity, \eqref{check2} is valid for all $f\in {\mathcal{H}}$. This completes the proof of the proposition. \end{proof} \begin{rem}\label{phibetter2} Proposition \ref{G} is in general not valid when $\phi_{\gamma}$ is replaced by $\psi_{\gamma}$ and $W(\gamma)$ is replaced by the constant value of $W(\Phi_{\gamma},\psi_{\gamma})$ on $I$ in the definition of the Green function $G_f(\cdot,\gamma)$. Indeed, for the proof of $\left(\bigl(L-\mu(\gamma)\bigr)G_f(\cdot,\gamma)\right)(-1)=f(-1)$ in the previous proposition, we use the fact that $\phi_{\gamma}$ is a solution of $\bigl(Lf\bigr)(x)=\mu(\gamma)f(x)$ in the end-point $x=-1$. This property fails in general to be true for $\psi_{\gamma}$, see Corollary \ref{psifoutind}{\bf (iii)}. \end{rem} Proposition \ref{G} plays a crucial role in determining the explicit form of the resolution of the identity $L=\int_{\ensuremath{\mathbb{R}}}tdE(t)$ for the self-adjoint operator $(L,{\mathcal{D}})$ on ${\mathcal{H}}$ since the spectral measure $E$ is related to the resolvent of $L$ by \begin{equation}\label{inversion} \begin{split} &\langle E\bigl((\mu_1,\mu_2)\bigr)f,g\rangle\\ &\,\,=\lim_{\delta\downarrow 0}\lim_{\epsilon\downarrow 0} \frac{1}{2\pi i}\int_{\mu_1+\delta}^{\mu_2-\delta} \left(\langle\bigl(L-(\mu+i\epsilon)\bigr)^{-1}f,g\rangle- \langle\bigl(L-(\mu-i\epsilon)\bigr)^{-1}f,g\rangle\right)d\mu, \end{split} \end{equation} where $\mu_1<\mu_2$ and $f\in {\mathcal{D}}$, $g\in {\mathcal{H}}$, see \cite[Theorem XII.2.10]{DS}. In the following two sections we use Proposition \ref{G} and \eqref{inversion} to give an explicit description of the continuous and discrete contributions to the spectral resolution $E$. \section{The continuous spectrum}\label{contmass} We start this section by proving that the closed interval $[-(1+a)^2,-(1-a)^2]$ is contained in the continuous spectrum $\sigma_c(L)$ of $(L,{\mathcal{D}})$. In section \ref{pointmass} we will see in fact that this interval is exactly equal to $\sigma_c(L)$ for $(a,b,c)\in V_z^{gen}$. Furthermore, we compute the spectral projection $P_c:=E\bigl([-(1+a)^2,-(1-a)^2]\bigr)$ explicitly, and give the Plancherel formula and inversion formula for the continuous part of the big $q$-Jacobi function transform. For $n\in \ensuremath{\mathbb{N}}$ and $x\in I$ we set $\varphi_\gamma^{(n)}(x)=\phi_{\gamma}^{(n)}(x)/\|\phi_{\gamma}^{(n)}\|$, where $\phi_{\gamma}^{(n)}\in {\mathcal{H}}$ is defined by \begin{equation*} \phi_{\gamma}^{(n)}(x)= \begin{cases} \phi_{\gamma}(x)\quad &\hbox{ if } \,\,\, x\in I\setminus [zq^{-n-1},\infty(z))_q,\\ 0 &\hbox{ if } \,\,\, x\in [zq^{-n-1},\infty(z))_q. \end{cases} \end{equation*} \begin{lem}\label{contspectrum} Let $\mu\in [-(1+a)^2,-(1-a)^2]$ and $\theta\in [0,\pi]$ such that $\mu=\mu(e^{i\theta})$. Then $\bigl(\varphi_{e^{i\theta}}^{(n)}\bigr)_{n\in\ensuremath{\mathbb{N}}}$ is a generalized eigenfunction of $\bigl(L,{\mathcal{D}}\bigr)$ with generalized eigenvalue $\mu$. Furthermore, the continuous spectrum $\sigma_c(L)$ of the self-adjoint operator $\bigl(L, {\mathcal{D}}\bigr)$ contains the interval $[-(1+a)^2, -(1-a)^2]$. \end{lem} \begin{proof} Recall that $\bigl(\varphi_{e^{i\theta}}^{(n)}\bigr)_{n\in\ensuremath{\mathbb{N}}}$ is a generalized eigenfunction of $(L,{\mathcal{D}})$ with generalized eigenvalue $\mu(e^{i\theta})$ if $\varphi_{e^{i\theta}}^{(n)}\in {\mathcal{D}}$, $\|\varphi_{e^{i\theta}}^{(n)}\|=1$ and $\lim_{n\to\infty}\bigl(L\varphi_{e^{i\theta}}^{(n)}- \mu(e^{i\theta})\varphi_{e^{i\theta}}^{(n)}\bigr)=0$ in ${\mathcal{H}}$. This can be checked for all $\theta\in [0,\pi]$ by an elementary computation using Proposition \ref{psifoutind}{\bf (ii)}. Let $\mu\in [-(1+a)^2,-(1-a)^2]$. Then $\mu$ is part of the spectrum $\sigma(L)$ of $(L,{\mathcal{D}})$ since there exists a generalized eigenfunction of $(L,{\mathcal{D}})$ with generalized eigenvalue $\mu$. Then $\mu$ is in the continuous spectrum $\sigma_c(L)$ of $(L,{\mathcal{D}})$ or in the point spectrum $\sigma_p(L)$ of $(L,{\mathcal{D}})$ since a self-adjoint operator does not have residual spectrum, see \cite[Theorem 13.27]{R}. It remains to show that $\mu\not\in \sigma_p(L)$. We have to distinguish between the cases $\mu\in \bigl(-(1+a)^2,-(1-a)^2\bigr)$ and $\mu=\mu(\pm 1)=-(1\mp a)^2$.\\ {\it Case 1:} $\mu\in \bigl(-(1+a)^2,-(1-a)^2\bigr)$, i.e. $\mu=\mu(e^{i\theta})$ with $\theta\in (0,\pi)$. It is a straightforward consequence of the asymptotic behaviour of $\Phi_{\gamma}$ to infinity, see \eqref{asymptoticsPhiB}, that any non-zero linear combination of the basis elements $\{\Phi_{e^{i\theta}}, \Phi_{e^{-i\theta}}\}$ of $V_{\mu(e^{i\theta})}^+$ does not lie in $M_{\mu(e^{i\theta})}$ (see Lemma \ref{l2} for the definition of $M_{\mu}$), hence $\mu\not\in \sigma_p(L)$.\\ {\it Case 2:} $\mu=\mu(\pm 1)=-(1\mp a)^2$. Observe that $\frac{\partial\Phi_{\gamma}}{\partial\gamma}|_{\gamma=\pm 1} \in V_{\mu(\pm 1)}^+$ since $\frac{\partial\mu}{\partial\gamma}(\gamma)|_{\gamma=\pm 1}=0$. Using that the asymptotics to infinity of $\frac{\partial\Phi_{\gamma}}{\partial\gamma}$ is given by \[\frac{\partial\Phi_{\gamma}}{\partial\gamma}(zq^{-m})=m\gamma^{-1} (a\gamma)^m\bigl(1+{\mathcal{O}}(q^m)\bigr),\qquad m\to\infty,\] we obtain that \[W\bigl(\Phi_{\pm 1}, \frac{\partial\Phi_{\gamma}}{\partial\gamma}|_{\gamma=\pm 1}\bigr)(x)= -aK\not=0,\qquad x\in I_+, \] compare with the proof of Lemma \ref{dep}. Combined with Lemma \ref{generaleigenfunction}{\bf (i)}, it follows that $\{\Phi_{\pm 1}, \frac{\partial\Phi_{\gamma}}{\partial\gamma}|_{\gamma=\pm 1}\}$ is a basis of $V_{\mu(\pm 1)}^+$. It is easy to see, using asymptotics to infinity, that any non-zero linear combination of these basis elements does not lie in $M_{\mu(\pm 1)}$, hence $\mu(\pm 1)\not\in \sigma_p(L)$. \end{proof} Let ${\mathcal{D}}_{fin}\subset{\mathcal{H}}$ be the linear subspace of functions $f:I\to\ensuremath{\mathbb{C}}$ with finite support. Observe that ${\mathcal{D}}_{fin}\subset {\mathcal{D}}\subset {\mathcal{H}}$ as dense subspaces. We define the big $q$-Jacobi function transform by \begin{equation}\label{F} \bigl({\mathcal{F}}f\bigr)(\gamma)= \langle f,\phi_{\gamma}\rangle=\int_{-1}^{\infty(z)} f(x){\overline{\phi_{\gamma}(x)}}\frac{d_qx}{p(x)}, \qquad f\in {\mathcal{D}}_{fin},\,\, \gamma\in \ensuremath{\mathbb{C}}^*. \end{equation} In this section, we will regard the big $q$-Jacobi function transform ${\mathcal{F}}f$ of $f\in {\mathcal{D}}_{fin}$ as a function on the unit circle $\ensuremath{\mathbb{T}}$. Observe that ${\mathcal{F}}f$ is $W$-invariant, where $W=\{\pm 1\}$ acts by $\bigl(\ensuremath{\epsilon}.g\bigr)(\gamma)=g\bigl(\gamma^{\ensuremath{\epsilon}}\bigr)$, $\ensuremath{\epsilon}\in W$. We define an absolutely continuous measure on $\ensuremath{\mathbb{T}}$ by \begin{equation}\label{mu} d\nu(\gamma)=\frac{1}{4\pi iK}\frac{d\gamma}{c(\gamma)c(\gamma^{-1})\gamma} =\frac{1}{4\pi iK}\frac{d\gamma}{|c(\gamma)|^2\gamma}, \end{equation} where $K$ is the positive constant defined by \eqref{Kconstant}. Observe that the measure $d\nu$ is a well defined measure on $\ensuremath{\mathbb{T}}$ since $(a,b,c)\in V$. In particular, possible zeros of the denominator of the weight function $1/|c(\cdot)|^2$ are compensated by zeros of the numerator. We first show that ${\mathcal{F}}$ extends uniquely to a partial isometry ${\mathcal{F}}: {\mathcal{H}}\to L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$, where $L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$ is the $L^2$-space of $W$-invariant functions with respect to $d\nu$, when $(a,b,c)\in V_z^{gen}$. We start with the following crucial consequence of \eqref{inversion} and Proposition \ref{G}. \begin{prop}\label{contE} Let $(a,b,c)\in V_z^{gen}$. Choose $-(1+a)^2\leq \mu_1< \mu_2\leq -(1-a)^2$ and let $0\leq \theta_2<\theta_1\leq\pi$ such that $\mu_j=\mu(e^{i\theta_j})$ \textup{(}$j=1,2$\textup{)}. Then \begin{equation}\label{contpart} \langle E\bigl((\mu_1,\mu_2)\bigr)f,g\rangle=\frac{1}{2\pi K} \int_{\theta_2}^{\theta_1} \bigl({\mathcal{F}}f\bigr)(e^{i\theta}) {\overline{\bigl({\mathcal{F}}g\bigr)(e^{i\theta})}} \frac{d\theta}{|c(e^{i\theta})|^2} \end{equation} for all $f,g\in {\mathcal{D}}_{fin}$. \end{prop} \begin{proof} For $\mu\in\ensuremath{\mathbb{C}}\setminus \ensuremath{\mathbb{R}}$ we write $\gamma[\mu]\in\ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$ for the unique complex number with modulus less than $1$ such that $\mu(\gamma[\mu])=\mu$. Fix $f,g\in {\mathcal{D}}_{fin}$. By a straightforward computation using Proposition \ref{G}, we have for $\mu\in \ensuremath{\mathbb{R}}$ and $\epsilon >0$ that \begin{equation}\label{crucfor} \begin{split} \langle \left(L-\bigl(\mu\pm i\epsilon\bigr)\right)^{-1}f,g\rangle= &{\underset{\stackrel{(x,y)\in I\times I}{x\leq y}}{\iint}} \frac{\phi_{\gamma[\mu\pm i\epsilon]}(x) \Phi_{\gamma[\mu\pm i\epsilon]}(y)}{W\bigl(\gamma[\mu\pm i\epsilon]\bigr)}\bigl(1-\frac{1}{2}\delta_{x,y}\bigr)\\ &\qquad.\bigl(f(x){\overline{g(y)}}+f(y){\overline{g(x)}} \bigr)\frac{d_qx}{p(x)}\frac{d_qy}{p(y)}, \end{split} \end{equation} where $\delta_{x,y}$ is the Kronecker-delta. Let $-(1+a)^2<\mu_1\leq\mu\leq\mu_2<-(1-a)^2$ and let $0<\theta_2\leq\theta\leq\theta_1<\pi$ such that $\mu=\mu(e^{i\theta})$ and $\mu_j=\mu(\theta_j)$ for $j=1,2$. Then we have \begin{equation}\label{gammalim} \lim_{\epsilon\downarrow 0}\gamma[\mu\pm i\epsilon]=e^{\mp i\theta}. \end{equation} Using the connection coefficients formula given in Theorem \ref{samenvatting}{\bf (iv)} and using the fact that $W(\gamma)^{-1}$ and $\Phi_{\gamma^{\pm 1}}(x)$ ($x\in I$) are regular at $\gamma\in \ensuremath{\mathbb{T}}\setminus \{\pm 1\}$, we obtain \begin{equation*} \begin{split} \lim_{\ensuremath{\epsilon}\downarrow 0}\left(\frac{\phi_{\gamma[\mu+i\ensuremath{\epsilon}]}(x) \Phi_{\gamma[\mu+i\ensuremath{\epsilon}]}(y)}{W\bigl(\gamma[\mu+i\ensuremath{\epsilon}]\bigr)}-\right. &\left.\frac{\phi_{\gamma[\mu-i\ensuremath{\epsilon}]}(x)\Phi_{\gamma[\mu-i\ensuremath{\epsilon}]}(y)} {W\bigl(\gamma[\mu-i\ensuremath{\epsilon}]\bigr)}\right)\\ =&\frac{\phi_{e^{i\theta}}(x)\Phi_{e^{-i\theta}}(y)} {W\bigl(e^{-i\theta}\bigr)}-\frac{\phi_{e^{i\theta}}(x)\Phi_{e^{i\theta}}(y)} {W\bigl(e^{i\theta}\bigr)}\\ =&\frac{1}{aK}\frac{\phi_{e^{i\theta}}(x)\phi_{e^{i\theta}}(y)} {c\bigl(e^{i\theta}\bigr)c\bigl(e^{-i\theta}\bigr) \bigl(e^{-i\theta}-e^{i\theta}\bigr)}. \end{split} \end{equation*} It follows now by symmetrization of the double $q$-Jackson integral that \begin{equation*} \begin{split} \lim_{\epsilon\downarrow 0}&\left(\langle \left(L-\bigl(\mu+i\epsilon\bigr)\right)^{-1}f,g\rangle- \langle \left(L-\bigl(\mu-i\epsilon\bigr)\right)^{-1}f,g\rangle\right)\\ &=\frac{1}{aK}{\underset{\stackrel{(x,y)\in I\times I}{x\leq y}}{\iint}} \frac{\phi_{e^{i\theta}}(x)\phi_{e^{i\theta}}(y) \bigl(f(x){\overline{g(y)}}+f(y){\overline{g(x)}}\bigr) \bigl(1-\frac{1}{2}\delta_{x,y}\bigr)}{|c(e^{i\theta})|^2 \bigl(e^{-i\theta}-e^{i\theta}\bigr)}\frac{d_qx}{p(x)}\frac{d_qy}{p(y)}\\ &=\frac{1}{aK} \frac{\bigl({\mathcal{F}}f\bigr)(e^{i\theta}) {\overline{\bigl({\mathcal{F}}g\bigr)(e^{i\theta})}}} {|c\bigl(e^{i\theta}\bigr)|^2\bigl(e^{-i\theta}-e^{i\theta}\bigr)}. \end{split} \end{equation*} The proposition follows now for all $-(1+a)^2<\mu_1<\mu_2<-(1-a)^2$ using \eqref{inversion} and changing the integration variable to $\theta$ using the map $\theta\mapsto \mu(e^{i\theta})$, see \eqref{eigenfunction}. The result now also holds when $\mu_1=-(1+a)^2$ or $\mu_2=-(1-a)^2$ since the spectral measure $E$ is countably additive. \end{proof} \begin{cor}\label{gevolg1} The big $q$-Jacobi function transform ${\mathcal{F}}$ uniquely extends to a continuous linear mapping ${\mathcal{F}}_c: {\mathcal{H}}\to L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$. If $(a,b,c)\in V_z^{gen}$, then ${\mathcal{F}}_c: {\mathcal{H}}\to L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$ factorizes through the orthogonal projection $P_c=E\bigl([-(1+a)^2,-(1-a)^2]\bigr)$ \textup{(}i.e. ${\mathcal{F}}_c={\mathcal{F}}_c\circ P_c$\textup{)}, and the restriction of ${\mathcal{F}}_c$ to the range ${\mathcal{R}}(P_c)$ of $P_c$ is an isometric isomorphism onto the range ${\mathcal{R}}({\mathcal{F}}_c)\subset L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$ of ${\mathcal{F}}_c$. \end{cor} \begin{proof} We assume first that $(a,b,c)\in V_z^{gen}$. In view of Proposition \ref{contpart} applied to the special case $\mu_1=\mu(-1)=-(1+a)^2$ and $\mu_2=\mu(1)=-(1-a)^2$, it suffices to observe that $E\bigl(\{\mu(\pm 1)\}\bigr)=0$, which is a consequence of Lemma \ref{contspectrum} and \cite[Theorem 13.27]{R}. Finally, observe that the inequality $\|{\mathcal{F}}f\|_2^2\leq \|f\|^2$ for $f\in {\mathcal{D}}_{fin}$ and $(a,b,c)\in V_z^{gen}$, where $\|.\|_2$ is the norm of $L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$, holds for all $(a,b,c)\in V$ by continuity. Hence ${\mathcal{F}}$ can be uniquely extended to a continuous linear map ${\mathcal{F}}_c: {\mathcal{H}}\to L_W^2\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)$ for all parameters $(a,b,c)\in V$. \end{proof} \begin{Def} We call ${\mathcal{F}}_c$ the continuous part of the big $q$-Jacobi function transform. \end{Def} Observe that the limit $\langle f,\phi_{\gamma}\rangle_l= \lim_{k,m\to\infty}\langle f,\phi_{\gamma}\rangle_{k;l,m}$ exists for all $\gamma\in \ensuremath{\mathbb{T}}$ since $\phi_{\gamma}$ is continuously differentiable at the origin, see Lemma \ref{oknul}. It follows that \begin{equation}\label{pointwise} \bigl({\mathcal{F}}_cf\bigr)(\gamma)= \lim_{l\to-\infty}\langle f,\phi_{\gamma}\rangle_l,\qquad \gamma\in\ensuremath{\mathbb{T}} \,\,\,\hbox{ a.e.} \end{equation} for $f\in {\mathcal{H}}$, where $\langle .,. \rangle_l$ ($l\in\ensuremath{\mathbb{Z}}$) is the truncated inner product \begin{equation}\label{trunc} \langle f,g\rangle_l=\int_{-1}^{zq^l}f(x){\overline{g(x)}}\frac{d_qx} {p(x)}. \end{equation} In the remainder of this section we show that ${\mathcal{F}}_c$ is surjective and we give an explicit formula for the inverse of the isometric isomorphism ${\mathcal{F}}_c: {\mathcal{R}}(P_c)\to L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu)$. The methods we employ are similar to the ones employed by G{\"o}tze \cite{G} and by Braaksma and Meulenbeld \cite{BM} for the classical Jacobi function transform, and by Kakehi \cite{K} and Kakehi, Masuda and Ueno \cite{KMU} for the little $q$-Jacobi function transform. \begin{lem}\label{Wphiphi} Let $\gamma,\delta\in \ensuremath{\mathbb{T}}$ with $\mu(\gamma)\not=\mu(\delta)$. For $k\in \ensuremath{\mathbb{Z}}_+$, $l,m\in \ensuremath{\mathbb{Z}}$ with $l<m$, we have \[ \langle \phi_{\gamma},\phi_{\delta}\rangle_{k;l,m}= \frac{W(\phi_{\gamma},\phi_{\delta})(zq^{l-1}) -W(\phi_{\gamma},\phi_{\delta})(zq^m) +W(\phi_{\gamma},\phi_{\delta})(-q^k)} {\mu(\gamma)-\mu(\delta)}. \] \end{lem} \begin{proof} For $\gamma,\delta\in\ensuremath{\mathbb{T}}$ we have that $\bigl(L\phi_{\gamma}\bigr){\overline{\phi_{\delta}}}- \phi_{\gamma}{\overline{\bigl(L\phi_{\delta}\bigr)}}= \bigl(\mu(\gamma)-\mu(\delta)\bigr) \phi_{\gamma}{\overline{\phi_{\delta}}}$ on $I$. The proof follows now from Lemma \ref{stoktermen} since $\phi_{\gamma}:I\to\ensuremath{\mathbb{C}}$ is real valued for $\gamma\in \ensuremath{\mathbb{T}}$. \end{proof} We define a linear map ${\mathcal{G}}_c: L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)\to {\mathcal{F}}(I)$ by \begin{equation}\label{Finvers} \bigl({\mathcal{G}}_cg\bigr)(x)=\int_{\ensuremath{\mathbb{T}}}g(\gamma)\phi_{\gamma}(x)d\nu(\gamma), \qquad x\in I. \end{equation} Observe that \begin{equation} \langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2 \rangle_{k;l,m}= \int_\ensuremath{\mathbb{T}}\int_\ensuremath{\mathbb{T}} g_1(\gamma){\overline{g_2(\gamma')}} \langle \phi_{\gamma}, \phi_{\gamma'}\rangle_{k;l,m}d\nu(\gamma) d\nu(\gamma') \end{equation} for $k\in \ensuremath{\mathbb{Z}}_+$, $l<m$ in $\ensuremath{\mathbb{Z}}$ and $g_1,g_2\in L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$. \begin{lem}\label{reduction} The limit $\langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle_l =\lim_{k,m\to\infty}\langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle_{k;l,m}$ exists for all $g_1,g_2\in L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$, and \begin{equation*} \begin{split} \langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle_l= &\int_\ensuremath{\mathbb{T}}\int_\ensuremath{\mathbb{T}} g_1(\gamma){\overline{g_2(\gamma')}} \langle \phi_{\gamma}, \phi_{\gamma'}\rangle_ld\nu(\gamma)d\nu(\gamma')\\ =&\int_\ensuremath{\mathbb{T}}\int_\ensuremath{\mathbb{T}} g_1(\gamma){\overline{g_2(\gamma')}} \frac{W\bigl(\phi_{\gamma},\phi_{\gamma'}\bigr)\bigl(zq^{l-1}\bigr)} {\bigl(\mu(\gamma)-\mu(\gamma')\bigr)} d\nu(\gamma)d\nu(\gamma'). \end{split} \end{equation*} \end{lem} \begin{proof} Observe that there exists a constant $K_l>0$ such that $|\langle \phi_{\gamma},\phi_{\gamma'}\rangle_{k;l,m}| \leq K_l$ for all $\gamma,\gamma'\in \ensuremath{\mathbb{T}}$ and for all $k,m\in\ensuremath{\mathbb{Z}}_+$ with $m>l$. The first equality follows then from Lebesgue's dominated convergence theorem. The second equality follows from Lemma \ref{Wphiphi}, using that $W\bigl(\phi_{\gamma},\phi_{\gamma'}\bigr)(0^+)= W\bigl(\phi_{\gamma},\phi_{\gamma'}\bigr)(0^-)$ since $\phi_{\gamma}$ is continuously differentiable at the origin, see Lemma \ref{oknul}. \end{proof} Finally, we determine the limit $\langle {\mathcal{G}}_cg_1,{\mathcal{G}}_cg_2\rangle=\lim_{l\to-\infty} \langle {\mathcal{G}}_cg_1,{\mathcal{G}}_cg_2\rangle_l$. The result is as follows. \begin{prop}\label{weaklimit} The limit $\langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle= \lim_{l\to-\infty}\langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle_l$ exists for all $g_1,g_2\in L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$, and \begin{equation}\label{i} \langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle= \int_{\ensuremath{\mathbb{T}}}g_1(\gamma){\overline{g_2(\gamma)}}d\nu(\gamma). \end{equation} In particular, ${\mathcal{G}}_c :L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)\to {\mathcal{H}}$ is an isometric isomorphism onto the range ${\mathcal{R}}\bigl({\mathcal{G}}_c\bigr)\subset {\mathcal{H}}$ of ${\mathcal{G}}_c$. \end{prop} \begin{proof} We only sketch the proof, since it is similar to the little $q$-Jacobi case, see \cite[Proposition 6.1]{KMU} and \cite[Proposition 7.4]{K}. Let $C_W(\ensuremath{\mathbb{T}})$ be the algebra of complex valued, continuous, $W$-invariant functions on $\ensuremath{\mathbb{T}}$. We fix $g_1,g_2\in C_W(\ensuremath{\mathbb{T}})$ such that $\{-1,1\}$ is not in the support of $g_1$ and $g_2$. It suffices to give a proof of \eqref{i} for such functions $g_1$ and $g_2$. We start with the second expression of $\langle {\mathcal{G}}_{l}g_1, {\mathcal{G}}_lg_2\rangle_l$ in Lemma \ref{reduction}, and replace $\phi_{\gamma}$ by its $c$-function expansion, see Theorem \ref{samenvatting}{\bf (iv)}. Using the estimate \[\sup_{\delta\leq \theta\not=\theta'\leq\pi-\delta} \left|\frac{R_k(e^{\pm i\theta})-R_k(e^{\pm i\theta'})} {e^{i\theta}-e^{i\theta'}}\right|={\mathcal{O}}(k(qa)^k),\qquad k\to\infty \] for $0<\delta<\pi/2$, where $R_k(\gamma)=\Phi_{\gamma}(zq^{-k})-(a\gamma)^k$ (cf. \eqref{asymptoticsPhiB}), which can easily be proved using the mean value theorem, we may replace in the expression of $\lim_{l\to-\infty}\langle {\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle_l$ the function $\Phi_{\gamma}(zq^{l-1})$ by its asymptotic value $(a\gamma)^{1-l}$ at $\infty$. Combined with \eqref{rinfty} it follows now from the bounded convergence theorem that \[ \lim_{m\to\infty}\langle {\mathcal{G}}_cg_1,{\mathcal{G}}_cg_2\rangle_{1-m} =\frac{-a}{4\pi^2K} \lim_{m\to\infty}\int_{0}^\pi \int_{0}^\pi\frac{g_1(e^{i\theta}){\overline{g_2(e^{i\theta'})}} s_m(\theta,\theta')d\theta d\theta'}{|c(e^{i\theta})|^2 |c(e^{i\theta'})|^2(\mu(e^{i\theta'})-\mu(e^{i\theta}))} \] provided that the limit in the right hand side of the equality exists, with \[s_m(\theta,\theta')=\sum_{\epsilon,\xi\in\{\pm 1\}} c(e^{i\epsilon\theta})c(e^{i\xi\theta'})e^{i(m-1)(\epsilon\theta+\xi\theta')} (e^{i\epsilon\theta}-e^{i\xi\theta'}). \] Since $c(\gamma)$ is continuous and non-zero on $\ensuremath{\mathbb{T}}\setminus \{\pm 1\}$, the Riemann-Lebesgue lemma implies that the contributions of the sums of $s_m$ with $\ensuremath{\epsilon}\xi>0$ tend to zero in the limit, so $s_m(\theta,\theta')$ may be replaced by \begin{equation*} \begin{split} t_m(\theta,\theta')= &c(e^{i\theta})c(e^{-i\theta'})e^{i(m-1)(\theta-\theta')}(e^{i\theta}- e^{-i\theta'})\\ &+c(e^{-i\theta})c(e^{i\theta'})e^{i(m-1)(\theta'-\theta)}(e^{-i\theta}- e^{i\theta'})\\ &=-4c(e^{i\theta})c(e^{-i\theta'})\sin\left(\frac{\theta+\theta'}{2}\right) \sin\left(\frac{(2m-1)(\theta-\theta')}{2}\right)\\ &+\bigl(c(e^{-i\theta})c(e^{i\theta'})-c(e^{i\theta})c(e^{-i\theta'})\bigr) e^{i(m-1)(\theta'-\theta)}(e^{-i\theta}-e^{i\theta'}). \end{split} \end{equation*} Applying the Riemann-Lebesgue lemma again and using \[\mu(e^{i\theta'})-\mu(e^{i\theta})=2a\bigl(\cos(\theta')-\cos(\theta)\bigr)= 4a\sin\left(\frac{\theta+\theta'}{2}\right) \sin\left(\frac{\theta-\theta'}{2}\right), \] we arrive at \[\lim_{m\to\infty}\langle{\mathcal{G}}_cg_1, {\mathcal{G}}_cg_2\rangle_{1-m}=\lim_{m\to\infty}\frac{1}{4\pi^2K} \int_0^\pi\int_0^\pi \frac{g_1(e^{i\theta}){\overline{g_2(e^{i\theta'})}}}{c(e^{-i\theta}) c(e^{i\theta'})}D_m(\theta,\theta')d\theta d\theta' \] provided that the limit in the right hand side of the equality exists, where $D_m$ is the Dirichlet kernel, \[D_m(\theta,\theta')= \frac{\sin\bigl((2m-1)(\theta-\theta')/2\bigr)} {\sin\bigl((\theta-\theta')/2\bigr)}.\] The result follows now from the well known $L^2$-properties of the Dirichlet kernel. \end{proof} Recall the notation $P_c$ for the orthogonal projection $E\bigl([-(1+a)^2,-(1-a)^2]\bigr)$. \begin{prop}\label{mainprop} {\bf (i)} ${\mathcal{F}}_c|_{\mathcal{R}({\mathcal{G}}_c)}: \mathcal{R}({\mathcal{G}}_c)\rightarrow L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$ is a surjective isometric isomorphism. Its inverse is given by ${\mathcal{G}}_c: L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)\rightarrow \mathcal{R}({\mathcal{G}}_c)$. {\bf (ii)} We have ${\mathcal{R}}({\mathcal{G}}_c)={\mathcal{R}}(P_c)$ for $(a,b,c)\in V_z^{gen}$. \end{prop} \begin{proof} {\bf (i)} It follows from Corollary \ref{gevolg1}, Lemma \ref{reduction} and Proposition \ref{weaklimit} that \begin{equation*} \begin{split} \int_{\ensuremath{\mathbb{T}}}\bigl({\mathcal{F}}_c\bigl({\mathcal{G}}_cf\bigr)\bigr)(\gamma) {\overline{g(\gamma)}}d\nu(\gamma) =&\lim_{l\to-\infty}\lim_{k,l\to\infty} \int_{\ensuremath{\mathbb{T}}}\langle {\mathcal{G}}_cf, \phi_{\gamma}\rangle_{k;l,m} {\overline{g(\gamma)}}d\nu(\gamma)\\ =&\lim_{l\to-\infty}\lim_{k,l\to\infty}\langle {\mathcal{G}}_cf, {\mathcal{G}}_cg\rangle_{k;l,m}=\int_{\ensuremath{\mathbb{T}}}f(\gamma){\overline{g(\gamma)}} d\nu(\gamma) \end{split} \end{equation*} for all $f,g\in L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$. Hence ${\mathcal{F}}_c\circ {\mathcal{G}}_c$ is the identity on $L^2_W\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$ and ${\mathcal{R}}\bigl({\mathcal{F}}_c\bigr)=L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$. Consequently, we have ${\mathcal{G}}_c\circ {\mathcal{F}}_c|_{{\mathcal{R}}({\mathcal{G}}_c)}= \hbox{Id}_{{\mathcal{R}}({\mathcal{G}}_c)}$. {\bf (ii)} Assume that $(a,b,c)\in V_z^{gen}$. Let $f\in {\mathcal{R}}({\mathcal{G}}_c)$. Observe that $\|f\|=\|{\mathcal{F}}_cf\|_2$ by the previous paragraph and by Proposition \ref{weaklimit}, where $\|.\|_2$ is the norm of $L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$. By Corollary \ref{gevolg1}, this implies that $f\in {\mathcal{R}}(P_c)$. Let $f\in {\mathcal{R}}(P_c)$. We have seen that ${\mathcal{F}}_c\bigl({\mathcal{R}}({\mathcal{G}}_c)\bigr)= L_W^2\bigl(\ensuremath{\mathbb{T}},d\nu\bigr)$, hence there exists a function $g\in {\mathcal{R}}({\mathcal{G}}_c)$ such that ${\mathcal{F}}_cf= {\mathcal{F}}_cg$. Now $f-g\in {\mathcal{R}}(P_c)$ by the previous paragraph, and ${\mathcal{F}}_c|_{{\mathcal{R}}(P_c)}$ is injective by Corollary \ref{gevolg1}, hence $f=g\in {\mathcal{R}}({\mathcal{G}}_c)$, as desired. \end{proof} \section{The point spectrum}\label{pointmass} In this section we determine the resolution of the identity $E$ on $\ensuremath{\mathbb{R}}\setminus [-(1+a)^2,-(1-a)^2]$. Observe that the kernel $K_{\gamma}(x,y)$ of the Green function is meromorphic as function of $\gamma\in {\mathbb{D}}$, where ${\mathbb{D}}=\{ v\in \ensuremath{\mathbb{C}} \, | \, 0<|v|< 1 \}$ is the punctured open unit disc in the complex plane. Since $\phi_{\gamma}(x)$ and $\Phi_{\gamma}(x)$ are analytic at $\gamma\in {\mathbb{D}}$, see Theorem \ref{samenvatting}{\bf (ii)} for $\Phi_{\gamma}$, we have that the poles of $K_{\gamma}(x,y)$ coincide with the poles of $W(\gamma)^{-1}$, which in turn coincide with the zeros of the map $\gamma\mapsto c(\gamma^{-1})$ $(\gamma\in {\mathbb{D}})$, where $c(\cdot)$ is the $c$-function defined by \eqref{cB}. In particular, the poles of $K_{\gamma}(x,y)$ in $\gamma\in {\mathbb{D}}$ are independent of $x,y\in I$ and are given by the set \begin{equation}\label{Sd} S=\left\{\frac{1}{q^ke} \,\, \left|\right. \,\, e\in \{a,b,c\}, k\in \ensuremath{\mathbb{Z}}_+: \,\, \frac{1}{q^{k}e}<1 \right\} \cup \left\{-\frac{abcz}{q^{k+1}} \,\, | \,\, k\in \ensuremath{\mathbb{Z}}: \,\, \frac{abcz}{q^{k+1}}<1 \right\}. \end{equation} Observe that these poles of $K_{\gamma}(x,y)$ are simple since at most one of the parameters $a,b,c$ is $\geq 1$. Furthermore, observe that $S\subset S_{reg}$ when $(a,b,c)\in V_z^{gen}$. \begin{lem}\label{Ltwo} The function $\phi_{\tilde{\gamma}}$ \textup{(}$\tilde{\gamma}\in S$\textup{)} is an eigenfunction of $(L,{\mathcal{D}})$ with eigenvalue $\mu(\tilde{\gamma})$. In particular, $\mu(S)$ is contained in the point spectrum $\sigma_p(L)$ of $(L,{\mathcal{D}})$. \end{lem} \begin{proof} Let $\tilde{\gamma}\in S$. Then $\phi_{\tilde{\gamma}}\in V_{\mu(\tilde{\gamma})}$ by Lemma \ref{oknul}. Furthermore, if $\tilde{\gamma}\not\in S_{reg}$, then \begin{equation}\label{phiatpositive} \phi_{\tilde{\gamma}}|_{I_+}= c(\tilde{\gamma})\Phi_{\tilde{\gamma}}\in M_{\mu(\tilde{\gamma})}, \end{equation} where the first equality follows from Proposition \ref{conncoef}. Since $c(\gamma)$ and $\Phi_{\gamma}\in V_{\mu(\gamma)}^+$ are regular at $\gamma\in S\cap S_{reg}$, it follows that \eqref{phiatpositive} is valid for all $\tilde{\gamma}\in S$. Hence $\phi_{\tilde{\gamma}}\in {\mathcal{D}}$ and $L\phi_{\tilde{\gamma}}=\mu(\tilde{\gamma})\phi_{\tilde{\gamma}}$ on $I$ for all $\tilde{\gamma}\in S$, as desired. \end{proof} \begin{rem}\label{phiateverywhere} If $(a,b,c)\in V_z^{gen}$, then $S\subset S_{reg}$ and $\phi_{\tilde{\gamma}}|_{I_+}=c(\tilde{\gamma})\Phi_{\tilde{\gamma}}$ by Theorem \ref{samenvatting}{\bf (iv)}, so formula \eqref{phiatpositive} extends for these parameter values to the whole $q$-interval $I$. \end{rem} Let $\tilde{\gamma}\in S$. It follows from Lemma \ref{Ltwo} that the linear functional $f\mapsto \bigl({\mathcal{F}}f\bigr)(\tilde{\gamma})$ with $f\in {\mathcal{D}}_{fin}$ is bounded, where ${\mathcal{F}}$ is the big $q$-Jacobi function transform defined by \eqref{F}, hence it uniquely extends to a continuous linear functional on ${\mathcal{H}}$. It is given explicitly by \begin{equation} \bigl({\mathcal{F}}f\bigr)(\tilde{\gamma})= \langle f,\phi_{\tilde{\gamma}}\rangle,\qquad f\in {\mathcal{H}}. \end{equation} We define the weight $d\nu\bigl(\{\tilde{\gamma}\}\bigr)$ for $\tilde{\gamma}\in S$ and $(a,b,c)\in V_z^{gen}$ by \begin{equation}\label{discrweight} d\nu\bigl(\{\tilde{\gamma}\}\bigr)=\frac{1}{K} \underset{\gamma=\tilde{\gamma}}{\hbox{Res}} \left(\frac{-1}{\gamma c(\gamma)c(\gamma^{-1})}\right) =\frac{1}{K}\underset{\gamma=\tilde{\gamma}^{-1}}{\hbox{Res}} \left(\frac{1}{\gamma c(\gamma)c(\gamma^{-1})}\right), \end{equation} where $K$ is the positive constant defined by \eqref{Kconstant}. Observe that the poles $\tilde{\gamma}\in S$ of the meromorphic function $\gamma\mapsto\bigl(c(\gamma)c(\gamma^{-1})\bigr)^{-1}$ are simple for parameters $(a,b,c)\in V_z^{gen}$. We derive explicit expressions for the discrete weights $d\nu\bigl(\{\tilde{\gamma}\}\bigr)$ by relating the continuous part of the Plancherel measure with the weight function $\Delta(x)=\Delta(x;t_0,t_1,t_2,t_3)$ of the Askey-Wilson polynomials, \begin{equation}\label{AskeyWilsonmeasure} \Delta(x)=\frac{\bigl(x^2,1/x^2;q\bigr)_{\infty}} {\prod_{j=0}^3\bigl(t_jx,t_j/x;q\bigr)_{\infty}}. \end{equation} Observe that \begin{equation}\label{relPlanchAW} \frac{1}{Kc(\gamma)c(\gamma^{-1})\gamma}=M \frac{\Delta(\gamma;e,f,-q/efgz,-efgz)}{\bigl(g\gamma, g/\gamma;q\bigr)_{\infty}\gamma}, \end{equation} where $\{e,f,g\}$ is an arbitrary permutation of $\{a,b,c\}$ (taking multiplicity into account), and where $M=M(a,b,c;z)$ is the positive constant \begin{equation}\label{M} M=\frac{1}{K}\bigl(ab,ac;q\bigr)_{\infty}^2\theta(-bcz,-bcz) =\frac{\bigl(ab,ac;q\bigr)_{\infty}^2\theta(-abz,-acz,-bcz)} {(1-q)z\theta(-1/z)}. \end{equation} For a simple pole $eq^k$ of $\Delta(x;t_0,t_1,t_2,t_3)$, where $e\in \{t_j\}_{j=0}^3$ and $k\in\ensuremath{\mathbb{Z}}_+$, we have the explicit formula \begin{equation}\label{discAW} \begin{split} {\underset{\gamma=eq^k}{\hbox{Res}}}&\left(\frac{\Delta(x;t_0,t_1,t_2,t_3)}{x} \right)= \frac{\bigl(e^{-2};q\bigr)_{\infty}} {\bigl(q,ef,f/e,eg,g/e,eh,h/e;q\bigr)_{\infty}}\\ &\qquad\qquad\qquad\qquad.\frac{\bigl(e^2,ef,eg,eh;q\bigr)_k} {\big(q,qe/f,qe/g,qe/h;q\bigr)_k}\frac{(1-e^2q^{2k})}{(1-e^2)} \left(\frac{q}{efgh}\right)^k, \end{split} \end{equation} see \cite[(7.5.22)]{GR}, where $\{f,g,h\}$ is such that $\{e,f,g,h\}= \{t_0,t_1,t_2,t_3\}$ (taking multiplicity into account). It is well known that the right hand side of \eqref{discAW} is well defined and positive for real parameters $t_i$ such that the $t_it_j$ ($i\not=j$) are strictly less than one. Combined with \eqref{discrweight}, \eqref{relPlanchAW} and \eqref{qshifttransform}, we obtain for $e\in \{a,b,c\}$ and $k\in \ensuremath{\mathbb{Z}}_+$ such that $1/q^ke<1$, \begin{equation}\label{normposdisc} \begin{split} d\nu\bigl(\{ 1/q^ke\}\bigr)= &M\frac{\bigl(e^{-2};q\bigr)_{\infty}} {\bigl(q,ef,f/e,eg,g/e;q\bigr)_{\infty}\theta(-fgz,-e^2fgz)}\\ &.\frac{\bigl(e^2,ef,eg;q\bigr)_k} {\bigl(q,qe/f,qe/g;q\bigr)_k}\frac{(1-e^2q^{2k})}{(1-e^2)} \left(\frac{-q^{(k+1)/2}}{fg}\right)^k, \end{split} \end{equation} where $f,g$ are such that $\{e,f,g\}=\{a,b,c\}$ (taking multiplicity into account). By the positivity of the residues of the Askey-Wilson measure, it follows that the right hand side of \eqref{normposdisc} is well defined, regular and strictly positive for parameters $(a,b,c)\in V$. Similarly, we obtain for $k\in \ensuremath{\mathbb{Z}}$ such that $abcz/q^{k+1}<1$, \begin{equation}\label{normstrange} \begin{split} &d\nu\bigl(\{-abcz/q^{k+1}\}\bigr)=\\ &\quad=\frac{M}{\bigl(q,q,-q/abz,-q/acz,-q/bcz, -a^2bcz/q,-ab^2cz/q,-abc^2z/q;q\bigr)_{\infty}}\\ &\qquad.\frac{\bigl(-q/abz,-q/acz,-q/bcz;q\bigr)_{k}} {\bigl(-1/a^2bc,-1/ab^2c,-1/abc^2;q\bigr)_{k}} \left(q^{2k}-\frac{a^2b^2c^2z^2}{q^2}\right) \left(\frac{-q^{(k+3)/2}}{a^2b^2c^2z}\right)^k. \end{split} \end{equation} The right hand side of \eqref{normstrange} is well defined, regular and strictly positive for parameters $(a,b,c)\in V$. We will sometimes abuse notation by writing $\nu\bigl(\{\tilde{\gamma}\}\bigr)$ for the discrete weight $d\nu\bigl(\{\gamma\}\bigr)$ ($\tilde{\gamma}\in S$). \begin{prop}\label{respositief} Let $(a,b,c)\in V_z^{gen}$. {\bf (i)} For $\mu_1<\mu_2<-(1+a)^2$ or $-(1-a)^2<\mu_1<\mu_2$ such that $\mu(S)\cap \bigl(\mu_1,\mu_2\bigr)=\emptyset$, we have $E\bigl((\mu_1,\mu_2)\bigr)=0$. {\bf (ii)} For $\tilde{\gamma}\in S$ and $f,g\in {\mathcal{H}}$ we have \[\langle E\bigl(\{\mu(\tilde{\gamma})\}\bigr)f,g\rangle =\bigl({\mathcal{F}}f\bigr)(\tilde{\gamma}) {\overline{\bigl({\mathcal{F}}g\bigr)(\tilde{\gamma})}} d\nu\bigl(\{\tilde{\gamma}\}\bigr). \] \end{prop} \begin{proof} Throughout the proof, we use the notations introduced in the proof of Proposition \ref{contE}. {\bf (i)} Let $f,g\in {\mathcal{D}}_{fin}$ and fix $\mu_1<\mu_2$ satisfying the properties as stated in {\bf (i)}. It suffices to prove that $\langle E\bigl((\mu_1,\mu_2)\bigr)f,g \rangle=0$. Observe that \eqref{crucfor} is still valid in the present setting, but that the analogue of \eqref{gammalim} is now given by \begin{equation}\label{gammalimdis} \lim_{\epsilon \downarrow 0}\gamma[\mu\pm i\epsilon]=\gamma, \end{equation} where $\mu_1<\mu<\mu_2$ and where $\gamma\in {\mathbb{D}}$ is the unique element satisfying $\mu=\mu(\gamma)$. By the condition $\mu(S)\cap \bigl(\mu_1,\mu_2\bigr)=\emptyset$, we have for $\gamma\in {\mathbb{D}}$ satisfying $\mu_1<\mu(\gamma)<\mu_2$ that \begin{equation}\label{ha} \begin{split} \lim_{\epsilon\downarrow 0}\langle \left(L-\bigl(\mu(\gamma)\pm i\epsilon\bigr)\right)^{-1}f,g\rangle= \underset{\stackrel{(x,y)\in I\times I}{x\leq y}}{\iint}& \frac{\phi_{\gamma}(x)\Phi_{\gamma}(y)}{W(\gamma)} \bigl(1-\frac{1}{2}\delta_{x,y}\bigr)\\ &.\bigl(f(x){\overline{g(y)}}+f(y){\overline{g(x)}} \bigr)\frac{d_qx}{p(x)} \frac{d_qy}{p(y)}. \end{split} \end{equation} By the bounded convergence theorem we may interchange the limit $\epsilon\downarrow 0$ and the integration in \eqref{inversion}. Combined with \eqref{ha}, this gives $\langle E\bigl((\mu_1,\mu_2)\bigr)f,g\rangle=0$, as desired. {\bf (ii)} Let $\tilde{\gamma}\in S$ and $f,g\in {\mathcal{D}}_{fin}$. Choose arbitrary $\mu_1<\mu_2$ such that $\mu(S)\cap \bigl(\mu_1,\mu_2\bigr)=\mu(\tilde{\gamma})$ and such that $[\mu_1,\mu_2]\cap [-(1+a)^2,-(1-a)^2]=\emptyset$. Then by the first part of the proposition, we have $E\big(\{\mu(\tilde{\gamma})\}\bigr)= E\bigl((\mu_1,\mu_2)\bigr)$. We compute now $\langle E\bigl((\mu_1,\mu_2)\bigr)f,g\rangle$ using \eqref{inversion}. We substitute \eqref{crucfor} in \eqref{inversion} and change the integration parameter $\mu$ in \eqref{inversion} to the $\gamma$-parameter using \eqref{eigenfunction}. Observe that for $\mu_1<\mu<\mu_2$ and $\epsilon>0$, we have that $\gamma[\mu+i\epsilon]$ (respectively $\gamma[\mu-i\epsilon]$) lies in the lower (respectively upper) half plane of $\ensuremath{\mathbb{C}}$. Since $W(\gamma)^{-1}$ has a simple pole in $\tilde{\gamma}$, it follows by Cauchy's Theorem that \begin{equation*} \begin{split} \langle E\bigl((\mu_1,\mu_2)\bigr)f,g\rangle =&\underset{\stackrel{(x,y)\in I\times I}{x\leq y}}{\iint} \bigl(f(x){\overline{g(y)}}+f(y){\overline{g(x)}}\bigr) \bigl(1-\frac{1}{2}\delta_{x,y}\bigr)\\ &.\phi_{\tilde{\gamma}}(x)\Phi_{\tilde{\gamma}}(y) \left(\frac{a}{\tilde{\gamma}}\bigl(\tilde{\gamma}^{-1}-\tilde{\gamma}\bigr) \underset{\gamma=\tilde{\gamma}}{\hbox{Res}}\bigl(W(\gamma)^{-1}\bigr) \right)\frac{d_qx}{p(x)}\frac{d_qy}{p(y)}, \end{split} \end{equation*} where the factor $\frac{a}{\tilde{\gamma}}(\tilde{\gamma}^{-1}-\tilde{\gamma})$ arises from changing the integration variable in \eqref{inversion} to $\gamma$ using the map $\gamma\mapsto \mu(\gamma)$, and from the fact that one has to change sign in order to get a positive oriented curve around $\tilde{\gamma}$. The proof is now completed using the explicit expression of $W(\gamma)$, using Remark \ref{phiateverywhere} and by symmetrizing the double $q$-Jackson integral. \end{proof} \begin{cor} Let $(a,b,c)\in V_z^{gen}$. The spectrum $\sigma(L)$ of the self-adjoint operator $\bigl(L,{\mathcal{D}}\bigr)$ is given by \[ \sigma(L)=[-(1+a)^2,-(1-a)^2]\cup\mu(S),\] where $\sigma_c(L)=[-(1+a)^2,-(1-a)^2]$ is the continuous spectrum and $\sigma_p(L)=\mu(S)$ is the point spectrum. \end{cor} \begin{proof} It follows from \cite[Theorem 13.27]{R}, Proposition \ref{contE}, Proposition \ref{mainprop}, Lemma \ref{Ltwo} and Proposition \ref{respositief} that $\sigma(L)= [-(1+a)^2,-(1-a)^2]\cup\mu(S)$. Observe now that $[-(1+a)^2,-(1-a)^2]\subset \sigma_c(L)$ by Corollary \ref{contspectrum} and $\mu(S)\subset \sigma_p(L)$ by Lemma \ref{Ltwo}. It follows that $\sigma_c(L)=[-(1+a)^2,-(1-a)^2]$ and $\sigma_p(L)=\mu(S)$, since $\sigma(L)$ is the disjoint union of $\sigma_c(L)$ and $\sigma_p(L)$, see \cite[Theorem 13.27]{R}. \end{proof} \begin{cor}\label{posPlanch} The functions $\phi_{\tilde{\gamma}} \in {\mathcal{D}}\subset {\mathcal{H}}$ \textup{(}$\tilde{\gamma}\in S$\textup{)} are mutually orthogonal in ${\mathcal{H}}$. Their quadratic norms are given by $\|\phi_{\tilde{\gamma}}\|^2=\nu\bigl(\{\tilde{\gamma}\}\bigr)^{-1}$. \end{cor} \begin{proof} Orthogonality is clear since the functions $\phi_{\tilde{\gamma}}$ ($\tilde{\gamma}\in S$) are eigenfunctions of the self-adjoint operator $(L,{\mathcal{D}})$ with mutually different eigenvalues $\mu(\tilde{\gamma})$ ($\tilde{\gamma}\in S$), see Lemma \ref{Ltwo}. It remains to derive the explicit expression for the quadratic norm $\|\phi_{\tilde{\gamma}}\|^2$. We first assume that $(a,b,c)\in V_z^{gen}$. Observe that $E\bigl(\{\mu(\tilde{\gamma})\}\bigr)\phi_{\tilde{\gamma}}= \nu\bigl(\{\tilde{\gamma}\}\bigr)\|\phi_{\tilde{\gamma}}\|^2 \phi_{\tilde{\gamma}}$ for $\tilde{\gamma}\in S$ by Proposition \ref{respositief}{\bf (ii)}. Since $\nu\bigl(\{\tilde{\gamma}\}\bigr)\|\phi_{\tilde{\gamma}}\|^2\not=0$ and $E\bigl(\{\mu(\tilde{\gamma})\}\bigr)$ is a projection, it follows that $\|\phi_{\tilde{\gamma}}\|^2=\nu\bigl(\{\tilde{\gamma}\}\bigr)^{-1}$ for $\tilde{\gamma}\in S$, as desired. The result now follows for $(a,b,c)\in V$ by continuity. \end{proof} \section{The Plancherel formula, inversion formula and the dual orthogonality relations}\label{PID} The explicit knowledge of the resolution of the identity $E$ of the self-adjoint operator $(L,{\mathcal{D}})$ on ${\mathcal{H}}$ leads directly to the Plancherel formula and the inversion formula for the big $q$-Jacobi function transform, which we formulate in this section explicitly. We show that the dual orthogonality relations imply orthogonality relations for the continuous dual $q^{-1}$-Hahn polynomials with respect to a one-parameter family of non-extremal weight functions. Furthermore, the dual orthogonality relations give explicit sets of functions which complete the continuous dual $q^{-1}$-Hahn polynomials to orthogonal bases of the corresponding $L^2$-space. We define the measure $d\nu(\cdot)=d\nu(\cdot;a,b,c;z)$ on $\ensuremath{\mathbb{C}}^*$ by \begin{equation} \int_{\ensuremath{\mathbb{C}}^*}f(\gamma)d\nu(\gamma)=\int_{\ensuremath{\mathbb{T}}}f(\gamma)d\nu(\gamma) +\sum_{\tilde{\gamma}\in S}f(\tilde{\gamma})d\nu(\{\tilde{\gamma}\}), \end{equation} where the measure $d\nu(\gamma)$ on $\ensuremath{\mathbb{T}}$ is defined by \eqref{mu}, $S\subset {\mathbb{D}}$ is the discrete set defined by \eqref{Sd} and the point mass $d\nu(\{\tilde{\gamma}\})$ for $\tilde{\gamma}\in S$ is defined by \eqref{normposdisc} for $\tilde{\gamma}>0$ and \eqref{normstrange} for $\tilde{\gamma}<0$. The measure $d\nu(\cdot)$ on $\ensuremath{\mathbb{C}}^*$ is well defined for $(a,b,c)\in V$. Indeed, for the absolutely continuous part of the measure the conditions on the parameters are such that the possible zeros of the denominator of the corresponding weight function $1/|c(\cdot)|^2$ are compensated by zeros of the numerator. For the discrete part of the measure it follows from the explicit expressions \eqref{normposdisc} and \eqref{normstrange} that $d\nu(\{\tilde{\gamma}\})$ is well defined and strictly positive for all $\tilde{\gamma}\in S$. Let $L^2_{W}\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)$ be the Hilbert space of $W$-invariant $L^2$-functions with respect to the measure $d\nu$. We define the big $q$-Jacobi function transform for functions $f\in {\mathcal{H}}$ such that $f(zq^{-k})=0$ for $k\gg 0$ by \begin{equation}\label{FourierB} g(\gamma):=\bigl({\mathcal{F}}f\bigr)(\gamma)=\langle f,\phi_{\gamma}\rangle= \int_{-1}^{\infty(z)}f(x){\overline{\phi_{\gamma}(x)}}\frac{d_qx}{p(x)}, \qquad \gamma\in\ensuremath{\mathbb{C}}^*, \end{equation} and we define for functions $g\in L^2_{W}\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)$ satisfying $g(-q^{k}abcz)=0$ for $k\gg 0$, \begin{equation}\label{FourierinversionB} f(x):=\bigl({\mathcal{G}}g\bigr)(x)= \int_{\ensuremath{\mathbb{C}}^*}g(\gamma)\phi_{\gamma}(x)d\nu(\gamma),\qquad x\in I. \end{equation} The results in section \ref{contmass} and section \ref{pointmass} lead to the following main theorem of this paper. \begin{thm}[The big $q$-Jacobi function transform]\label{main} Let $z>0$ and $(a,b,c)\in V$. The maps ${\mathcal{F}}$ and ${\mathcal{G}}$ uniquely extend to surjective isometric isomorphisms \begin{equation*} {\mathcal{F}}: {\mathcal{H}}\to L^2_{W}\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr),\qquad {\mathcal{G}}: L^2_{W}\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)\to {\mathcal{H}}. \end{equation*} Furthermore, ${\mathcal{G}}={\mathcal{F}}^{-1}$, hence \eqref{FourierB} and \eqref{FourierinversionB} give the big $q$-Jacobi function transform pair \textup{(}interpreted in the suitable $L^2$-sense\textup{)}. \end{thm} \begin{proof} We write $\bigl( .,.\bigr)$ for the inner product of $L^2_{W}\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)$. Suppose that $(a,b,c)\in V_z^{gen}$ and let $f,g\in {\mathcal{D}}_{fin}$. Applying Proposition \ref{contE} and Proposition \ref{respositief}, we obtain $\langle f,g\rangle=\langle E(\ensuremath{\mathbb{R}})f,g\rangle= \bigl( {\mathcal{F}}f, {\mathcal{F}}g\bigr)$. In order to extend this result to parameters $(a,b,c)$ in $V$, we show that $\langle f,g\rangle$ and $\bigl({\mathcal{F}}f,{\mathcal{F}}g\bigr)$ depend continuously on $(a,b,c)\in V$. This is clear for $\langle f,g\rangle$, while for $\bigl({\mathcal{F}}f,{\mathcal{F}}g\bigr)$ it is clear except for the term \[ \sum_{\tilde{\gamma}\in S: \tilde{\gamma}<0} \bigl({\mathcal{F}}f\bigr)(\tilde{\gamma}){\overline{\bigl( {\mathcal{F}}g\bigr)(\tilde{\gamma})}}d\nu(\tilde{\gamma}) = \sum_{k\in\ensuremath{\mathbb{Z}}}\bigl({\mathcal{F}}f\bigr)(-q^kabcz){\overline{\bigl( {\mathcal{F}}g\bigr)(-q^kabcz)}}d\nu(-q^kabcz), \] where we use the convention that $d\nu(-q^kabcz)=0$ if $-q^kabcz\not\in S$ for the right hand side. The continuity of this term follows by Lebesgue's dominated convergence theorem, using the asymptotics \begin{equation}\label{asingamma} \phi_{-q^kabcz}(x)={\mathcal{O}}(x^{-k}),\quad d\nu\bigl(-q^kabcz\bigr)={\mathcal{O}}(z^kq^{k(k-1)/2}),\qquad k\to\infty, \end{equation} where $x\in I$, which hold uniformly for $(a,b,c)$ in compacta of $V$. To compute the asymptotics \eqref{asingamma} for $\phi_{\gamma}(x)$ with $x\in I_+$ we used \eqref{phiatpositive}, and for $x\in I_-$ we used the formula \begin{equation}\label{pf} \phi_{\gamma}(x)= \frac{\bigl(q/ab,q/ac,q\gamma^2;q\bigr)_{\infty}} {\bigl(q\gamma/a,q\gamma/b,q\gamma/c;q\bigr)_{\infty}} \Phi_{\gamma}^{-}(x), \end{equation} see Lemma \ref{collaps} (observe that the right hand side of \eqref{pf} is well defined for $\gamma<0$ with $|\gamma|<a^{-1}$ and can be uniquely extended by continuity to $(a,b,c)\in V$). It follows that $\langle f,g\rangle=\bigl({\mathcal{F}}f,{\mathcal{F}}g\bigr)$ for $f,g\in {\mathcal{D}}_{fin}$ and $(a,b,c)\in V$, hence ${\mathcal{F}}$ uniquely extends to an isometric isomorphism ${\mathcal{F}}: {\mathcal{H}}\to L^2_{W}\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)$ onto its image for all $(a,b,c)\in V$. Let $f$ and $g$ be $W$-invariant continuous functions on $\ensuremath{\mathbb{C}}^*$ with compact support. We have $\bigl({\mathcal{G}}f\bigr)(x)=\bigl({\mathcal{G}}_cf\bigr)(x)+ \bigl({\mathcal{G}}_df\bigr)(x)$ for $x\in I$ with ${\mathcal{G}}_c$ given by \eqref{Finvers} and \[ \bigl({\mathcal{G}}_df\bigr)(x)=\sum_{\tilde{\gamma}\in S} f(\tilde{\gamma})\phi_{\tilde{\gamma}}(x)d\nu\bigl(\{\tilde{\gamma}\}\bigr), \] which is a finite sum by the assumptions on $f$. Assume that $(a,b,c)\in V_z^{gen}$, then it follows from Proposition \ref{mainprop}{\bf (ii)} that ${\mathcal{G}}_cf\in {\mathcal{R}}(P_c)$. Furthermore, it follows from the proof of Corollary \ref{posPlanch} that ${\mathcal{G}}_df\in {\mathcal{R}}(P_d)=\bigl({\mathcal{R}}(P_c)\bigr)^{\bot}$, where $P_d=E\bigl(\mu(S)\bigr)$. Hence ${\mathcal{G}}f\in {\mathcal{H}}$ and $\langle {\mathcal{G}}f, {\mathcal{G}}g\rangle= \langle {\mathcal{G}}_cf, {\mathcal{G}}_cg\rangle +\langle {\mathcal{G}}_df,{\mathcal{G}}_dg\rangle$. It follows from Proposition \ref{mainprop} and Corollary \ref{posPlanch} that \begin{equation}\label{partPlanch} \langle {\mathcal{G}}_cf, {\mathcal{G}}_cg\rangle= \int_{\gamma\in \ensuremath{\mathbb{T}}}f(\gamma){\overline{g(\gamma)}}d\nu(\gamma), \qquad \langle {\mathcal{G}}_df, {\mathcal{G}}_dg\rangle= \sum_{\tilde{\gamma}\in S}f(\tilde{\gamma}){\overline{g(\tilde{\gamma})}} d\nu\bigl(\{\tilde{\gamma}\}\bigr). \end{equation} This implies that $\langle {\mathcal{G}}f, {\mathcal{G}}g\rangle=\bigl(f,g\bigr)$. It follows from Proposition \ref{mainprop}{\bf (i)} and Corollary \ref{posPlanch} that \eqref{partPlanch} is valid for all $(a,b,c)\in V$. Furthermore, by continuity arguments, we have $\langle {\mathcal{G}}_cf,{\mathcal{G}}_dg\rangle=0$ for all $(a,b,c)\in V$. Hence, ${\mathcal{G}}$ uniquely extends to an isometric isomorphism ${\mathcal{G}}: L_{W}^2\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)\to {\mathcal{H}}$ onto its image for all $(a,b,c)\in V$. A direct computation now shows that \[\bigl( {\mathcal{F}}f,g\bigr)=\langle f,{\mathcal{G}}g\rangle, \qquad \forall f\in {\mathcal{H}}, \,\, \forall g\in L_{W}^2\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr) \] for $(a,b,c)\in V$, cf. the proof of Proposition \ref{mainprop}. This implies that ${\mathcal{R}}\bigl({\mathcal{F}}\bigr)=L_{W}^2\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)$, ${\mathcal{R}}\bigl({\mathcal{G}}\bigr)={\mathcal{H}}$ and ${\mathcal{G}}={\mathcal{F}}^{-1}$, which completes the proof of the theorem. \end{proof} In the remainder of this section we derive a one-parameter family of non-extremal orthogonality measures for the continuous dual $q^{-1}$-Hahn polynomials, and explicit sets of functions which complement the polynomials to orthogonal bases of the corresponding Hilbert spaces. We write $t_0=1/a$, $t_1=1/b$, $t_2=1/c$. The condition $(a,b,c)\in V$ is then equivalent to the conditions \begin{equation}\label{con5} t_i>0,\qquad t_it_j>1 \,\, (i\not=j) \end{equation} on the parameters $(t_0,t_1,t_2)$. We define polynomials $p_k(\gamma)=p_k(\gamma;t_0,t_1,t_2;q^{-1})$ ($k\in\ensuremath{\mathbb{Z}}_+$) in $\gamma+\gamma^{-1}$ by \begin{equation}\label{contdual} p_k(\gamma)=\phi_{\gamma}(-q^k;t_0^{-1},t_1^{-1},t_2^{-1})= {}_3\phi_2\left( \begin{array}{c} q^k, t_0\gamma, t_0/\gamma \\ t_0t_1, t_0t_2 \end{array} ; q^{-1},q^{-1} \right),\qquad k\in\ensuremath{\mathbb{Z}}_+. \end{equation} Here we use \cite[Exercise 1.4(i)]{GR} to obtain the second equality. Observe that the second equality of \eqref{contdual} shows that $\{p_k(\cdot;t_0,t_1,t_2;q^{-1}) \}_{k\in\ensuremath{\mathbb{Z}}_+}$ are exactly the continuous dual $q^{-1}$-Hahn polynomials, i.e. Askey-Wilson polynomials in base $q^{-1}$ with one of the four parameters equal to zero. For $z>0$, we define a measure $d\sigma_z(.)=d\sigma_z(.;t_0,t_1,t_2;q^{-1})$ on $\ensuremath{\mathbb{C}}^*$ by \begin{equation} \int_{\ensuremath{\mathbb{C}}^*}f(\gamma)d\sigma_z(\gamma)= \frac{1}{M}\int_{\ensuremath{\mathbb{C}}^*}f(\gamma)d\nu(\gamma;t_0^{-1},t_1^{-1},t_2^{-1};z). \end{equation} Explicitly, we have \begin{equation} \int_{\ensuremath{\mathbb{C}}^*}f(\gamma)d\sigma_z(\gamma)= \frac{1}{4\pi i}\int_{\ensuremath{\mathbb{T}}}f(\gamma)w_z(\gamma)\frac{d\gamma}{\gamma} +\sum_{\tilde{\gamma}\in S_z}f(\tilde{\gamma}) \underset{\gamma=\tilde{\gamma}}{\hbox{Res}}\left(\frac{w_z(\gamma)}{\gamma} \right), \end{equation} where \begin{equation*} \begin{split} S_z=S^{-1}=&\left\{\frac{q^k}{e} \, | \, e\in \{t_0,t_1,t_2\}, k\in \ensuremath{\mathbb{Z}}_+: \frac{q^k}{e}>1\right\}\\ &\qquad\quad\cup \left\{ \frac{-q^{k}t_0t_1t_2}{z} \, | \, k\in \ensuremath{\mathbb{Z}}: \frac{q^{k}t_0t_1t_2}{z}>1 \right\} \end{split} \end{equation*} and with weight function $w_z(.)=w_z(.;t_0,t_1,t_2;q^{-1})$ given by \begin{equation} w_z(\gamma)=\frac{\bigl(\gamma^2,1/\gamma^{2};q\bigr)_{\infty}} {\theta\bigl(-z\gamma/t_0t_1t_2, -z/t_0t_1t_2\gamma\bigr)\prod_{j=0}^2 \bigl(\gamma/t_j, 1/t_j\gamma;q\bigr)_{\infty}} \end{equation} for parameters $(t_0,t_1,t_2)$ such that the poles of $w_z(\gamma)$ at $\gamma\in S_z$ are simple. Finally, we define $W$-invariant functions $r_k^z(\gamma)=r_k^z(\gamma;t_0,t_1,t_2;q^{-1})$ for $k\in\ensuremath{\mathbb{Z}}$ by \begin{equation}\label{contdualextra} \begin{split} r_k^z(\gamma)&=\phi_{\gamma}(zq^k;t_0^{-1},t_1^{-1},t_2^{-1})\\ &=\frac{\bigl(\gamma/t_0, 1/t_1t_2, -q^kz/t_0t_1t_2\gamma;q\bigr)_{\infty}} {\bigl(1/t_0t_1, 1/t_0t_2, -q^kz/t_1t_2;q\bigr)_{\infty}} {}_3\phi_2\left( \begin{array}{c} 1/t_1\gamma, 1/t_2\gamma, -q^kz/t_1t_2 \\ 1/t_1t_2, -q^kz/t_0t_1t_2\gamma \end{array} ; q,\gamma/t_0 \right) \end{split} \end{equation} where the second equality holds for $\gamma\in \ensuremath{\mathbb{C}}^*$ with $|\gamma/t_0|<1$, see \eqref{ancontB}. \begin{thm}\label{Hahntheorem} Let $z>0$ and fix parameters $t_i$ satisfying the conditions \eqref{con5}. Then, $\{p_k\}_{k\in\ensuremath{\mathbb{Z}}_+}\cup \{r_k^z\}_{k\in\ensuremath{\mathbb{Z}}}$ is an orthogonal basis of the Hilbert space $L_W^2\bigl(\ensuremath{\mathbb{C}}^*,d\sigma_z\bigr)$. The quadratic norms of the basis elements are given by \begin{equation*} \begin{split} \int_{\ensuremath{\mathbb{C}}^*}|p_k(\gamma)|^2d\sigma_z(\gamma)&=\frac{\theta\bigl(-z\bigr)} {\theta\bigl(-z/t_0t_1,-z/t_0t_2, -z/t_1t_2\bigr)}\\ &\qquad. \frac{1}{\bigl(q,1/t_0t_1,1/t_0t_2,1/t_1t_2;q\bigr)_{\infty}} \frac{\bigl(q,1/t_1t_2;q\bigr)_k}{\bigl(1/t_0t_1,1/t_0t_2;q\bigr)_k}q^{-k},\\ \int_{\ensuremath{\mathbb{C}}^*}|r_k^z(\gamma)|^2d\sigma_z(\gamma)&= \frac{\theta\bigl(-1/z\bigr)} {\theta\bigl(-z/t_0t_1,-z/t_0t_2,-z/t_1t_2\bigr)}\\ &\qquad. \frac{1}{\bigl(1/t_0t_1,1/t_0t_2;q\bigr)_{\infty}^2} \frac{\bigl(-q^kz/t_0t_1,-q^kz/t_0t_2;q\bigr)_{\infty}} {\bigl(-q^kz/t_1t_2,-q^{1+k}z;q\bigr)_{\infty}}q^{-k}. \end{split} \end{equation*} \end{thm} \begin{proof} It follows from Theorem \ref{main} that \[ \{ \gamma\mapsto \phi_{\gamma}(-q^k) \, | \, k\in\ensuremath{\mathbb{Z}}_+\}\cup \{\gamma\mapsto \phi_{\gamma}(zq^k)\, | \, k\in\ensuremath{\mathbb{Z}}\} \] is an orthogonal basis of $L_W^2\bigl(\ensuremath{\mathbb{C}}^*,d\nu\bigr)$, and that their quadratic norms are given by \begin{equation*} \begin{split} \int_{\ensuremath{\mathbb{C}}^*}|\phi_{\gamma}(-q^k)|^2d\nu(\gamma)&=\frac{p(-q^k)}{(1-q)q^k}, \qquad k\in\ensuremath{\mathbb{Z}}_+,\\ \int_{\ensuremath{\mathbb{C}}^*}|\phi_{\gamma}(zq^k)|^2d\nu(\gamma)&=\frac{p(zq^k)}{(1-q)zq^k}, \qquad k\in\ensuremath{\mathbb{Z}}. \end{split} \end{equation*} The theorem follows now immediately by setting $t_0=1/a$, $t_1=1/b$ and $t_2=1/c$ and using the explicit expressions \eqref{pr} and \eqref{M} of the function $p(\cdot)$ and the constant $M$. \end{proof} \begin{rem} We remarked in section \ref{section1} that the big $q$-Jacobi function transform is associated with harmonic analysis on the $SU(1,1)$ quantum group. Analogous considerations for the quantum group of plane motions lead to the so-called big $q$-Hankel transform, see \cite{Koelink2} and \cite{CKK}. The corresponding function theoretic aspects of the big $q$-Hankel transform are discussed in detail in \cite{CKK}. The dual orthogonality relations for the big $q$-Hankel transform have a similar interpretation as the dual orthogonality relations for the big $q$-Jacobi function transform, namely, they give orthogonality relations for Moak's $q$-Laguerre polynomials with respect to a one-parameter family of non-extremal orthogonality measures, as well as explicit sets of functions which complement the $q$-Laguerre polynomials to orthogonal bases of the associated Hilbert spaces, see \cite[Theorem 4.1]{CKK}. \end{rem} \section{Big $q$-Jacobi polynomials: functional analytic approach} \label{polsection} We show in this section how the orthogonality relations and the quadratic norm evaluations for the big $q$-Jacobi polynomials can be derived from a functional analytic approach. The arguments are closely related to the ones used for the big $q$-Jacobi function transform, so we merely sketch the steps and indicate the main differences. To avoid confusion with previous notations, we label definitions in this section with a subscript (or superscript) $\wp$ (indicating that it is connected with the polynomial case). The conditions on the parameters $a,b,c$ are now taken to be \begin{equation}\label{condparB} ab,qa/b,ac,qa/c<1,\qquad bc<0. \end{equation} We consider the second order $q$-difference equation $L$ acting on the space ${\mathcal{F}}(I_{\wp})$ of complex-valued functions $f:I_p\to\ensuremath{\mathbb{C}}$, where $I_{\wp}=[-1,-q/bc]_q=\{-q^k\}_{k\in\ensuremath{\mathbb{Z}}_+}\cup \{-q^{k+1}/bc\}_{k\in\ensuremath{\mathbb{Z}}_+}$. At the end-points $x=-1$ and $x=-q/bc$ this should be read as $\bigl(Lf\bigr)(x)=A(x)(f(qx)-f(x))$. We can write this in self-adjoint form, similarly as was done in Lemma \ref{self-adjointform} for the non-compact case, with the same functions $p$ and $r$, see \eqref{pr}. Observe that $p(x)>0$ for all $x\in I_{\wp}$ by the conditions \eqref{condparB} on the parameters. We define ${\mathcal{H}}_{\wp}=\{f:I_{\wp}\to\ensuremath{\mathbb{C}} \,\, | \,\, \|f\|_{\wp}^2=\langle f,f\rangle_{\wp}<\infty\}$, where \[ \langle f,g\rangle_{\wp}=\int_{-1}^{-q/bc}f(x){\overline{g(x)}} \frac{d_qx}{p(x)}. \] It is well known that the big $q$-Jacobi polynomials, which are explicitly given by \eqref{connpol}, form an orthogonal basis of ${\mathcal{H}}_{\wp}$, see \cite{AA}. This fact and the evaluation of the corresponding quadratic norms have been derived in \cite{AA} using the $q$-binomial formula \cite[(1.3.2)]{GR} and the $q$-Pfaff-Saalsch{\"u}tz formula \cite[(1.7.2)]{GR}. In this section we derive these results by functional analytic methods. Similarly as in the non-compact setting, we truncate the inner product by \[\langle f,g\rangle_{\wp,k,l}= \left(\int_{-1}^{-q^{k+1}}+\int_{-q^{l+2}/bc}^{-q/bc}\right) f(x){\overline{g(x)}}\frac{d_qx}{p(x)} \] for $k,l\in\ensuremath{\mathbb{Z}}_+$. The analogue of Lemma \ref{stoktermen} is then given by \[\langle Lf,g\rangle_{\wp,k,l}-\langle f,Lg\rangle_{\wp,k,l}= W(f,{\overline{g}})(-q^k)-W(f,{\overline{g}})(-q^{l+1}/bc), \] with the Wronskian as defined in \eqref{Wronskian}. Let $\alpha\in\ensuremath{\mathbb{T}}$, then we write ${\mathcal{D}}_{\wp,\alpha}$ for the functions $f\in {\mathcal{H}}_{\wp}$ such that $Lf\in {\mathcal{H}}_{\wp}$ and $f(0^+)=\alpha f(0^-)$, $f'(0^+)=\alpha f'(0^-)$, where now $f(0^-)=\lim_{k\to\infty}f(-q^k)$ and $f(0^+)=\lim_{k\to\infty}f(-q^k/bc)$, etc. (cf. section \ref{sdomain}). By similar arguments as in section \ref{sdomain}, we have \begin{prop} The operator $(L,{\mathcal{D}}_{\wp,\alpha})$ on ${\mathcal{H}}_{\wp}$ is self-adjoint for all $\alpha\in \ensuremath{\mathbb{T}}$. \end{prop} We denote ${\mathcal{D}}_{\wp}={\mathcal{D}}_{\wp,1}$. Now observe that $\phi_{\gamma}, \psi_{\gamma}\in {\mathcal{D}}_{\wp}$ for $\gamma\in\ensuremath{\mathbb{C}}^*$ by Lemma \ref{oknul} (see \eqref{phiB} and \eqref{psiB} for the definition of $\phi_{\gamma}$ and $\psi_{\gamma}$, respectively). Furthermore, from the arguments as given in section \ref{eigenf} it follows that $\bigl(L\phi_{\gamma}\bigr)(x)=\mu(\gamma)\phi_{\gamma}(x)$ for $x\in [-1,-q/bc)_q=I_{\wp}\setminus \{-q/bc\}$ and $\bigl(L\psi_{\gamma}\bigr)(x)= \mu(\gamma)\psi_{\gamma}(x)$ for $x\in (-1,-q/bc]_q=I_{\wp}\setminus \{-1\}$, where $\mu(\gamma)$ is given by \eqref{eigenfunction}. The Wronskian $W_{\wp}(\gamma)=W(\psi_{\gamma},\phi_{\gamma})\in {\mathcal{F}}(I_{\wp})$ can again be seen to be constant on $I_{\wp}$ (cf. Lemma \ref{generaleigenfunction}{\bf (iii)}), and an explicit expression of the Wronskian $W_{\wp}(\gamma)$ is given by \[ W_{\wp}(\gamma)=(1-q)\frac{\bigl(a\gamma,a/\gamma;q\bigr)_{\infty}\theta(bc)} {\bigl(ab,ac,qa/b,qa/c;q\bigr)_{\infty}}. \] Indeed, observe that Proposition \ref{Wpsiphi} is also valid for the present choice \eqref{condparB} of parameter values by analytic continuation. In particular, the functions $\phi_{\gamma}$ and $\psi_{\gamma}$ in ${\mathcal{D}}_{\wp}$ are linearly independent if and only if $\gamma\not\in S_{pol}$, where $S_{pol}=\{\gamma_n^{\pm 1}\}_{n\in\ensuremath{\mathbb{Z}}_+}$, $\gamma_n=aq^n$. Hence $\bigl(L\phi_{\gamma}\bigr)(-q/bc)\not=\mu(\gamma)\phi_{\gamma}(-q/bc)$ and $\bigl(L\psi_{\gamma}\bigr)(-1)\not=\mu(\gamma)\psi_{\gamma}(-1)$ if $\gamma\not\in S_{pol}$, cf. Corollary \ref{psifoutind}{\bf (iii)}. Let $\mu=\mu(\gamma)\in \ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$. We define the Green kernel $K_{\gamma}^{\wp}(x,y)$ for $x,y\in I_{\wp}$ by \begin{equation*} K_{\gamma}^{\wp}(x,y)= \begin{cases} W_{\wp}(\gamma)^{-1}\psi_{\gamma}(x)\phi_{\gamma}(y), \qquad &y\leq x,\\ W_{\wp}(\gamma)^{-1}\phi_{\gamma}(x)\psi_{\gamma}(y), \qquad &y\geq x. \end{cases} \end{equation*} We have a well defined linear map ${\mathcal{H}}_{\wp}\to {\mathcal{F}}(I_{\wp})$ which maps $f\in {\mathcal{H}}_{\wp}$ to \begin{equation} G_{f}^{\wp}(x,\gamma)=\langle f, {\overline{K_{\gamma}^{\wp}(x,\cdot)}}\rangle_{\wp}, \qquad x\in I_{\wp}. \end{equation} By similar arguments as in the proof of Proposition \ref{G}, we derive that for $f\in {\mathcal{H}}_{\wp}$ and for $\gamma\in\ensuremath{\mathbb{C}}^*$ such that $\mu(\gamma)\in \ensuremath{\mathbb{C}}\setminus\ensuremath{\mathbb{R}}$, \begin{equation}\label{inverpol} G_f^{\wp}(\cdot,\gamma)=\bigl(L-\mu(\gamma).\hbox{Id}\bigr)^{-1}f. \end{equation} For the proof of \eqref{inverpol}, we have to consider case 2 of the proof of Proposition \ref{G} twice, namely for the end-point $-1$ as well as for the end-point $-q/bc$. The arguments go through for both end-points, since $\phi_{\gamma}$ (respectively $\psi_{\gamma}$) satisfies the eigenvalue equation $(Lf)(x)=\mu(\gamma)f(x)$ in the end-point $x=-1$ (respectively $x=-q/bc$). We are now in a position to compute the resolution of the identity $E_{\wp}$ for the self-adjoint operator $\bigl(L,{\mathcal{D}}_{\wp}\bigr)$ on ${\mathcal{H}}_{\wp}$ using \eqref{inversion}. For the moment it is convenient to assume that the parameters also satisfy the condition $a^2\not\in \{q^k\}_{k\in \ensuremath{\mathbb{Z}}}$. This condition can be removed later on by continuity. We keep the notations of the proof of Proposition \ref{contE}. Choose $f,g\in {\mathcal{H}}_{\wp}$ with finite support. Then we have for $\mu\in\ensuremath{\mathbb{R}}$ and $\epsilon>0$, that \begin{equation}\label{polcrucfor} \begin{split} \langle \left(L-\bigl(\mu\pm i\epsilon\bigr)\right)^{-1}f,g\rangle_{\wp}= &{\underset{\stackrel{(x,y)\in I_{\wp}\times I_{\wp}}{x\leq y}}{\iint}} \frac{\phi_{\gamma[\mu\pm i\epsilon]}(x) \psi_{\gamma[\mu\pm i\epsilon]}(y)}{W_{\wp}\bigl(\gamma[\mu\pm i\epsilon]\bigr)}\bigl(1-\frac{1}{2}\delta_{x,y}\bigr)\\ &\qquad.\bigl(f(x){\overline{g(y)}}+f(y){\overline{g(x)}} \bigr)\frac{d_qx}{p(x)}\frac{d_qy}{p(y)}. \end{split} \end{equation} Let $\xi>0$ such that $\mu\bigl(S_{pol}\bigr)\cap [-(1+a)^2-\xi, -(1-a)^2+\xi]=\emptyset$. Using \eqref{gammalim}, \eqref{gammalimdis} and the invariance of $\phi_{\gamma}$, $\psi_{\gamma}$ and $W_{\wp}(\gamma)$ under $\gamma\leftrightarrow\gamma^{-1}$, we obtain from \eqref{polcrucfor} that for all $-(1+a)^2-\xi<\mu<-(1-a)^2+\xi$, \[ \lim_{\epsilon\downarrow 0}\left( \langle \bigl(L-(\mu+i\epsilon)\bigr)^{-1}f,g\rangle_{\wp} -\langle \bigl(L-(\mu-i\epsilon)\bigr)^{-1}f,g\rangle_{\wp}\right)=0. \] It follows now from \eqref{inversion} that $E_{\wp}\bigl([-(1+a)^2,-(1-a)^2]\bigr)=0$. For $-\infty<\mu_1<\mu_2<-(1+a)^2$ or $-(1-a)^2<\mu_1<\mu_2<\infty$ such that $(\mu_1,\mu_2)\cap \mu(S_{pol})=\emptyset$, we have $E_{\wp}\bigl((\mu_1,\mu_2)\bigr)=0$, cf. the proof of Proposition \ref{respositief}{\bf (i)}. Setting $\mu_n=\mu(\gamma_n)$ for $n\in \ensuremath{\mathbb{Z}}_+$, we have, due to the simple pole of $W_{\wp}(\gamma)^{-1}$ in $\gamma=\gamma_n$, \begin{equation*} \begin{split} \langle E_{\wp}\bigl(\{\mu_n\}\bigr)f,g\rangle_{\wp} =&\underset{\stackrel{(x,y)\in I_{\wp}\times I_{\wp}}{x\leq y}}{\iint} \bigl(f(x){\overline{g(y)}}+f(y){\overline{g(x)}}\bigr) \bigl(1-\frac{1}{2}\delta_{x,y}\bigr)\\ &.\phi_{\gamma_n}(x)\psi_{\gamma_n}(y) \left(\frac{a}{\gamma_n}\bigl(\gamma_n^{-1}-\gamma_n\bigr) \underset{\gamma=\gamma_n}{\hbox{Res}}\bigl(W_{\wp}(\gamma)^{-1}\bigr) \right)\frac{d_qx}{p(x)}\frac{d_qy}{p(y)} \end{split} \end{equation*} for $f,g\in {\mathcal{H}}_{\wp}$ with finite support, cf. the proof of Proposition \ref{respositief}{\bf (ii)}. Using \eqref{connpol} to rewrite $\psi_{\gamma_n}$ as a multiple of $\phi_{\gamma_n}$, and by symmetrizing the double $q$-Jackson integral, we obtain \begin{equation}\label{almostthere} \langle E_{\wp}(\{\mu_n\})f,g\rangle_{\wp}={\mathcal{N}}_{\wp}(n)^{-1} \langle f,\psi_{\gamma_n}\rangle_{\wp}\langle \psi_{\gamma_n},g\rangle_{\wp}, \end{equation} where \begin{equation*} \begin{split} {\mathcal{N}}_{\wp}(n)=&\left(\frac{a}{\gamma_n} \bigl(\gamma_n^{-1}-\gamma_n\bigr) \frac{\bigl(qa/b,qa/c;q\bigr)_n}{\bigl(ab,ac;q\bigr)_{n}} \left(\frac{bc}{q}\right)^n \underset{\gamma=\gamma_n}{\hbox{Res}}\bigl(W_{\wp}(\gamma)^{-1}\bigr) \right)^{-1}\\ &=\frac{(1-q)\bigl(q,bc,q/bc,q^{2n+1}a^2;q\bigr)_{\infty} \bigl(q,q^{n}a^2;q\bigr)_n} {\bigl(qa/b,qa/c,q^{n}ab,q^{n}ac;q\bigr)_{\infty} \bigl(qa/b,qa/c;q\bigr)_n}\left(-\frac{q^{(3-n)/2}}{bc}\right)^n. \end{split} \end{equation*} Observe that \eqref{almostthere} holds for all $f,g\in {\mathcal{H}}_{\wp}$ by continuity. We can now immediately recover the orthogonality relations and quadratic norm evaluations for the big $q$-Jacobi polynomials by applying well known properties of resolution of identities to $E_{\wp}$. The result is as follows. \begin{thm}[\cite{AA}] If the parameters $(a,b,c)$ satisfy the conditions \eqref{condparB}, then the polynomials $\{\psi_{\gamma_n}\}_{n\in\ensuremath{\mathbb{Z}}_+}$ form an orthogonal basis of ${\mathcal{H}}_{\wp}$, and their quadratic norms are given by \[ \|\psi_{\gamma_n}\|_{\wp}^2={\mathcal{N}}_{\wp}(n),\qquad n\in\ensuremath{\mathbb{Z}}_+. \] \end{thm} \bibliographystyle{amsplain}
\section{Introduction} The Hubble Space Telescope (HST) Archival Pure Parallel Program, started in 1997 shortly after the second Hubble servicing mission, was implemented to provide the maximum amount of science possible with HST and its primary cameras: the Wide--Field Planetary Camera 2 (WFPC2), the Near-Infrared Camera and Multi--Object Spectrograph (NICMOS), and the Space Telescope Imaging Spectrograph (STIS). During observations of a target by the primary instrument, the other instruments make a preplanned series of observations in nearby fields, the positions of which are determined by the focal plane offset of the instrument and the roll angle of HST. The resulting observations are placed in the Hubble Data Archive at the Space Telescope Science Institute, and are made immediately available to the astronomical community. The STIS parallels include broadband camera images and full field slitless spectroscopy. Full details of the STIS parallel survey (SPS) capabilities and techniques can be found in Gardner et al. (1998). The imaging mode has a very broad bandpass (FWHM $\approx$ 5000\AA) peaking at 5500\AA\ and ranging from $\approx$ 2500\AA\ to 10300\AA. The full field spectroscopy mode has a wavelength range of 5200\AA\ to 10300\AA, and a dispersion of 4.9\AA\ per pixel. In an imaging mode observation of 2000 seconds, point sources as faint as ${\rm m_V}=28$ can be detected with a S/N of 5, and the spectra can achieve a limiting magnitude of 22 under the same circumstances. Approximately 6000 STIS parallel images and spectra were taken in the first year of the parallel program, comprising more than 1000 separate fields. Of these, over 300 images and spectra are of the Large Magellanic Cloud (LMC). These data represent a unique opportunity to peer deeply into the LMC and investigate many of its physical properties. Planetary nebulae (PNe, singular PN) provide a way to study several important properties of the LMC, including kinematics (Vassiliadis, Meatheringham \& Dopita 1992), dynamical and chemical evolution and can also be used as distance indicators (for example, Feldmeier, Ciardullo \& Jacoby 1997 and references therein). Nearly 300 PNe have been found in the LMC (Leisy et al. 1997). LMC PNe are of particular importance in the determination of the PNe luminosity function, since they all lie at approximately the same distance. Unfortunately, the great distance to the LMC means that most PNe are very faint and unresolvable by ground based telescopes, and the typically crowded LMC fields make studying the PNe very difficult. The high spatial resolution of HST and low sky background of Earth orbit provide an advantage for studying LMC PNe (for example, Dopita et al. 1996). In particular, STIS can record spatially resolved spectra of the larger, fainter PNe. Here we report on the serendipitous STIS parallel observations of the planetary nebula M94--20, first discovered by Morgan (1994) in a ground--based emission line survey of the LMC. The STIS camera mode images clearly resolve the nebula to be $\approx$2 arcseconds in diameter and also detect the central star. The spectra provide identification of several emission lines in the nebula. The low resolution of the STIS spectra together with the relatively large angular extent of the nebula blends the nebular diagnostic emission lines, but the images and spectra are still a useful tool in investigating extragalactic PNe and can be used to plan followup observations, both ground and space--based. \section{Observations} Six images and two spectra were taken of the field containing the planetary nebula M94-20. The total exposure times were 1100 and 1200 seconds for the images and spectra, respectively. The observations were taken as part of the STIS Archival Pure Parallel Program, HST Program ID 7783, S. Baum, Principal Investigator. WFPC2 was the prime instrument during the STIS parallels, observing the LMC as part of the HST program ``Star Formation History of the Large Magellanic Cloud'', HST ID 7382, Smecker--Hane, Principal Investigator. The primary target was LMC-DISK1 at $\alpha = 05^h 11^m 14\farcs92$ \ and \ $ \delta =-71^{\circ} 15^m 41\farcs62$. STIS was located approximately five arcminutes north of WFPC2 during the observations. The STIS parallel images were formatted as 1024x1024 arrays (one pixel = 0\farcs 051), while the spectra were automatically binned onchip in the y--direction to 1024x512 (one pixel = 0\farcs 051 in the spectral direction, and 0\farcs 102 in the spatial direction). The observations were taken over two orbits, and there was a 0\farcs8 (15.5 STIS CCD pixels) telescope slew to the southwest between the two sets of observations. The offsets between images were determined by cross--correlating a small subsection in each, which were then used to shift and combine the images. The offsets were applied with appropriate binning to the spectra which were shifted and combined as well. The images and full field spectra were processed using STIS Investigation Definition Team pipeline calibration software which performs basic data reduction steps such as bias and dark current subtraction. The observation particulars are shown in Table 1. For the analysis using the imaging mode, the sky background was subtracted by taking a simple median of the area near the PN. The background in the full field spectral image was subtracted from the spectrum using a column--by--column median, employing sigma--clipping to remove the positive bias of the stellar contamination. \section{Discussion} \subsection{The Nebula} Figure 1 shows the full field processed image containing M94--20. The PN is located at the top of the image, centered roughly in the horizontal (x--axis) direction. North is to the left and East is down; we display the image in this way so that it has the same sense as the spectrum, which disperses light along the x--direction. Though close to the detector edge, the nebula is located fully inside the image. Note the small irregular galaxy located 9 arcseconds to the north of M94--20, and another located 22 arcseconds to the south. Although both lie in the dispersion direction of the PN, neither interferes significantly with the spectrum. The bright diagonal line in the image is a diffraction spike from a star located just off the detector, and the circular features near bright stars are internal reflections in STIS. The inset shows a close--up of just the nebula. The PN, unresolved in the discovery survey (Morgan 1994), is clearly resolved in the STIS image. The measured properties of the nebula and central star (see Section 3.2) are listed in Table 2. The nebula is slightly elliptical, with a mean diameter of roughly 2 arcseconds, making this one of the largest PNe in the LMC. Most LMC PNe are smaller than 1 arcsecond in diameter, with the notable exception of LMC--SMP72, a bipolar PN measuring approximately 2 x 3 arcseconds (Dopita et al. 1996). M94--20 has a bright elliptical rim, and the outermost parts of the nebula also show some faint structure, reminiscent of the double--elliptical structure of NGC 6543 (aka the ``Cat's Eye''). One of the outer ellipses is aligned with the inner rim, while the other has a position angle of $\sim 145^{\circ}$. The large angular extent of the PN implies that this is an evolved object. A subarray of the spectrum containing M94--20 is shown in Figure 2 ({\it top}). The extracted subarray covers the full spectral range of the original spectrum, but only 32 pixels ($3\farcs3$) in the spatial direction. The stars in the image appear as sources of continuum, while M94--20 is clearly an emission--line object. Figure 2 ({\it middle}) shows a closeup of the nebular spectrum from 6000\AA\ to 7200\AA\ as well as the positions of the line images using ellipses with major and minor axes corresponding to those measured in the camera mode image ({\it bottom}). The bright elliptical patch corresponds to the lines of H$\alpha$ and [NII] in M94--20. At this resolution, the [NII] 6548\AA\ and 6584\AA\ lines are separated from H$\alpha$ by only 3 and 4 pixels, respectively. The nebula itself is about 40 pixels or ten times that size, so the three spectral line images overlap. The blended nebular spectral image is about 4 pixels wider than the nebular image in the camera mode, also indicating that more than one line is present. The [OI] line at 6300\AA\ is also clearly seen. The [SII] 6717, 6731\AA\ emission line images are barely detected in the spectrum. We also note that we may have a detection of [SIII] 9069\AA\ and Pa$\epsilon$,[SIII] at 9545\AA, which does not reproduce well in Figure 2 but can be seen very faintly in the original data. We flux calibrated the brighter emission lines using the absolute sensitivity of STIS for a point source (Collins and Bohlin 1998), correcting the sensitivity curve for a slitless spectrum of an extended line emission object, and masking out the continuum spectra of nearby stars that overlapped the PN spectrum. We find that the total observed H$\alpha$ + [NII] 6548, 6584\AA\ flux is $\rm{7.3 x 10^{-15}\ erg/sec/cm^2}$. D. Morgan and Q. Parker (1998, private communication) made followup observations of many of the PNe found by Morgan (1994), and for M94--20 detected [OIII] 4959, 5007\AA\ and H$\alpha$. They report the [OIII] flux to be $\rm{2.7 x 10^{-14}\ erg/sec/cm^2}$. The value of the ratio of H$\alpha$ + [NII] 6548, 6584\AA\ to [OIII] 4959, 5007\AA\ in LMC PNe is typically 3--4 (for example, Vassiliadis, Dopita, Morgan and Bell 1992); which is somewhat lower than but consistent with these measurements. The typical value of the ratio of [OI] 6300\AA\ to H$\alpha$ in LMC PNe is approximately 0.07. We find the [OI] 6300\AA\ flux is $\rm{1.2 x 10^{-15}\ erg/sec/cm^2}$, which again gives a ratio higher than but consistent with the typical value. Morgan and Parker also find an upper limit to the H$\beta$ flux of $\rm{1.2 x 10^{-14}\ erg/sec/cm^2}$, making M94--20 a relatively faint LMC PN (Vassiliadis, Dopita, Morgan and Bell 1992). We note that our measurements may also be upper limits, since background subtraction is difficult due to the stellar spectra superposed on the PN spectrum. The detection of [SII] 6717, 6731\AA\ is very weak and is complicated by their proximity to the H$\alpha$ + [NII] 6548, 6584\AA\ lines. We subtracted a linear fit to the slope of the wing of the H$\alpha$ + [NII] lines to find the total flux of the [SII] lines, and get an upper limit of $\rm{5 x 10^{-16}\ erg/sec/cm^2}$. The emission lines from the PN are remarkably featureless spatially, given the obvious structure in the camera mode image. The spectral mode of STIS has a blue cutoff at $\sim$5300\AA, while the imaging mode goes down to $\sim$2000\AA. The imaging mode will therefore detect the lines of H$\beta$, the [OII] doublet at 3727 and 3729\AA, and the lines of [OIII] at 4959, 5007\AA. The emission lines seen in the imaging mode but not in the spectroscopic mode must account for these features. \subsection{The Central Star} The central star, as well as two other superposed stars, are clearly visible in the camera mode image. Relative to the Fine Guidance Sensor guide stars, we find the J2000 coordinates of the central star are $\rm{\alpha = 5^h 11^m 10\farcs 64 \ and \ \delta = -71^{\circ} 10^m 26\farcs 54}$, in good agreement with those found in Leisy et al. (1997). The spectrum of the star is too weak to determine a temperature, so we used the camera mode sensitivity of STIS to flux calibrate blackbody models of 20000K, 40000K and 100000K. The calibrated models were then folded into a Johnson V filter bandpass to find the V magnitude of the star. We found the central star magnitude to be $\rm{m_{V}} = 26.0 \pm 0.2$. The sensitivity of STIS in the imaging mode drops rapidly with decreasing wavelength, so the calculated magnitude does not depend strongly on the temperature of the star. Given the magnitude and a distance to the LMC of 55 kpc, we find that the luminosity of the central star is 0.09 solar, and ${\rm log(T_{eff}/T_{\odot}) = 4.6}$ for a typical white dwarf radius of ${\rm 5 x 10^8 cm}$. The low temperature indicates this may be a relatively old white dwarf, which is consistent with the evolved nature of the PN. \subsection{Other PNe} There are approximately 300 known PNe in the LMC (Leisy et al. 1997). The number density is highest near the Bar and tapers off rapidly with distance (Morgan 1994). M94--20 lies over a degree from the Bar, near the edge of the LMC. Over 300 observations comprising 70 separate fields have been taken in the LMC by STIS. Of these, 29 also have associated full field spectra, enhancing the ability to detect candidate PNe (i.e., emission line objects) clearly. The detection limit of an observation depends most strongly on the brightness and the size of the nebula in a given emission line and the integration time. The 1200 second exposure of the M94--20 observations is fairly typical of parallels, and noting that the [SII] 6717, 6731\AA\ emission lines in the nebula are barely detected, we expect to detect any nebula the size of M94--20 brighter than about ${\rm 2-3 x 10^{-15}\ erg/sec/cm^2}$ in a line for a typical LMC PN. A smaller nebula would be detectable at fainter limits, with the brightness scaling inversely with the area. We note here again that M94--20 is a relatively faint, large LMC PN. No other previously catalogued LMC PNe have been observed by STIS in the parallel survey. We have processed and examined the other fields in the LMC and also 20 STIS fields in the Small Magellanic Cloud (SMC) and have found no other extended PNe to within the detection limits, catalogued or otherwise, although large HII regions are common and easily detected. Unresolved or partially resolved PNe are difficult to identify in this case because most of the full--field spectra in the LMC and SMC are either single observations or have one cosmic ray split (that is, two observations without telescope movement), making the differentiation of sharp spectral features from cosmic rays and hot pixels difficult. We note here that in many other fields, STIS observations have two or more cosmic ray splits, making unresolved PN detection relatively easier. We have also searched the relevant WFPC2 parallel fields; no known LMC or SMC PNe are located within these parallel survey fields. Discovering previously unrecorded PNe using WFPC2 parallels is more difficult due to the lack of unambiguous spectral information in the broad passbands used. Although the odds of finding PNe in the random STIS parallels are low for the LMC or SMC, we are encouraged by the observations of M94--20. The ability of STIS to get deep, spatially resolved spectra makes it an excellent instrument for tagging objects for further deep ground--based spectroscopy. STIS parallels will continue to be taken in the Magellanic Clouds, and it is only a matter of time before additional PNe are observed. Interestingly, PNe in nearby galaxies can also be detected in STIS parallels, although this becomes a difficult process as PNe at that distance are no longer resolved spatially. However, it also means that the larger volumes of space on the scale of the host galaxy are observed as well. Ground--based surveys using narrow passband filters centered on [OIII] 5007\AA\ and H$\alpha$ have found many PN candidates in M31 and M33 (Ciardullo et al. 1989; Bohannan, Conti \& Massey 1985). We made a preliminary search of STIS parallel fields in M31 and NGC 205 which yielded several emission lines objects. At least one of these objects is a planetary nebula, even though only a small fraction of the surface area of those galaxies has been observed. A similar search in M33 fields has also resulted in several detections. These observations are awaiting ground--based followups to get spectra with wavelength coverage that includes ${\rm H\beta}$ and [OIII] 5007\AA\ to help distinguish PNe from HII regions. Future STIS parallels will undoubtedly find more PNe, especially if the narrow [OII] 3727\AA\ and [OIII] 5007\AA\ filters are used in imaging mode in conjunction with the G430L grating, which has a bandpass that includes these emission lines. We note finally that the use of STIS parallels goes well beyond that of finding (or adding to our knowledge of pre--existing) PNe. Moderate redshift galaxy counts (Gardner et al. 1998), research on stellar population counts, low mass stars (Plait 1999), globular clusters in nearby galaxies, and many other fields can benefit from a relatively simple search of the parallel archive. Since the data already exist and are publically accessible, the STIS parallels are a excellent tool that should be exploited to their fullest extent. The authors wish to acknowledge the help of Nick Collins, Bob Hill, Jon Gardner and David Morgan for giving their advice during this work.
\section{INTRODUCTION} Beaming of relativistic ejecta in GRBs has been postulated by many authors in order to ease the GRB energy budget (see, \eg, M\'esz\'aros, Rees, \& Wijers~1998 and refs. therein). There are basically two ways to verify the beaming observationally: one is statistical and is based on counting the afterglow like transient sources and comparing their rate with the GRB rate, and the second one is related to the beaming effects predicted to be imprinted in the afterglow light curves of individual objects (Rhoads~1997). Applying the first method to the X-ray transient sources, Grindlay~(1999) found that results are consistent with no beaming differentiating the GRB and X-transient rates. However, as was pointed out by Woods \& Loeb~(1999), the conclusive results about excess (or its lack) of X-ray transients over GRBs must wait for much more sensitive future instruments. This is because statistically significant contribution to the excess of X-ray transients over GRBs is expected to be provided only by weak X-ray transients, those representing afterglows phases when the bulk Lorentz factor of the radiating shell drops below the inverse of its angular size. Similar studies can be performed also in optical and radio band (Rhoads~1997; Woods \& Loeb~1999). In individual objects, the beaming related effects are expected to be imprinted in the optical and X-ray afterglow light-curves. The lateral expansion of the shocked, relativistic plasma causes that at some moment the front of the blast wave starts to increase faster than due to the cone-outflow (Rhoads~1997). Due to this the blast wave begins to decelerate faster than without the sideways outflows and this produces a break in the light curve, the sooner the larger the beaming factor is. Such a break is claimed to be present in the light curve of GRB~990123, the most energetic GRB up to date (Kulkarni \ea~1999). Sari, Piran, \& Halpern~(1999) speculate that afterglows with very steep light curves are highly beamed. Possibly the break in such objects is not recorded because it took place before the optical follow-ups. As now, all theoretical studies of the light-curve breaks are analytical and are based on: power-law approximation of the blast wave dynamics, broken power-law approximation of radiation spectra and on ``on-axis'' relation between the observed flux and the emitted flux (Rhoads~1977, 1999; Kulkarni \ea~1999; Sari, Piran, \& Halpern~1999). In this paper, we treat the dynamics using the prescription given by Blandford and McKee~(1976). The evolution of radiation spectrum is calculated exactly, by computing the time evolution of electrons from continuity equation and by computing the observed luminosity through integrating the emitted radiation over the ``$t= {\rm const}$'' surfaces. Our results show that the change of the light curve slope is significant, but smaller than predicted analytically. And, what is more important, the light curves steepen very slowly, so that it is very difficult to talk about the specific time location of the break. In order to demonstrate better the beaming effect, we compare our results with the spherical case. We also show how the light curve should look like if there is no lateral expansion. In \S 2 we collect equations, which are used to compute the blast wave speed, evolution of electrons, and afterglow light-curves. In \S 3 we present results of our numerical studies of afterglows produced by beamed ejecta, and, in \S 4 we compare them with simple analytical estimations. \section{BASIC EQUATIONS} \subsection{Dynamics} The deceleration of a blast wave is described by the following equations (Blandford \& McKee~1976; Chiang \& Dermer~1998): \be {d\Gamma \over dm} = - {\Gamma^2 -1 \over M} , \label{eq1} \ee \be {dM \over dr} = {dm \over dr} [\Gamma - \epsilon_{rad}\epsilon_e (\Gamma-1)] , \label{eq2} \ee and \be dm/dr = \Omega_j r^2 \rho = 2\pi r^2 (1-cos \theta_j) \rho , \label{eq3} \ee where $\Gamma$ is the bulk Lorentz factor of the blast wave, $M$ is the total mass including internal energy, $r$ is the distance from the central engine to the blast wave, $dm$ is the rest mass swept up in the distance $dr$, $\rho$ is the mass density of the external medium, $\epsilon_e$ is the fraction of dissipated energy converted to relativistic electrons, $\epsilon_{rad}$ is the fraction of electron energy which is radiated, and $\theta_j$ is the angular size of the blast wave. This angular size is not constant but increases due to thermal expansion (Rhoads~1997), \be \theta_j \equiv {a\over r} = \theta_{j0} + {v_l' \over c \Gamma}, \label{eq4} \ee where the speed of the lateral expansion, $v_l'$, is assumed by Rhoads~(1999) to be equal to the sound speed in the relativistic plasma, $c_s = c / \sqrt{3}$, but considered by Sari \ea~(1999) to be relativistic. Noting, that the plasma in the blast wave is continuously loaded by the fresh gas which initially doesn't have any lateral bulk speed, one can expect that in reality $v_l'$ does not reach relativistic value and sets up somewhere between $c_s$ and $\beta_{\Gamma}c$, and in general depends on $r$, and on $\theta_j$. \subsection{Electron energy distribution} We assume that the electrons are injected with the power low energy distribution \be Q = K \gamma^{-p} , \label{eq5} \ee with the minimum energy of injected electrons \be \gamma_m = {\epsilon_e (\Gamma-1) m_p \over m_e} \,{p-2 \over p-1} . \label{eq6} \ee The maximum energy of injected electrons for a given magnetic field, $B'$, is assumed to be given by (de~Jager~\ea~1996): \be \gamma_{max} \simeq 4 \times 10^7 {\left ( {B' \over 1{\rm G}} \right )}^{-1/2} \ee Normalization of the injection function, $K$, is provided by \be L_{e,inj}' \equiv \int_{\gamma_m}^{\gamma_{max}} Q \gamma m_e c^2 ~d\gamma = \epsilon_e {dE_{acc}' \over dt'} \, ,\label{eq7} \ee where \be {dE_{acc}' \over dt'} = {dr \over dt'} {dE_{acc}' \over dr} = {dr \over dt'} {dm \over dr} c^2 (\Gamma-1) = \Omega_j r^2 \rho \beta_{\Gamma} \Gamma (\Gamma -1) c^3 , \label{eq8} \ee is the rate of accreted kinetic energy, $dr = c \beta_{\Gamma} \Gamma dt' $, $\beta_{\Gamma} = \sqrt {\Gamma^2 -1}/\Gamma $, and $t'$ is the time measured in the blast wave comoving frame. The evolution of the electron energy distribution is given by the continuity equation \be {\partial N_{\gamma} \over \partial r} = {\partial \over \partial \gamma} \left(N_{\gamma} {d\gamma \over dr}\right) + Q , \label{eq9} \ee where \be {d\gamma \over dr} = - f(r)\gamma^2 - g{\gamma \over r} , \label{eq10} \ee are the electron energy losses. In both equations above, the derivatives over comoving time $t'$ have been replaced by the derivatives over the distance $r$, according to the relation $\partial /\partial t' = c \beta_{\Gamma} \Gamma \partial /\partial r\, $. The first term on the rhs of Eq.~(\ref{eq10}) represents synchrotron plus Compton energy losses, \ie, \be f(r) = {\sigma_T \over 6 m_e c^2 } {B'^2 \over \beta_{\Gamma} \Gamma} (1 + u_s'/u_B') , \label{eq11} \ee where $u_B' = B'^2/8\pi$ is the magnetic energy density, and $u_s'$ is the energy density of the synchrotron radiation, both as measured in the blast wave frame. The second term on the rhs of Eq.~(\ref{eq10}) represents the adiabatic losses. The parameter $g$ depends on the geometry of the expansion; for 2-dimensional (lateral) expansion $g =2/3$, and for 3-dimensional expansion $g=1$. We calculate the magnetic field following Chiang \& Dermer~(1999) \be u_B' \equiv {{B'}^2 \over 8 \pi} = \epsilon_B \kappa \rho c^2 \Gamma^2 , \label{eq12} \ee where $\kappa$ is the compression ratio and $\epsilon_B$ parameterizes the departure of the magnetic field intensity from its equipartition value. \subsection{Synchrotron spectrum} The evolution of the synchrotron spectrum in the blast wave frame is given by \be L_{syn,\nu'}'(r) = \int N_{\gamma}(r) P(\nu', \gamma) d\gamma , \label{eq13} \ee where $P(\nu', \gamma)$ is the power spectrum of synchrotron radiation of a single electron in isotropic magnetic field (see, \eg, Chiaberge \& Ghisellini~1999). The apparent monochromatic synchrotron luminosity as a function of time (a light curve) is calculated from \be L_{syn,\nu}(t, \theta_{obs}) = \int\!\!\!\int_{\Omega_j} {L_{syn,\nu'}'[r(\tilde \theta)] {\cal D}^3 \over \Omega_j } d\cos {\tilde \theta} d \tilde \phi , \label{eq14} \ee where ${\cal D} = 1 /\Gamma(1-\beta_{\Gamma} \cos{\tilde\theta})$ is the Doppler factor of the blast wave at the angle $\tilde\theta$. The coordinates ($\tilde\theta, \tilde \phi$) are chosen so that the observer is located at $\tilde \theta=0$ ($\theta=\theta_{obs}$) and the jet axis is at $\tilde \theta = \theta_{obs}$. The integral is taken over the surfaces \be t = \int {(1-\beta_{\Gamma} \cos \tilde\theta) \over c \beta_{\Gamma}} \, dr = {\rm const} , \label{eq15} \ee enclosed within the blast wave boundaries, $\Omega_j$. \subsection{Inverse-Compton radiation} We assume hereafter that cooling of relativistic electrons is dominated by synchrotron radiation, \ie\ that $u_s' \ll u_B'$. This condition will be verified and discussed in Appendix A. \section{RESULTS} We have used the following model parameters of the afterglow model in our calculations: initial energy per solid angle, $E_0/\Omega_{j0} = 10^{54} \, {\rm ergs \, s}^{-1} / 4 \pi$; $\Gamma_0 = 300$; $\theta_{j0} = 0.2$; $\kappa = 4$; $\rho = m_p/1 {\rm cm}^3$; $\epsilon_e = 0.1$ (the quasi-adiabatic case); $\epsilon_B = 0.03$; $p=2.4$. The parameters were not chosen to fit any specific observations, but rather to demonstrate the difference between simple analytical predictions and self-consistent numerical calculations regarding the beaming effects in a light-curves. In Fig.~1 we present the dependence of a bulk Lorentz factor of the blast wave on its distance from the central engine. We show three solutions for three different values of $v_l'$: $v_l'=0$ (thin line); $v_l'=c/\sqrt {3}$ (solid line); and $v_l'=c$ (dotted line). For $v_l'=0$ ($\to \theta_j={\rm const}$) and $r_0 \ll r \ll r_{nr}$, the bulk Lorentz factor is well approximated by \be \Gamma \simeq \Gamma_0 \left(r_0/r \right)^{3/2} , \label{eq16} \ee where \be r_0 \simeq \left( 3 E_0 \over \Gamma_0^2 \rho c^2 \Omega_j \right)^{1/3} \simeq 1.2 \times 10^{17} {\rm cm} \,, \label{eq17} \ee is the radius where deceleration of the GRB ejecta by sweeping of interstellar gas starts to be efficient, and \be r_{nr} \simeq r_0 \left (\Gamma_0\over 2 \right)^{2/3} \simeq 3.4 \times 10^{18} {\rm cm} \, , \label{eq18} \ee is the radius above which the blast wave becomes nonrelativistic. Fig.~1 demonstrates that steepening of the $\Gamma(r)$ curves due to lateral outflow is very smooth, without any sharp break like the one predicted analytically to take place at a distance, at which $\Gamma$ drops below $1/\theta_{j0}$, \ie\ at \be r_D \simeq r_0 (\Gamma_0 \theta_{j0})^{2/3} \simeq 1.8 \times 10^{18} {\rm cm} \, . \label{eq19} \ee In Fig.~2 we show the radial dependence of the rate of kinetic energy accreted by the blast wave, $ dE_{acc}'/ dt' $ (see Eq.~\ref{eq8}). As one could expect, the larger the lateral outflow speed, the larger the accretion rate is. The steepening of curves at large $r$ is due to transition from the relativistic regime ($\Gamma >2$) to the nonrelativistic regime, where $dE_{acc}' /dt'$ is significantly reduced, and becomes $\propto (\Gamma-1)$ (see Eq.~\ref{eq8}). When divided by $\epsilon_e$, curves in Fig.~2 illustrate also the $r$ dependence of injection luminosity of relativistic electrons (see Eq.~\ref{eq7}). In Fig.~3 we show time evolution of the electron energy distribution, $N_{\gamma}$, multiplied by $\gamma^2$. The curves are calculated at such values of the radius $r$, from which the signal produced on the axis $\tilde \theta=0$ is reached by the observer $t=1$,$ 10$, $10^2$, ..., $10^7$ seconds after ``the signal'' from $r=0$. The relation between $r$ and $t$ is \be t = \int_0^r {(1-\beta_{\Gamma})\over c \beta_{\Gamma}}\, dr \simeq \int_0^r {1 \over 2 c \Gamma^2} \, dr . \label{eq20} \ee The peak positions of the $N_{\gamma}\gamma^2$ curves mark Lorentz factor of those electrons which carry most of leptonic energy at a given distance. For injection spectral index $2 < p < 3$, the peak is located at $\gamma_m$ given by Eq.~(\ref{eq6}), and this is the case in our model. Another characteristic energy is \be \gamma_c = 6.1 \times 10^{20} {m_p\over \epsilon_B \kappa \rho } {1\over r \Gamma} , \label{eq21} \ee below which the time scale of electron energy losses due to synchrotron radiation is longer than the dynamical time scale. We present the dependence of $\gamma_m$ and $\gamma_c$ on the radius $r$ in Fig.~4. We can see from Fig.~4, that for the first 5 curves presented in the Fig.~3, $\gamma_m > \gamma_c$. In this case, in accordance with the analytical predictions, the electron spectra at $\gamma > \gamma_m$ are well described by the power-law function, $N_{\gamma} \propto \gamma^{-s}$, with the index $s=p+1$. For $\gamma_c < \gamma < \gamma_m$, analytical crude estimations predict $s=2$, which in our plot should be represented by horizontal lines. This, however, is expected to be true only for $\gamma_c \ll \gamma < \gamma_m$. In our model the ratio $\gamma_m/\gamma_c$ is not large enough to provide space for $s=2$ and there is a smooth transition to very hard low energy tail reached by electrons due to adiabatic losses. For $\gamma_c > \gamma_m$, which is the case for the top 4 curves in the Fig.~3, the predicted electron spectra should have a slope $s=p+1$ for $\gamma> \gamma_c$, and $s=p$ for $\gamma_m < \gamma < \gamma_c$. The former is seen, but the latter, again, due to narrow range between $\gamma_c$ and $\gamma_m$ doesn't apply. Instead, there the log-energy distribution is curved, smoothly joining the high energy portion of the electron spectrum with its low energy adiabatic part. Let us note, that very steep low energy tails of the curves on top of the plot result from the fact that there is not enough time for electrons to drift adiabatically to lower energies. Note also, that the details of the low energy parts of the electron energy distribution are not important, because contribution of electrons from these parts to the observed radiation is negligible. In Fig.~5 we present the observed radiation spectra computed for the same sequence of $t$, as electron energy distributions shown in Fig.~3. It should be pointed out, however, that unlike in simple analytical calculations, they are computed by integration of electron radiation from $t={\rm const}$ surfaces (see Eq.~\ref{eq15}), \ie\ taking into account light travel differences between photons emitted at different $\tilde \theta$'s. The observed radiation spectra are peaked around $h \nu \sim \Gamma \gamma_m^2 B/B_{cr} m_e c^2$, where rhs quantities are calculated for $r$ given by Eq.~(\ref{eq15}) and $B_{cr} = 2 \pi m_e^2 c^3/h e \simeq 4.4 \times 10^{13}$ Gauss. As we can see from Fig.~5, the high energy and the low energy parts of the observed radiation spectra are well described by power-law functions $L_{\nu} \propto \nu^{-\alpha}$, with $\alpha = p/2 = 1.1$ and $\alpha = -1/3$, respectively. The former is produced, as predicted analytically, by electrons with $\gamma >$ Max$[\gamma_m; \gamma_c]$, the latter represents the low energy synchrotron radiation of electrons with energies $\gamma <$ Min$[\gamma_c;\gamma_m]$. These high and low energy spectrum portions are joined very smoothly without showing any intermediate piece of the power-law spectrum. This smoothing of the observed radiation spectra results mainly from the fact that the observed radiation at any given moment is contributed by radiation from $t={\rm const}$ surfaces, \ie\ from different radii. The light curves, computed for $\nu = 4.2 \times 10^{14}$Hz and $\nu = 2.5 \times 10^{17}$Hz, are shown in Fig.~6. In this calculation we used $v_l'= c/\sqrt{3}$ and considered two different locations of the observer, $\tilde \theta=0$ and $\tilde \theta= 0.28$. The latter case is for the observer located outside the initial ejecta cone. In order to demonstrate better the beaming effect and its dependence on $v_l'$, we plot in Fig.~7 four optical light curves; three for different values of $v_l'$: $ 0$, $c/\sqrt{3}$, and $c$, and the fourth one for the spherical outburst. We can see, that for models with lateral expansion the steepening of the light curves is extended over more than two time decades and, down to nonrelativistic regime, doesn't reach analytically predicted slope $\beta = p$ ($L_{\nu} \propto t^{-\beta}$). Sharp break is found only for $v_l'=0$ model, and, it emerges shortly after $t(r_D)$, as theoretically predicted. We should note here, that our calculations of the model with the outflows expansion are not fully consistent, because the Doppler factor includes only the radial component of the bulk motion. This, however, is expected to affect only the results of the $v_l=c$ model, where we overestimate the radiation contributed from the blast wave edge. \section{DISCUSSION AND CONCLUSIONS} The afterglows provide exceptional opportunity to study whether and how much the GRB ejecta are beamed. As predicted by Rhoads~(1997), the beamed outflows should diverge from the cone geometry while decelerated by sweeping up the external gas. The sideways outflow of the shocked relativistic plasma increases the front of the blast wave leading to a faster deceleration. Rhoads~(1997) showed, using simple analytical analyses, that this should be imprinted in the light curve as a break around $t(r_D)$, \ie\ when $\Gamma$ drops below $(c_l/c)/\theta$. There the light curve should steepen, changing the slope from $\beta =(3p-2)/4$ (for $\gamma > {\rm Max}[\gamma_c; \gamma_m]$) or from $\beta = 3(p-1)/4$ (for $\gamma_m < \gamma < \gamma_c$) (Sari, Piran, \& Narayan~1998) to $\beta=p$ (Rhoads~1999). Our numerical results agree only qualitatively with these predictions. The steepening does occur, however, the slope change is smaller (to about $2.0$ instead of $p=2.4$) and is extended over more than two decades of the observed time. There are two reasons why, contrary to simple analytical estimations, the distinct break does not emerge in our calculations. First, as is shown in Fig.~1, the dynamics of the blast wave is affected by the lateral outflow very smoothly over the whole deceleration phase, and not just around $r_D$ (note, that at $r_D$ the blast wave area is already almost $4$ times larger than it would be without the sideways expansion). Second, the observed radiation at any given moment $t$ is contributed by the plasma which at larger $\theta$ emits radiation from smaller $r$, and noting, that at smaller $r$ plasma is moving faster and radiating stronger than at larger $r$, the contribution of the off-axis plasma to the observed radiation is larger than in the case of radiation contribution taken from $r={\rm const}$ surfaces as calculated analytically. The steepening of the light curve is predicted also for the beamed ejecta without the lateral outflows. In this case, because no change of dynamics and small light travel effects at $r >r_b$ (note that there the Doppler cone becomes narrower than the ejecta cone), the break is very well located, just around the time the $\Gamma$ drops below $1/\theta_j$, and the light curve steepens by $\Delta \beta = 3/4$, in accordance with analytical predictions (see, \eg, M\'esz\'aros \& Rees~1999). It should be emphasized, however, that our treatment of the dynamics with the sideways expansion is based on the approximation, that at any $r$ the material is uniformly distributed across the blast wave. In reality, the lateral outflow can create $\theta $ dependent structure, with the density of the swept material and the radial bulk Lorentz factor decreasing sideways, and in this case the break in the light curve may become more prominent. 2D-hydro relativistic simulations are required to verify this. \acknowledgments This project was partially supported by NSF grant No. AST-9529175, ITP/NSF grant No. PHY94-07194, and the Polish KBN grant No. 2P03D 00415. MS and RM thank Fellows of ITP/UCSB for hospitality during the visit and participating in the program ``Black Hole Astrophysics''. RM thanks NASA for support under the Long Term Space Astrophysics grant NASA-NAG-6337. MS acknowledges financial support of NASA/RXTE and ASTRO-E observing grants to USRA (G. Madejski, PI).
\section{Introduction} The study of the radical $J(V)$ for a vertex operator algebra $V$ was initiated in [DLMM], where we defined the radical $J(V)$ and determined $J(V)$ in the case $V$ is CFT type (see Section 3 for the definition of CFT type vertex operator algebra). Let $M$ be an admissible $V$-module (see [DLM2] and below for the definition). The $M$-radical $J_M(V)$ of $V$ consists of vectors $v\in V$ such that $o(v)=0$ on $M$ where $o(v)=v_{{\rm wt} v-1}$ if $v$ is homogeneous and $o(u+v)=o(u)+o(v).$ In the case that $M=V,$ $J_V(V)$ is exactly $J(V)$ which was determined to be $(L(0)+L(-1))V+J(V)_1$ in [DLMM] where $J(V)_1$ is the weight one subspace of $J(V).$ It turns out that a similar result is true for $J_M(V).$ We show in this paper that $J_M(V)=(L(0)+L(-1))V+J_M(V)_{(0,1)}$ where $J_M(V)_{(0,1)}$ is the intersection of $J_M(V)$ with $V_0\oplus V_1.$ Although the method for determining $J_M(V)$ is similar to that for determining $J_V(V)$ in [DLMM], the argument here is more complicated. The reason is that $V$ has a vacuum ${\bf 1}$ but $M$ does not have in general. We expect that the concept of $M$-radical $J_M(V)$ of $V$ will play a very important role in the theory of vertex operator algebra. The second main result in this paper is a criterion for irreducibility of an admissible $V$-module $M$ (see Proposition \ref{p5.3}). The result says that $M$ is irreducible if and only if each homogeneous subspace is irreducible $\hat V(0)$-module or $S_M(V)$-module (see Section 4 for the definition of $\hat V(0)$ and $S_M(V)$). We also formulate this result in terms of the theory of associative algebras $A_n(V)$ developed in [DLM4]. This result is important in the study of dual pair associated to a vertex operator algebra and an automorphism group (cf. [DLM1]). Both results are extended to twisted modules. In particular we also define the radical $J_V(M)$ for an admissible $g$-twisted $V$-module for an automorphism $g$ of $V$ of finite order and determine $J_V(M)$ precisely. A similar criterion of irreducibility of $M$ is obtained too in terms of certain Lie algebra $S_M(V^0)$ (see Section 5) and associative algebra $A_{g,n}(V)$ [DLM5]. \section{Preliminary} Let $(V,Y,{\bf 1},\omega)$ be a vertex operator algebra (see [B] and [FLM]). We shall use commuting formal variables $z,z_0,z_1,z_2.$ We shall also use delta-function $\delta(z)=\sum_{n\in\mathbb Z}z^n$ whose elementary properties can be found in [FLM]. First recall from [FLM], [Z], [DLM2] the definitions of weak module, admissible module, and ordinary module for a vertex operator algebra $V.$ A {\em weak module} $M$ for $V$ is a vector space equipped with a linear map $$\begin{array}{l} V\to (\mbox{End}\,M)[[z^{-1},z]]\label{map}\\ v\mapsto\displaystyle{Y_M(v,z)=\sum_{n\in \mathbb Z}v_nz^{-n-1}\ \ (v_n\in\mbox{End}\,M)} \mbox{ for }v\in V\label{1/2} \end{array}$$ satisfying the following conditions for $u,v\in V$, $w\in M$: \begin{eqnarray*}\label{e2.1} & &v_nw=0\ \ \ \mbox{for}\ \ \ n\in \mathbb Z \ \ \mbox{sufficiently\ large};\label{vlw0}\\ & &Y_M({\bf 1},z)=1;\label{vacuum} \end{eqnarray*} \begin{equation}\label{jacobi} \begin{array}{c} \displaystyle{z^{-1}_0\delta\left(\frac{z_1-z_2}{z_0}\right) Y_M(u,z_1)Y_M(v,z_2)-z^{-1}_0\delta\left(\frac{z_2-z_1}{-z_0}\right) Y_M(v,z_2)Y_M(u,z_1)}\\ \displaystyle{=z_2^{-1}\delta\left(\frac{z_1-z_0}{z_2}\right) Y_M(Y(u,z_0)v,z_2)}. \end{array} \end{equation} Here and below $(z_i-z_j)^n$ for $n\in \mathbb C$ is to be expanded in nonnegative powers of the second variable $z_j.$ This completes the definition. We denote this weak module by $(M,Y_M)$ (or briefly by $M$). An {\em ordinary} $V$-{\em module} is a weak $V$-module which carries a $\mathbb C$-grading $$M=\bigoplus_{\lambda \in{\mathbb C}}M_{\lambda} $$ such that $\dim M_{\lambda }$ is finite and $M_{\lambda +n}=0$ for fixed $\lambda $ and $n\in {\mathbb Z}$ small enough. Moreover one requires that $M_{\lambda }$ is the $\lambda $-eigenspace for $L(0):$ $$L(0)w=\lambda w=(\mbox{wt}\,w)w, \ \ \ w\in M_{\lambda }$$ where $L(0)$ is the component operator of $Y_M(\omega,z)=\sum_{n\in\mathbb Z}L(n)z^{-n-2}.$ An {\em admissible} $V$-{\em module} is a weak $V$-module $M$ which carries a ${\mathbb Z}_{+}$-grading $$M=\bigoplus_{n\in {\mathbb Z}_{+}}M(n)$$ ($\mathbb Z_+$ is the set all nonnegative integers) such that if $r, m\in {\mathbb Z} ,n\in {\mathbb Z}_{+}$ and $a\in V_{r}$ then $$a_{m}M(n)\subseteq M(r+n-m-1).$$ Note that any ordinary module is an admissible module. A vertex operator algebra $V$ is called {\em rational} if any admissible module is a direct sum of irreducible admissible modules. It was proved in [DLM3] that if $V$ is rational then there are only finitely many inequivalent irreducible admissible modules and each irreducible admissible module is an ordinary module. The following proposition can be found in [L2] and [DM]. \begin{prop}\label{p1} Any irreducible weak $V$-module $M$ is spanned by $\{u_nw|u\in V,n\in\mathbb Z\}$ where $w\in W$ is any fixed nonzero vector. \end{prop} Let $M$ be a weak $V$-module. We define the $M$-radical of $V$ to be \begin{equation}\label{3.1} J_M(V)=\{v\in V|o(v)|_M=0\} \end{equation} where $o(v)=v_{{\rm wt} v-1}$ for homogeneous $v\in V$ and $o(u+v)=o(u)+o(v).$ If $M=V$ this is precisely the definition of radical of $V$ given in [DLMM]. ($J_V(V)$ was denoted by $J(V)$ in [DLMM]). If $M=\oplus_{n\geq 0}M(n)$ is an admissible module then $o(v)M(n)\subset M(n)$ for all $n\in\mathbb Z.$ Recall from [DLMM] that $V$ is of CFT type if $V$ is simple and $V=\oplus_{n\geq 0}$ with $V_0$ one-dimensional. It was proved in [DLMM] that if $V$ is of CFT type then $J_V(V)$ is equal to $(L(0)+L(-1))V+J_V(V)_1$ where $J_V(V)_1=V_1\cap J(V).$ Here we prove a similar result for $J_M(V)$ for any admissible module $M$ with the same assumption on $V.$ \section{Determination of $J_M(V)$} We need several lemmas. \begin{lem}\label{l4.a1} Let $V$ be a simple vertex operator algebra and $M$ a weak $V$-module. Let $u\in V$ such that the vertex operator $Y_M(u,z)$ on $M$ involves only either finitely many positive powers or finitely many negative powers of $z$ then $u\in V_0.$ \end{lem} \noindent {\bf Proof:} \, The proofs in the two cases are similar. We only deal with the case that $Y_M(u,z)$ involves only finitely many positive powers of $z.$ We first prove that $$Y_M(u,z_1)Y_M(v,z_2)=Y_M(v,z_2)Y_M(u,z_1)$$ for all $v\in V.$ By (7.24) of [DL] (also see [FLM]) there exists a nonnegative integer $n$ such that \begin{equation}\label{new} (z_1-z_2)^nY_M(u,z_1)Y_M(v,z_2)=(z_1-z_2)^nY_M(v,z_2)Y_M(u,z_1). \end{equation} Since each factor in (\ref{new}) involves only finitely many positive powers of $z_1$ we multiply (\ref{new}) by $(z_1-z_2)^{-n}$ to obtain $Y_M(u,z_1)Y_M(v,z_2)=Y_M(v,z_2)Y_M(u,z_1).$ From $[Y_M(\omega,z_1),Y_M(u,z_2)]=0$ we see that $$0=[L(-1),Y_M(u,z)]=Y_M(L(-1)u,z).$$ From the Jacobi identity (\ref{jacobi}) we have the associator formula (see Chapter 8 of [FLM]): for $a,b\in V$ and $w\in M$ there exists a nonnegative integer $n,$ which depends on $a$ and $w$ only, such that $$(z_0+z_2)^nY_M(Y(a,z_0)b,z_2)w=(z_0+z_2)^nY_M(a,z_0+z_2)Y_M(b,z_2)w.$$ So if $b=L(-1)u$ then $Y_M(b,z_2)=0$ on $M$ and $$(z_0+z_2)^nY_M(Y(a,z_0)b,z_2)w=0$$ or that $$Y_M(Y(a,z_0)b,z_2)w=0.$$ This shows that $Y_M(a_mb,z)=0$ on $M$ for any $a\in V$ and $m\in\mathbb Z.$ Assume that $b\ne 0.$ Since $V$ is simple then the span of $a_mb$ for $a\in V$ and $m\in \mathbb Z$ is the whole $V$ by Proposition \ref{p1}. As a result we have $Y_M(v,z)=0$ for every $v\in V.$ This is a contradiction as $Y_M({\bf 1},z)=id_M.$ Thus $b=L(-1)u=0.$ Since $L(-1): V_n\to V_{n-1}$ is injective if $n\ne 0$ (cf. [L1] and [DLiM]) we immediately have that $u\in V_0,$ as required. \mbox{ $\square$} \begin{lem}\label{l4.a2} Let $V$ be a vertex operator algebra of CFT type. Let $v\in V_n$ with $n\geq 2$ such that $L(1)v=0.$ Then $v\not\in J_M(V).$ \end{lem} \noindent {\bf Proof:} \, Assume that $v\in J_M(V).$ Then $o(v)=v_{n-1}=0$ on $M.$ Using the relation $[L(-1),v_m]=-mv_{m-1}$ we see that $v_k=0$ for $0\leq k\leq n-1.$ Thus for any $u\in V$ we have $$0=[v_i,u_{-i}]=\sum_{t=0}^i\binom{i}{t}(v_tu)_{-t}$$ for $i=0,...,n-1.$ This shows that $$(v_iu)_{-i}=0$$ for $i=0,...,n-1.$ Using the relation $[L(-1),a]=-ma_{m-1}$ repeatedly for $a\in V$ gives $$(v_iu)_k=0$$ for $i=1,..,n-1$ and $k\leq -i.$ Thus $Y_M(v_iu,z)$ involves only finitely many positive powers of $z.$ It follows from Lemma \ref{l4.a1} that $v_iu\in V_0$ for $i=1,..n-1.$ If $u$ is homogeneous of weight $s\geq 0$ then the weight $v_{n-1}u$ again is $s.$ Thus if $s>0$ then $v_{n-1}u=0.$ If $s=0$ then $u$ is a multiple of ${\bf 1}$ and again $v_{n-1}u=0.$ Thus $v\in J_V(V).$ On the other hand $J_V(V)=J_V(V)_1+(L(-1)+L(0))V$ (Theorem 1 of [DLMM]). It is clear that $v\not\in J_V(V).$ This is a contradiction. \mbox{ $\square$} \begin{lem}\label{l4.a3} Let $V$ be a vertex operator algebra of CFT type and $M$ a weak $V$-module. Let $v\in V_1$ such that $o(v)$ is a constant on $M.$ Then $v\in J_V(V).$ Moreover if $M$ is irreducible then $o(v)$ is a constant on $M$ if and only if $v\in J_V(V).$ \end{lem} \noindent {\bf Proof:} \, For any $u\in V$ and $n\in \mathbb Z$ we have $$0=[o(v),Y_M(u,z)]=[v_0,Y_M(u,z)]=Y_M(v_0u,z)$$ As in the proof of Lemma \ref{l4.a1} we conclude that $v_0u=0$ for all $u\in V.$ That is, $v\in J_V(V).$ If $v\in J_V(V)$ then again we have $[o(v),Y_M(u,z)]=Y_M(v_0u,z)=0.$ If $M$ is irreducible then $M$ has countable dimension. Let ${\rm Hom}_V(M,M)$ denote the set of all $V$-homomorphism from $M$ to itself. Then ${\rm Hom}_V(M,M)$ is a division ring over $\mathbb C.$ Let $w\in M$ be any nonzero vector. Then $f\mapsto f(w)$ gives a bijection from ${\rm Hom}_V(M,M)$ to ${\rm Hom}_V(M,M)w$ which has countable dimension. Thus ${\rm Hom}_V(M,M)$ has countable dimension. Since any division ring over $\mathbb C$ with countable dimension is $\mathbb C$ itself (cf. [DLM3]) we conclude that ${\rm Hom}_V(M,M)=\mathbb C.$ we now see immediately that $o(v)$ is a constant on $M.$ \mbox{ $\square$} We can now determine the radical $J_M(V)$ precisely. \begin{thm}\label{t3.1} Suppose that $V$ is a vertex operator algebra of CFT type. Then for any admissible $V$-module $M$ we have $$J_M(V)=(L(0)+L(-1))V+J_M(V)_{(0,1)}$$ where $J_M(V)_{(0,1)}=(V_0+V_1)\cap J_M(V).$ Moreover, if $a=a^0+a^1\in J_M(V)_{(0,1)}$ with $a^i\in V_i$ then $a^1\in J_V(V).$ That is, the image of the projection of $J_M(V)_{(0,1)}$ into $V_1$ is contained in $J_V(V).$ \end{thm} \noindent {\bf Proof:} \, The proof of this theorem is similar to that of Theorem 1 of [DLMM]. The conclusion $(L(0)+L(-1))V+J_M(V)_{(0,1)}\subset J_M(V)$ is clear. First we recall a result from [DLiM] (Corollary 3.2). As a module for $sl(2,\mathbb C)=\<L(-1),L(0),L(1)\>,$ $V$ is a direct sum of highest weight modules $X(\mu)$ with highest weights $\mu>$ ($\mu>0$), the trivial module and the projective cover $P(1)$ of $X(1).$ Thus for any $x\in J_M(V)$ we can write $$x=\sum_{n=0}^mL(-1)^nu^n$$ where each $u^n$ either is in $V_1$ or satisfies $L(1)u^n=0.$ We assume that $u^m\ne 0.$ We prove by induction on $m$ that $x$ lies in $J_M(V)_{(0,1)}+(L(0)+L(-1))V.$ Suppose first that $m=0.$ Then $x=u^0.$ Write $x=\sum_{i\geq 0}x^i$ where $x^i\in V_i$ and $L(1)x^i=0$ if $i\ne 1.$ Since $o(x)=0$ on $M$ we have $$0=[L(1),o(x)]=\sum_{i\geq 0}[L(1),o(x^i)]=\sum_{i\geq 2}2(i-(i-1)/2-1)x^i_i$$ on $M$ where we have used the fact that $[L(1),o(x^i)]=0$ for $i\leq 1$ (see Lemma 2.5 of [DLMM]). That is, $$\sum_{i\geq 2}(i-1)x^i_i=0.$$ Thus $$0=\sum_{i\geq 2}(i-1)[L(1),[L(-1),x_i^i]]=\sum_{i\geq 2}(i-1)[L(1),-ix^i_{i-1}]=-\sum_{i\geq 2}(i-1)^2ix_i^i$$ on $M.$ Continuing in this way we get $$\sum_{i\geq 2}(i-1)^ki^{k-1}x^i_i=0$$ for all $k\geq 1.$ It follows that each $x^i_i=0$ for all $i\geq 2.$ Using the relation $[L(-1),u_n]=-nu_{n-1}$ shows inductively that $x^i_j=0$ for $j=0,...,i.$ Thus $x^i\in J_M(V).$ If $x^i\ne 0$ then $x^i$ is not in $J_M(V)$ by Lemma \ref{l4.a2}. This is a contradiction. Thus $x^i=0$ for all $i\geq 2.$ So $x=x^0+x^1.$ Since $x^0$ is a multiple of ${\bf 1}$ we see that $o(x^1)=-o(x^0)$ is a constant on $M.$ By Lemma \ref{l4.a3} $x^1\in J_V(V).$ This proves the result for $m=0.$ For $m>0$ set $a=L(-1)^{m-1}u^m$ and $b=\sum_{n=0}^{m-1}L(-1)^nu^n.$ Thus $x=L(-1)a+b.$ From $(L(0)+L(-1))a\in J_M(V)$ we have $$0=o(x)=o(L(-1)a)+o(b)=-o(L(0)a)+o(b)=o(b-L(0)a).$$ Note that $L(0)a=(m-1)L(-1)^{m-1}u^m+L(-1)^{m-1}L(0)u^m$ so that $$b-L(0)a=\sum_{n=0}^{m-2}L(-1)^nu^n-L(-1)^{m-1}((m-1)u^m+L(0)u^m-u^{m-1})$$ lies in $J_M(V).$ Since either $L(0)u^m\in V_1$ or $L(1)L(0)u^m=0,$ we conclude by induction that $b-L(0)a$ lies in $J_M(V)_{(0,1)}+(L(0)+L(-1))V.$ But then the same is true for $x=b -L(0)a+(L(0)+L(-1))a.$ This completes the proof of the theorem. \mbox{ $\square$} \section{A criterion for irreducibility} In this section we give a criterion for irreducibility of an admissible module for an arbitrary vertex operator algebra $V$ which we do not assume to be simple. We consider the quotient space $$\hat V={\mathbb C}[t,t^{-1}]\otimes V/D{\mathbb C}[t,t^{-1}]\otimes V$$ where $D=\frac{d}{dt}\otimes 1+1\otimes L(-1).$ Denote by $v(n)$ the image of $t^n\otimes v$ in $\hat V$ for $v\in V$ and $n\in \mathbb Z.$ Then $\hat V$ is $\mathbb Z$-graded by defining the degree of $v(n)$ to be ${\rm wt} v-n-1$ if $v$ is homogeneous. Denote the homogeneous subspace of degree $n$ by $\hat V(n).$ The space $\hat V$ is, in fact, a $\mathbb Z$-graded Lie algebra with bracket $$[a(m), b(n)]=\sum_{i=0}^{\infty}\binom{m}{i}a_ib(m+n-i)$$ for $a,b\in V$ (see [B], [L2] and [DLM3]). Note that $\hat V(0)$ is a subalgebra of $\hat V$ and is isomorphic to $V/(L(-1)+L(0))V$ whose Lie bracket is given by $$[a,b]=\sum_{n=0}^{{\rm wt} a-1}\binom{{\rm wt} a-1}{n}a_nb$$ for homogeneous $a,b\in V.$ Let $M$ be an admissible $V$-module. Then the map from $\hat V$ to ${\rm End} M$ by sending $v(m)$ to $v_m$ is a Lie algebra homomorphism (cf. [L2] and [DLM3]). In particular, the restriction of this map to $\hat V(0)$ gives a Lie algebra homomorphism from $ V/(L(-1)+L(0))V$ to ${\rm End} M.$ The kernel of this map is exactly the $M$-radical $J_M(V).$ Set $$S_M(V)=V/J_M(V).$$ Then $S_M(V)$ is a quotient Lie algebra of $V/(L(-1)+L(0))V$ by Theorem \ref{t3.1} and acts on $M$ faithfully. \begin{lem}\label{la1} Let $V$ be a finite dimensional vertex operator algebra. Then $V=V_0$ is a commutative associative algebra such that $Y(a,z)b=ab$ for $a,b\in V.$ \end{lem} \noindent {\bf Proof:} \, Since $L(-1)$ is injective on $\sum_{n>0}V_n$ (see [L1] and [DLiM]) we observe that $\sum_{n>0}V_n=0.$ In particular, $\omega=0$ and $L(0)=0.$ This shows that $V=V_0.$ It is clear now that $a_n=0$ for $a\in V_0$ and $n\ne -1.$ This implies that $Y(a,z)b=a_{-1}b.$ The reader can verify that $ab=a_{-1}b$ defines a commutative associative algebra structure on $V_0$ (see [B] and [L2]). \mbox{ $\square$} \begin{lem}\label{l1} Let $V$ be a vertex operator algebra and $M=\oplus_{n\geq 0}M(n)$ an admissible $V$-module with $M(0)\ne 0.$ Then $M$ is not equal to $\oplus_{n=0}^kM(n)$ for any $k\geq 0$ unless $V$ is finite dimensional. \end{lem} \noindent {\bf Proof:} \, If $M=\oplus_{n=0}^kM(n)$ such that $M(k)\ne 0.$ Take a nonzero $u\in M(k).$ Then from the definition of admissible module $L(-1)u\in M(k+1)=0.$ Thus $u$ is a vacuum-like vector and the submodule $W$ of $M$ generated by $u$ is isomorphic to the adjoint module $V$ [L1]. Since $L(-2)u\in M(n+2)=0$ we see that $\omega=0$ and $V=V_0$ is finite dimensional. \mbox{ $\square$} Now we use Proposition \ref{p1} and Lemma \ref{l1} to give a criterion for irreducibility of an admissible module. \begin{prop}\label{p5.3} Let $V$ be a vertex operator algebra with $\omega\ne 0.$ An admissible $V$-module $M=\oplus_{n\geq 0}M(n)$ with $M(0)\ne 0$ is irreducible if and only if each $M(n)$ is an irreducible $\hat V(0)$-module, or each $M(n)$ is an irreducible $S_M(V)$-module. \end{prop} \noindent {\bf Proof:} \, We have already mentioned that $M$ is a module for $\hat V$ under the action $v(n)\mapsto v_n$ where $Y_M(v,z)=\sum_{n\in\mathbb Z}v_nz^{-n-1}$ are vertex operators on $M$ for $v\in V.$ First we assume that $M$ is irreducible. By Proposition \ref{p1} $M=\hat V\cdot w$ for any nonzero vector $w$ of $M$. Now take $w\in M(n).$ Then $M(k)=\hat V(k-n)\cdot w.$ In particular, $M(n)=\hat V(0)\cdot w.$ Thus $M(n)$ is an irreducible $\hat V(0)$-module. Conversely suppose each $M(n)$ is an irreducible $\hat V(0)$-module. From the proof of Lemma 1.2.1 of [Z] we see that $L(0)$ acts on each $M(n)$ as a scalar. Let $W$ be any nonzero submodule of $M.$ Then $$W=\oplus_{n\geq 0}W(n)$$ where $W(n)=M(n)\cap W.$ From Lemma \ref{l1} and the injectivity of $L(-1)$ on $M(n)$ for all large $n$ (cf. [L1] and [DLiM]) we see that $W(n)\ne 0$ for all large $n.$ Note that each $W(n)$ is a submodule of $M(n)$ for $\hat V(0).$ So $W(n)=M(n)$ for all large $n$ as $M(n)$ is irreducible $\hat V(0)$-module. If $W\ne M$ then the quotient $M/W$ is an admissible $V$-module with only finitely many homogeneous subspaces. This is a contradiction by Lemma \ref{l1} unless $V$ is finite dimensional. Thus $W=M$ and $M$ is irreducible. \mbox{ $\square$} \begin{rem} {\rm In the case $V=V_0$ is a commutative associative algebra, the assertion in Proposition \ref{p5.3} is false. For example, if we take $V=\mathbb C,$ then $M=\sum_{n\geq 0}M(n)$ is a $V$-module with each $M(n)=V.$ Clearly, each $M(n)$ is an irreducible $S_M(V)$-module but $M$ is not irreducible under $V.$ } \end{rem} The result discussed Proposition \ref{p5.3} can be also formulated in terms of the theory of associative algebra $A_n(V)$ developed in [DLM4]. Let $O_n(V)$ be the linear span of all $u\circ_n v$ and $L(-1)u+L(0)u$ where for homogeneous $u\in V$ and $v\in V,$ $$u\circ_n v={\rm Res}_{z}Y(u,z)v\frac{(1+z)^{{\rm wt} u+n}}{z^{2n+2}}.$$ Define the linear space $A_n(V)$ to be the quotient $V/O_{n}(V).$ We also define a second product $*_n$ on $V$ for $u$ and $v$ as above: $$u*_nv=\sum_{m=0}^{n}(-1)^m\binom{m+n}{n}{\rm Res}_zY(u,z)\frac{(1+z)^{{\rm wt}\,u+n}}{z^{n+m+1}}v.$$ Extend linearly to obtain a bilinear product on $V$ which coincides with that of Zhu [Z] if $n=0.$ The following theorem was proved in [DLM4]; In the case $n=0$ it was proved previously in [Z]. \begin{thm}\label{theorem:2.1} Let $M=\sum_{n\geq 0}M(n)$ be an admissible $V$-module with $M(0)\ne 0.$ Then {\rm (i)} The product $*_n$ induces an associative algebra structure on $A_n(V)$ with the identity ${\bf 1}+O_n(V).$ Moreover $\omega+O_n(V)$ is a central element of $A_n(V).$ {\rm (ii)} The identity map on $V$ induces an onto algebra homomorphism from $A_n(V)$ to $A_m(V)$ for $0\leq m\leq n.$ {\rm (iii)} The map $u\mapsto o(u)$ gives a representation of $A_n(V)$ on $M(i)$ for $0\leq i\leq n.$ Moreover, $V$is rational if and only if $A_n(V)$ are finite dimensional semisimple algebras for all $n.$ \end{thm} Note that both the actions of $A_{n}(V)$ and $S_M(V)$ on $\sum_{0\leq m\leq n}M(m)$ are given by $v\mapsto o(v).$ Combining Proposition \ref{p5.3} and Theorem \ref{theorem:2.1} immediately gives \begin{thm}\label{t6} Assume that the Virasoro element $\omega$ of $V$ is nonzero. Then an admissible $V$-module $M=\oplus_{n\geq 0}M(n)$ with $M(0)\ne 0$ is irreducible if and only if each $M(n)$ is an irreducible $A_n(V)$-module. \end{thm} \section{Twisted case} This section is an analogue of Section 4 for a twisted module $M.$ We will omit a lot of details and refer the reader to the previous sections when it is clear how the corresponding proofs and arguments before carry out in this case. First we give definitions of various twisted modules following [FLM] and [DLM3]. Let $g$ be an automorphism of $V$ of order $T.$ Then we have eigenspace decomposition $V=\sum_{k=0}^{T-1}V^k$ where $V^k=\{v\in V|gv=e^{\frac{-2\pi i k}{T}}v\}.$ Then $V^0$ is a vertex operator subalgebra of $V$ with the same Virasoro vector. A {\em weak} $g$-{\em twisted} $V$-{\em module} $M$ is a vector space equipped with a linear map $$\begin{array}{l} V\to ({\rm End}\,M)\{z\}\\ v\mapsto\displaystyle{ Y_M(v,z)=\sum_{n\in\mathbb Q}v_nz^{-n-1}\ \ \ (v_n\in {\rm End}\,M)} \end{array}$$ such that for all $0\leq r\leq T-1,$ $u\in V^r$, $v\in V,$ $w\in M$, \begin{eqnarray*} & &Y_M(u,z)=\sum_{n\in \frac{r}{T}+\mathbb Z}u_nz^{-n-1} \\ & &u_lw=0\ \ \ \mbox{for}\ \ \ l>>0\\ & &Y_M({\bf 1},z)=id_M; \end{eqnarray*} $$ \begin{array}{c} \displaystyle{z^{-1}_0\delta\left(\frac{z_1-z_2}{z_0}\right) Y_M(u,z_1)Y_M(v,z_2)-z^{-1}_0\delta\left(\frac{z_2-z_1}{-z_0}\right) Y_M(v,z_2)Y_M(u,z_1)}\\ \displaystyle{=z_2^{-1}\left(\frac{z_1-z_0}{z_2}\right)^{-r/T} \delta\left(\frac{z_1-z_0}{z_2}\right) Y_M(Y(u,z_0)v,z_2)}. \end{array} $$ It is clear if $g=1$ this reduces the definition of weak module in Section 3. An {\em ordinary} $g$-{\em twisted $V$-module} is a weak $g$-twisted $V$-module $M$ with a $\mathbb C$-grading induced by the eigenvalues of $L(0):$ $$M=\bigoplus_{\lambda \in{\mathbb C}}M_{\lambda} $$ where $M_{\lambda }=\{w\in M|L(0)w=\lambda w\},$ $\dim M_{\lambda }$ is finite and for fixed $\lambda ,$ $M_{\frac{n}{T}+\lambda }=0$ for all small enough integers $n.$ An {\em admissible $g$-twisted $V$-module} is a weak $\frac{1}{T}{\mathbb Z}$-graded $g$-twisted $V$-module $M$ $$M=\bigoplus_{n=0}^{\infty}M(n/T)$$ such that $M(0)\ne 0$ and that $v_{m}M(n/T)\subseteq M(n/T+{\rm wt} v-m-1)$ for homogeneous $v\in V.$ Clearly, an ordinary $g$-twisted $V$-module is an admissible $g$-twisted $V$-module. \begin{rem} {\rm From the definition we see that any weak (admissible, ordinary) $g$-twisted $V$-module is a weak (admissible, ordinary) $V^0$-module.} \end{rem} Let $M$ be an admissible $g$-twisted $V$-module. For homogeneous $v\in V$ we denote $o(v)=v_{{\rm wt} v-1}$ on $M$ and extend it linearly to whole $V,$ as before. Then it is immediate from the definition that $o(v)=0$ for $v\in V^1\oplus\cdots \oplus V^{T-1}.$ Since $M$ is an admissible $V^0$-module we consider the $M$-radical of $V^0$ given in (\ref{3.1}). By Theorem \ref{t3.1} we have \begin{thm} Suppose that $V$ is a vertex operator algebra of CFT type. Then for any admissible $g$-twisted $V$-module $M$ we have $$J_M(V^0)=(L(0)+L(-1))V^0+J_M(V^0)_{(0,1)}$$ Moreover, if $a=a^0+a^1\in J_M(V^0)_{(0,1)}$ with $a^i\in V^0_i$ then $a^1\in J_{V^0}(V^0).$ \end{thm} Proposition \ref{p5.3} still holds in this case. \begin{prop}\label{p6.3} Let $V$ be a simple vertex operator algebra with $\omega\ne 0.$ An admissible $g$-twisted $V$-module $M=\oplus_{n\geq 0}M(n/T)$ with $M(0)\ne 0$ is irreducible if and only if each $M(n/T)$ is an irreducible $S_M(V^0)$-module. \end{prop} \noindent {\bf Proof:} \, If $M$ is irreducible then one can show that the analogue of Proposition \ref{p1} is true. That is, $M=\{u_nw|u\in V,n\in\frac{1}{T}\mathbb Z\}$ for any nonzero $w\in M.$ Thus for any nonzero $w\in M(n/T)$ we have $\{o(v)w|v\in V^0\}=M(n/T).$ So $M(n/T)$ is an irreducible $S_M(V^0)$-module. Note that for each $k=0,...,T-1,$ $M^k=\oplus_{n\geq 0}M(n+k/T)$ is an admissible $V^0$-module. If all $M(n/T)$ are irreducible $S_M(V^0)$-modules then $M^k$ is an irreducible admissible $V^0$-modules for $k=0,...,T-1$ by Proposition \ref{p5.3}. Using the associativity of vertex operators on $M$ we show that if $Y_M(v,z)w=0$ for some nonzero $v\in V$ and $w\in M$ then $Y_M(u,z)=0$ for all $u\in V$ (cf. Proposition 11.9 of [DL]: here we use the assumption that $V$ is simple). Since $\sum_{n=0}^{\infty}M(n)$ is nonzero ($M(0)\ne 0$ by assumption) we see from the associativity of vertex operators on $M$ that $\{u_n\sum_{n=0}^{\infty}M(n)|u\in V^k,n\in \mathbb Z\}$ is a nonzero $V^0$-submodule of $M^k.$ Thus $$\{u_n\sum_{n=0}^{\infty}M(n)|u\in V^k,n\in \mathbb Z\}=M^k.$$ In particular, $M^k$ is nonzero for all $k.$ Clearly $M^k$ and $M^i$ for $i\ne k$ are inequivalent $V^0$-modules. Let $0\ne w\in M^k.$ Then $\{u_nw|u\in V^i,n\in\mathbb Q\}$ is an admissible $V^0$-submodule of $M^{i+k}$ (where $i+k$ is understood modulo $T$) and thus must be equal to $M^{i+k}$ for all $i.$ That is, $\{u_nw|u\in V,n\in\mathbb Q\}=M$ and $M$ is an irreducible admissible $g$-twisted $V$-module. \mbox{ $\square$} As in the untwisted case, we can also formulate Proposition \ref{p6.3} in terms of theory of associative algebra $A_{g,n}(V)$ developed in [DLM3] and [DLM5]. Let $V$ and $g$ be before. Fix $n=l+\frac{i}{T}\in\frac{1}{T}\mathbb Z$ with $l$ a nonnegative integer and $0\leq i\leq T-1.$ For $0\leq r\leq T-1$ we define $\delta_i(r)=1$ if $i\geq r$ and $\delta_i(r)=0$ if $i<r$. We also set $\delta_i(T)=1.$ Let $O_{g,n}(V)$ be the linear span of all $u\circ_{g,n} v$ and $L(-1)u+L(0)u$ where for homogeneous $u\in V^r$ and $v\in V,$ $$u\circ_{g,n} v={\rm Res}_{z}Y(u,z)v\frac{(1+z)^{{\rm wt} u-1+\delta_i(r)+l+r/T}}{z^{2l +\delta_{i}(r)+\delta_{i}(T-r) }}.$$ Define the linear space $A_{g,n}(V)$ to be the quotient $V/O_{g,n}(V).$ Then $A_{g,n}(V)$ is the untwisted associative algebra $A_n(V)$ as defined in Section 4 if $g=1$ and is $A_g(V)$ in [DLM3] if $n=0.$ We also define a second product $*_{g,n}$ on $V$ for $u$ and $v$ as above: $$u*_{g,n}v=\sum_{m=0}^{l}(-1)^m\binom{m+l}{l}{\rm Res}_zY(u,z)\frac{(1+z)^{{\rm wt}\,u+l}}{z^{l+m+1}}v$$ if $r=0$ and $u*_{g,n}v=0$ if $r>0.$ Extend linearly to obtain a bilinear product on $V.$ Recall from [DLM3] that $V$ is called $g$-{\em rational} if any admissible $g$-twisted $V$-module is completely reducible. The following theorem was given in [DLM5]. \begin{thm} \label{theorem:2.2} Let $M=\sum_{n=0}^{\infty}M(n/T)$ be an admissible $g$-twisted $V$-module with $M(0)\ne 0.$ We have {\rm (i)} The product $*_{g,n}$ induces an associative algebra structure on $A_{g,n}(V)$ with the identity ${\bf 1}+O_{g,n}(V).$ Moreover $\omega+O_{g,n}(V)$ is a central element of $A_{g,n}(V).$ {\rm (ii)} The identity map on $V$ induces an onto algebra homomorphism from $A_{g,n}(V)$ to $A_{g,m}(V)$ for $m,n\in\frac{1}{T}\mathbb Z$ and $0\leq m\leq n.$ {\rm (iii)} The map $u\mapsto o(u)$ gives a representation of $A_{g,n}(V)$ on $M(i)$ for $i\in \frac{1}{T}\mathbb Z$ and $0\leq i\leq n.$ Moreover, $V$ is $g$-rational if and only if $A_{g,n}(V)$ are finite dimensional semisimple algebra for all $n.$ \end{thm} Clearly, both the actions of $A_{g,n}(V)$ and $S_M(V^0)$ on $\sum_{0\leq m\leq n}M(m)$ are induced by $v\mapsto o(v).$ Combining Proposition \ref{p6.3} and Theorem \ref{theorem:2.2} gives an analogue of Theorem \ref{t6}. \begin{thm}\label{t7} Assume that the Virasoro element $\omega$ of $V$ is nonzero. Then an admissible $g$-twisted $V$-module $M=\oplus_{n\geq 0}M(n)$ with $M(0)\ne 0$ is irreducible if and only if each $M(n)$ is an irreducible $A_{g,n}(V)$-module. \end{thm}
\section{Introduction} During the last decade the study of topological gauge theories provided deep insights into the topology and geometry of low dimensional manifolds. The central feature of topological field theories is that the observables only depend on the global structure of the space-time manifold on which the model is defined. There are in fact two different types of topological field theories: whether the whole gauge fixed action or just the gauge fixing part can be written as a BRS variation the model is of Witten-type or of Schwarz-type, respectively. An example of Witten-type models is the topological Yang-Mills theory, representatives of Schwarz-type models are Chern-Simons and BF theories. A common feature of many such field theories is the appearance of a so-called vector-like supersymmetry in the flat space-time limit. It is due to the energy-momentum tensor being BRS-exact. Its generators when anticommuted with the BRS operator close on translations. We are in particular interested in a BF-model defined on a two dimensional space-time manifold. Such a model was shown to be equivalent to a two dimensional gravity, which has been discussed in connection with the Liouville theory \cite{2dgravity}. Typically in two space-time dimensions the propagators of massless scalar fields are ill-defined \cite{balasin, olivier2}. Due to the singular behaviour of the ghost propagator at long distances, an infrared regulator mass has to be introduced \cite{blasi}. The infrared and ultraviolet finiteness of the two dimensional BF-model was already discussed in the realm of algebraic renormalization \cite{blasi}. The present work is devoted to the investigation of an enlarged model with the inclusion of a topological matter interaction \cite{2dgravity} in the context of the algebraic renormalization program \cite{olivier2}. The resulting model is characterized by an enlarged BRS symmetry. Moreover, we show the existence of a vector-like supersymmetry, which fact simplifies the investigation of the infrared and ultraviolet renormalizability of this model. The paper is organized as follows. In section 2 we describe the model at the classical level and we display its BRS transformations as well as the vector-like supersymmetry transformations. In section 3 we prove the finiteness of the model. Finally, in section 4, we show that the model is anomaly free. \section{Definition of the model at the classical level} Let us first consider the BF model on a two dimensional flat manifold ${\cal M}$ endowed with an Euclidean metric $\eta_{\mu\nu}$. This field model possesses the following metric independent action: \begin{equation} \S^{(1)}_{inv}=\frac{1}{2}\int_{{\cal M}}d^2x~ \varepsilon^{\mu\nu}F^a_{\mu\nu}\phi^a\,, \eqn{action1} where $\varepsilon^{\mu\nu}$ is the completely antisymmetric Levi--Civita tensor (with $\varepsilon^{12}=+1$), $\phi^a$ is a scalar field, and the field strength $F^a_{\mu\nu}$ is given by \begin{equation} F^a_{\mu\nu}=\partial_\mu A^a_\nu -\partial_\nu A^a_\mu + f^{abc}A^b_\mu A^c_\nu\,. \eqn{8345ljkdfg} $A_\mu^a$ stands for the gauge field with gauge index $a$. $f^{abc}$ are the structure constants of the gauge group which is assumed to be a compact and simple Lie group with all fields belonging to its adjoint representation. The generators of the Lie algebra are chosen to be anti-hermitian, such that $[T^a,T^b]=f^{abc} T^c$ and $Tr(T^a T^b)= \d^{ab}$.\\ The action $\S^{(1)}_{inv}$ is invariant under the following infinitesimal gauge transformations \begin{eqnarray} \d_{(\theta)} A^a_\mu &=& ({\partial_\mu} \theta^a + f^{abc} A^b_\mu\theta^c) \equiv (D_\mu\theta)^a\,, \nonumber\\ \d_{(\theta)} \phi^a &=& -f^{abc}\theta^b\phi^c\,, \eqan{gauge1} where $\theta^a$ is a local parameter, and $D_\mu$ is the covariant derivative. This model has already been studied in much detail \cite{blasi}. In the present work we enlarge the model by introducing a set of $N$ vector fields $B_\mu^{a\a}$ and $N$ scalars $X^{a\a}$ \cite{2dgravity}, where the index $\a$ takes all values from 1 to $N$. The contribution of these new fields represents a matter interaction. It is given by the metric independent local functional \begin{equation} \S_{N}=\int_{\cal M} d^2x ~ \varepsilon^{\mu\nu}(D_\mu B^\a_\nu)^a X^a_\a \,. \eqn{sdkfjgh} Without loss of generality, we restrict ourselves to the case $N=1$ implying that (\ref{sdkfjgh}) now reads \begin{equation} \S^{(2)}_{inv}=\int_{\cal M} d^2x ~ \varepsilon^{\mu\nu}(D_\mu B_\nu)^a X^{a} \,. \eqn{action2} The full action $\S_{inv}=\S^{(1)}_{inv}+ \S^{(2)}_{inv}$ is in fact invariant under an additional symmetry \cite{2dgravity} given by the transformations \begin{eqnarray} \d_{(\psi)}\phi^a&=&-f^{abc}\psi^b X^{c}\,,\nonumber\\ \d_{(\psi)} B_\mu^{a}&=&D_\mu\psi^{a}\,, \eqan{gauge2} where $\psi^a$ is an infinitesimal parameter.\\ As usual, the quantisation procedure requires a gauge fixing. In the present case we use a Landau-type gauge. Due to the fact that we have two gauge symmetries we need two sets of ghost fields with corresponding Lagrange multiplier fields: ($c^a$,$\l^{a}$) are the Faddeev-Popov ghost fields with corresponding antighost fields ($\bar{c}^a$,$\bar{\l}^{a}$), and ($b^a$,$d^{a}$) are Lagrange multiplier fields enforcing the Landau gauge conditions. The gauge fixed action is then given by $\Sigma_{inv}+\Sigma_{gf}$, where \begin{eqnarray} \S_{gf}&=&s\int_{\cal M} d^2x (\bar{c}^a\partial_\mu A^{a\mu }+ \bar{\l}^{a}\partial_\mu B^{a\mu})=\nonumber\\ &=&\int_{\cal M} d^2x[b^a\partial_\mu A^{a\mu}+d^{a}\partial_\mu B^{a\mu}]+ \nonumber\\ &+& \int_{\cal M} d^2x [-\bar{c}^a \partial_\mu (D^\mu c)^a -\bar{\l}^{a} \partial_\mu (D^\mu \l)^a + f^{abc} \bar{\l}^{a} \partial_\mu(c^b B^{c\mu})]\,. \eqan{actiongf} The next step is to promote the two gauge symmetries to the nilpotent and nonlinear BRS-symmetry of the gauge fixed action \begin{eqnarray} sA_\mu^a&=&(D_\mu c)^a\,,\nonumber\\ sB_\mu^{a}&=& (D_\mu \l)^{a}- f^{abc} c^b B_\mu^{c}\,,\nonumber\\ s\phi^a &=& -f^{abc} (c^b \phi^c+\l^{b} X^c)\,,\nonumber\\ sX^{a}&=& -f^{abc} c^b X^{c}\,,\nonumber\\ sc^a&=&-\frac{1}{2}f^{abc} c^b c^c\,,\nonumber\\ s\l^{a}&=&-f^{abc} c^b \l^{c}\,,\nonumber\\ s\bar{c}^a&=& b^a\,,\qquad sb^a=0\,,\nonumber\\ s\bar{\l}^{a}&=& d^{a}\,,\qquad sd^{a}=0\,,\nonumber\\ s^2&=&0\,. \eqan{brst} The canonical dimensions and Faddeev-Popov charges of all fields introduced so far are listed in (Table \ref{fp1}). \begin{table}[h] \begin{center} \begin{tabular}{|l|r|r|r|r|r|r|r|r|r|r|r|} \hline & $A^a_\mu$ & $B^a_\mu$ & $\phi^a$ & $X^a$ & $c^a$ & $\l^a$ & $\bar{c}^a$ & $\bar{\l}^a$ & $b^a$ & $d^a$ & $\partial_\mu$ \\ \hline dim & 1 & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 \\ \hline $\phi\pi$ & 0 & 0 & 0 & 0 & 1 & 1 & -1 & -1 & 0 & 0 & 0\\ \hline \end{tabular}\end{center}\caption{Dimensions and Faddeev-Popov charges of the fields.}\label{fp1} \end{table} \\ Note, that the BRS-exact gauge fixing term introduces a metric. As a consequence the energy-momentum tensor is BRS-exact as well and the model possesses a further symmetry carrying a vectorial index: \begin{equation} \begin{array}{l@{$\,=\,$}ll@{$\,=\,$}l} \d_\mu A_\nu^a & 0\,, & \d_\mu B_\nu^{a} & 0\,, \\ \d_\mu \phi^a & -\varepsilon_{\mu\nu} \partial^\nu\bar{c}^a\,, & \d_\mu X^{a} & -\varepsilon_{\mu\nu} \partial^\nu\bar{\l}^{a}\,,\\ \d_\mu c^a & A_\mu^a \,, & \d_\mu \l^{a} & B_\mu^{a}\,,\\ \d_\mu \bar{c}^a & 0\,, & \d_\mu \bar{\l}^{a} & 0 \,,\\ \d_\mu b^a & \partial_\mu\bar{c}^a\,, & \d_\mu d^{a} & \partial_\mu\bar{\l}^{a} \,.\\ \end{array} \eqn{vsusy} The transformations (\ref{vsusy}) define the so-called vector-like supersymmetry \footnote{We will see in due course that the algebra of $\d_\mu$ with the BRS operator $s$ closes on translations.} \cite{oli}. The invariant action plus the gauge fixing part is indeed invariant under (\ref{vsusy}), \begin{equation} \d_\mu (\S_{inv}+\S_{gf})=0\,. \eqn{invar} Additionally, the symmetries (\ref{brst}) and (\ref{vsusy}) give rise to the following on-shell algebra: \begin{eqnarray} \{s,s\} &=& 0\,,\nonumber\\ {\{s,\d_\mu\}} &=& \partial_\mu + \hbox{equations of motion}\,,\nonumber\\ {\{\d_\mu,\d_\nu\}} &=& 0\,. \eqan{algebra} In order to write down the Slavnov identity, which expresses the symmetry content of the model with respect to BRS, we couple the nonlinear BRS transformations to BRS-invariant external sources leading to \begin{equation} \S_{ext}=\int_{\cal M} d^2x[\O^{\mu a}(s A_\mu^a)+L^a(sc^a)+\r^a(s\phi^a)+ \sigma} \renewcommand{\S}{\Sigma^{\mu a}(sB^a_{\mu})+\L^{a}(s\l^a)+Y^{a}(sX^a)]\,, \eqn{actionext} where $(\O^{\mu a},L^a,\r^a,\sigma} \renewcommand{\S}{\Sigma^{\mu a},\L^{a},Y^{a})$ are the external sources \footnote{Clearly, the presence of these external sources breaks the symmetry (\ref{vsusy}). This fact will be discussed later.} whose canonical dimensions and Faddeev-Popov charges can be read off (Table \ref{fp2}).\\ \begin{table}[h] \begin{center} \begin{tabular}{|l|r|r|r|r|r|r|} \hline & $\O^{\mu a}$ & $L^a$ & $\r^a$ & $\sigma} \renewcommand{\S}{\Sigma^{\mu a}$ & $\L^{a}$ & $Y^{a}$\\ \hline dim & 1 & 2 & 2 & 1 & 2 & 2 \\ \hline $\phi\pi$ & -1 & -2 & -1 & -1 & -2 & -1 \\ \hline \end{tabular}\end{center} \caption{Dimensions and Faddeev-Popov charges of the external sources.} \label{fp2} \end{table} Typically the propagators of massless scalar fields are ill-defined in two space-time dimensions \cite{balasin, olivier2}. In particular, the analysis of the infrared problem connected with the ghost-antighost propagator \begin{equation} \langle \bar{c}^a c^b \rangle = \d^{ab} \frac{1}{k^2}\,, \eqn{x12x} requires the introduction of a regulator mass $m$ \cite{blasi} such that \begin{equation} \langle \bar{c}^a c^b \rangle_m = \d^{ab} \frac{1}{k^2+m^2}\,. \eqn{propc} Keeping the algebraic structure (\ref{algebra}) amounts to adding\footnote{In fact, it would be sufficient to add to the action the quantity \begin{eqnarray} \S_\tau &=& \int_{\cal M} d^2x \left ( -(\t_1+m^2)\bar{c}^a c^a - \t_2(b^ac^a+ \frac{1}{2}f^{abc} \bar{c}^a c^b c^c) \right) \nonumber\\ &=& s\int_{\cal M} d^2x(\t_2 \bar{c}^a c^a )\,, \nonumber \eqan{exp4laining} which guarantees (\ref{propc}) and the BRS invariance of the total action. However, to maintain also the vector supersymmetry (allowing only linear breaking terms in the corresponding Ward identity as well as keeping the algebraic structure (\ref{algebra}), see below) we need the additional expression $\int_{\cal M} d^2x \ s(\tau^\mu_4 \bar c^a A_\mu^a)$ containing the external sources $\tau^\mu_3$ and $\tau^\mu_4$.} the following expression \cite{blasi} to the action \begin{eqnarray} \S_m&=& \int_{\cal M} d^2x \left ( -(\t_1+m^2)\bar{c}^a c^a - \t_2(b^ac^a+ \frac{1}{2}f^{abc} \bar{c}^a c^b c^c) + \t_3^\mu \bar{c}^a A_\mu^a + \right.\nonumber \\ &+& \left.\frac{}{}\t_4^\mu(b^a A_\mu^a - \bar{c}^a(D_\mu c)^a)\right)=\nonumber\\ &=& s\int_{\cal M} d^2x[\t_2 \bar{c}^a c^a + \t_4^\mu \bar{c}^a A_\mu^a]\,, \eqan{actionm} The quantities $(\t_1,\t_2,\t_3^\mu,\t_4^\mu)$ are the new external sources with the following BRS transformation laws \begin{equation} \begin{array}{ll} s\t_2= -(\t_1+m^2), ~~~~&~~~~ s \t_1=0, \\ s\t_4^\mu=\t_3^\mu, ~~~~&~~~~ s \t_3^\mu = 0\,, \\ \end{array} \eqn{brsextm} and with dimensions and Faddeev Popov charges as given in (Table \ref{fp3}). \begin{table}[h] \begin{center} \begin{tabular}{|l|c|c|c|c|} \hline & $\t_1$ & $\t_2$ & $\t_3^\mu$ & $\t_4^\mu$\\ \hline dim & 2 & 2 & 1 & 1 \\ \hline $\phi\pi$ & 0 & -1 & 1 & 0 \\ \hline \end{tabular} \end{center} \caption{Dimensions and Faddeev-Popov charges of the external sources $\tau$.} \label{fp3} \end{table} \noindent The same strategy can be used to regularize the propagator \begin{equation} \langle \bar{\l}^{a} \l^b \rangle = \d^{ab} \frac{1}{k^2}\,. \eqn{yxy} Correspondingly, we have to add the following BRS exact expression to the action \begin{eqnarray} \S_m'&=& \int_{\cal M} d^2x \left ( - ( \tau_{1}+m^2) \bar{\l}^{a}\l^a - \tau_2(d^{a} \l^{a}+ f^{abc} \bar{\l}^{a} c^b \l^c) + \tau_3^\mu \bar{\l}^{a} B_\mu^a + \right.\nonumber \\ &+& \left.\frac{}{}\tau_4^\mu[d^{a} B_\mu^a - \bar{\l}^{a} (D_\mu \l)^a+ f^{abc}\bar{\l}^{a} c^b B_\mu^c]\right)\nonumber\\ &=& s\int_{\cal M} d^2x[\tau_2\bar{\l}^{a} \l^{a} + \tau_4^\mu\bar{\l}^{a} B_\mu^{a}]\, . \eqan{actionM} It follows that the propagator (\ref{yxy}) is regularized according to \begin{equation} \langle \bar{\l}^{a} \l^b \rangle_m = \d^{ab} \frac{1} {k^2+ m^2}\,. \eqn{propl} The total action, which is just the vertex functional at the classical level, becomes \begin{equation} \S^{(0)}=\S_{inv}+\S_{gf}+ \S_{ext} + \S_{{\cal I} {\cal R}}, \eqn{totaction} with \begin{equation} \S_{{\cal I} {\cal R}}= \S_m + \S_m' = s \int_{\cal M} d^2x \left( \t_2 (\bar{c}^ac^a + \bar \l^a \l^a) + \t_4^\mu(\bar{c}^a A^a_\mu + \bar \l^a B^a_\mu) \right). \eqn{ljf898} We are now ready to write down the non-linear functional Slavnov identity corresponding to the BRS invariance of the total action \begin{eqnarray} {\cal S}(\S^{(0)}) &=& \int_{\cal M} d^2x \left( \frac{\d\S^{(0)}}{\d\O^{a\mu}}\frac{\d\S^{(0)}}{\d A_\mu^a}+ \frac{\d\S^{(0)}}{\d L^a}\frac{\d\S^{(0)}}{\d c^a }+ \frac{\d\S^{(0)}}{\d\r^a}\frac{\d\S^{(0)}}{\d\phi^a}+ \frac{\d\S^{(0)}}{\d\sigma} \renewcommand{\S}{\Sigma^{a\mu}}\frac{\d\S^{(0)}}{\d B^a_{\mu}}+ \frac{\d\S^{(0)}}{\d\L^{a}}\frac{\d\S^{(0)}}{\d\l^a}+ \right. \nonumber \\ &+& \left. \frac{\d\S^{(0)}}{\d Y^{a}}\frac{\d\S^{(0)}}{\d X^a}+ b^a \frac{\d\S^{(0)}}{\d \bar{c}^{a}} + d^a \frac{\d\S^{(0)}}{\d \bar{\l}^{a}} -(\t_1+m^2)\frac{\d \S^{(0)}}{\d\t_2}+ \t_3^\mu\frac{\d \S^{(0)}}{\d\t_4^\mu} \right)=0\,. \eqan{slavnovid} For later use, we derive from (\ref{slavnovid}) the linearized version ${\cal S}_{\S^{(0)}}$ of the non-linear BRS-operator: \begin{eqnarray} {{\cal S}_{\S^{(0)}}} &=& \int_{\cal M} d^2x \left(\frac{\d\S^{(0)}}{\d\O^{a\mu}} \frac{\d}{\d A_\mu^a}+ \frac{\d\S^{(0)}}{\d A_\mu^a}\frac{\d}{\d\O^{a\mu}}+ \frac{\d\S^{(0)}}{\d L^a}\frac{\d}{\d c^a }+ \frac{\d\S^{(0)}}{\d c^a }\frac{\d}{\d L^a}+ \frac{\d\S^{(0)}}{\d\r^a}\frac{\d}{\d\phi^a}+ \frac{\d\S^{(0)}}{\d\phi^a}\frac{\d}{\d\r^a}+ \right. \nonumber\\ &+& \frac{\d\S^{(0)}}{\d\sigma} \renewcommand{\S}{\Sigma^{a\mu}}\frac{\d}{\d B^a_{\mu}}+ \frac{\d\S^{(0)}}{\d B^a_{\mu}}\frac{\d}{\d\sigma} \renewcommand{\S}{\Sigma^{a\mu}}+ \frac{\d\S^{(0)}}{\d\L^{a}}\frac{\d}{\d\l^a}+ \frac{\d\S^{(0)}}{\d\l^a}\frac{\d}{\d\L^{a}}+ \frac{\d\S^{(0)}}{\d Y^{a}}\frac{\d}{\d X^a}+ \frac{\d\S^{(0)}}{\d X^a}\frac{\d}{\d Y^{a}}+ \nonumber\\ &+& \left. b^a \frac{\d}{\d \bar{c}^{a}} + d^{a} \frac{\d}{\d \bar{\l}^{a}} -(\t_1+m^2)\frac{\d}{\d\t_2}+\t_3^\mu\frac{\d}{\d\t_4^\mu} \right)\,. \eqan{slavnovop} As already mentioned, the introduction of external sources breaks the vector-like supersymmetry. This fact is expressed by the following broken Ward-identity (WI) for the symmetry (\ref{vsusy}): \begin{equation} {\cal V}_\mu\S^{(0)} = \Delta_\mu^{cl}\,, \eqn{invar2} where ${{\cal V}_\mu}$ is the vector--like supersymmetry Ward operator \begin{eqnarray} {\cal V}_\mu &=& \int_{\cal M} d^2x \Bigg( \varepsilon_{\mu\nu}\r^a \frac{\d}{\d A_\nu^a} -\varepsilon_{\mu\nu}[\O^{\nu a}+\partial^\nu\bar{c}^a -\t_4^\nu\bar{c}^a]\frac{\d}{\d\phi^a} +A_\mu^a\frac{\d}{\d c^a }+(\partial_\mu \bar{c}^a)\frac{\d}{\d b^a} + L^a \frac{\d}{\d \O^{\mu a}} \nonumber \\ &+& B_\mu^a \frac{\d}{\d \l^a} +\varepsilon_{\mu\nu} Y^a \frac{\d}{\d B^a_\nu}- \varepsilon_{\mu\nu} [\sigma} \renewcommand{\S}{\Sigma^{\nu a}+\partial^\nu\bar{\l}^a -\t_4^\nu\bar{\l}^a]\frac{\d}{\d X^a} +(\partial_\mu \bar{\l}^a)\frac{\d}{\d d^a}+ \L^a \frac{\d}{\d \sigma} \renewcommand{\S}{\Sigma^{\mu a}} \nonumber\\ &-& (\partial_\mu\t_2)\frac{\d}{\d\t_1}+[\partial_\mu\t_4^\nu-\d^\nu_\mu(\t_1+m^2)] \frac{\d}{\d\t_3^\nu}-\t_2\frac{\d}{\d\t_4^\mu} \Bigg). \eqan{ward} Fortunately, the breaking term $\Delta_\mu^{cl}$ is linear in the quantum fields and 'insertions' of linear quantum fields are not renormalized by quantum corrections. It is given by \begin{eqnarray} \Delta_\mu^{cl} &=& \int_{\cal M} d^2x \left( L^a \partial_\mu c^a-\r^a \partial_\mu\phi^a-\O^{\nu a} \partial_\mu A_\nu^a-\varepsilon_{\mu\nu} \r^a \partial^\nu b^a + \L^{a}\partial_\mu \l^a\right. \nonumber \\ &-& \left.Y^{a} \partial_\mu X^a -\sigma} \renewcommand{\S}{\Sigma^{\nu a}\partial_\mu B_{\nu}^a- \varepsilon_{\mu\nu} Y^{a}\partial^\nu d^a+\varepsilon_{\mu\nu}\r^a\t_3^\nu\bar{c}^a+\varepsilon_{\mu\nu}\r^a\t_4^\nu b^a \right.\nonumber\\ &+& \left.\frac{}{}\varepsilon_{\mu\nu} Y^{a}\t_3^\nu\bar{\l}^a+\varepsilon_{\mu\nu} Y^a \t^\nu_4 d^a \right)\,. \eqan{break} Furthermore, the total action $\S^{(0)}$ turns out to be constrained by: \\ (i) 2 gauge conditions, \begin{equation} \begin{array}{rcl} \dfrac{\d \S^{(0)}}{\d b^a} &=& \partial^\mu A^a_\mu - \t_2 c^a + \t_4^\mu A^a_\mu, \\ \\[0.1mm] \dfrac{\d \S^{(0)}}{\d d^a} &=& \partial^\mu B^a_\mu - \t_2 \l^a + \t^\mu_4 B^a_\mu. \end{array} \eqn{g.con} (ii) 2 ghost equations, obtained by commuting the gauge conditions with the Slavnov identity \cite{olivier2}, \begin{equation} \begin{array}{rcl} {{\cal G}}_1\S^{(0)}&=&\dfrac{\d \S^{(0)}}{\d \bar c^a} + (\partial^\mu + \t^\mu_4) \dfrac{\d \S^{(0)}}{\d \O^{a\mu}} + \t_2 \dfrac{\d \S^{(0)}}{\d L^a} = -(\t_1 + m^2)c^a - \t^\mu_3 A^a_\mu, \\ \\[0.1mm] {{\cal G}}_2\S^{(0)}&=&\dfrac{\d \S^{(0)}}{\d \bar \l^a} + (\partial^\mu + \t^\mu_4) \dfrac{\d \S^{(0)}}{\d \sigma} \renewcommand{\S}{\Sigma^{a\mu}} + \t_2 \dfrac{\d \S^{(0)}}{\d \L^a} = -(\t_1 + m^2)\l^a - \t^\mu_3 B^a_\mu. \end{array} \eqn{a.eq} \noindent (iii) 2 antighost equations, \\ \begin{equation} \begin{array}{rcl} \bar{{\cal G}}_1^a \S^{(0)} &=& \Delta^a_1, \\ \bar{{\cal G}}_2^a \S^{(0)} &=& \Delta^a_2, \end{array} \eqn{lkiri} where \begin{equation} \bar{{\cal G}}^a_1 = \int_{\cal M} d^2x \bigg( \frac{\d }{\d c^a} - f^{abc} \bar c^b \frac{\d}{\d b^c} - f^{abc} \bar \l^b \frac{\d}{\d d^c} \bigg). \eqn{lksdj1} The corresponding breaking $\Delta^a_1$ is linear in the quantum fields \begin{equation} \Delta^a_1 = \int_{\cal M} d^2x \bigg\lbrace f^{abc} \bigg(- \O^{b\mu} A_\mu^c + L^b c^c - \r^b \phi^c - \sigma} \renewcommand{\S}{\Sigma^{b\mu} B^c_\mu + \L^b \l^c - Y^b X^c \bigg) + (\t_1 + m^2) \bar c^a + \t_2 b^a \bigg\rbrace. \eqn{lksdj2} For the second antighost equation we have \begin{equation} \bar{{\cal G}}_2^a = \int_{\cal M} d^2x \left( \frac{\d}{\d \l^a} - f^{abc} \bar{\l}^b \frac{\d}{\d b^c} \right), \eqn{anti2} such that \begin{equation} \Delta^a_2 = -\int_{\cal M} d^2x \Big( f^{abc}(\r^b X^c - \L^b c^c + \sigma} \renewcommand{\S}{\Sigma^{b\mu}A^c_\mu) - (\t_1 + m^2)\bar{\l}^a - \t_2d^a \Big)\,. \eqn{j45nm} Now, for an arbitrary functional $\Gamma$, depending on the same fields as the total action $\S^{(0)}$, the corresponding linearized Slavnov operator ${\cal S}_\Gamma$, the Ward operator for the vector-like supersymmetry ${\cal V}_\mu$, and the two antighost operators $\bar{{\cal G}}^a_1$, $\bar{{\cal G}}^a_2$ yield the following nonlinear algebra: \begin{equation} \begin{array}{rcl} {\cal S}_\Gamma {\cal S} (\Gamma) &=& 0, \\ {\cal S}_\Gamma ({\cal V}_\mu \Gamma - \Delta_\mu^{cl}) + {\cal V}_\mu {\cal S}(\Gamma) &=& {\cal P}_\mu \Gamma, \\ \left\{ {\cal V}_\mu, {\cal V}_\nu \right\} \Gamma & = & 0, \\ {\cal S}_\Gamma ( \bar {\cal G}^a_1 \Gamma - \Delta^a_1 ) + \bar {\cal G}^a_1 {\cal S}(\Gamma) &=& {\cal H}^a \Gamma, \\ {\cal S}_\Gamma ( \bar {\cal G}^a_2 \Gamma - \Delta^a_2 ) + \bar {\cal G}^a_2 {\cal S}(\Gamma) &=& {\cal K}^a \Gamma, \\ {\cal V}_\mu ( \bar {\cal G}^a_1 \Gamma - \Delta^a_1 ) + \bar {\cal G}^a_1 ( {\cal V}_\mu \Gamma - \Delta^{cl}_\mu ) &=& 0, \\ {\cal V}_\mu ( \bar {\cal G}^a_2 \Gamma - \Delta^a_2 ) + \bar {\cal G}^a_2 ( {\cal V}_\mu \Gamma - \Delta^{cl}_\mu ) &=& 0. \end{array} \eqn{mnybcxgkjaj} ${\cal P}_\mu$ is the Ward operator for translations \begin{equation} {\cal P}_\mu = \int_{\cal M} d^2x \sum_{\Phi_i}{\partial_\mu} \Phi_i \frac{\d}{\d \Phi_i}, \eqn{trans} where $\Phi_i$ represents collectively all the fields introduced so far. For the operator ${\cal H}^a$ we have to consider only the fields possessing a gauge index (represented by $\Theta^a$) such that \begin{equation} {\cal H}^a= \int d^2x \sum_\Theta \left(-f^{abc}\Theta^b\frac{\d} {\d\Theta^c}\right)\,. \eqn{455ahfg9283} The operator ${\cal K}^a$ is given by \begin{eqnarray} {\cal K}^a &=& - \int d^2x f^{abc} \Big(X^b\frac{\d}{\d\phi^c} + A_\mu^b\frac{\d}{\d B_\mu} + d^b\frac{\d}{\d b^c}+c^b\frac{\d}{\d\l^c}+\bar{\l}^b\frac{\d}{\d\bar{c}^c} \nonumber\\ &+& \sigma} \renewcommand{\S}{\Sigma^{\mu b}\frac{\d}{\d\Omega^{\mu c}}+\rho^b\frac{\d}{\d Y^c}+ \L^b\frac{\d}{\d L^c}\Big)\,. \eqan{kdfj543kj} Now, it is easy to check that the classical action is invariant under the symmetries expressed by ${\cal H}^a$ and ${\cal K}^a$, i.e., \begin{equation} {\cal H}^a \S^{(0)} = {\cal K}^a \S^{(0)} = 0\,. \eqn{ja7t843kjsd} If the functional $\S^{(0)}$ is a solution of the Slavnov identity, ${\cal S}(\S^{(0)})=0$, of the Ward identities (\ref{invar2}) and (\ref{ja7t843kjsd}), and of the two antighost equations (\ref{lkiri}), then, from the above nonlinear algebra, we get the following linear algebra: \begin{equation} \label{algebraend} \begin{array}{rcl} \left\{ {\cal S}_{\S^{(0)}}, {\cal S}_{\S^{(0)}} \right\} & = & 0, \\ \left\{ {\cal S}_{\S^{(0)}}, {\cal V}_\mu \right\} & = & {\cal P}_\mu, \\ \left\{ {\cal V}_\nu, {\cal V}_\mu \right\} & = & 0, \\ \left\{ {\cal S}_{\S^{(0)}}, \bar {\cal G}^a_1 \right\} & = & {\cal H}^a, \\ \left\{ {\cal S}_{\S^{(0)}}, \bar {\cal G}^a_2 \right\} & = & {\cal K}^a, \\ \left\{ {\cal V}_\mu, \bar {\cal G}^a_1 \right\} & = & 0, \\ \left\{ {\cal V}_\mu, \bar {\cal G}^a_2 \right\} & = & 0, \\ \end{array} \end{equation} So far we have regularized the infrared divergent propagators and analyzed the symmetries of the model as well as derived the constraints which the total action obeys. In the remaining part of the paper we will extend our analysis to the quantum level. \section{Search for counterterms} We devote this section to the discussion of the stability problem which amounts to analyze all possible invariant counterterms for the total action. In a first step one has to modify the classical action as \begin{equation} \S'=\S^{(0)}+\Delta, \eqn{huiz} where $\Delta$ stands for appropriate invariant counterterms. The total classical action $\S^{(0)}$ is a solution of the Slavnov identity (\ref{slavnovid}), the vector--like supersymmetry WI (\ref{invar2}), the two gauge conditions (\ref{g.con}), the two ghost equations (\ref{a.eq}), and the two antighost equations (\ref{lkiri}) as well as the two Ward identities (\ref{ja7t843kjsd}). The perturbation $\Delta$ is an integrated, local and Lorentz invariant polynomial of dimension 2 and vanishing ghost number. \\ By studying the stability of the theory we are looking for the most general deformation of the classical action such that the functional $\S'$ still obeys all the constraints listed above. Then the quantity $\Delta$ is subject to the following set of constraints \begin{equation} \dfrac{\d \Delta}{\d b^a} = 0, \eqn{cstr1} \begin{equation} \dfrac{\d \Delta}{\d d^a} = 0, \eqn{cstr2} \begin{equation} \dfrac{\d \Delta}{\d \bar c^a} + (\partial^\mu + \t^\mu_4) \dfrac{\d \Delta}{\d \O^{a\mu}} + \t_2 \dfrac{\d \Delta}{\d L^a} = 0, \eqn{cstr3} \begin{equation} \dfrac{\d \Delta}{\d \bar \l^a} + (\partial^\mu + \t^\mu_4) \dfrac{\d \Delta}{\d \sigma} \renewcommand{\S}{\Sigma^{a\mu}} + \t_2 \dfrac{\d \Delta}{\d \L^a} = 0, \eqn{cstr4} \begin{equation} {\cal S}_{\S^{(0)}} \Delta = 0, \eqn{cstr5} \begin{equation} {\cal V}_\mu \Delta = 0, \eqn{cstr6} \begin{equation} {\cal P}_\mu\Delta=0\,, \eqn{cstr9} \begin{equation} \int_{\cal M} d^2x \frac{\d \Delta}{\d c^a} = 0, \eqn{cstr7} \begin{equation} \int_{\cal M} d^2x \frac{\d \Delta}{\d \l^a} = 0, \eqn{cstr8} \begin{equation} {\cal H}^a \Delta = 0, \eqn{kjaabf3452nb} \begin{equation} {\cal K}^a \Delta = 0. \eqn{jadhhh345b8f} The first two equations (\ref{cstr1}) and (\ref{cstr2}) signify that $\Delta$ is independent of the two Lagrange multiplier fields $b^a$ and $d^a$. The equation (\ref{cstr3}) implies that $\Delta$ depends on the fields $\O^{a\mu}$, $\bar{c}^a$ and $L^a$ only through the two combinations \begin{equation} \begin{array}{rcl} \tilde \O^{a\mu} &=& \O^{a\mu} + \partial^\mu \bar c^a - \t^\mu_4 \bar c^a, \\ \tilde L^a &=& L^a + \t_2 \bar c^a\,. \end{array} \eqn{comb1} Correspondingly, we deduce from equation (\ref{cstr4}) that the fields $\sigma} \renewcommand{\S}{\Sigma^{a\mu}$, $\bar{\l}^a$ and $\L^a$ appear in the expression of $\Delta$ only through the two combinations \begin{equation} \begin{array}{rcl} \tilde \sigma} \renewcommand{\S}{\Sigma^{a\mu} &=& \sigma} \renewcommand{\S}{\Sigma^{a\mu} + \partial^\mu \bar \l^a - \t^\mu_4 \bar \l^a, \\ \tilde \L^a &=& \L^a + \t_2 \bar \l^a. \end{array} \eqn{comb2} Finally, the constraints (\ref{cstr5}), (\ref{cstr6}) and (\ref{cstr9}) may be combined in a single Ward-operator $\d$, \begin{equation} \d = {\cal S}_{\S^{(0)}} + \xi^\mu {\cal V}_\mu + \varepsilon^\mu {\cal P}_\mu - \int_{\cal M} d^2x ~ \xi^\mu \frac{\partial }{\partial \varepsilon^\mu}. \eqn{delta} It is easy to check the nilpotency of the above defined operator $\d$. The vectors $\xi^\mu$ and $\varepsilon^\mu$ are constant vectors of ghost numbers $+2$ and $+1$, respectively. Clearly, we get \begin{equation} \d \Delta = 0, \eqn{co} which constitutes a cohomology problem. The nilpotency property of the operator $\d$ implies immediately that any expression of the form $\d \hat \Delta$ is a solution of (\ref{co}), where $\hat \Delta$ is a local integrated polynomial of dimension 2 and ghost number $-1$. Therefore, the general solution of (\ref{co}) is of the form \begin{equation} \Delta = \Delta_c + \d \hat \Delta\,, \eqn{mxcipwe345} where $\Delta_c$ is the nontrivial solution whereas $\d \hat \Delta$ is called the trivial solution. \\ The form of the trivial counterterm is restricted by dimension and ghost number requirements. Since the fields $(\phi, X)$ both have dimension and ghost number zero, an arbitrary combination of these fields may appear many times in the counterterm. We denote the most general and possible combination of $(\phi, X)$ by $f_\alpha[\phi,X]$, such that \begin{equation} \label{falpha} f_\alpha[\phi,X]=\sum_{\{n_i\},\{m_i\}=0}^\infty \b^\a_{n_i,m_i}\Bigg( \prod_{i=0}^\infty \phi^{n_{i}} X^{m_{i}}\Bigg)\,, \end{equation} where $\{n_i\}$ and $\{m_i\}$ are understood as $\{n_0,n_1,\dots\}$ and $\{m_0,m_1,\dots\}$, respectively. $\b^\a_{n_i,m_i}$ are, of course, constant coefficients to be determined. Actually, the most general trivial counterterm $\d\hat{\Delta}$ reads \begin{eqnarray}\label{counter} \d \hat{\Delta} &=& \d \int_{\cal M} d^2 x \hbox{Tr} \left(\frac{}{} \rho f_1 + Y f_2 + \t_2 f_3 + \tilde{\O}^\nu f_4 A_\nu f_5+ \varepsilon_{\mu\nu}\tilde{\O}^\mu f_6 A^\nu f_7+ \tilde{\sigma} \renewcommand{\S}{\Sigma}^\nu f_8 A_\nu f_9 \right.\nonumber\\ &+& \varepsilon_{\mu\nu}\tilde{\sigma} \renewcommand{\S}{\Sigma}^\mu f_{10} A^\nu f_{11} + \tilde{\O}^\nu f_{12} B_\nu f_{13} + \varepsilon_{\mu\nu}\tilde{\O}^\mu f_{14} B^\nu f_{15} + \tilde{\sigma} \renewcommand{\S}{\Sigma}^\nu f_{16} B_\nu f_{17} \nonumber\\ &+& \varepsilon_{\mu\nu}\tilde{\sigma} \renewcommand{\S}{\Sigma}^\mu f_{18} B^\nu f_{19}+ (\partial^\nu \tilde{\O}_\nu) f_{20} + \varepsilon_{\mu\nu}(\partial^\mu \tilde{\O}^\nu) f_{21} + (\partial^\nu \tilde{\sigma} \renewcommand{\S}{\Sigma}_\nu) f_{22} + \varepsilon_{\mu\nu} (\partial^\mu \tilde{\sigma} \renewcommand{\S}{\Sigma}^\nu) f_{23} \nonumber\\ &+& \t_4^\nu f_{24}\tilde{\O}_\nu f_{25} + \varepsilon_{\mu\nu}\t_4^\mu f_{26}\tilde{\O}^\nu f_{27} + \t_4^\nu f_{28} \tilde{\sigma} \renewcommand{\S}{\Sigma}_\nu f_{29} + \varepsilon_{\mu\nu}\t_4^\mu f_{30}\tilde{\sigma} \renewcommand{\S}{\Sigma}^\nu f_{31} \nonumber\\ &+& \left.\frac{}{} \tilde{L}f_{32} c f_{33} + \tilde{\L}f_{34} c f_{35} + \tilde{L}f_{36} \l f_{37} + \tilde{\L} f_{38}\l f_{39} \right) \,. \end{eqnarray} In fact the expression (\ref{counter}) may depend on the vector parameters $\xi^\mu$ and $\varepsilon^\mu$ which are not present in the total action (\ref{totaction}). For this reason we have to arrange for the trivial counterterm to be independent\footnote{ Our aim is to renormalize the theory defined by (\ref{totaction}) which is independent of $\xi^\mu$ and $\varepsilon^\mu$.} of these two constant vectors. In other words, we require that $\hat \Delta$ has to be invariant under the vector--like supersymmetry and translation Ward operators. A lengthy and detailed analysis results in the expression for the trivial counterterm given below: \begin{eqnarray} \label{nmxc54nmf9ksd} {\cal S}_{\S^{(0)}} {\cal C}_t&=& {\cal S}_{\S^{(0)}} \int_{\cal M} d^2 x \hbox{Tr} \left(\frac{}{} \b^1\,(\tilde{\O}^\nu A_\nu+\rho\phi-\tilde{L}c)+\b^2\, (\tilde{\sigma} \renewcommand{\S}{\Sigma}^\nu A_\nu+Y\phi-\tilde{\L}c) ~+ \right.\nonumber\\ &+& \left.\b^3\,(\tilde{\O}^\nu B_\nu+\rho X-\tilde{L}\l)+ \b^4\,(\tilde{\sigma} \renewcommand{\S}{\Sigma}^\nu B_\nu+Y X -\tilde{\L}\l)\right)\,, \end{eqnarray} where the $\b^i$, $i=1,...,4$ are arbitrary constants. \\ Now we turn to the computation of the nontrivial counterterms $\Delta_c$ in (\ref{mxcipwe345}). In a first step we introduce a filtering operator ${\cal N}$ such that \begin{equation} {\cal N} = \int_{\cal M} d^2x \Bigg( \sum_f f \frac{\d}{\d f} \Bigg), \eqn{filter} where we have assigned to each field homogeneity degree one. The quantity $f$ in (\ref{filter}) stands for all fields (including also $\varepsilon^\mu$ and $\xi^\mu$). The operator ${\cal N}$ leads to the decomposition of the operator $\d$ as \begin{equation} \d = \d_0 + \d_1\,. \eqn{decomp} The nilpotency of the operator $\d$ implies now that \begin{equation} \d_0^2 = \lbrace \d_0,\ \d_1 \rbrace = \d_1^2 = 0. \eqn{d0nilpot} The operator $\d_0$ does not increase the homogeneity degree, whereas $\d_1$ increases the homogeneity degree by one unit. Now it is evident that from $\d \Delta = 0$ we get \begin{equation} \d_0 \Delta = 0, \eqn{co2} where \begin{equation}\begin{array}{rcl} \d_0 &=& \displaystyle{\int}_{\cal M} \left( d c^a \frac{\d }{\d A^a} + d A^a \frac{\d } {\d \r^a} + d \phi^a \frac{\d }{\d \hat \O^a} + d \hat \O^a \frac{\d }{\d \hat L^a} \right) + \nonumber\\ \\ &+& \displaystyle{\int}_{\cal M} \left( d \l^a \frac{\d }{\d B^a} + d B^a \frac{\d } {\d \hat Y^a} + d X^a \frac{\d }{\d \hat \sigma} \renewcommand{\S}{\Sigma^a} + d \hat \sigma} \renewcommand{\S}{\Sigma^a \frac{\d }{\d \hat \L^a} \right) + \nonumber\\ \\ &+& \displaystyle{\int}_{\cal M} d^2 x \bigg( - \t_1 \frac{\d }{\d \t_2} + \t^\mu_3 \frac{\d }{\d \t^\mu_4} - \xi^\mu \frac{\d }{\d \varepsilon^\mu} \bigg)\,. \eqacn{delta0} The first two parts of the expression of $\d_0$ are given in terms of forms where, \begin{equation} \begin{array}{rcl} A^a &=& A_\mu^a dx^\mu , \\ \hat \O^a &=& \varepsilon_{\mu\nu} {\tilde\Omega}^{a\mu} dx^\nu , \\ \hat L^a &=& \frac{1}{2} \varepsilon_{\mu\nu} \tilde L^a dx^\mu dx^\nu , \\ \hat \r^a &=& \frac{1}{2} \varepsilon_{\mu\nu} \r^a dx^\mu dx^\nu , \\ B^a &=& B_\mu^a dx^\mu , \\ \hat \sigma} \renewcommand{\S}{\Sigma^a &=& \varepsilon_{\mu\nu} \tilde \sigma} \renewcommand{\S}{\Sigma^{a\mu} dx^\nu , \\ \hat \L^a &=& \frac{1}{2} \varepsilon_{\mu\nu} \tilde \L^a dx^\mu dx^\nu , \\ \hat Y^a &=& \frac{1}{2} \varepsilon_{\mu\nu} Y^a dx^\mu dx^\nu , \\ \end{array} \eqn{forms} and $d$ is the exterior derivative $d = dx^\mu {\partial_\mu}$.\\ By looking to the expression (\ref{delta0}) one easily recognizes that the fields $\t_1, \t_2, \t^\mu_3, \t^\mu_4, \varepsilon^\mu$ and $\xi^\mu$ transform as $\d_0$ doublets which means that they are out of the cohomology of $\d_0$ \cite{brandt}. In order to solve the cohomology problem (\ref{co2}) we write $\Delta$ as an integrated local polynomial: \begin{equation} \Delta=\int_{\cal M} f^0_2, \eqn{z123z} such that $f^0_2$ has form degree 2 and ghost number $0$. The use of Stoke's theorem, the algebraic Poincare lemma \cite{brandt} and the anticommutator relation $\lbrace \d_0, ~ d \rbrace =0$ lead to the set of descent equations \begin{equation} \begin{array}{rcl} \d_0 f^0_2 + d f^1_1 &=& 0, \\ \d_0 f^1_1 + d f^2_0 &=& 0, \\ \d_0 f^2_0 &=& 0. \end{array} \eqn{ab.c} Due to dimension and ghost number requirements the most general expression for $f^2_0$ is given by \begin{equation} f_0^2 = \sum_{n_{ij},m_{ij}, \atop l_{pq},k_{pq}=0}^\infty ~\sum_{j,q=1}^\infty \a_{jq} Tr \bigg \lbrack ~c \prod_{i=1}^\infty (\phi^{n_{ij}} X^{m_{ij}}) ~c \prod_{p=1}^\infty (\phi^{l_{pq}} X^{k_{pq}}) \bigg \rbrack ~~+ \eqn{solct} \[ ~~~ + \sum_{\hat n_{ij},\hat m_{ij}, \atop \hat l_{pq},\hat k_{pq}=0}^\infty ~\sum_{j,q=1}^\infty \b_{jq} Tr \bigg \lbrack ~c \prod_{i=1}^\infty (\phi^{\hat n_{ij}} X^{\hat m_{ij}}) ~\l \prod_{p=1}^\infty (\phi^{\hat l_{pq}} X^{\hat k_{pq}}) \bigg \rbrack ~~+ \] \[ ~~~ + \sum_{\bar n_{ij},\bar m_{ij}, \atop \bar l_{pq},\bar k_{pq}=0}^\infty ~\sum_{j,q=1}^\infty \gamma_{jq} Tr \bigg \lbrack ~\l \prod_{i=1}^\infty (\phi^{\bar n_{ij}} X^{\bar m_{ij}}) ~\l \prod_{p=1}^\infty (\phi^{\bar l_{pq}} X^{\bar k_{pq}}) \bigg \rbrack ~~. \] $\a_{jq},\b_{jq}$ and $\gamma_{jq}$ stand for constant and field independent coefficients. The upper indices of the fields $\phi$ and $X$ are just integer exponents required by locality. To solve the descent equations (\ref{ab.c}) we follow the same strategy as in \cite{oli}. We define the operator \begin{equation} \bar \d_0 = \displaystyle{\int}_{\cal M} \Big( 2 \hat \r \frac{\d}{\d A} + A \frac{\d}{\d c} + \hat \O \frac{\d}{\d \phi} + 2 \hat L \frac{\d}{\d \hat \O} +2 \hat Y \frac{\d}{\d B} + B \frac{\d}{\d \l} + \hat \sigma} \renewcommand{\S}{\Sigma \frac{\d}{\d X} + 2 \hat \L \frac{\d}{\d \hat \sigma} \renewcommand{\S}{\Sigma} \Big), \eqn{843ghfgz8} which, when commuted with the operator $\d_0$, gives translations \begin{equation} \lbrack \d_0 , \bar \d_0 \rbrack = \int_{\cal M} \Big( d \Psi \frac{\d}{\d \Psi} \Big)\,. \eqn{aksdh4kjhasd8} Recall that we are now working in the space generated by the fields which belong to the cohomology of $\d_0$. These fields are denoted by $\Psi$. In other words, the operator $\d_0$ appearing in equation (\ref{aksdh4kjhasd8}) is restricted to this space where the $\d_0$ doublets are absent.\\ One can easily show \cite{oli} that the solution of the descent equations (\ref{ab.c}) is given by \begin{equation} f^0_2 = \frac{1}{2} \bar \d_0 \bar \d_0 f^2_0 . \eqn{solsol234} $\displaystyle{\int_{\cal M}} f^0_2~$ is then the nontrivial solution of the cohomology problem (\ref{co2}). A direct investigation shows that each monomial in (\ref{solct}) leads (after applying on it $\bar{\d_0}^2$) to a polynomial depending on the ghost fields $c$ and $\l$. But this is forbidden by the two constraints (\ref{cstr7}) and (\ref{cstr8}) which are valid at each homogeneity degree. In other words the nontrivial solution of the $\d_0$ cohomology in the space constrained by (\ref{cstr7}) and (\ref{cstr8}) is zero. From this we deduce\footnote{ Recall \cite{brandt} \cite{olivier2} that the cohomology of the nilpotent operator $\d$, that is the space of all nontrivial solutions of (\ref{co}), is isomorphic to a subspace of the cohomology of $\d_0$.} that the cohomology of the whole operator $\d$ is empty. So, the most general solution of (\ref{co}) takes the form \begin{equation} \Delta = {\cal S}_{\S^{(0)}} {\cal C}_t \, . \eqn{kafhgi342z8df} The restriction coming from the two antighost equations (\ref{cstr7}) and (\ref{cstr8}) eventually implies the vanishing of all constant coefficients in (\ref{nmxc54nmf9ksd}). \\ Thus we have shown that the constraint system (\ref{cstr1}--\ref{cstr8}) forbids any deformations of the classical action. Furthermore, if the symmetries of the model are anomaly free, then the symmetry content of the model at the classical level is also valid in the presence of quantum corrections. The analysis of anomalies is the subject of the next section. \section{Search for anomalies} In order to describe possible breaking of the symmetries which characterize the model, one has to apply the quantum action principle (QAP) \cite{olivier2}. The latter allows to describe symmetry breaking in the following way: \begin{equation} \d\Gamma=\tilde{\Delta} \, , \eqn{symbr} where $\Gamma$ is the full vertex functional given by a power series in $\hbar$. The QAP requires that the breaking $\tilde{\Delta}$ is a local, integrated, Lorentz-invariant polynomial of dimension 2 and ghost number 1. The nilpotency of $\d$ leads again to a cohomology problem \begin{equation} \d\tilde{\Delta}=0\,, \eqn{cohpr} which implies the solution, \begin{equation} \tilde\Delta = \d\tilde{\tilde{\Delta}}+\AA\,, \eqn{sllsdkjaekrihtpoe} where $\AA\not=\d\tilde{\AA}$. The anomaly candidate $\AA$, as a solution of (\ref{cohpr}), has to obey\footnote{ A non trivial statement is that the anomaly candidate has to fulfill the antighost equations. Due to the algebra (\ref{algebraend}) we deduce \begin{equation} \begin{array}{rcl} \lbrace \d, \bar {\cal G}^a_1 \rbrace &=& {\cal H}^a, \\ \lbrace \d, \bar {\cal G}^a_2 \rbrace &=& {\cal K}^a. \end{array} \eqn{kfg9g9fido9} Furthermore, the quantum generating functional of vertex functions $\Gamma = \S^{(0)} + {\cal O} (\hbar)$ fulfills the two antighost equations and the two WI's (\ref{ja7t843kjsd}), see below. This fact together with (\ref{symbr}) and (\ref{kfg9g9fido9}) shows that the anomaly candidate must obey the antighost equations.} the constraints (\ref{cstr1}) --- (\ref{cstr4}) as well as (\ref{cstr7}) and (\ref{cstr8}). In other words we have to solve the cohomology problem (\ref{cohpr}) in the same space of functionals as the problem (\ref{co}), Hence, $\AA$ depends only on the fields: $A^a$, $c^a$, $\r^a$, ${\varphi}^a$, $\hat \O^a$, $\hat L^a$, $\l^a$, $B^a$, $\hat Y^a$, $X^a$, $\hat \sigma} \renewcommand{\S}{\Sigma^a$ and $\hat \L^a$. In terms of forms the functional $\AA$ is a local integrated polynomial of form degree 2 and ghost number 1 \begin{equation} \AA = \int_{\cal M} f^1_2. \eqn{lksdfh33} By using the strategy of the previous section we get the following set of descent equations \begin{equation} \begin{array}{rcl} \d_0 f^1_2 + d f^2_1 &=& 0, \\ \d_0 f^2_1 + d f^3_0 &=& 0, \\ \d_0 f^3_0 &=& 0. \end{array} \eqn{ab.a} The last equation in (\ref{ab.a}) has the general solution \begin{equation} f_0^3 = \sum_{n_{ij},m_{ij},l_{pq}, \atop k_{pq},t_{gy},h_{gy}=0}^\infty ~\sum_{j,q,y=1}^\infty \a_{jqy} Tr \bigg \lbrack ~c \prod_{i=1}^\infty (\phi^{n_{ij}} X^{m_{ij}}) ~c \prod_{p=1}^\infty (\phi^{l_{pq}} X^{k_{pq}}) ~c \prod_{g=1}^\infty (\phi^{t_{gy}} X^{h_{gy}}) \bigg \rbrack ~~+ \eqn{solan} \[ ~~~ + \sum_{\hat n_{ij},\hat m_{ij},\hat l_{pq}, \atop \hat k_{pq},\hat t_{gy},\hat h_{gy}=0}^\infty ~\sum_{j,q,y=1}^\infty \b_{jqy} Tr \bigg \lbrack ~c \prod_{i=1}^\infty (\phi^{\hat n_{ij}} X^{\hat m_{ij}}) ~c \prod_{p=1}^\infty (\phi^{\hat l_{pq}} X^{\hat k_{pq}}) ~\l \prod_{g=1}^\infty (\phi^{\hat t_{gy}} X^{\hat h_{gy}}) \bigg \rbrack ~~+ \] \[ ~~~ + \sum_{\bar n_{ij},\bar m_{ij},\bar l_{pq}, \atop \bar k_{pq},\bar t_{gy},\bar h_{gy}=0}^\infty ~\sum_{j,q,y=1}^\infty \gamma_{jqy} Tr \bigg \lbrack ~c \prod_{i=1}^\infty (\phi^{\bar n_{ij}} X^{\bar m_{ij}}) ~\l \prod_{p=1}^\infty (\phi^{\bar l_{pq}} X^{\bar k_{pq}}) ~\l \prod_{g=1}^\infty (\phi^{\bar t_{gy}} X^{\bar h_{gy}}) \bigg \rbrack ~~+ \] \[ ~~ +\sum_{\tilde n_{ij},\tilde m_{ij},\tilde l_{pq}, \atop \tilde k_{pq},\tilde t_{gy},\tilde h_{gy}=0}^\infty ~\sum_{j,q,y=1}^\infty \pi_{jqy} Tr \bigg \lbrack ~\l \prod_{i=1}^\infty (\phi^{\tilde n_{ij}} X^{\tilde m_{ij}}) ~\l \prod_{p=1}^\infty (\phi^{\tilde l_{pq}} X^{\tilde k_{pq}}) ~\l \prod_{g=1}^\infty (\phi^{\tilde t_{gy}} X^{\tilde h_{gy}}) \bigg \rbrack, \] where the quantities $\a_{jqy},\b_{jqy},\gamma_{jqy}$ and $\pi_{jqy}$ are constant coefficients.\\ By using the same arguments as before (see the last section) one can prove that all constant coefficients appearing in (\ref{solan}) must vanish. Of course, this is due to the constraints (\ref{cstr7}) and (\ref{cstr8}). Let us, for instance consider the special case where \begin{equation} f_0^3 = \a Tr(c)^3 + \b Tr(\l)^3 + \gamma Tr(c^2 \l), \eqn{bla1} It leads to the anomaly candidate \begin{equation} \AA = Tr \int_{\cal M} \bigg( 3\a(\hat \r c^2 + A^2 c) + 3~\b(\hat Y \l^2 + B^2 \l) + \gamma ( \hat \r \lbrace c,~ \l \rbrace + \lbrace c,~ A \rbrace B + A^2 \l + \hat Y c^2 ) \bigg), \eqn{aaa} where $\a$, $\b$ and $\gamma$ are constant coefficients.\\ However, it is easy to verify that this anomaly candidate (\ref{aaa}) does not obey the two antighost equations (\ref{cstr7}) and (\ref{cstr8}) unless $\a=\b=\gamma=0$. This means in particular that the nontrivial solution of (\ref{cohpr}) is zero. Then, the Slavnov identity as well as the WI for the vector-like supersymmetry transformations are anomaly free. Hence, they are valid at the full quantum level. Concerning the two antighost equations, one can prove their validity at the quantum level by simply following the arguments of \cite{bps}. Furthermore, the two gauge conditions, the two ghost equations and the two WI's (\ref{ja7t843kjsd}) are also valid at the quantum level. This can be proven by simply using the strategy of \cite{olivier2}. \\ As a conclusion, the model we analyzed in this paper is anomaly free and ultraviolet as well as infrared finite at all orders of perturbation theory.
\section{Introduction} Let $X$ be a compact connected Riemann surface of genus $g$, with $g\geq 2$. Let ${\cal M}_{\xi} := {\cal M}(n,\xi)$ denote the moduli space of stable vector bundles $E$ of rank $n$, with $n\geq 2$, over $X$, such that the line bundle ${\bigwedge}^n E$ is isomorphic to a fixed holomorphic line bundle $\xi$ over $X$. The degree $d= \mbox{deg}(\xi)$ and $n$ are assumed to be coprime. We also assume that if $g=2$, then $n\neq 2,3$, and if $g=3$, then $n\neq 2$. The moduli space ${\cal M}_{\xi}$ is a connected smooth projective variety over $\Bbb C$, and for fixed $n$, the moduli space ${\cal M}_{\xi}$ is isomorphic to ${\cal M}_{{\xi}'}$ if ${\xi}'$ is another holomorphic line bundle with $\mbox{deg}(\xi) = \mbox{deg} ({\xi}')$. We take $\xi$ to be of the form $L^{\otimes d}$, where $L$ is a holomorphic line bundle over $X$ such that $L^{\otimes (2g-2)}$ is isomorphic to the canonical line bundle $K_X$. The Picard group $\mbox{Pic}({\cal M}_{\xi})$ is isomorphic to $\Bbb Z$. The anticanonical line bundle $K^{-1}_{{\cal M}_{\xi}}$ is isomorphic to ${\Theta}^{\otimes 2}$, where $\Theta$ is the ample generator of $\mbox{Pic}({\cal M}_{\xi})$, known as the {\it generalized theta line bundle}. Let $D$ be a smooth divisor on ${\cal M}_{\xi}$ such that the holomorphic line bundle ${\cal O}_{{\cal M}_{\xi}}(D)$ over ${\cal M}_{\xi}$ is isomorphic to $K^{-1}_{{\cal M}_{\xi}}$. Such a divisor is a connected simply connected smooth projective variety with trivial canonical line bundle. In other words, $D$ is a Calabi-Yau variety. If we move the triplet $(X,L,D)$, in the space of all triplets $(X',L',D')$, where $D'$ is a smooth Calabi-Yau hypersurface on a moduli space of stable vector bundles, of the above type, over $X'$, then we get deformations of the complex manifold $D$, simply by associating the complex manifold $D'$ to any triplet $(X',L' ,D')$. The Kodaira-Spencer infinitesimal deformation map for this family gives a homomorphism from the tangent space of the moduli space of triplets $(X,L,D)$, of the above type, into $H^1(D,\, T_D)$, the space parametrizing the infinitesimal deformations of the complex manifold $D$. The main result here, [Theorem 3.3], says \medskip {\it The above Kodaira-Spencer infinitesimal deformation map is an isomorphism.} \medskip Consequently, there is an exact sequence $$ 0 \, \longrightarrow \, \mbox{Hom}(l, H^0 ({\cal M}_{\xi},\, K^{-1}_{{\cal M}_{\xi}})/l) \, \longrightarrow \, H^1(D,\, T_D) \, \longrightarrow \, H^1(X,\, T_X) \, \longrightarrow \, 0 \, , \leqno{(1.1)} $$ where $l \subset H^0 ({\cal M}_{\xi},\, K^{-1}_{{\cal M}_{\xi}})$ is the one dimensional subspace defined by $D$. The above inclusion map $$ \mbox{Hom}(l, H^0 ({\cal M}_{\xi},\, K^{-1}_{{\cal M}_{\xi}})/l) \, \longrightarrow \, H^1(D,\, T_D) $$ corresponds to the deformations of $D$ obtained by moving the hypersurface of the {\it fixed} variety ${\cal M}_{\xi}$, i.e., $X$ is kept fixed, and the projection $H^1(D,\, T_D) \longrightarrow H^1(X,\, T_X)$ in (1.1) is the forgetful map from the space of infinitesimal deformations of the triplet $(X,L,D)$ to the space of infinitesimal deformations of $X$. From the above exact sequence (1.1) it follows immediately that $$ \dim H^1(D,\, T_D) \, = \, 3g-4 \, + \, \dim H^0({\cal M}_{\xi}, \, K^{-1}_{{\cal M}_{\xi}}) \, . $$ We note that the dimension of any $H^0({\cal M}_{\xi}, \, {\Theta}^{\otimes k})$, in particular that of $H^0({\cal M}_{\xi}, \, K^{-1}_{{\cal M}_{\xi}})$, is given by the Verlinde formula. Let ${\cal U}_D$ denote the restriction to $X\times D$ of a Poincar\'e vector bundle over $X\times {\cal M}_{\xi}$. For any $x\in X$, the vector bundle over $D$, obtained by restricting ${\cal U}_D$ to $x\times D$, is denoted by $({\cal U}_D)_x$. The following result is used in the proof of Theorem 3.3. \medskip {\it For any $x\in X$, the vector bundle $({\cal U}_D)_x$ is stable with respect to any polarization on $D$. Moreover, the infinitesimal deformation map $$ T_xX \, \longrightarrow \, H^1(D,\, \mbox{Ad}(({\cal U}_D))_x)\, , $$ for the family ${\cal U}_D$ of vector bundles over $D$ parametrized by $X$, is an isomorphism.} \medskip This result was proved in \cite[Theorem 2.5]{BB} under the assumption that $n\geq 3$. Here it is extended to the rank two case [Theorem 2.1]. \section{Restriction of the universal vector bundle} We continue with the notation of the introduction. The anticanonical line bundle $K^{-1}_{{\cal M}_{\xi}} := {\bigwedge}^{\rm top} T_{{\cal M}_{\xi}}$ is isomorphic to ${\Theta}^{\otimes 2}$ \cite[page 69, Theorem 1]{Ra}, where the generalized theta line bundle $\Theta$ is the ample generator of the Picard group $\mbox{Pic}({\cal M}_{\xi})$; the Picard group is isomorphic to ${\Bbb Z}$. Let $D \subset {\cal M}_{\xi}$ be a smooth divisor, satisfying the condition that the line bundle ${\cal O}_{{\cal M}_{\xi}}(D)$ is isomorphic to $K^{-1}_{{\cal M}_{\xi}}$. Let $$ {\tau} \, : \hspace{.1in} D \hspace{.1in} \longrightarrow \hspace{.1in} {\cal M}_{\xi} $$ denote the inclusion map. Using the Poincar\'e adjunction formula, we have $K_D \, \cong \, {\tau}^* K_{{\cal M}_{\xi}}\bigotimes {\tau}^* {\cal O}_{{\cal M}_{\xi}} (D)$. In view of the assumption ${\cal O}_{{\cal M}_{\xi}}(D) \cong K^{-1}_{{\cal M}_{\xi}}$, the canonical line bundle $K_D$ is trivial. Since the divisor $D$ is ample, it is connected. Since the moduli space ${\cal M}_{\xi}$ is simply connected, the divisor $D$ is also simply connected. Therefore, $D$ is a Calabi-Yau variety. Fix a Poincar\'e vector bundle $\cal U$ over $X\times {\cal M}_{\xi}$. In other words, for any $m\in {\cal M}_{\xi}$, the vector bundle over $X$ obtained by restricting $\cal U$ to $X\times m$ is represented by the point $m$. Let ${\rm Ad}({\cal U})$ denote the rank $n^2-1$ vector bundle over $X\times {\cal M}_{\xi}$ defined by the trace zero endomorphisms of ${\cal U}$. The vector bundle $(\mbox{Id}_X \times {\tau})^* {\cal U}$ (respectively, $(\mbox{Id}_X \times {\tau})^* {\rm Ad}({\cal U})$) over $X\times D$ will be denoted by ${\cal U}_D$ (respectively, ${\rm Ad}({\cal U}_D)$). For any fixed $x\in X$, let ${\cal U}_x$ denote the vector bundle over ${\cal M}_{\xi}$ obtained by restricting $\cal U$ to $x\times {\cal M}_{\xi}$. The vector bundle over $D$ obtained by restricting ${\cal U}_D$ (respectively, ${\rm Ad}({\cal U}_D)$) to $x\times D$ will be denoted by $({\cal U}_D)_x$ (respectively, ${\rm Ad}({\cal U}_D)_x$). Since $H^2(D,\, {\Bbb Z}) \, = \, {\Bbb Z}$, the stability of a vector bundle over $D$ does not depend on the choice of polarization needed to define the degree of a coherent sheaf over $D$. \medskip \noindent {\bf Theorem\, 2.1.}\, {\it For any point $x\in X$, the vector bundle $({\cal U}_D)_x$ over $D$ is stable. Moreover, the infinitesimal deformation map $$ T_xX \, \longrightarrow \, H^1(D,\, {\rm Ad}({\cal U}_D)_x) $$ for the family ${\cal U}_D$ of vector bundles over $D$ parametrized by $X$, is an isomorphism.} \medskip {\it Proof.}\, If $n\geq 3$ and also $g\geq 3$, then the theorem has already been proved in \cite[Theorem 2.5]{BB}. Take a point $x\in X$. We start, as in the proof of Theorem 2.5 of \cite{BB}, by considering the exact sequence $$ 0 \, \longrightarrow \, \mbox{Ad}({\cal U})_x\otimes {\cal O}_{{\cal M}_{\xi}}(-D) \, \longrightarrow \, \mbox{Ad}({\cal U})_x \, {\buildrel{F}\over\longrightarrow} \, {\tau}_*\mbox{Ad}({\cal U}_D)_x \, \longrightarrow \, 0 \, , $$ over $X\times {\cal M}_{\xi}$, where $F$ denotes the restriction map. This yields the long exact sequence $$ H^1({\cal M}_{\xi}, \, \mbox{Ad}({\cal U})_x\otimes {\cal O}_{{\cal M}_{\xi}}(-D)) \hspace{.1in} \longrightarrow \hspace{.1in} H^1({\cal M}_{\xi}, \, \mbox{Ad}({\cal U})_x) $$ $$ \longrightarrow \hspace{.1in} H^1(D,\, \mbox{Ad}({\cal U}_D)_x) \hspace{.1in} \longrightarrow \hspace{.1in} H^2({\cal M}_{\xi}, \, \mbox{Ad}({\cal U})_x\otimes {\cal O}_{{\cal M}_{\xi}}(-D)) $$ of cohomologies. If we consider $\cal U$ as a family of vector bundles over ${\cal M}_{\xi}$ parametrized by $X$, then the infinitesimal deformation map $$ T_x X \hspace{.1in} \longrightarrow \hspace{.1in} H^1({\cal M}_{\xi}, \, \mbox{Ad}({\cal U})_x) $$ is an isomorphism \cite[page 392, Theorem 2]{NR}. In view of the above long exact sequence, to prove that the infinitesimal deformation map is surjective it suffices to establish the following lemma. \medskip \noindent {\bf Lemma\, 2.2.}\, {\it If $i=1,2$, then the following vanishing of cohomology $$ H^i({\cal M}_{\xi}, \, {\rm Ad}({\cal U}_x)\otimes {\cal O}_{{\cal M}_{\xi}}(-D)) \hspace{.1in} = \hspace{.1in} 0 $$ is valid.} \medskip {\it Proof of Lemma 2.2.}\, This lemma was proved in \cite[Lemma 2.1]{BB} under the assumption that $n\geq 3$. So in the proof we will assume that $n=2$ and $g\geq 4$. Let $p$ denote, as in \cite[Section 3]{BB}, the natural projection of the projective bundle ${\Bbb P}({\cal U}_x)$ over ${\cal M}_{\xi}$ onto ${\cal M}_{\xi}$. Let $T^{\mbox{rel}}_p$ denote the relative tangent bundle for the projection $p$ from ${\Bbb P}({\cal U}_x)$ to ${\cal M}_{\xi}$. Since $R^1p_* T^{\mbox{rel}}_p =0$ and $p_* T^{\mbox{rel}}_p \cong {\rm Ad}({\cal U}_x)$, for any $i =0,1,2$, the isomorphism $$ H^i({\cal M}_{\xi}, \, {\rm Ad}({\cal U}_x)\otimes {\cal O}_{{\cal M}_{\xi}}(-D)) \, = \, H^i(U,\, p^*K_{{\cal M}_{\xi}}\otimes T^{\mbox{rel}}_{p}) $$ is obtained from the Leray spectral sequence for the map $p$. If $E$ is a stable vector bundle of rank two and degree one over $X$, then the vector bundle $E'$ over $X$ obtained by performing an elementary transformation $$ 0 \, \longrightarrow\, E' \, \longrightarrow\, E \, \longrightarrow\, L_x \, \longrightarrow\, 0 \, , $$ where $L_x$ is a one dimensional quotient of the fiber $E_x$, is semistable. Therefore, we have a morphism, which we will denote by $q$, from ${\Bbb P}({\cal U}_x)$ to the moduli space ${\cal M}_{\xi (-x)}$. Here $\xi (-x)$ denotes the line bundle $\xi\bigotimes {\cal O}_X(-x)$, and ${\cal M}_{\xi (-x)}$ is the moduli space of semistable vector bundles over $X$ of rank two and determinant $\xi (-x)$. Define $U \subset {\Bbb P}({\cal U}_x)$ to be the inverse image, under that map $q$, of the stable locus of ${\cal M}_{\xi (-x)}$. The line bundle $T^{\mbox{rel}}_{p}$ is isomorphic to the relative canonical bundle $K^{\mbox{rel}}_{q}$ \cite[page 85]{Ty}, \cite{NR}. Therefore, to prove the lemma it suffices to show that $$ H^i(U,\, p^*K_{{\cal M}_{\xi}}\otimes K^{\mbox{rel}}_{q}) \, = \, 0 \, , \leqno{(2.3)} $$ where $i =0,1,2$. Using the isomorphism of $T^{\mbox{rel}}_{p}$ with $K^{\mbox{rel}}_{q}$, from $$ q^*K_{{\cal M}_{\xi (-x)}}\bigotimes K^{\mbox{rel}}_{q} \, \cong \, K_U \, \cong \, p^*K_{{\cal M}_{\xi}} \bigotimes K^{\mbox{rel}}_{p} \leqno{(2.4)} $$ we have $$ p^*K_{{\cal M}_{\xi}}\bigotimes K^{\mbox{rel}}_{q} \, \cong \, q^*K_{{\cal M}_{\xi (-x)}} \bigotimes \left(K^{\mbox{rel}}_{q} \right)^{\otimes 3} \, . $$ Since the restriction of the line bundle $p^*K_{{\cal M}_{ \xi}} \bigotimes K^{\mbox{rel}}_{p}$ to a fiber of the map $q$ has strictly negative degree, using the above isomorphism, and the projection formula, we have $$ H^i(U,\, p^*K_{{\cal M}_{\xi}}\otimes K^{\mbox{rel}}_{q}) \, = \, H^{i-1}\left({\cal M}_{\xi (-x)}, \, K_{{\cal M}_{\xi (-x)}}\bigotimes R^1q_* \left(K^{\mbox{rel}}_{q}\right)^{\otimes 3}\right) \, , \leqno{(2.5)} $$ where $i = 0,1,2$. The map $q$ is smooth fibration ${\Bbb C}{\Bbb P}^1$ fibration over an open subset $U'$ of ${\cal M}_{\xi (-x)}$. The assumption that the genus of $X$ at least four, ensures that the codimension of the complement of $U'$ is at least four. Therefore, by using the Hartog type theorem for cohomology, the isomorphism (2.5) is established. Setting $i=0$ in (2.5), we conclude that $H^0(U,\, p^*K_{{\cal M}_{\xi}} \bigotimes K^{\mbox{rel}}_{q}) = 0$. The following proposition is needed for our next step. \medskip \noindent {\bf Proposition\, 2.6.}\, {\it Let $W$ be a holomorphic vector bundle of rank two over a complex manifold $Z$, and let $f : {\Bbb P}(V) \longrightarrow Z$ be the corresponding projective bundle. Then there are canonical isomorphisms $$ R^1f_* K^{\otimes 3}_f \, \cong \, S^4(W)\bigotimes \left({\bigwedge}^2 W^*\right)^{\otimes 2} \, \cong \, R^0f_* T^{\otimes 2}_f \, , $$ where $K_f$ (respectively, $T_f$) is the relative canonical (respectively, anticanonical) line bundle.} \medskip {\it Proof of Proposition 2.6.}\, To construct the isomorphisms, let $V$ be a complex vector space of dimension two. Choosing a basis of $V$, we identify the tangent bundle $T_{{\Bbb P}(V)}$ with ${\cal O}_{{\Bbb P}(V)}(2)$, and also obtain an identification of the line ${\bigwedge}^2 V^*$ with $\Bbb C$. Since, $$ H^0({\Bbb P}(V),\, {\cal O}_{{\Bbb P}(V)}(m))\, = \, S^m(V) \, , $$ we have an isomorphism of $H^0({\Bbb P}(V),\, T^{\otimes 2}_{{\Bbb P}(V)})$ with $S^4(V)\bigotimes \left({\bigwedge }^2 V^*\right)^{\otimes 2}$. Now it is a straight forward computation to check that this isomorphism is $GL(V)$ invariant, i.e., it does not depend on the choice of a basis of $V$. Therefore, this pointwise construction of a canonical isomorphism of vector spaces induces an isomorphism $$ R^0f_* T^{\otimes 2}_f \, \cong \, S^4(W)\bigotimes \left({\bigwedge }^2 W^*\right)^{\otimes 2} $$ between vector bundles. To obtain the other isomorphism in the statement of the proposition, first note that by the Serre duality we have $H^0({\Bbb P}(V),\, T^{\otimes 2}_{{\Bbb P}(V)}) = H^1({\Bbb P}(V),\, K^{\otimes 3}_{{\Bbb P}(V)})^*$. Now the canonical identification of $S^4(W)\bigotimes \left({\bigwedge}^2 W^*\right)^{\otimes 2}$ with its dual, namely $S^4(W^*)\bigotimes \left({\bigwedge}^2 W\right)^{\otimes 2}$, gives the other isomorphism. This completes the proof of Proposition 2.6.$\hfill{\Box}$ \medskip The isomorphisms in Proposition 2.6 are {\it canonical isomorphisms}, i.e., they are compatible with the pull back of $W$ using any map $Z' \longrightarrow Z$, and furthermore, the isomorphisms are compatible with substituting $W$ by $W\bigotimes L$, where $L$ is a holomorphic line bundle over $Z$. Combining Proposition 2.6 with (2.5), and using the projection formula, we get that if $i = 0,1,2$, then $$ H^i(U,\, p^*K_{{\cal M}_{\xi}}\bigotimes K^{\mbox{rel}}_{q}) \, = \, H^{i-1}\left({\cal M}_{(-x)}, \, K_{{\cal M}_{\xi (-x)}}\bigotimes q_* \left(T^{\mbox{rel}}_{q}\right)^{\otimes 2}\right) \leqno{(2.7)} $$ $$ = \, \hspace{.1in} H^{i-1}\left(U, \, q^* K_{{\cal M}_{\xi (-x)}}\bigotimes \left(T^{\mbox{rel}}_{q}\right)^{\otimes 2}\right) \, . $$ Indeed, the first isomorphism in (2.7) is a consequence of (2.5) and Proposition 2.6, and since $R^1p_* \left(T^{\mbox{rel}}_{q}\right)^{\otimes 2} = 0$, the second isomorphism in (2.7) is valid. Although there is no universal vector bundle over $X\times {\cal M}_{\xi (-x)}$, the properties of the isomorphism $R^1f_* K^{\otimes 3}_f \cong R^0f_* T^{\otimes 2}_f$ in Proposition 2.6 that were explained earlier, evidently ensure that the isomorphism in (2.7) is valid. More precisely, the pointwise construction of the isomorphism between $R^1q_* \left(K^{\mbox{rel}}_{q}\right)^{\otimes 3}$ and $q_* \left(T^{\mbox{rel}}_{q}\right)^{\otimes 2}$ gives an isomorphism of vector bundles. Using (2.4), and the earlier mentioned fact that $T^{\mbox{rel}}_{p} \cong K^{\mbox{rel}}_{q}$, we obtain that $$ q^* K_{{\cal M}_{\xi (-x)}}\bigotimes \left(T^{\mbox{rel}}_{q}\right)^{\otimes 2} \, \cong \, p^*K_{{\cal M}_{\xi}}\bigotimes \left(K^{\mbox{rel}}_{p}\right)^{\otimes 3} \, . $$ Since the restriction of $\left(K^{\mbox{rel}}_{p} \right)^{\otimes 3}$ to a fiber of $p$ has strictly negative degree, we have $p_*\left(K^{\mbox{rel}}_{p} \right)^{\otimes 3} = 0$. Consequently, the above isomorphism simplifies the terms in (2.7) to give the following isomorphism $$ H^{i-1}\left(U, \, q^* K_{{\cal M}_{\xi (-x)}}\bigotimes \left(T^{\mbox{rel}}_{q}\right)^{\otimes 2}\right)\, = \, H^{i-2}\left({\cal M}_{\xi},\, K_{{\cal M}_{\xi}} \bigotimes R^1p_*\left(K^{\mbox{rel}}_{p}\right)^{\otimes 3}\right) \leqno{(2.8)} $$ where $i = 0,1,2$. Note that we obtain $H^1(U,\, p^*K_{{\cal M}_{\xi}}\bigotimes K^{\mbox{rel}}_{q} ) = 0$ by setting $i=1$ in (2.8). In order to complete the proof of the lemma we need to show that $$ H^2(U,\, p^*K_{{\cal M}_{\xi}} \bigotimes K^{\mbox{rel}}_{q}) \, = \, 0 \, . \leqno{(2.9)} $$ To prove the above statement first observe that using (2.8), and setting $i=2$, we have the following isomorphism $$ H^2(U,\, p^*K_{{\cal M}_{\xi}} \bigotimes K^{\mbox{rel}}_{q}) \, = \, H^0\left({\cal M}_{\xi},\, K_{{\cal M}_{\xi}}\bigotimes R^1p_*\left(K^{\mbox{rel}}_{p} \right)^{\otimes 3}\right)\, . \leqno{(2.10)} $$ Now using Proposition 2.6 we have $$ H^0\left({\cal M}_{\xi},\, K_{{\cal M}_{\xi}}\otimes R^1p_*\big(K^{\mbox{rel}}_{p} \big)^{\otimes 3}\right) \, = \, H^0\left({\cal M}_{\xi}, \, K_{{\cal M}_{\xi}} \otimes S^4({\cal U}_x)\otimes \big({\bigwedge}^2 {\cal U}^*_x\big)^{\otimes 2}\right)\, , $$ where ${\cal U}_x$, as defined earlier, is the vector bundle over ${\cal M}_{\xi}$ obtained by restricting the Poincar\'e bundle $\cal U$ to the subvariety $x\times {\cal U}_{\xi} \subset X\times {\cal U}_{\xi}$. The vector bundle ${\cal U}_x$ is known to be stable. Consequently, the vector bundle $$ S^4({\cal U}_x)\bigotimes \left({\bigwedge}^2 {\cal U}^*_x\right)^{\otimes 2} $$ is semistable. Now, since the vector bundle $S^4({\cal U}_x) \bigotimes \left({\bigwedge}^2 {\cal U}^*_x\right)^{\otimes 2}$ is the dual of itself, its degree is zero. On the other hand, the degree of $K_{{\cal M}_{\xi}}$ is strictly negative. From these it follows that the vector bundle $$ K_{{\cal M}_{\xi}}\bigotimes S^4({\cal U}_x)\bigotimes \left({\bigwedge}^2 {\cal U}^*_x\right)^{\otimes 2} $$ does not admit any nonzero section, since it is semistable of strictly negative degree. In view of (2.10), this establishes the assertion (2.9). Therefore, the assertion (2.3) is valid. This completes the proof of the lemma.$\hfill{\Box}$ \medskip Since we have established, in Lemma 2.2, the rank two analog of Lemma 2.1 of \cite{BB}, the proof of the stability of the vector bundle $({\cal U}_D)_x$ for rank at least three, as given in \cite[Theorem 2.5]{BB}, is also valid for the rank two case if $g\geq 4$. We note that \cite[Theorem 2.5]{BB} was proved under the assumption that $g\geq 3$. However, the proof remains valid for $g=2$ if the condition that the rank is at least four is imposed. Under this condition, the codimension of the subvariety over which the map $q$ fails to be smooth and proper is sufficiently large in order to be able to apply the analog Hartog's theorem, which has been repeatedly used, for the cohomologies in question. This completes the proof of Theorem 2.1.$\hfill{\Box}$ \medskip In view of the above Lemma 2.2, all the results established in Section 2 of \cite{BB} for rank $n \geq 3$ remain valid for rank two and $g \geq 4$. \section{Computation of the infinitesimal deformations} Let $X$ be a compact connected Riemann surface of genus $g$, with $g\geq 2$. Take a holomorphic line bundle $\xi$ over $X$ of degree $d$. Let ${\cal M}_{\xi} := {\cal M}(n,\xi)$ denote the moduli space of stable vector bundles $E$ of rank $n$ over $X$, with ${\bigwedge}^n E = \xi$. For another line bundle ${\xi}'$ of degree $d$, the variety ${\cal M}(n,{\xi}')$ is isomorphic to ${\cal M}(n,\xi)$. Indeed, if $\eta$ is a line bundle over $X$ with ${\eta}^{\otimes n} = {\xi}'\otimes {\xi}^*$, then the map defined by $E \longmapsto E\otimes\eta$ is an isomorphism from ${\cal M}(n,\xi)$ to ${\cal M}(n,{\xi}')$. Therefore, we can rigidify (infinitesimally) the choice of $\xi$ by the following procedure. Fix a line bundle $L$ of degree one over $X$ such that $L^{\otimes (2-2g)}$ is isomorphic to the tangent bundle $T_X$. We fix $\xi$ to be $L^{\otimes d}$. We will assume that the integers $n$ and $d$ are coprime, and $n\geq 2$. We will further assume that if $g=2$ then $n\neq 2,3$, and if $g=3$, then $n\neq 2$. The above numerical assumptions are made in order to ensure that the assertion in Theorem 2.1 is valid for ${\cal M}_{\xi}$. Take a smooth divisor $D$ on ${\cal M}_{\xi}$ such that ${\cal O}_{{\cal M}_{\xi}}(D) = K^{-1}_{{\cal M}_{\xi}}$. Consider the exact sequence of sheaves $$ 0 \, \longrightarrow \, {\cal O}_{{\cal M}_{\xi}} \, \longrightarrow \, {\cal O}_{{\cal M}_{\xi}}(D) \, \longrightarrow \, {\tau}_*N_D \, \longrightarrow \, 0 $$ over ${\cal M}_{\xi}$, where $N_D$ is the normal bundle of the divisor $D$, and ${\tau}$ is the inclusion map of $D$ into ${\cal M}_{\xi}$. Since $$ H^1({\cal M}_{\xi}, \, {\cal O}_{{\cal M}_{\xi}}) \, = \, 0 \, , $$ using the exact sequence of cohomologies, the space of sections $H^0(D,\, N_D)$ gets identified with the quotient vector space $H^0({\cal M}_{\xi},\, {\cal O}_{{\cal M}_{\xi}}(D))/{\Bbb C}$. Let $\cal S$ denote the space of all divisors $D'$ on ${\cal M}_{\xi}$ such that $D'$ is homologous to $D$, i.e., they are represented by the same element in $H^2({\cal M}_{\xi}, \, {\Bbb Z})$. Therefore, $\cal S$ is identified with ${\Bbb P}H^0({\cal M}_{\xi},\, K^{-1}_{{\cal M}_{\xi}})$. The tangent space to $\cal S$, at the point $[D'] \in {\cal S}$ representing a divisor $D'$, has the following identification $$ T_{[D']}{\cal S} \, = \, H^0(D',\, N_{D'}) \, = \, H^0({\cal M}_{\xi}, \, {\cal O}_{{\cal M}_{\xi}}(D'))/{\Bbb C} \, . $$ Let $\cal N$ denote the moduli space of triplets of the form $(X,L,D)$, where $X$, $L$ and $D$ are as above (the line bundle $L$ is a $(2g-2)$-th root of $K_X$). So $\cal N$ is an open subset of moduli space of triplets of the form $(X,L,\alpha)$, where $\alpha$ is a linear subspace of $H^0({\cal M}_{\xi},\, K^{-1}_{{\cal M}_{\xi}})$ of dimension one. The space $\cal N$ parametrizes a family of Calabi-Yau varieties, simply by associating the Calabi-Yau variety $D$ to any triplet $(X,L, D) \in {\cal N}$. Take a point $\gamma \,:= \, (X,L,D)$ in the moduli space $\cal N$. Associated to this family is the homomorphism $$ F \, : \, T_{\gamma} {\cal N} \, \longrightarrow \, H^1(D,\, T_D) \leqno{(3.1)} $$ that maps the tangent space $T_{\gamma}{\cal N}$ of $\cal N$ at $\gamma$ to the space of infinitesimal deformation of the complex manifold $D$. In other words, this homomorphism $F$ sends any tangent vector $v \in T_{\gamma} {\cal N}$ to the corresponding Kodaira-Spencer infinitesimal deformation class of $D$ for the above family parametrized by $\cal N$. The vector space $T_{\gamma} {\cal N}$ fits naturally into the short exact sequence $$ 0 \, \longrightarrow \, H^0(D,\, N_D) \, \longrightarrow \, T_{\gamma}{\cal N} \, \longrightarrow \, H^1(X,\, T_X) \, \longrightarrow \, 0 \, , \leqno{(3.2)} $$ where the projection $T_{\gamma} {\cal N} \longrightarrow H^1(X,\, T_X)$ corresponds to the forgetful map, which sends any point $(X',L',D') \in {\cal N}$ to the point represented by $X'$ in the moduli space of Riemann surfaces; the inclusion $H^0(D,\, N_D) \longrightarrow T_{\gamma} {\cal N}$ in (3.2) corresponds to the obvious homomorphism $T_{[D]} {\cal S} \longrightarrow T_{\gamma} {\cal N}$, where $\cal S$, as before, is ${\Bbb P}H^0({\cal M}_{\xi},\, K^{-1}_{{\cal M}_{\xi}})$, the space of anticanonical divisors on ${\cal M}_{\xi}$. \medskip \noindent {\bf Theorem\, 3.3.}\, {\it The Kodaira-Spencer infinitesimal deformation map $F$ constructed in (3.1) is an isomorphism of the tangent space $T_{\gamma} {\cal N}$ with $H^1(D,\, T_D)$.} \medskip {\it Proof.}\, We start by considering the exact sequence $$ 0 \, \longrightarrow \, T_D \, \longrightarrow \, {\tau}^*T_{{\cal M}_{\xi}} \, \longrightarrow \, N_D \, \longrightarrow \, 0 $$ of vector bundles over $D$, where $N_D$ is the normal bundle of $D$, and ${\tau}$, as before, is the inclusion map of $D$ into ${\cal M}_{\xi}$. This gives us the exact sequence $$ H^0(D,\, {\tau}^*T_{{\cal M}_{\xi}}) \longrightarrow \,H^0(D,\, N_D) \longrightarrow H^1(D,\, T_{D}) \longrightarrow H^1(D,\, {\tau}^*T_{{\cal M}_{\xi}}) \longrightarrow H^1(D,\, N_D) \leqno{(3.4)} $$ of cohomologies. Since the canonical line bundle $K_D$ is trivial, and $N_D \cong {\tau}^* K^{-1}_{{\cal M}_{\xi}}$ is ample, the Kodaira vanishing theorem gives $$ H^1(D,\, N_D) \, = \, 0 \, . \leqno{(3.5)} $$ Therefore, the homomorphism $H^1(D,\, T_{D}) \longrightarrow H^1(D,\, {\tau}^*T_{{\cal M}_{\xi}})$ in (3.4) is surjective. Our next aim is to show that $$ H^0(D,\, {\tau}^*T_{{\cal M}_{\xi}}) \hspace{.1in} = \hspace{.1in} 0 \, , \leqno{(3.6)} $$ which would be the first step in turning (3.4) into the short exact sequence (1.1) that we are seeking. For that purpose, consider the vector bundle ${\cal U}_D$ over $X\times D$ obtained by restricting a Poincar\'e bundle. Let $\phi$ (respectively, $\psi$) denote the projection of $X\times D$ to $X$ (respectively, $D$). The vector bundle $R^1 {\psi}_* {\rm Ad}({\cal U}_D)$ over $D$ is naturally isomorphic to ${\tau}^* T_{{\cal M}_{\xi}}$. Also, ${\psi}_* {\rm Ad}({\cal U}_D) = 0$, as the vector bundle $({\cal U}_D)_x$ is stable, hence {\it simple}, for every $x\in X$ [Theorem 2.1]. The vector bundle ${\rm Ad}({\cal U}_D)$, as in Section 2, is the subbundle of ${\rm End}({\cal U}_D)$ consisting of trace zero endomorphisms. Now, using the Leray spectral sequence for the projection $\psi$, the isomorphism $$ H^0(D,\, {\tau}^*T_{{\cal M}_{\xi}}) \, = \, H^1(X\times D,\, {\rm Ad}({\cal U}_D)) $$ is obtained. The vector bundle $({\cal U}_D)_x$ over $D$, defined in Section 2, has been proved to be stable in Theorem 2.1. So, we have $H^0(D,\, {\rm Ad}(({\cal U}_D)_x)) = 0$ for every $x \in X$. Consequently, the isomorphism $$ H^1(X\times D,\, {\rm Ad}({\cal U}_D))\, = \, H^0(X,\, R^1{\phi}_* {\rm Ad}({\cal U}_D)) $$ is obtained. Now, from the second part of Theorem 2.1 we have a natural isomorphism $$ R^1{\phi}_* {\rm Ad}({\cal V}_D) \hspace{.1in} = \hspace{.1in} T_X $$ obtained using the Poincar\'e bundle. Finally, since $H^0(X,\, T_X) =0$, the assertion in (3.6) is an immediate consequence of the above isomorphism. Using (3.5) and (3.6), the exact sequence in (3.4) reduces to $$ 0 \, \longrightarrow \, H^0(D,\, N_D) \, \longrightarrow \, H^1(D,\, T_{D})\, \longrightarrow \, H^1(D,\, {\tau}^*T_{{\cal M}_{\xi}}) \, \longrightarrow \, 0 \, . \leqno{(3.7)} $$ The comparison of (3.7) with (3.2) shows that the next step has to be computation of $H^1(D,\, {\tau}^*T_{{\cal M}_{\xi}})$. Consider the short exact sequence $$ 0 \, \longrightarrow \, T_{{\cal M}_{\xi}}\bigotimes {\cal O}_{{\cal M}_{\xi}}(-D) \, \longrightarrow \, T_{{\cal M}_{\xi}} \, \longrightarrow \, {\tau}_*{\tau}^* T_{{\cal M}_{\xi}} \, \longrightarrow \, 0 $$ of sheaves over ${\cal M}_{\xi}$. We know that $H^2({\cal M}_{\xi},\, T_{{\cal M}_{\xi}}) = 0$ \cite[page 391, Theorem 1.a]{NR}. Also, we have (3.6). Consequently, the exact sequence yields the long exact sequence $$ 0 \, \longrightarrow \, H^1({\cal M}_{\xi},\, T_{{\cal M}_{\xi}}\bigotimes K_{{\cal M}_{\xi}})\, \longrightarrow \, H^1({\cal M}_{\xi},\, T_{{\cal M}_{\xi}}) \leqno{(3.8)} $$ $$ \, \longrightarrow \, H^1(D,\, {\tau}^*T_{{\cal M}_{\xi}})\, \longrightarrow \, H^2({\cal M}_{\xi},\, T_{{\cal M}_{\xi}}\bigotimes K_{{\cal M}_{\xi}}) \, \longrightarrow \, 0 $$ of cohomologies; note that $H^i({\cal M}_{\xi},\, {\tau}_*{\tau}^* T_{{\cal M}_{\xi}}) = H^i(D,\, {\tau}^*T_{{\cal M}_{\xi}})$. It was proved in \cite{NR} that the Kodaira-Spencer deformation map for ${\cal M}_{\xi}$, as the Riemann surface $X$ moves in the moduli space of Riemann surfaces, is an isomorphism of $H^1({\cal M}_{\xi},\, T_{{\cal M}_{\xi}})$ with $H^1(X,\, T_X)$. Therefore, comparing (3.2) with (3.7), and using the exact sequence (3.8), we conclude that in order to complete the proof of the theorem, it suffices to establish the following statement~: if $i=1,2$, then $$ H^i({\cal M}_{\xi},\, T_{{\cal M}_{\xi}}\bigotimes K_{{\cal M}_{\xi}})\, = \, 0 \, . \leqno{(3.9)} $$ Indeed, (3.9) implies that $H^1(D,\, {\tau}^*T_{{\cal M}_{\xi}}) = H^1({\cal M}_{\xi},\, T_{{\cal M}_{\xi}}) = H^1(X,\, T_X)$. To prove (3.9), let $\delta$ denote the dimension of the variety ${\cal M}_{\xi}$. The Serre duality gives the following isomorphism $$ H^i({\cal M}_{\xi},\, T_{{\cal M}_{\xi}}\bigotimes K_{{\cal M}_{\xi}})\, = \, H^{\delta -i}({\cal M}_{\xi},\, {\Omega}^1_{{\cal M}_{\xi}})^* \, = \, H^{1,\delta -i}({\cal M}_{\xi})^* \, . \leqno{(3.10)} $$ (Here $H^{j,k}({\cal M}_{\xi}) \, := \, H^k({\cal M}_{\xi},\, {\Omega}^j_{{\cal M}_{\xi}})$.) To finish the proof of the statement (3.9) we need to use some properties of the Hodge structure of the cohomology algebra $H^*({\cal M}_{\xi}, \, {\Bbb C})$, which will be recalled now. Fix a Poincar\'e bundle $\cal U$ over $X\times {\cal M}_{\xi}$. Let $c_k \, := \, c_k({\cal U}) \, \in \, H^{k,k}(X\times {\cal M}_{\xi})$ denote the $k$-th Chern class of $\cal U$. For any ${\alpha} \in H^{i,j}(X)$, we have $$ \lambda (k,\alpha) \, :=\, \int_X c_k\cup f^*\alpha \, \in \, H^{k+i-1,k+j-1}({\cal M}_{\xi})\, , \leqno{(3.11)} $$ where $f$ denotes the obvious projection of $X\times {\cal M}_{\xi}$ onto $X$, and $\int_X$ is the Gysin map for this projection, which is constructed by integrating differential forms on $X\times {\cal M}_{\xi}$ along the fibers of the projection $f$. The collection of all these cohomology classes $\{\lambda (k,\alpha)\}$, constructed in (3.11), generate the cohomology algebra $H^*({\cal M}_{\xi},\, {\Bbb C})$ \cite[page 581, Theorem 9.11]{AB}. On the other hand, we know that the following $$ H^{0,1}({\cal M}_{\xi}) \hspace{.1in} = \hspace{.1in} 0 $$ is valid. With these properties of $H^*({\cal M}_{\xi},\, {\Bbb C})$ at our disposal, we are in a position to prove that the algebra generated by the cohomology classes $\{\lambda (k,\alpha)\}$ cannot have a nonzero element in $H^{1,\delta -1}({\cal M}_{\xi})$ or $H^{1,\delta -2} ({\cal M}_{\xi})$, where $\delta = {\dim}_{\Bbb C} {\cal M}_{\xi}$. To prove the above assertion, suppose that $$ {\omega} \, = \, {\omega}_1\wedge {\omega}_2\wedge \cdots \wedge {\omega}_l $$ is a nonzero element in $H^{1,\delta -1}({\cal M}_{\xi})\bigoplus H^{1,\delta -2}({\cal M}_{\xi})$, where ${\omega}_j \in \{\lambda (k,\alpha)\}$ for all $j \in [1,l]$. We will see that ${\omega}_j \in H^{0,1}({\cal M}_{\xi})$ for at least one $j \in [1,l]$. Since $H^{0,1}({\cal M}_{\xi}) =0$, this would prove that $\omega = 0$. First observe that in (3.11), we have $k+i-1 \geq k-1$ and $k+j-1 \leq k$, as ${\dim}_{\Bbb C} X =1$. In other words, we have $$ (k+j-1) - (k+i-1) \hspace{.1in} \leq \hspace{.1in} 1 \, . \leqno{(3.12)} $$ Let ${\omega}_i \in H^{a_i,b_i}({\cal M}_{\xi})$, where $i\in [1,l]$. Then $a_i \leq 1$, and consequently from (3.12) the inequality $b_i\leq 2$ is obtained. Furthermore, $a_j \neq 0$ for at most one $j\in [1,l]$. If $a_i =0$, then $b_i\leq 1$; but the possibility $b_i =1$ is ruled out as $H^{0,1}({\cal M}_{\xi}) = 0$. Therefore, all ${\omega}_i$ except one is a scalar. Now, if $a_j=1$, then from (3.12) we have $b_j\leq 2$. On the other hand, we have $\delta -2 > 2 \geq b_j$. Consequently, we conclude that $\omega = 0$. Since the cohomology classes $\lambda (k,\alpha)$ are of pure type, i.e., $$ \lambda (k,\alpha) \hspace{.1in} \in \hspace{.1in} H^{a,b}({\cal M}_{\xi}) $$ for some integers $a$ and $b$, it is easy to see that for any $i\geq 0$, the cohomology group $H^{i}({\cal M}_{\xi},\, {\Bbb C})$ is generated, as a complex vector space, by completely decomposable elements, i.e., elements of the type $\omega$ considered above. Therefore, we have $H^{1,\delta -1}({\cal M}_{\xi}) = 0 = H^{1,\delta -2}({\cal M}_{\xi})$. In view of (3.10), this completes the proof of the statement (3.9). We already noted that the statement (3.9) completes the proof of the theorem.$\hfill{\Box}$ \medskip As a consequence of Theorem 3.3 we get that $$ \dim H^1(D,\, T_D) \, = \, 3g-4 \, + \, \dim H^0({\cal M}_{\xi}, \, {\Theta}^{\otimes 2}) \, . $$ The dimension of $H^0({\cal M}_{\xi}, \, {\Theta}^{\otimes 2})$ is given by the Verlinde formula. \medskip \noindent {\bf Remark\, 3.13.}\, From \cite[page 760, Proposition 1]{Be}, coupled with \cite[page 759, Th\'eor\`eme 1]{Be}, it follows that $H^0(D,\, T_D) =0$. We note that this is also an immediate consequence of (3.6). We have $H^2(D,\, T_D) \,= \, H^{m-1,2}(D) \, = \, H^{m,3}({\cal M}_{\xi})$, where $m= \dim D$; the second isomorphism is obtained from the Lefschetz hyperplane theorem \cite[page 156]{GH}. The earlier proof that $H^{1,\delta -i}({\cal M}_{\xi}) =0$ for $i=1,2$, easily extends to prove that $H^{1,\delta -3}({\cal M}_{\xi}) =0$. Therefore, we have $H^2(D,\, T_D) =0$. However, by a theorem due to Bogomolov-Kawamata-Tian-Todorov it is already known that the deformations of a Calabi-Yau variety are unobstructed.
\section{Motivation for Measuring Quark Asymmetries} The accurate determination of the \mbox{forward-backward} asymmetries, $A_{\rm FB}$, of quarks serves to test the structure of Standard Model ({\sffamily SM})\cite{SM} couplings to fermions. They also probe radiative corrections to the {\sffamily SM} and consequently allow greater precision when predicting unknown parameters of the model. These are increasingly used to constrain uncertainties on the mass of the Higgs boson\cite{lepewwg}. Global fits to electroweak data assume the {\sffamily SM} structure of Z couplings to leptons ($e, \, \mu \, \tau$) and both up ($u, \, c$) and \mbox{down-type} ($d, \, s, \, b$) quarks. Given recent, highly accurate, lepton measurements from the $\tau$ polarisation\cite{taupol} and purely leptonic \mbox{forward-backward} asymmetries\cite{lepewwg}, a similar precision in the quark sector is needed to confirm the internal consistency of the model. The suite of complimentary measurements described here provide such a precision for both up and \mbox{down-type} quark families. Performing these measurements at {\sffamily LEP 1} offers several advantages. The sensitivity of initial state couplings to the effective weak mixing angle, ${\rm sin^2}\theta_{\rm w}^{\rm eff}$, is compounded by large, measurable asymmetries from quark final states close to the Z. Heavy quark asymmetries in particular are especially favourable, as flavour and direction of the final state quark can be tagged with greater ease than is the case with lighter quarks. \section{Definitions and Experimental Issues} In the {\sffamily SM}, the differential \mbox{cross-section} for the process $e^+ \, e^- \rightarrow f^+ \,f^-$ can be written as~: \begin{equation} \frac{1}{\sigma} \: \frac{{\rm d} \, \sigma^f}{ {\rm d} \, cos \, \theta} \: = \: \frac{3}{8} \; ( 1 + cos^2 \theta) \; + \; A_{\rm FB}^f \, cos \, \theta \end{equation} where $A_{\rm FB}^f$ defined to be the \mbox{forward-backward} asymmetry for fermion flavour, $f$. It can be expressed as~: \begin{equation} A_{\rm FB}^f \: = \: \frac{3}{4} \, {\cal A}_e \; {\cal A}_f \end{equation} where ${\cal A}_f$ is the polarisation of the fermion concerned~: \begin{equation} {\cal A}_f \: = \: \frac{ 2 x}{1 \, + \, x^2 } \: = \: 1 \, - \, \frac{ 2 q }{I_3^f} \, ( {\rm sin^2} \theta_{\rm W}^{\rm eff} \: + \: {\cal C}_f ) \end{equation} where $x$ is the ratio of the vector and axial couplings of the fermion to the Z. This final form, separates the terms containing sensitivity to parameters of the {\sffamily SM}, such as $( m_{\rm t}, \, m_{\rm H})$ through ${\rm sin^2} \theta_{\rm W}^{\rm eff}$, from vertex corrections in ${\cal C}_f$. The latter are typically of the order of $\sim 1\%$ for $b$ quarks\cite{physics:at:lep}. For hadronic decays of quarks, the precise direction of the final state fermion is not accessible experimentally, and so the direction of the thrust axis is usually signed according to methods of correlating the charge of the quark its final decay products. Asymmetries are of the order of $\sim 10\%$ for $b$ and $c$ quarks but are diluted by several effects. These are caused primarily by the correlation method mistagging the quark charge, or by $B^0 \bar{B^0}$ mixing or cancellations between other quark backgrounds. Consequently, the methods of measurement described here represent different compromises between rates of charge mistag, and the \mbox{efficiency$+$purity} of the flavour tagging procedure. With the increasing sophistication of analyses, several methods of tagging the charge and flavour of decaying quarks are available. These are applied either singly or in combinations. Minor complications arise when interpretating the results of these analyses, in the form of corrections to pure electroweak predictions of heavy quark asymmetries. For example, quark mass corrections to the electroweak process are generally small, {\em eg.} representing shifts of 0.05\% in the calculation of $A_{\rm FB}^b$, and are well understood theoretically\cite{physics:at:lep}. Larger and more problematic corrections arise from hard gluon emission in the final state. These \mbox{so-called} {\sffamily QCD} corrections are mass, flavour and analysis dependent and so their treatment and associated uncertainties must be handled with care. \section{Latest Techniques for Measuring $A_{\rm FB}^b$} \subsection{Semileptonic Decays of Heavy Quarks} \label{semilepton} The ``classical'' method of measuring heavy quark asymmetries relies on differences in the momentum and transverse momentum, $(p, \, p_{\rm T})$, spectra of leptons arising from semileptonic $b$ and $c$ quark decays. This method benefits from an unambiguous charge and flavour tag in the case of unmixed $b$ and $c$ hadrons. It however suffers from several disadvantages when used on a more general sample. The methods reliance on a pair of correlated inputs, such as lepton $(p, \, p_{\rm T})$ alone, leads to a dependence of the method on the precise values of the branching ratios, \mbox{BR($b\rightarrow l$)} and \mbox{BR($b\rightarrow c \rightarrow l$)}, semileptonic decay modelling and $c$ branching fractions assumed in Monte Carlo simulations. Some uncertainties are constrained by full use of measurements from lower energy experiments, according to the prescription detailed in\cite{lepewwg}. However, the extrapolation of such measurements to {\sffamily LEP} energies, and into the different environment of a fragmented jet containing a \mbox{$b$-hadron}, leaves residual systematic uncertainties. As the purity of the sample, and its charge mistag, are determined by Monte Carlo simulation alone these residual uncertainties are correlated between the 4 {\sffamily LEP} experiments. The relatively small value of the ($b\rightarrow l$) branching fraction means that an overall $b$ tagging efficiency of the order of $\sim 13\%$ is typically achieved with reasonable purities of {\em eg.} $\sim 80\%$ in the case of\cite{aleph:leptons}. Effects of $B^0 \bar{B^0}$ mixing, ``cascade'' $b$ decays ($b \rightarrow c \rightarrow l$), backgrounds from $c$ decays ($c \rightarrow l$) and other light quark sources, including detector misidentification, all serve to dilute the otherwise excellent mistag rate. The systematic impact of such effects can be severely reduced by fitting simultaneously for the the \mbox{time-integrated} mixing parameter, $\bar{\chi}$\footnote{$\bar{\chi}$ is defined to be the probability that a $B$ meson has oscillated to a $\bar{B}$ meson by the time of its decay.}, together with the $b$ and $c$ asymmetries in a ``global'' fit. Otherwise both $\bar{\chi}$ and $A_{\rm FB}^c$ \hspace{0.5mm} are inputs to the determination of $A_{\rm FB}^b$. Their values and uncertainties are either fixed by experiment or, in the case of $A_{\rm FB}^c$, set to their {\sffamily SM} \hspace{0.5mm} expectation and the dependence on $A_{\rm FB}^b$ used as input to subsequent electroweak ({\sffamily EW}) fits\cite{lepewwg}. Uncertainties can be further minimised by adding input information which discriminates between $b$ and $c$ events. In such analyses\cite{delphi:leptons,opal:leptons} information coming from lifetime tags, using silicon vertex detectors ({\sffamily VDET's}), and event shapes amplify the discrimination from semileptonic $(p, \, p_{\rm T})$ spectra. The {\sffamily LEP} experiments make use of such additional inputs, and extra fit quantities, to varying degrees as is summarised in Table~\ref{table-lepton fits}. \begin{table}[h] \begin{center} \begin{tabular}{|c|cccc|} \hline {\em Inputs} & {\em lepton $(p, \, p_{\rm T})$} & {\em lifetime} & {\em jet shapes} & $\bar{\chi}$ \\ {\em Expt.} & {\bf ADLO} & {\bf DO} & {\bf DO} & {\bf DL} \\ \hline {\em Outputs} & $A_{\rm FB}^b$ & $\bar{\chi}$ & $A_{\rm FB}^c$ & \\ {\em Expt.} & {\bf ADLO} & {\bf AO} & {\bf DO} & \\ \hline \end{tabular} \caption{\em Additional inputs and fit parameters for the semileptonic decay measurements of $A_{\rm FB}^b$ from the 4 {\sffamily LEP} collaborations {\sffamily ALEPH, DELPHI, L3} and {\sffamily OPAL} ({\sffamily ADLO}).} \label{table-lepton fits} \end{center} \end{table} Analyses which make use of the classical semileptonic $(p, \, p_{\rm T})$ spectra, and complement it with lifetime tags in this way, are the most powerful, statistically and with greater systematic control. For example, the {\sffamily OPAL} measurement\cite{opal:leptons} uses lepton $(p, \, p_{\rm T})$ and \mbox{event-shape} information as inputs to a neural net ({\sffamily NN}) $b$ tagging algorithm. In addition, it utilises a largely orthogonal $c$ tag, based on lifetime information of jets in the event, combined with the impact parameter significance and detector identification criteria of the lepton. The output of these two neural nets is shown in Figure~\ref{figure-opal nnout} \begin{figure}[h] \begin{center} \mbox{\epsfig{file=opal.ps,width=8.6cm}} \caption{\em Two-dimensional summary plot of the $c$ tag neural net output vs. that of the $b$ tag {\sffamily NN} output for each of the expected lepton sources in the {\sffamily OPAL} semileptonic asymmetry analysis. The areas of displayed pixels is proportional to the bin contents.} \label{figure-opal nnout} \end{center} \end{figure} where the strong separation between sources of leptons in hadronic events is clearly evident. The separation between $b$ and $c$ lepton sources, and the more limited distinction between those and other background sources, enables both an precise determination of $A_{\rm FB}^b$, $\bar{\chi}$ and the most accurate measurement of $A_{\rm FB}^c$ from the same sample of events. The net gain of such a method is an approximate $\sim 25\%$ improvement in the statistical sensitivity of $A_{\rm FB}^b$ and a $\sim 25\%$ improvement in that of $A_{\rm FB}^c$. A summary of the current results for $A_{\rm FB}^b$ \hspace{0.5mm} from semileptonic measurements at {\sffamily LEP} is given in Table~\ref{table-basyms}. \begin{table}[h] \begin{center} \begin{tabular}{|l|llll|} \hline \multicolumn{5}{|c|}{\em Semileptonic Measurements of $A_{\rm FB}^b$} \\ \hline {\em Experiment} & $A_{\rm FB}^b$ & {\em stat.} & {\em syst.} & {\em total} \\ \hline {\sffamily ALEPH } & 0.0965 & $\pm$0.0044 & $\pm$0.0026 & $\pm$0.0051 \\ {\sffamily DELPHI} & 0.0979 & $\pm$0.0065 & $\pm$0.0029 & $\pm$0.0071 \\ {\sffamily L3 } & 0.0963 & $\pm$0.0065 & $\pm$0.0035 & $\pm$0.0074 \\ {\sffamily OPAL } & 0.0910 & $\pm$0.0044 & $\pm$0.0020 & $\pm$0.0048 \\ \hline \multicolumn{5}{|c|}{\em Lifetime and Jetcharge Measurements of $A_{\rm FB}^b$} \\ \hline {\em Experiment} & $A_{\rm FB}^b$ & {\em stat.} & {\em syst.} & {\em total} \\ \hline {\sffamily ALEPH } & 0.1040 & $\pm$0.0040 & $\pm$0.0032 & $\pm$0.0051 \\ {\sffamily DELPHI} & 0.0979 & $\pm$0.0047 & $\pm$0.0021 & $\pm$0.0051 \\ {\sffamily L3 } & 0.0855 & $\pm$0.0118 & $\pm$0.0056 & $\pm$0.0131 \\ {\sffamily OPAL } & 0.1004 & $\pm$0.0052 & $\pm$0.0044 & $\pm$0.0068 \\ \hline \end{tabular} \caption{\em Summary of latest measurements of $A_{\rm FB}^b$ from the 4 {\sffamily LEP} experiments.} \label{table-basyms} \end{center} \end{table} \subsection{Lifetime Tagging and Jetcharge Measurements} An alternative, complementary technique to measure $A_{\rm FB}^b$ \hspace{0.5mm} is based on lifetime information from silicon vertex detectors. This is then combined with a fully inclusive charge correlation method, referred to as the ``jetcharge'' technique. This method was initially pioneered using samples of untagged hadronic events containing all types of quark flavours accessible at these energies\cite{aleph:qfb}. As a consequence of the low semileptonic branching ratios, such inclusive measurements are almost entirely uncorrelated from semileptonic measurements and so can either be combined or used as a \mbox{cross-check} for consistency of measurements between different methods. The jetcharge method is based upon the correlation between leading particles in a jet, with that of the parent quark. A hemisphere based jetcharge estimator is formed using a summation over particle charges, $q$, weighted by their momentum, $\vec{p}$~: \begin{equation} Q_{\rm F} \: = \: \frac{ \sum_i^{ \vec{p_i} \cdot \vec{T} > 0 } \mid \vec{p_i} \cdot \vec{T} \mid^{\kappa} \, q_i } { \sum_i^{ \vec{p_i} \cdot \vec{T} > 0 } \mid \vec{p_i} \cdot \vec{T} \mid^{\kappa} } \, , \label{equation-charge summation} \end{equation} and analogously for $Q_{\rm B}$. The $\kappa$ parameter is used to optimise the measurement sensitivity. The charge flow between hemispheres, namely $Q_{\rm FB} = Q_{\rm F} - Q_{\rm B}$ is then used to sign the direction of the thrust axis. Currently all {\sffamily LEP} collaborations use this method, and the above formalism. The method benefits from many of the systematic studies performed in untagged samples\cite{aleph:qfb}, especially to understand the degree of charge correlation between hemispheres in background events and their light quark parents. The recent $A_{\rm FB}^b$ \hspace{0.5mm} analysis carried out by {\sffamily DELPHI} also illustrates this method\cite{delphi:jetcha}. In addition to jetcharge information, {\sffamily OPAL} makes uses of a weighted vertex charge method\cite{opal:jetcha}. This quantity is a weighted sum of the charges of tracks in a jet which contains a tagged secondary vertex, {\em ie~:} \begin{equation} q_{\rm vtx} \: = \: \sum_{{\rm tracks}=i} \, \omega_i \, q_i \end{equation} where $q_i$ is the charge of each track,$i$ and $\omega_i$ is related to the probability that the tracks comes from the secondary vertex relative to that it came from the primary. The latter probabilities are determined using impact parameter, momentum and multiplicity information. An estimate of the accuracy of the $q_{\rm vtx}$ charge estimator is derived from its variance. Selecting hemispheres with $\mid q_{\rm vtx} \mid > 1.4 \, \times \, \sigma_{\rm q} + 0.2$ largely removes neutral $B^0$ mesons, and those with poorly measured vertex charges. This also leads to a severe reduction in the size of the event sample, leaving only $\sim 13,000$ events out of a total, untagged input of roughly 4 million hadronic events. Hence, the contribution of the vertex charge measurement, when combined with the jetcharge determination of $A_{\rm FB}^b$, is relatively low. More significant improvements in statistical precision and control of systematic uncertainties can be obtained from the variety of new techniques summarised in Table~\ref{table-new tech}. \begin{table}[h] \begin{center} \begin{tabular}{|l|l|} \hline {\em Expt.} & {\em Improvement} \\ \hline {\bf A L} &{\rm Fit data as a function of tag purity.} \\ \hline {\bf A L} &{\rm Increase acceptance beyond central region.} \\ \hline {\bf ADLO} &{\rm Fit asymmetry as a function of angle.} \\ \hline {\bf A} &{\rm Fit data using range of $\kappa$ values.} \\ \hline {\bf ADO} &{\rm Extract $b$ quark mistag factor from data.} \\ \hline {\bf AD} &{\rm Extract lighter quark mistags from data.} \\ \hline \end{tabular} \caption{\em Recent improvements to the lifetime and jetcharge method for measuring $A_{\rm FB}^b$ from the 4 {\sffamily LEP} collaborations {\sffamily ALEPH, DELPHI, L3} and {\sffamily OPAL} (ADLO).} \label{table-new tech} \end{center} \end{table} Experiments make use of these techniques to varying degrees, the most significant of which being the improvement in statistical sensitivity gained by fitting to the asymmetry as a function of angle. With increased statistical precision, comes the need for improved systematic control. The most important of these being the extraction of the charge mistag factor for $b$ quark from data. {\sffamily ALEPH, DELPHI} and {\sffamily OPAL} now perform this extraction while {\sffamily ALEPH} also extracts it as a function of the polar angle of the thrust axis. This takes into account particle losses close to the edge of detector acceptance. The output distributions from the {\sffamily ALEPH} measurement are shown in Figure~\ref{figure-alout}. \begin{figure}[h] \begin{center} \mbox{\epsfig{file=qfb_2.eps,width=8.6cm}} \end{center} \caption{\em Extracted fit variables from the {\sffamily ALEPH} lifetime+jetcharge measurement of $A_{\rm FB}^b$. Columns represent the $b$ charge mistag factor and sample $b$ purity as a function of angle in samples with increasing $b$ purity. Filled and open circles represent $\kappa$ values of 0.5 and 2.0 in equation~(4) respectively.} \label{figure-alout} \end{figure} The asymmetry measurement is made separately in each bin of polar angle, $\kappa$ and $b$ sample purity before being combined. Some points of interest in such new analyses include the increased statistical power arising from the bins at low angles, even those where the thrust axis lies outside the {\sffamily VDET} acceptance. The large value of the asymmetry in these regions compensating for the low tagging efficiencies. The increase in sample $b$ purity at large $cos \, \theta$ for high $b$ purity samples, is due to the loss of tracks at the edge of the {\sffamily VDET} acceptance. This affects $b$ events the least, as more tracks with large $p_{\rm T}$ to the thrust axis continue to tag the event. This is true to a much lesser extent for lighter quark flavours. Measurements of the $b$ asymmetry using the lifetime and jetcharge method are also summarised in Table~\ref{table-basyms}. It is interesting to note that such measurements, whilst providing similar sensitivity to $A_{\rm FB}^b$ as semileptonic $b$ decays, as yet do not provide the possibility of analogous measurements of $A_{\rm FB}^c$. \section{Latest Techniques for Measuring the $c$ Asymmetry} In contrast to the incremental progress in the field of $b$ asymmetries, measurements of the corresponding quantity in $c$ decays have improved dramatically in recent years. Of the 4 {\sffamily LEP} collaborations, {\sffamily DELPHI, L3} and {\sffamily OPAL} have determined the $c$ asymmetry as an output of global semileptonic fits to $b$ and $c$ decays\cite{delphi:leptons,l3:leptons,opal:leptons} whereas {\sffamily ALEPH, DELPHI} and {\sffamily OPAL} have performed the same measurement using fully reconstructed samples of $D$ meson decays\cite{aleph:D,delphi:D,opal:D}. As semileptonic measurements are discussed in Section~\ref{semilepton}, only the latter are described here. The method of exclusively reconstructing $D$ decays aim to use as many channels as possible by reconstructing the $D^0$ through its decay to $K^-\pi^+$. The $D^0$ is generally reconstructed\cite{aleph:D} by taking all 2 and 4 track combinations and a $\pi^0$ candidate with zero total charge. Those combinations with odd charges are then used to form possible $D^+$ candidates. Each experiment reconstructs a different subset out of a total of 9 different decay modes. The dominant channels however are the $D^+ \rightarrow K^- \pi^+ \pi^+$ and $D^{*+}$ modes, with all modes offering some statistical power. Similarly, each experiment has different selection criteria in each mode, depending on the momentum and particle identification resolutions of the detectors. These differences lead to widely varying efficiencies and \mbox{signal-to-background} ratios. For example, the {\sffamily DELPHI} mass difference distributions for 4 of the 8 selected modes are shown in Figure~\ref{figure-delphidmass}. \begin{figure}[h] \begin{center} \mbox{\epsfig{file=delphid.ps,width=8.6cm}} \end{center} \caption{\em Difference distributions between the mass of the $D^{*+}$ and the $D^0$ candidate for different decay modes.} \label{figure-delphidmass} \end{figure} An important advantage of such measurements is the ability to determine the background asymmetry, $A_{\rm FB}^{\rm bkg}$, from data, \mbox{mode-by-mode} using information from sidebands. A major difficulty is encountered when trying to correct for the substantial fraction of events due to $B^0 \rightarrow D$ decays for $B^0-\bar{B^0}$ mixing. Each $D$ mode is corrected using an ``effective'' $\bar{\chi}$ depending on the expected fractions of $B_{\rm d}^0$ and $B_{\rm s}^0$ decays contributing to the mode concerned. These effective factors are determined from\cite{pdg} using Monte Carlo simulation and so give rise to systematic uncertainties. The observed asymmetry, $A_{\rm FB}^{\rm obs}$, is then found using~: \begin{equation} A_{\rm FB}^{\rm obs} \: = \: f_{\rm sig} f_c A_{\rm FB}^c \, + \, f_{\rm sig} ( 1 - f_c) A_{\rm FB}^b \, + \, ( 1 - f_{\rm sig} ) A_{\rm FB}^{\rm bkg} \end{equation} where $f_{\rm sig}$ is the fraction of signal of \mbox{signal+background} events and $f_c$ is the fraction of events containing a true $D$ meson which are due to $c$ quark events. As far as possible, the sample $c$ purities, $f_c$, are determined from data using lifetime, mass and \mbox{event-shape} information in both hemispheres of events containing a $D$ tag. Each experiment makes use of lifetime information, with both {\sffamily DELPHI} and {\sffamily OPAL} using the $D$ momentum and \mbox{jet-shape} information respectively in addition, so as to disentangle the substantial contamination from $b$ decays. The {\sffamily DELPHI} experiment does so in the context of a simultaneous fit to both $b$ and $c$ asymmetries. However, besides constraining systematic uncertainties, the precision available of $A_{\rm FB}^b$ is negligible compared to methods discussed previously. The $c$ asymmetry measurements described here are summarised in Table~\ref{table-casyms} and indicate that, despite complex systematics, such measurements remain primarily limited by the low efficiencies and purities of the $D$ meson reconstruction. \begin{table}[h] \begin{center} \begin{tabular}{|l|llll|} \hline \multicolumn{5}{|c|}{\em Semileptonic Measurements of $A_{\rm FB}^c$} \\ \hline {\em Experiment} & $A_{\rm FB}^b$ & {\em stat.} & {\em syst.} & {\em total} \\ \hline {\sffamily DELPHI} & 0.0770 & $\pm$0.0113 & $\pm$0.0071 & $\pm$0.0133 \\ {\sffamily L3 } & 0.0784 & $\pm$0.0370 & $\pm$0.0250 & $\pm$0.0446 \\ {\sffamily OPAL } & 0.0595 & $\pm$0.0059 & $\pm$0.0053 & $\pm$0.0079 \\ \hline \multicolumn{5}{|c|}{\em Exclusive D Tag Measurements of $A_{\rm FB}^c$} \\ \hline {\em Experiment} & $A_{\rm FB}^b$ & {\em stat.} & {\em syst.} & {\em total} \\ \hline {\sffamily ALEPH } & 0.0630 & $\pm$0.0090 & $\pm$0.0030 & $\pm$0.0095 \\ {\sffamily DELPHI} & 0.0658 & $\pm$0.0093 & $\pm$0.0042 & $\pm$0.0102 \\ {\sffamily OPAL } & 0.0630 & $\pm$0.0120 & $\pm$0.0055 & $\pm$0.0132 \\ \hline \end{tabular} \caption{\em Summary of latest measurements of $A_{\rm FB}^c$ from the 4 {\sffamily LEP} experiments.} \label{table-casyms} \end{center} \end{table} Systematic errors vary widely, with the \mbox{time-dependence} of the background asymmetry remaining merely one of many dominant sources depending on decay mode and experiment. \section{Radiative Corrections to Asymmetry Measurements} Several small corrections must be made to the $b$ and $c$ asymmetries extracted from the described analyses. In the case of $A_{\rm FB}^b$, {\sffamily QED} corrections for {\sffamily ISR} and {\sffamily FSR} are relatively minor, amounting to -0.0041 and -0.00002 respectively. Similarly, corrections for pure $\gamma$ exchange and $\gamma-Z$ interference diagrams give rise to a correction of +0.0003. Such corrections, are general in nature and so apply equally to all analyses. A more difficult set of corrections involves those needed to correct for the presence of hard gluon radiation which can distort the angular distribution of the final state quarks when compared with the pure electroweak process. Estimates of such corrections to heavy quark asymmetries have been computed to first and second order in $\alpha_{\rm s}$ both numerically\cite{lampe} and, most recently, analytically\cite{neerven} in different scenarios for either $c$, $b$ or massless quarks. A common procedure for correcting and ascribing systematic uncertainties for the {\sffamily LEP} heavy quark asymmetries has been developed\cite{qcdcorr}. The more recent analytical calculations indicate several discrepancies when compared with the numerical results. These remain to be resolved. Current systematic uncertainties are determined using a procedure of comparing the effects between first and second order in {\sffamily QCD} and by switching between massless quarks, and assumptions for the $c$ and $b$ quark masses. Further difficulties arise when considering the application of such corrections to individual analyses. Theoretical calculations are typically based on the direction of the outgoing quark, whereas the analyses described here use the thrust direction. Further, the sensitivity to hard gluon radiation of data, containing either a lepton of a given $(p, \, p_{\rm T})$, a reconstructed $D$ meson or purely inclusive events, varies dramatically. Effects of \mbox{non-perturbative} {\sffamily QCD} and \mbox{higher-order} effects during hadronisation must also be evaluated. The latter render {\sffamily QCD} corrections both detector {\em and} analyses dependent, {\em eg.} event shape selections implying an implicit dependence on the strength of gluon emission. The correction to be applied to a given analysis is derived from~: \begin{equation} A_{\rm FB}^{b,c} \: = \: ( 1 - {\cal C}_{b,c} {\cal S}_{b,c} ) \mid_{\rm no \ QCD} \end{equation} where ${\cal C}_{b,c}$ represents the {\sffamily QCD} correction at parton-through-to-hadron level, and ${\cal S}_{b,c}$ is the analysis dependent modification. Examples of the magnitude of the {\sffamily QCD} corrections at the theoretical and experimental levels are shown in Table~\ref{table-qcdcorr} for the cases of $A_{\rm FB}^b$, determined using semileptonic and lifetime+jetcharge analyses\cite{qcdcorr}. \begin{table}[h] \begin{center} \begin{tabular}{|lc|c|c|} \hline & & {\em Lepton Analyses} & {\em Lifetime+Jetcharge} \\ \hline ${\cal S}_{b}$ & min. & 0.52 $\pm$0.06 &0.24 $\pm$0.46 \\ & max. & 0.74 $\pm$0.07 &0.36 $\pm$0.32 \\ \hline ${\cal C}_{b}$ & min. & 1.54 $\pm$0.28 &0.71 $\pm$1.36 \\ & max. & 2.19 $\pm$0.37 &1.07 $\pm$0.96 \\ \hline \end{tabular} \caption{\em Summary of {\sffamily QCD} corrections to $A_{\rm FB}^b$ for the different analysis methods.} \label{table-qcdcorr} \end{center} \end{table} The constants are evaluated in terms of Monte Carlo simulations, before and after experimental cuts. The hadronisation dependence of corrections is included as a systematic uncertainty by comparing results from both the {\tt HERWIG}\cite{herwig} and {\tt JETSET}\cite{jetset} models. It is seen from comparing parton and hadron level that the effect of hadronisation is to reduce the magnitude of the {\sffamily QCD} correction\footnote{It is thought that \mbox{non-perturbative} colour reconnection effects during the shower may be responsible.}. It is important to note that the corrections for lifetime+jetcharge measurements are negligible. Significant corrections are observed for both semileptonic,and $D$ tag measurements of both $A_{\rm FB}^b$ and $A_{\rm FB}^c$. Jetcharge measurements are immune to such corrections as the \mbox{$b$-quark} charge mistag factor is defined using Monte Carlo with respect to the original $b\bar{b}$ quark pair orientation, prior to gluon or final state photon radiation, parton shower, hadronisation and $B^0 \bar{B^0}$ mixing. All these effects are therefore included, by construction, in the analyses, as far as they are properly modelled in the {\tt JETSET}\cite{jetset} hadronisation model. \section{Conclusion and Perspectives} With the completion of {\sffamily LEP} \mbox{data-taking} at energies close to the Z resonance in 1995, the 4 experiments ({\sffamily ALEPH, DELPHI, L3} and {\sffamily OPAL}), have accumulated large samples of hadronic events. From this data, the \mbox{forward-backward} asymmetry of the $b$ quark has emerged as the most sensitive single test of the {\sffamily SM} at {\sffamily LEP}. The complementary $c$ asymmetry measurements offer additional precison and a new window on couplings in the \mbox{down-type} quark family. The precision from semileptonic measurements of $A_{\rm FB}^b$ is now matched by that of \mbox{lifetime+jetcharge} measurements. The electroweak sensitivity of $A_{\rm FB}^c$ measurements now equals that obtained from combined quark asymmetries measured in untagged samples, highlighting the benficial effect of flavour tagging. However, in light of these measurements great sensitivity to the couplings of the {\sffamily SM}, the continuing discrepancy between electroweak results from {\sffamily LEP} and {\sffamily SLD}\cite{altarelli} make it essential to understand whether it is due to statistical fluctuations or systematic effects. Separating {\sffamily LEP} measurements of $A_{\rm FB}^b$ into those from the two dominant techniques, and conservatively ignoring correlated systematic uncertainties, indicates that there is at most a $1.2\sigma$ discrepancy between semileptonic and \mbox{lifetime+jetcharge} measurements. This is insufficient to explain the {\sffamily LEP-SLD} discrepancy but indicates that care must be taken when considering common systematics in leptonic decay modelling and fragmentation uncertainties. Further improvements in both $b$ and $c$ asymmetries are possible, as both sets of measurements are still dominated by statistics. For semileptonic analyses, the benefits of using both lifetime and lepton information are emphasised. In the case of the lifetime and jetcharge method, these are most likely to come in the form of improved $b$ tagging efficiencies and extensions of tagging to lower angles. The situation for improvments to measurements of $A_{\rm FB}^c$ is more difficult as the number of available modes is exhausted, and efficient methods of tagging $c$ events remain to be discovered. At this point, without the prospect of significant, further {\sffamily LEP} \mbox{data-taking} at the Z, it is important to focus upon the latest techniques which offer the greatest sensitivity to the couplings of the {\sffamily SM} combined with systematic control. The measurements described here obtain combined precisions on the $b$ and $c$ asymmetries of 2.2\% and 7.1\% respectively. Hence, the goal of acheiving similar precision on the Z couplings to quarks, as that obtained for leptons, has been reached.
\section{Introduction} \begin{sloppypar} A model of Computation and Complexity over a ring was developed in~\cite{BSS} and ~\cite{BCSS}, generalizing the classical $\mathcal{NP}$-completeness theory~\cite{GAREY-JOHNSON}. Of particular interest is the model of Complexity over the ring $\mathbb C$ of complex numbers. \end{sloppypar} \par In the model of complexity over $\mathbb C$, a machine is allowed to input, to output and to store complex numbers, to compute polynomials and to branch on equality (See the textbook~\cite{BCSS} for background). This model shares some of the features of the classical (Turing) model of computation (There is a discussion in~\cite{MM}). It is known~\cite{KOIRAN97,LECTURE} that the hypothesis $\mathcal {BPP} \not \supseteq \mathcal{NP}$ in the Turing setting implies $\pnp$ over $\mathbb C$. ($\mathcal {BPP}$ stands for Bounded Probability Polynomial Time. If $\mathcal {BPP}$ would happen to contain $\mathcal{NP}$, then there would be polynomial time randomized algorithms for such tasks as factorizing large integers or breaking most modern cryptographic systems). \medskip \par In \cite{Shub-Smale, BCSS, SmaleXXI}, the hypothesis $\pnp$ over the Complex numbers was related to a number-theoretical conjecture. Define a straight-line program as a list \[ s_0 = 1\ ,\ s_1 = x\ ,\ s_2\ , \ \cdots\ , \ s_{\tau} \] where $s_i$ is, for $i\ge 2$, either $s_j+s_k$, $s_j-s_k$ or $s_j s_k$, for some $j, k < i$. Each $s_i$ is thus a polynomial in $x$. The straight-line program is said to {\em compute} the polynomial $s_{\tau}(x)$. \par Given a polynomial $f \in \mathbb Z [x]$, the quantity $\tau(f)$ is defined as the smallest $\tau$ such that there exists a straight-line program $s_0, \cdots, s_{\tau}$ computing $f(x)$. For instance, $\tau(x^{2^n}-1) = 2+n$. Similarly, if $g \in \mathbb Z[x_1, \cdots , x_n]$, then $\tau(g)$ is the minimal length of a straight-line program $s_0=1, s_1=x_1, \cdots, s_n=x_n, s_{n+1}, \cdots, s_{\tau}=g(x)$. \begin{tauconjecture} There is a constant $a > 0$ such that for any univariate polynomial $f \in \mathbb Z[x]$, \[ n(f) < \tau(f)^a \] where $n(f)$ is the number of integer zeros of $f$, without multiplicity. \end{tauconjecture} \medskip \par It is known~\cite{BCSS} that the $\tau$-Conjecture for polynomials implies $\pnp$ over $\mathbb C$. A main step towards this result is the fact that, if the $\tau$-Conjecture is true, then the polynomials \[ p_d(x) = (x-1) (x-2) \cdots (x-d) \] are {\em ultimately hard to compute}. This means that there cannot be constants $a$ and $b$ such that, for any degree $d$, for some non-zero polynomial $f$ (depending on $d$), we would have \[ \tau \left(\ p_d (x) f(x) \ \right) < a \left( \log_2 d \right) ^b \] \par Therefore, all non-zero multiples of $p_d$ are hard to compute, hence the wording ultimately hard. \medskip \par The goal of this paper is to define a new complexity class $\mathcal{UP}$, of {\em ultimate polynomial time} problems. This class will contain $\mathcal P \cap \mathcal K$, where $\mathcal P$ is the class of problems decidable in polynomial time and $\mathcal K$ is the class of problems definable without constants (See~\cite{KOIRAN97b} and Definition~\ref{K} below). Moreover: \begin{theorem} \label{th1} The implications (a) $\Rightarrow$ (b) $\Rightarrow$ (c) $\Rightarrow$ (d) are true: \begin{itemize} \item [(a)] The $\tau$-conjecture for polynomials. \item [(b)] $\forall d$, $p_d$ is ultimately hard to compute. \item [(c)] $\upnp$ over $\mathbb C$. \item [(d)] $\pnp$ over $\mathbb C$. \end{itemize} \end{theorem} \par The implication (a) $\Rightarrow$ (b) $\Rightarrow$ (d) appears in ~\cite{BCSS}, the hypothesis (c) in-between is new. It is at least as likely as the $\tau$-conjecture, while still implying $\pnp$. \par We will also show a $\mathcal{NP}$-hardness result for the class $\mathcal{UP}$: there is a structured problem $\nullstellensatz \in \mathcal{NP} \cap \mathcal K$, such that: \begin{theorem} \label{th2} $\upnp$ over $\mathbb C$ if and only if $\nullstellensatz \not \in \mathcal{UP}$ over $\mathbb C$. \end{theorem} \par The problem $\nullstellensatz$ is precisely the (structured) Hilbert Nullstellensatz, known to be $\mathcal{NP}$-complete over $\mathbb C$~(\cite{BCSS}). \medskip \par This paper was written while the author was visiting Mathematical Sciences Research Institute in Berkeley. The author also wishes to thank Pascal Koiran and Steve Smale for their comments and suggestions. \section{Background and Notations} Recall from $\cite{BCSS}$ that $\mathbb C^{\infty}$ is the disjoint union \[ \mathbb C^{\infty} = \bigsqcup_{i=0, 1, \cdots} \mathbb C^i \] This means that there is a well-defined {\em size} function, \[ \begin{array}{crcl} \size: & \mathbb C^{\infty} & \rightarrow & \mathbb N \\ & x & \mapsto & \size(x) = i \text{\ such that \ } x \in \mathbb C^i \end{array} \] \par A decision problem $X$ is a subset of $\mathbb C^{\infty}$. It is in the class $\mathcal P$ if and only if there is a machine $M$ over $\mathbb C$, that terminates for any input $x$ in time bounded by a polynomial on $\size(x)$, and such that \[ M(x) = 0 \ \Leftrightarrow \ x \in X \] where $M(x)$ is the result of running $M$ with input $x$. Without loss of generality we may assume that $M(x) \in \{ 0;1 \}$. \medskip \par Under some circumstances, it is possible to assume that the machine $M$ above has only coefficients 0 or 1 (This is called a {\em constant-free} machine). However, one may have to replace problem $X$ over $\mathbb C$ by problem $X \cap \mathbb Z$ over $\mathbb Z$, with unit cost. (This is the contents of Propositions~3 and~9 of Chapter~7 of ~\cite{BCSS}). In order to avoid this technical complication and keep the same problem over $\mathbb C$, we will follow another approach to Elimination of Constants. \par This approach was introduced by Koiran in ~\cite{KOIRAN97b}. The idea is to consider only machines for a subclass of problems. This subclass will contain most of the interesting examples, while precluding pathological cases such as $X = \{ \pi \}$. \begin{definition}[Koiran] \label{K} A problem $L$ is said to be {\em definable without constants} if for each input size $n$ there is a formula $F_n$ in the first order theory of $\mathbb C$ such that $0$ and $1$ are the only constants occurring in $F_n$, and for any $x \in \mathbb C^n$, $x \in L$ if and only if $F_n(x)$ is true (there is no restriction on the size of $F_n$. \end{definition} \par For future reference, we quote below Theorem~2 of~\cite{KOIRAN97b}. The original statements of both Definition~\ref{K} and Theorem~\ref {ELIM} are actually more general (for any algebraically closed field of characteristic 0). \begin{theorem}[Koiran] \label{ELIM} Let $L \subseteq K^\infty$ be a problem which is definable without constants. If $L \in \mathcal P$, $L$ can be recognized in polynomial time by a constant-free machine. \end{theorem} \par The class of all the problems definable without constants will be denoted by $\mathcal K$. \medskip \par We will need crucially in the sequel the notion of a {\em structured problem}. A structured problem is a pair $(X, X\yes)$, $X\yes \subseteq X \subseteq \mathbb C^{\infty}$. A non-structured problem $X$ can always be written as the structured problem $(\mathbb C^{\infty}, X)$. The class $\mathcal{UP}$ will be meaningful only as a class of structured problems. But first of all, recall that \begin{definition} A structured problem $(X, X\yes)$ belongs to the class $\mathcal {P}$ if and only if $X \in \mathcal P$ and $X\yes \in \mathcal P$. \end{definition} \begin{definition} A structured problem $(X, X\yes)$ belongs to the class $\mathcal {K}$ if and only if $X \in \mathcal K$ and $X\yes \in \mathcal K$. \end{definition} \begin{definition} A structured problem $(X, X\yes)$ belongs to the class $\mathcal {NP}$ if and only if: \begin{itemize} \item [(1)] The problem $X$ belongs to the class $\mathcal P$. \item [(2)] There is a machine $M$ with input $x,g$ such that \[ x \in X \text{ and } \exists g \in \mathbb C^{\infty} \text { s.t. } M(x,g)=0 \ \Leftrightarrow \ x \in X\yes \] \item [(3)] Furthermore, there is a polynomial $p$ such that, for all $x \in X\yes$, there is $g \in \mathbb C^{\infty}$ such that $M(x,g)=0$ and the running time of $M$ with input $x,g$ is no more than $p(\size(x))$. \end{itemize} \end{definition} \begin{example}\label{ex1} Let $\hn$ be the class of all lists $(m, n, f_1, \cdots,f_m)$ where $f_1, \cdots,$ $f_m$ are polynomials in $n$ variables. Each polynomial $f = \sum f_I x^I$ is represented sparsely by a list of monomials $(S, m_1, \cdots, m_S)$, where each monomial is a list $(f_I, I_1, \cdots, I_n)$. \par An important convention to have in mind: integers appearing in the definition of a problem should be represented in bit representation. In this case, $m, n, S, I_j$ are all lists of zeros and ones. Complex values are represented by one complex number. With this convention, $\hn$ is clearly in the class $\mathcal P$. \par We also define $\hn \yes$ as the subset of polynomial systems in $\hn$ that have a common root over~$\mathbb C$. \par The definition above of the structured problem $\nullstellensatz$ can be translated into first order constant-free formulae over $\mathbb C$. Therefore, $\nullstellensatz \in \mathcal K$. It is also $\mathcal{NP}$-complete over the complex numbers (Theorem~1 in Chapter~5 of ~\cite{BCSS}). \end{example} \begin{example}\label{ex2} Let \begin{eqnarray*} X &=& \left\{ (m,x) \in \mathbb N \times \mathbb C \right\} \\ X\yes &=& \left\{ (m,x) \in X \text{ such that } x \in \{ 1, 2, \cdots, m \} \right\} \\ \end{eqnarray*} with the convention that $m$ is in bit representation, while $x$ is a complex number. Hence, $\size((m,x)) = O(1 + \lceil \log_2 (m) \rceil)$. Then the problem $(X, X\yes)$ is in $\mathcal {NP}$ over $\mathbb C$. The machine $M(x,g)$ can be constructing by guessing the bit decomposition $g_i$ of $x$, and computing $x - \sum g_i 2^i$. \par Again, $(X,X\yes)$ is definable without constants. \end{example} \section{Construction of the class $\mathcal {UP}$} \par In Chapter 7 of~\cite{BCSS}, it is proved that if the problem $(X, X\yes)$ from Example~\ref{ex2} would happen to belong to the class $\mathcal {P}$, then condition (b) in Theorem ~\ref{th1} would be false. Therefore (b) implies \pnp over $\mathbb C$. \medskip \par The class $\mathcal{UP}$ will be constructed by abstracting the same reasoning. The construction relies on some geometric properties of structured problems in $\mathcal P$. The notation that follows will be used in the sequel: \par Let $(X, X\yes)$ be a structured problem with $X \in \mathcal P$. We denote by $X \cap \mathbb C^i$ the set $\{ x \in X: \size(x) = i \}$ of size $i$ instances of the problem. Then we write $\overline {X \cap \mathbb C^i}$ for its Zariski closure over $\mathbb C$. We can define a new object associated to $X$ as: \[ \overline{X} = \bigsqcup_{i=0,1, \cdots} \overline {X \cap \mathbb C^i} \] \par We can think of $\overline{X}$ as the {\em closure} of $X$, indeed it is the smallest `closed' problem containing $X$. Remark that in Examples~\ref{ex1} and~\ref{ex2}, we have respectively $X=\overline{X}$ and $\hn=\overline{\hn}$. \par We can also decompose each Zariski-closed set $\overline {X \cap \mathbb C^i}$ into a finite union of irreducible components (affine varieties). Thus it makes sense to write $\overline{X}$ as the countable union: \[ \overline{X} = \bigcup X_j \] where each $X_j$ is an affine variety lying in some $\mathbb C^s$, where $s = \size (x), x \in X_j$. We can further define: \begin{eqnarray*} X_j\yes &=& X_j \cap X\yes \\ X_j\no &=& X_j \setminus X\yes \end{eqnarray*} \par \begin{figure} \centerline{ \resizebox{4cm}{7cm}{\includegraphics{up3.eps}} \hspace{2cm} \begin{minipage}[b]{4cm} {\sf \footnotesize This is Problem $(X,X\yes)$ from Example~\ref{ex2}, restricted to the inputs $(m_0, m_1, x)$ of size $3$. $X$ is represented by the four (complex !) lines and $X\yes$ by the dots. Each of the complex lines is irreducible, and hence corresponds to a different $X_i$.} \vspace{1cm} \end{minipage} \hspace{2cm} } \caption{\label{fig1} $(X,X\yes)$ from Example~\ref{ex2}} \end{figure} \par (See Figure~\ref{fig1}). Using this notation, \begin{definition}\label{dup} The class $\mathcal{UP}$ is the class of all structured problems $(X, X\yes)$ such that $X \in \mathcal{P}$ and for all $X_i$, there is a non-zero polynomial $f_i \in \mathbb Z[x_1, \cdots, x_{s_i}]$, where $s_i = \size (x)$ for $x \in X_i$, with the following properties: \begin{itemize} \item [(1)] $\tau(f_i)$ is polynomially bounded in $S_i$. \item [(2)] $X_i\yes \subseteq Z(f)$ or $X_i\no \subseteq Z(f)$ \end{itemize} \end{definition} \medskip \par \begin{proposition}\label{pup} $\mathcal P \cap \mathcal K \subseteq \mathcal {UP}$ \end{proposition} \begin{proof}[Proof of Proposition~\ref{pup}] Let $(X,X\yes)$ be in $\mathcal P \cap \mathcal K$. Let $M = M(x)$ be the machine that recognizes $x \in X\yes$ in polynomial time, where the input $x$ is assumed to be in $\overline{X}$. Although it is possible that an $x \in X_i$ is not in $X$, it is still possible to recognize $x \in X\yes$ in polynomial time. Indeed, $X$ is also in $\mathcal P$. The machine $M(x)$ will check $x \in X$ and $x \in X\yes$. \par Now we apply elimination of constants (Theorem~\ref{ELIM}), and choose $M$ to be constant-free. \par The nodes of the machine $M$ are supposed to be numbered. Given an input $x$, the {\em path} followed by input $x$ is the list of nodes traversed during the computation of $M(x)$. \par When the input is restricted to one of the affine varieties $X_i$'s, we can define the canonical path (associated to $X_i$ as the path followed by the generic point of $X_i$. This corresponds to the following procedure: \par At each decision node, at time $T$, branch depends upon an equality $F^T(x) = 0$, where $x$ is the original input. The polynomial $F$ can be computed within the machine running time. In case $F^T(x) = 0$ for all $x \in X_i$, we follow the Yes-path and say that this branching is trivial. \par If not, we follow the no-path and say that this branching is non-trivial. The fact that $X_i$ is a variety is essential here, since it guarantees that only a codimension $\ge 1$ subset of inputs may eventually follow the Yes-path at this time. \par The set of inputs that do NOT follow the canonical path can be described as the zero-set of \[ f_i = \prod F^T \] where the product ranges over the non-trivial branches only. The polynomial $f_i$ can be computed in at most twice the running time of the machine $M$ restricted to $X_i$. By hypothesis, this is polynomial time in the size of $x \in X_i$. \par Since we assumed that $M$ returns only $0$ or $1$, the set of the inputs that follow the canonical path (i.e. $Z(f_i)$) is either all in $X_i\yes$ or all in its complementary $X_i\no$. \par There are now two possibilities. First possibility, $X_i\yes$ has measure zero in $X_i$, and therefore it must be contained in $Z(f_i)$. Second possibility, $X_i\yes$ has non-zero measure, hence it contains the complementary of $Z(f_i)$, and hence $X_i\no$ is a subset of $Z(f_i)$. \end{proof} \section {Proof of the Theorems} \begin{proof}[Proof of Theorem~\ref{th1}] \medskip \ {} \par (a) $\Rightarrow$ (b) is trivial, refer to ~\cite{BCSS} Chapter 7. \medskip \par (b) $\Rightarrow$ (c): Let $(X,X\yes)$ be the problem in Example~\ref{ex2}. Since $X_i\no$ is generic in $X_i$, all inputs in $X_i\yes$ should escape the canonical path. Hence, if $f_d$ is the polynomial that defines the canonical path, $f_d(i) = 0$ for $i=1,2, \cdots, d$. But then it cannot be evaluated in time polylog($d$), by hypothesis (b). Hence, under the assumption (b), the problem $(X,X\yes)$ is not in $\mathcal {UP}$. It does belong to $\mathcal {NP} \cap \mathcal K$, so $\upnp$. \medskip \par (c) $\Rightarrow$ (d) : Using Theorem~\ref{th2}, Condition (c) implies that $\nullstellensatz \not \in \mathcal{UP}$. However, since $\nullstellensatz \in \mathcal K$, Proposition~\ref{pup} implies $\nullstellensatz \not \in \mathcal P$. Hence \pnp over $\mathbb C$. \end{proof} \begin{proof}[Proof of Theorem~\ref{th2}] Let $(X, X\yes) \in \mathcal{NP} \cap \mathcal{K}$ and assume that $\nullstellensatz \in \mathcal{UP}$. We have to show that $(X, X\yes) \in \mathcal{UP}$. \par For each $X_i$, one can embed $(X_i, X_i\yes)$ into some $(\hn_i, \hn_i\yes)$ as follows: \par Let $M = M(x)$ be the deterministic polynomial time machine to recognize $X$, and let $N=N(x,g)$ be the non-deterministic polynomial time machine to recognize $X\yes$. We can assume without loss of generality that $M$ and $N$ are constant-free (Theorem~\ref{ELIM}). \par Let $T$ be the maximum running time of $M$ and $N$ when the input is restricted to $X_i$. Let $\phi(x)$ be the combined Register Equations of machines $M$ and $N$ for time $T$ (Theorem~2 in Chapter~3 of~\cite{BCSS}). Thus, $\phi(x)$ is a system of polynomial equations with integer coefficients and indeterminate coefficients $x_1, x_2, \cdots$. The polynomial system $\phi(x)$ can be constructed in polynomial time from $x$, and the size of $\phi(x)$ is polynomially bounded by the size of $x$. \medskip \par We claim that $\phi(X_i)$ is contained in some $\hn_j$, and that in that case $\phi(X_i\yes) \subseteq \hn_j\yes$ and $\phi(X_i\no) \subseteq \hn_j\no$. \par Indeed, $X_i \subseteq \mathbb C^s$ for some $s$, and $\phi(\mathbb C^s) \subseteq HN_j$ for some $j$. Then $x \in X_i$ belongs to $X\yes$ if and only if the corresponding $\phi(x)$ has a solution over $\mathbb C$. \medskip \par We now distinguish two cases: \par Case 1: $\hn_j\yes$ has measure zero in $\hn_j$. Thus $\hn_j\yes \subseteq Z(\hat f_j)$ for an easy-to-compute polynomial $\hat f_j$. In that case, since $X_i\yes$ gets mapped into $\hn_j\yes$, the composition $f_i = \hat f_j \circ \phi$ gives the polynomial associated to $X_i$. \par Case 2: $\hn_j\no$ has measure zero in $\hn_j$. Thus $\hn_j\no \subseteq Z(\hat f_j)$ for an easy-to-compute polynomial $\hat f_j$. In that case, since $X_i\no$ gets mapped into $\hn_j\no$, $f_i = \hat f_j \circ \phi$ is the polynomial associated to $X_i$. \end{proof} \section{Ultimate Complexity} Let $(Y,Y\yes)$ be a problem over $\mathbb C$, definable without constants and with $Y$ semi-decidable (i.e. $Y$ is the halting set of some machine). The closure $\overline Y$ is well-defined and can be written as a countable union of irreducible varieties $Y_i$. \par For any machine $M$ to solve $(Y,Y\yes)$, one can produce a family of polynomials $f_i$, vanishing on the set of inputs that follow the canonical-path of $M$ restricted to $Y_i$. As in item (2) of Definition~\ref{dup}, we have \[ Y_i\yes \subseteq Z(f_i) \text{\ or\ } Y_i\no \subseteq Z(f_i) \] \par Also, for each input size $s$, one has a finite number of indices $i$ corresponding to components i $Y_i \subseteq \overline Y$ of size-$s$ input. We can thus maximize over those indices $i$: \[ u_M (s) = \max_{i:Y_i \subseteq \mathbb C^{i}} \tau(f_i) \] \par This invariant may be called `ultimate running time', and is a lower bound (up to a constant) for the worst-case running time of $M$. As with ordinary complexity theory, one can define the `ultimate complexity' class of a problem as the class of functions $u: \mathbb N \rightarrow \mathbb R$ such that $\exists M, c>0: \forall x u_M(x) \le c u(x)$ and $M$ recognizes $(Y, Y\yes)$. This provides notions such as `ultimate logarithmic time' or `ultimate exponential time'. \par In~\cite{M99}, a similar construction is used to obtain lower bounds for some specific decision problems. Those problems, however, had a very simple geometric structure (for each `input size', $X\yes$ was a finite set in $\mathbb C$). The motivation of this paper was to extend some of the ideas therein and in Chapter~7 of~\cite{BCSS} to non-codimension-1 problems. \bibliographystyle{plain}
\subsection*{Two-dimensional random tilings} Let us now move to planar systems, where we will mainly consider two illustrative examples, namely the classical random tilings consisting of dominoes and lozenges. Because of their symmetries, we call them crystallographic random tilings. Though still only two-dimensional, they are of practical relevance because of the existence of so-called T-phases (see \cite{Baake} and references therein) which are irregular planar layers stacked periodically in the third direction due to a very anisotropic growth mechanism. It is thus appropriate to investigate the diffraction spectrum of a single layer obtaining then the complete spectrum once again as a product measure, compare the previous Section. Unfortunately, already the treatment of planar systems is a lot more involved than in the 1D case. Although we will have to deal ``only'' with the action of $\mathbb Z^2$, we cannot directly apply standard results of ergodic theory as above, because we first have to establish the ergodicity of the measures involved. Even for the two simple systems we shall discuss below, the rigorous classification of invariant measures is only in its infancy, see \cite{BP} and the discussion in \cite{Ken1}. Fortunately, the investigation of invariant equilibrium states, which form a subclass, is well developed \cite{Georgii,Israel,Simon}. If combined with certain results of statistical mechanics \cite{Ruelle}, this allows for a determination of extremal states, which will be unique in our examples. They are ergodic and thus admit the application of Birkhoff's pointwise ergodic theorem in its version for $\mathbb Z^2$-action, see e.g.\ \cite[Thm.\ 2.1.5]{Keller}. Let us summarize the key features in a way adapted to our later examples. \subsubsection*{Preliminaries} A {\em tiling} $\omega$ of a region\footnote{We tacitly assume that any such region is sufficiently nice, i.e.\ it should be compact, measurable and simply connected.} $\Lambda\subset\mathbb R^d$ (with positive volume ${\rm vol}(\Lambda)$) is a countable covering of $\Lambda$ by tiles, i.e.\ by bounded closed sets homeomorphic to balls, having pairwise disjoint interiors and non-vanishing overlap with the region $\Lambda$. In our case, the tiles are translates of finitely many prototiles. We deal with free boundary conditions in the sense that the tiles may protrude beyond the boundary. Thus, the boundary of $\Lambda$ does not impose any restrictions of the kind known from fixed boundary conditions or exact fillings of given patches, compare \cite{Ken2}. Two coverings of the same region $\Lambda$ are called {\em equivalent} if they are translates of one another. For further conceptual details, we refer to \cite{Richard}. The examples discussed below belong to the class of polyomino tilings, where the prototiles are combinations of several elementary cells of a given periodic graph $G$. These tilings can be described as polymer models \cite{Richard}, and as dimer models in our case. \begin{figure}[ht] \centerline{\epsfysize=3cm \epsfbox{figs/dual.eps}} \caption{\label{dual}Representation of a random tiling on the dual cell complex.} \end{figure} A {\em dimer} is a diatomic molecule occupying two connected sites of a graph. A graph is close-packed if all sites are occupied precisely once. In Figure~\ref{dual}, the one-to-one correspondence between the tiling on a (periodic) graph and its close-packed dimer configuration on the dual cell complex (the so-called Delone complex) is illustrated for the domino tiling. The scatterers (Dirac unit measures) are placed in the centre of the tiles, resp.\ dimers. In the space $\Omega$ of all tilings, two elements are close if they agree on a large neighbourhood of the origin. $\Omega$ is compact in this topology. The group $\mathbb Z^2$ of translations acts continuously on $\Omega$ (in an appropriate parametrization) because of the periodicity of the underlying graph. We use a grand-canonical setup where we assign equal (zero) interaction energy and a finite {\em chemical potential} $\mu_i$ or activity $z_i=e^{\mu_i}$ to each of the $M$ different prototiles (setting the inverse temperature $\beta=1$). For fixed prototile numbers $n_1,\dots,n_M$, let us denote the number of nonequivalent $\Lambda$-patches that use $n_i$ prototiles of type $i$ by $g_{\Lambda}(n_1,\dots,n_M)$. The {\em grand-canonical partition function} is given by the following configuration generating function ($\mu=(\mu_1^{},\dots,\mu_M^{})$) \begin{equation} \mathcal{Z}_{\Lambda}^{}(\mu) \; = \; \sum_{n^{}_1,\dots,n^{}_M}g^{}_{\Lambda}(n_1,\dots,n_M)\, z_1^{n^{}_1} \cdot\ldots\cdot z_M^{n^{}_M}. \end{equation} To adapt the usual language of statistical mechanics, let $\omega\in\Omega_{\Lambda}$ be a tiling of a finite region $\Lambda$ which is now positioned relative to a fixed lattice, $\mathbb Z^2$ say. An {\em interaction} is the translation invariant assignment of a continuous function $\Phi(\omega)\in C_{\omega}$ to every $\omega$, so that a shift of $\Lambda$ by $a\in\mathbb Z^2$ results in $\Phi(\omega+a)=(\tau_a \Phi)(\omega)$, where $\tau_a$ denotes the corresponding shift (cf.\ \cite{Ruelle,Simon}). To avoid repeated counting of contributions to the interaction energy, $\Phi$ represents only the basic interaction of tiles in $\omega$ that are not already included in the interaction of subsystems. For $\omega\subset\Omega^{}_{\Lambda}$, we now define the Hamiltonian $H_{\Lambda}^{\Phi}(\omega)$ by \begin{equation} H_{\Lambda}^{\Phi}(\omega) \; = \; \sum_{\omega' < \, \omega}\Phi(\omega') \, , \end{equation} where the sum runs over all sub-patches of $\omega$. This is well defined due to the restrictions on $\Phi$ mentioned before. Let $\mathcal{B}$ be the Banach space (with norm $\|.\|_{\infty}$) of interactions $\Phi$ subject to the restriction \begin{equation} |||\Phi||| \; := \; \sum_{\omega\ni 0} \frac{\|\Phi(\omega)\|_{\infty}}{|\omega|} \; < \; \infty \, , \end{equation} where the sum is over all tilings covering the origin and $|\omega|$ denotes the number of tiles (cf.\ also \cite[App.~B]{Israel}). Let us now, for simplicity, assume that each tiling $\omega$ of $\Lambda$ has the same total number of tiles, $|\omega|=N^{}_{\Lambda}$. The {\em pressure\/} (negative grand-canonical potential) per tile is then given by \begin{equation} p^{}_{\Lambda}(\Phi) \; = \; \frac{1}{N^{}_{\Lambda}} \sum_{\omega\in\Omega_{\Lambda}}e^{-H_{\Lambda}^{\Phi}(\omega)} \; = \; \frac{1}{N^{}_{\Lambda}}\log\mathcal{Z}_{\Lambda}^{}(\Phi) \, . \end{equation} If the extra assumption is not fulfilled, $N^{}_{\Lambda}$ is the average number of tiles. This is a reasonable definition as long as we take the limit $\Lambda\to\infty$ in the sense of van Hove, see \cite{Ruelle} for details on this concept. But then, for $\Phi\in\mathcal{B}$, the thermodynamic limit $p(\Phi)=\lim_{\Lambda\to\infty}p_{\Lambda}(\Phi)$ exists and is a convex function of $\Phi$, compare \cite{Israel,Ruelle,Simon}. The interaction in our models is simply the self-energy ($-\mu_i$ for prototiles of type $i$) and thus included in $\mathcal{B}$. A {\em state} $\nu$ of the infinite system is a Borel probability measure on $\Omega$. A state $\nu$ is called {\em invariant} or {\em translation invariant} if it is invariant under the action of $\mathbb Z^2$. The set of all invariant measures, $\mathcal{M}^I$, is compact and forms a simplex. Following Ruelle's presentation \cite[Ch.~7.3]{Ruelle}, let $D$ be the set of all $\Phi\in\mathcal{B}$ such that the graph of $p$ has a unique tangent plane at the point $(\Phi,p(\Phi))$. If $\Phi\in\mathcal{B}$, there exists a unique linear functional $\alpha^{\Phi}$ in the dual $\mathcal{B}^*$ of $\mathcal{B}$ such that \begin{equation} \label{alpha} p(\Phi+\Psi) \; \geq \; p(\Phi)-\alpha^{\Phi}(\Psi) \, . \end{equation} The {\em mean densities} $\rho_i^{\Lambda}(\mu)=\frac{\langle n_i\rangle_{\Lambda}}{{\rm vol}(\Lambda)}$, $i=1,\dots,M$, of the different prototiles in the ensemble can be computed as functional derivative of $p_{\Lambda}(\mu)$ with respect to $\mu$ ($\langle\,.\,\rangle_{\Lambda}$ denotes the finite-size average for given chemical potentials $\mu_1,\dots,\mu_M$). For $\mu\in\mathcal{B}$ and $\Lambda\to\infty$ in the sense of van Hove, we have\footnote{Usually one defines $A_{\Phi}(\omega)=\sum_{\omega' < \, \omega} \frac{\Phi(\omega')}{|\omega'|}\in\mathcal{B}^*$. Restricting ourselves to the tiling models where the chemical potentials are the only interactions, we may identify $A_{\mu}$ and $\mu$.} \begin{equation} \lim_{\Lambda\to\infty}\sum_{i=1}^M\rho_i^{\Lambda}(\mu)\tilde{\mu}_i \; = \; \alpha^{\mu}(\tilde{\mu}) \end{equation} for all $\tilde{\mu}\in\mathcal{B}$, and \begin{equation} \lim_{\Lambda\to\infty}\rho_i^{\Lambda} \; = \; \rho_i^{} \; = \; \frac{\partial p(\mu_1,\dots,\mu_M)}{\partial\mu_i} \, , \end{equation} with $\sum_{i=1}^M \rho_i=1$. Since the chemical potentials and also the conjugate (mean) densities do not form an independent set of macroscopical parameters, we may choose an independent subset $\rho_1^{},\dots,\rho_k^{}$ by setting $\mu_{k+1}^{}=\dots=\mu_M^{}=0$. This normalization of $p$ leaves the densities invariant. The {\em entropy} per tile of a finite region tiling is defined as \begin{equation} s(\nu^{}_{\Lambda}) \; = \; -\frac{1}{N^{}_{\Lambda}}\, \sum_{\omega\in\Omega_{\Lambda}} \nu^{}_{\Lambda}(\omega)\log \nu^{}_{\Lambda}(\omega) \end{equation} where $\nu^{}_{\Lambda}$ is the restriction of $\nu$ to $\Omega_{\Lambda}$. For translation invariant measures and $\mu\in\mathcal{B}$, the infinite volume limit exists, giving $\nu^{}_{\Lambda}\to\nu$ for $\Lambda\to\infty$ taken in an appropriate way, and the functional $s$ is affine upper semicontinuous \cite[Ch.~7.2]{Ruelle}. According to Gibbs' variational principle \cite[Ch.~7.4]{Ruelle}, the pressure can then be calculated as \begin{equation} p(\mu) \; = \; \sup_{\nu\in\mathcal{M}^I}[s(\nu)-\nu(\mu)]\, . \end{equation} The measure for which this supremum is attained is called {\em equilibrium measure}. We may formulate the weak Gibbs phase rule \cite[Ch.~7.5]{Ruelle}. \begin{theorem} Let $D\subset\mathcal{B}$ defined as above. \begin{enumerate} \item If $\mu\in D$, the function $\nu\mapsto s(\nu)-\nu(\mu)$ reaches its maximum $p(\mu)$ at exactly one point $\nu^{\mu}\in\mathcal{M}^I$. \item If $\mu\in D$ and $\alpha^{\mu}\in\mathcal{B}^*$ is defined by (\ref{alpha}), then, for all $\tilde{\mu}\in\mathcal{B}$, \begin{equation} \nu^{\mu}(\tilde{\mu}) \; = \; \alpha^{\mu}(\tilde{\mu}) \end{equation} so that $\nu^{\mu}$ is the infinite volume equilibrium state corresponding to the chemical potential $\mu$. \item If $\mu\in D$, $\nu^{\mu}$ is a $\mathbb Z^2$-ergodic state and may thus be interpreted as pure thermodynamic phase. \hfill $\square$ \end{enumerate} \end{theorem} In what follows, we calculate the ensemble average of the correlations. Since the diffraction image is taken from a single member of the ensemble, the above theorem ensures that the {\em typical} member is self-averaging as long as the pressure is differentiable (no first order phase transition). Calculating the diffraction of our models consists essentially in calculating the corresponding dimers autocorrelation which we will base upon previous work of Fisher and Stephenson \cite{FS} and of Kenyon \cite{Ken1}. Kasteleyn \cite{Kast2} has shown that for any finite planar graph with even number of sites, and also for any periodic graph with a fundamental cell of an even number of sites, a Pfaffian\footnote{A Pfaffian is basically the square root of the determinant of an even antisymmetric matrix, see \cite[App.\ E]{Thompson} or \cite[Ch.\ IV.2]{McCoy} for an introduction.} can be constructed which is equal to the dimer generating function. For this, one has to orientate the graph in such a way that every configuration is counted with the correct sign. In addition, every bond is weighted with the corresponding dimer activity $z_i$. The configuration function is then given by the Pfaffian of the activity-weighted adjacency matrix $\bs{A}$. Although Kasteleyn's proof applies to arbitrary graphs, the calculations simplify considerably when restricted to periodic simply connected graphs as in our case. If we define occupation variables for a bond between sites ${\bs k}$ and ${\bs k}'$ \begin{equation} \eta^{}_{\bs{kk}'} \; = \; \begin{cases} 1, \qquad\text{bond (${\bs k},{\bs k}'$) occupied,} \\ 0, \qquad\text{otherwise,} \end{cases} \end{equation} we can state \cite{FS} \begin{prop} Let $G$ be an infinite simply connected periodic graph with close-packed dimer configuration where each dimer orientation has density $\rho_i^{}>0$. Let $\bs{A}$ be the invertible weighted adjacency matrix. If the dimer autocorrelation (joint occupation probability) exists, it is given by \begin{eqnarray} \label{auto2d} P_{\alpha\beta} & = & \langle \eta^{}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}} \eta^{}_{\bs{k}_{\beta}^{}\bs{k}'_{\beta}}\rangle \\ & = & \langle\eta^{}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}\rangle\langle \eta^{}_{\bs{k}_{\beta}^{}\bs{k}'_{\beta}} \rangle- A_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}} A_{\bs{k}_{\beta}^{}\bs{k}'_{\beta}} (A^{-1}_{\bs{k}_{\alpha}^{}\bs{k}_{\beta}^{}} A^{-1}_{\bs{k}'_{\alpha} \bs{k}'_{\beta}} -A^{-1}_{\bs{k}_{\alpha}^{}\bs{k}'_{\beta}} A^{-1}_{\bs{k}'_{\alpha}\bs{k}_{\beta}^{}})\, . \nonumber \end{eqnarray} \end{prop} {\sc Proof}: Let $\bar{\eta}_{\bs{kk}'}^{}=1-\eta^{}_{\bs{kk}'}$. Obviously, \begin{equation}\label{auto2dgen} \langle \eta^{}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}} \eta^{}_{\bs{k}_{\beta}^{} \bs{k}'_{\beta}}\rangle \; = \; \langle\eta^{}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}\rangle\langle \eta^{}_{\bs{k}_{\beta}^{}\bs{k}'_{\beta}} \rangle + \langle \bar{\eta}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}} \bar{\eta}_{\bs{k}_{\beta}^{}\bs{k}'_{\beta}}\rangle-\langle \bar{\eta}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}\rangle\langle \bar{\eta}_{\bs{k}_{\beta}^{}\bs{k}'_{\beta}} \rangle\, . \end{equation} The first term on the RHS of (\ref{auto2dgen}) depends only on the densities $\rho_{\alpha}^{}$ resp.\ $\rho_{\beta}^{}$ of the dimers that can occupy the bond $(\bs{k}_{\alpha}^{}\bs{k}'_{\alpha})$ resp.\ $(\bs{k}_{\beta}^{}\bs{k}'_{\beta})$. In the case of $\bs{k}_{\alpha}^{}$ and $\bs{k}'_{\alpha}$ being connected by a bond of type $i$ and only one bond of this type leading to each site, this would result in $\langle \eta^{}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}\rangle=\rho_{i}$, because we normalize with respect to the total number of dimers (and not to the number of sites or bonds as in \cite{FS}). It remains to prove the equivalence of the second terms of (\ref{auto2d}) and (\ref{auto2dgen}). This was shown in \cite{FS}, so we will just give an outline here. Here, $\langle\bar{\eta}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}\rangle$ is the (weighted) sum of all dimer configurations where the bond $(\bs{k}_{\alpha}^{}\bs{k}'_{\alpha})$ is {\em not} occupied, divided by the total (weighted) sum of configurations $\mathcal{Z}$. Since $\mathcal{Z}=\text{Pf}(\bs{A})$, we define $\text{Pf}(\bs{\tilde{A}})=\text{Pf}(\bs{A}+\bs{E})$ as the ``perturbed'' Pfaffian counting precisely all configurations where $(\bs{k}_{\alpha}^{}\bs{k}'_{\alpha})$ is not occupied, i.e.\ all elements of $\bs{E}$ are zero except $E_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}=-E_{\bs{k}'_{\alpha}\bs{k}_{\alpha}^{}}= -A_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}$. Note that $\bs{A},\bs{\tilde{A}}$ and $\bs{E}$ are skew-symmetric matrices. So $\langle\bar{\eta}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}}\rangle =\text{Pf}(\bs{\tilde{A}})/\mathcal{Z} =\text{Pf}(\mathbb I+\bs{A}^{-1}\bs{E})=\text{Pf}(\bs{E})\; \text{Pf} (\bs{E}^{-1}+\bs{A}^{-1})$, where the last equality holds only if $\bs{E}$ is invertible, $\mathbb I$ denotes the unit matrix of appropriate dimension. The same applies to $\langle \bar{\eta}_{\bs{k}_{\alpha}^{}\bs{k}'_{\alpha}} \bar{\eta}_{\bs{k}_{\beta}^{} \bs{k}'_{\beta}}\rangle$ but with four nonvanishing elements of $\bs{E}$. The evaluation of the Pfaffian reduces to the calculation of a small determinant and yields the desired result. \hfill $\square$ \smallskip The calculation of $\mathcal{Z}$ and ${\bs A}^{-1}$ can be simplified considerably by imposing periodic boundary conditions. Usually, for the partition function for a toroidal graph, one needs four determinants differing from that with free boundary only in exactly these boundary elements \cite{Kast}. But in the infinite volume limit, these modifications do not change the value of a determinant (given by the product of the eigenvalues) as can be derived from the following result of Ledermann \cite{Led} \begin{lemma} If in a Hermitian matrix the elements of $r$ rows and their corresponding columns are modified in any way whatever, provided the matrix remains Hermitian, then the number of eigenvalues that lie in any given interval cannot increase or decrease by more than $2r$. \hfill $\square$ \end{lemma} If $N=mn$ is the number of sites or elementary cells, the number of eigenvalues per unit length, which is $\mathcal{O}(N)$, changes only by $\mathcal{O}(\sqrt{N})$, which is negligible in the limit as $N\to\infty$ (for details see also \cite{Montroll}). The same argument holds for $\bs{A}^{-1}$. How can we calculate the elements of $\bs{A}^{-1}$? Since $\bs{A}$ is the adjacency matrix of a graph made up of a periodic array of elementary cells with toroidal boundary conditions and is therefore cyclic, it can be reduced to the diagonal form $\bs{\Lambda}=\text{diag}\{\lambda_{\bs j}\}$ by a Fourier-type similarity transformation with matrix elements $S_{\bs{kk}'}= (mn)^{-1/2}\exp(2\pi i (k_1^{} k'_1/m +k_2^{} k'_2/n))$. $\bs{A}^{-1}$ is now determined by \begin{equation} \label{diag} A^{-1}_{\bs{kk}'} \; = \; \left(\bs{S\Lambda}^{-1}\bs{S}^{-1}\right)_{\bs{kk}'} \; = \; \sum_{\bs{j}=(1,1)}^{(m,n)} S_{\bs{kj}}^{} \lambda_{\bs{j}}^{-1}S^{\dagger}_{\bs{k}'\bs{j}} \, . \end{equation} In the infinite volume limit, the sums approach integrals (Weyl's Lemma), and by introducing $\varphi_1^{}=2\pi i j_1^{}/m$ etc. we obtain \begin{equation} \label{inverse} A^{-1}_{\bs{kk}'} \; = \; \frac{1}{4 \pi^2}\int_0^{2 \pi} \int_0^{2 \pi}\lambda^{-1}(\varphi_1^{},\varphi_2^{}) e^{-i(\varphi_1^{}(k'_1-k_1^{})+\varphi_2^{}(k'_2-k_2^{}))} d\varphi_1^{} d\varphi_2^{}. \end{equation} Let us illustrate this by two examples, see \cite{Hoeffe} for another case. \subsubsection*{Domino tiling} A domino is a $2$ by $1$ or $1$ by $2$ rectangle, whose vertices have integer coordinates in the plane. A tiling of the plane with dominoes is equivalent to a close-packed dimer configuration on the square lattice $\mathbb Z^2$. This model exhibits no phase transition. We assume finite, positive activities in order to have non-vanishing tile densities. The degenerate case will be treated separately. Labelling the sites by $\bs{k}=(k_1,k_2)$, and adopting from the various equivalent possibilities (see \cite{FS} for details) the technically most convenient choice of complex weights, one obtains the weighted adjacency matrix \begin{align} A(k_1,k_2;k_1+1,k_2) & = -A(k_1+1,k_2;k_1,k_2)=z_1, \nonumber\\ A(k_1,k_2;k_1,k_2+1) & = -A(k_1,k_2+1;k_1,k_2)=iz_2, \\ A(k_1,k_2;k_1',k_2') & = A_{\bs{kk}'} = 0, \qquad\text{otherwise}, \nonumber \end{align} with eigenvalues \begin{equation} \lambda(\varphi_1^{},\varphi_2^{}) \; = \; 2i\left(z_1^{} \sin \varphi_1^{}+i z_2^{}\sin\varphi_2^{}\right) \, . \end{equation} After inserting in (\ref{inverse}) and separating real and imaginary parts, one gets ($\bs{k}'-\bs{k}={\bs r}=(x,y)\in\mathbb Z^2$) \begin{equation} \label{dominv} A^{-1}_{\bs{kk}'} \; = \; \frac{1}{2 \pi^2}\int_0^{\pi}\int_0^{\pi}\frac{M(x,y|\varphi_1^{}, \varphi_2^{})}{z_1^2 \sin^2_{}\varphi_1^{}+z_2^2 \sin^2\varphi_2^{}}d\varphi_1^{} d\varphi_2^{} \end{equation} with \begin{equation} M(x,y|\varphi_1^{},\varphi_2^{}) \; = \; \begin{cases} 0, & \text{($x,y$ same parity)} \\ -z_1^{}\sin\varphi_1^{}\sin x\varphi_1^{}\cos y\varphi_2^{}, &\text{($x$ odd, $y$ even)} \\ -i z_2^{}\sin\varphi_2^{}\sin y\varphi_2^{}\cos x\varphi_1^{}, &\text{($y$ odd, $x$ even)} \end{cases} \end{equation} (compare with \cite{FS}). We introduce the abbreviation for the {\em coupling function} \cite{Ken1} \begin{equation} [x,y] \; = \; A^{-1}(k_1,k_2;k_1',k_2')\, , \end{equation} obeying $[x,y]=-[-x,-y]$. We place the scatterers in the centres of the tiles or equivalently of the dimers. With (\ref{auto2d}) and (\ref{auto2dgen}), the joint occupation probability of two horizontal dimers with scatterers at distance ${\bs r}$ in the centre of an infinite lattice is given by \begin{equation} P_{11}({\bs r}) \; = \; \frac{\rho_1^2}{4} + c^{}_{11}(\bs r) \; = \; \frac{\rho_1^2}{4}-z_1^2\left([x,y]^2-[x-1,y][x+1,y]\right) \, , \end{equation} (similarly for $P_{22}$) where only $c^{}_{11}(\bs r)$ (resp.\ $c^{}_{22}(\bs r)$) depends on the distance ${\bs r}$. For a pair of mutually perpendicular dimers, the possible distance vectors of the scatterers ${\bs r}+{\bs a}$, ${\bs a}=(-1/2,1/2)^t$, are odd half-integer. The joint occupation probability is \begin{align} P_{12}\left({\bs r}+{\bs a}\right) &\; = \; P_{21}\left({\bs r}-{\bs a}\right) \nonumber\\ &\; = \; \frac{\rho_1^{}\rho_2^{}}{4}-iz_1 z_2 \left([x,y][x-1,y+1]-[x,y+1][x-1,y]\right) \, . \end{align} The non-constant part is $c^{}_{12}(\bs r)=c^{}_{21}(\bs r)$. Note that either the first or second part of the term in brackets vanishes because of the parity of $x$ and $y$. The full autocorrelation for the positions of the scatterers is thus given by \begin{equation}\label{domauto} \gamma^{}_{\omega} \; = \; \sum_{{\bs r}\in\mathbb Z^2_{}} \left(P_{11}\left({\bs r}\right)+P_{22}\left({\bs r}\right)\right) \delta_{{\bs r}} +\delta_{\bs a}^{}*\sum_{{\bs r}\in\mathbb Z^2_{}} \left(P_{12}\left({\bs r}+{\bs a}\right)+P_{21}\left({\bs r}+{\bs a} \right)\right)\delta_{{\bs r}} \, . \end{equation} \begin{theorem} \label{thm4} Under the above assumptions, with $\rho_1 \rho_2>0$, the diffraction spectrum of the domino tiling exists with probabilistic certainty and consists of a pure point and an absolutely continuous part, i.e.\ $\hat{\gamma}_{\omega}^{}=(\hat{\gamma}_{\omega}^{})_{pp}^{}+ (\hat{\gamma}_{\omega}^{})_{ac}^{}$, with \begin{equation} \left(\hat{\gamma}_{\omega}\right)_{pp} \; = \; \frac{1}{4}\sum_{(h,k)\in\mathbb Z^2}\left(\rho_1+ (-1)^{h+k}\rho_2\right)^2 \delta_{(h,k)} \, . \end{equation} In particular, there is no singular continuous part. Furthermore, $\hat{\gamma}_{\omega}^{}$ is periodic with lattice of periods $\{ (h,k)\in\mathbb Z^2\mid h+k \mbox{ even}\}$. \end{theorem} {\sc Proof}: The point spectrum can be calculated directly by taking the Fourier transform of the constant part of (\ref{domauto}). This requires Poisson's summation formula and the convolution theorem, which leads to the phase factor $(-1)^{h+k}$ and to the periodicity claimed. The non-constant part of $\gamma_{\omega}$ is determined by $c^{}_{11}$, $c^{}_{12}$ and $c^{}_{22}$. To establish our claim, we will show that the formal Fourier series $\sum_{\bs{r}\in\mathbb Z^2} c^{}_{ij}(\bs r) e^{-2\pi i \bs{k}\cdot\bs{r}}$ actually converge to $L^1$-functions and thus represent absolutely continuous measures. The real coefficients $c^{}_{ij}(\bs r)$ are essentially products of the form $[x,y]^2$. One integration in (\ref{dominv}) may be performed explicitly. With standard asymptotic methods involving the Laplace transform, see the Appendix of \cite{Young}, one obtains the asymptotic behaviour in the limit of large $x$ and $y$ as \begin{equation} [x,y] \; \sim \; \begin{cases} -\frac{1}{\pi}\frac{z_2^{} x}{(z_2^{} x)^2+(z_1^{} y)^2}, & \text{($x$ odd, $y$ even)}, \\ -\frac{i}{\pi}\frac{z_1^{} y}{(z_2^{} x)^2+(z_1^{} y)^2}, & \text{($y$ odd, $x$ even)}. \end{cases} \end{equation} Thus, we obtain $0\leq c_{ij}(\bs{r}) =\mathcal{O}\left(\frac{1}{(x^2+y^2+1)^2} \right)<\infty$, since $x$ and $y$ have different parity. Note that the implied constant still depends on $z^{}_1$ and $z^{}_2$. Now, e.g.\ by referring to Eqs. (6.1.126) and (6.1.32) of \cite{Hansen}, one can see that $\sum_{y=0}^{\infty}\sum_{x=0}^{\infty}\frac{1}{(x^2+y^2+1)^2}$ converges (it actually is even less than 2 in value). Consequently, by Cauchy's double series theorem, $\sum_{\bs{r}\in\mathbb Z^2_{}}\left(c_{ij}(\bs{r})\right)^2$ converges absolutely. So, the $c_{ij}(\bs{r})$ can be seen as functions in $\ell^2(\mathbb Z^2)$ and, by the Riesz-Fischer Theorem \cite[Thm.\ 23.3]{Heuser}, each of the Fourier series $\sum_{\bs{r}\in\mathbb Z^2} c^{}_{ij}(\bs r) e^{-2\pi i \bs{k}\bs{r}}$ converges to a function in $L^2(\mathbb R^2/\mathbb Z^2)$ in the $L^2$-norm. The limit is independent of the order of summation, and convergence is actually also pointwise, almost everywhere. H\"older's inequality gives $L^2\left(\mathbb R^2/\mathbb Z^2\right)\subset L^1\left(\mathbb R^2/\mathbb Z^2\right)$, and combining these periodic functions with the appropriate phase shifts as implied by (\ref{domauto}) results in a function with lattice of periods $\Gamma=\{ (h,k)\in\mathbb Z^2\mid h+k \mbox{ even}\}$ which is certainly in $L^1(\mathbb R^2/\Gamma)$, so the Radon-Nikodym theorem \cite{RS} leads to the result stated. \hfill $\square$ \smallskip {\sc Remark}: One may assign complex weights $h_1,h_2$ to the scatterers on the two dominoes without changing the spectral type. This results in the pure point part \begin{equation} \left(\hat{\gamma}_{\omega}\right)_{pp} \; = \; \frac{1}{4}\sum_{(h,k)\in\mathbb Z^2}\left| \rho_1 h_1+ (-1)^{h+k}\rho_2 h_2 \right|^2 \delta_{(h,k)} \, . \end{equation} This applies analogously to the next example. \begin{figure}[ht] \centerline{\epsfysize=6cm \epsfbox{figs/dompatch.eps} \hspace{1cm} \epsfysize=6cm \epsfbox{figs/domfou.eps}} \caption{\label{domino} Typical tiling for $\rho_2^{}=0.3$ (left) and its diffraction image (right). The scatterers are located at the centre of the tiles.} \end{figure} One has to be aware that the pure point part (\ref{domauto}) does not display the correct (statistical) symmetry of the system. Away from the point of maximum entropy (which is $\rho_1^{}=\rho_2^{}=1/2$), the tiling is no longer fourfold symmetric, as still indicated by the point part, but the twofold symmetry is only displayed in the diffuse background. This can be seen in Figure~\ref{domino}. The $pp$ part is calculated from the exact expression and the Bragg peaks are represented by white circles with area proportional to the intensity. The $ac$ part was calculated numerically by means of standard FFT, because this is simpler than using the exact expression for the correlation functions. In the special case of only one domino orientation remaining, the scatterers distribution is a Dirac comb on a rectangular lattice. Using Poisson's summation formula, the diffraction spectrum is a Dirac comb on the reciprocal rectangular lattice. In such a limit, the diffuse background accumulates at the extra positions and converges vaguely to a point measure that completes the square lattice arrangement to the proper rectangular one. \subsubsection*{Lozenge tiling} A lozenge is a rhombus with side 1, smaller angle $\pi/3$, and vertices in the triangular lattice $\Gamma=A_2^{}/\sqrt{2}$ with minimal distance $1$. This tiling can be mapped on a dimer configuration on the honeycomb packing. \begin{figure}[ht] \centerline{\epsfysize=4cm \epsfbox{figs/hexgitter.eps}} \caption{\label{hexlattice}(a)~Lozenge tiling and the dimer configuration of the honeycomb packing. (b)~Elementary cell for the weighted adjacency matrix.} \end{figure} The different tile densities are nonvanishing if $z_i^{}>|z_j^{}-z_k^{}|$, $i,j,k$ pairwise different \cite{Richard}; at equality, the system undergoes a phase transition of Kasteleyn type \cite{Kast2} with only one lozenge orientation remaining. This trivial case shall be excluded in the sequel. With Dirac unit measures on the tile centres, the support of the scatterers is given by a Kagom\'e grid of minimal vertex distance $1/2$. The adjacency matrix then has entries that are $2$ by $2$ matrices themselves, describing the elementary cells of the packing (see Figure \ref{hexlattice}). By wrapping the graph on a torus, we may transform it to block diagonal form with elements \begin{equation} \bs{\lambda}(\varphi_1,\varphi_2) \; = \; \begin{pmatrix} 0 & -(z_1^{}e^{-i\varphi_1}_{}+z_2^{}e^{-i\varphi_2}_{}+z_3^{}) \\ z_1^{}e^{i\varphi_1}_{}+z_2^{}e^{i\varphi_2}_{}+z_3^{} & 0 \end{pmatrix}\, . \end{equation} If we introduce a coordinate system with $\hat{x}=(1,0)^t_{}$ and $\hat{y}=1/2(1,\sqrt{3}\,)^t_{}$, we can use the above notation for the difference vectors of the elementary cells. Denoting the left and right site of the elementary cell by $L$ and $R$ we see that $[x,y]_{LL}^{}=[x,y]_{RR}^{}=0$. In the infinite size limit, the remaining matrix elements are given by \begin{equation} [x,y|z_1^{},z_2^{},z_3^{}]_{LR} \; = \; \frac{1}{4 \pi^2}\int_0^{2 \pi}\int_0^{2 \pi}\frac{e^{i(\varphi_1^{}x+\varphi_2^{} y)}} {z_1^{}e^{-i \varphi_1^{}}+z_2^{}e^{-i \varphi_2^{}}+z_3^{}} d\varphi_1^{}d\varphi_2^{} \end{equation} (compare with \cite{Ken1}). Let $v=e^{-i \varphi_1^{}}$ and $w=e^{-i\varphi_2^{}}$. Then \begin{equation} \label{lozcoup} [x,y|z_1^{},z_2^{},z_3^{}]_{LR} \; = \; \frac{1}{4 \pi^2}\int_{S^1\times S^1}\frac{v^{-x}w^{-y}} {z_1^{} v+z_2^{} w+z_3^{}}\frac{dv}{iv}\frac{dw}{iw} \, . \end{equation} As was already observed by Kenyon \cite{Ken1} for the isotropic case, the coupling function has all the symmetries of the graph: interchanging $v$ and $w$ or the substitution $(v,w)\to(w^{-1}_{},vw^{-1}_{})$ and combinations of these let the integral invariant, i.e.\ \begin{equation} \label{sym} \begin{array}{rcl} [x,y\mid z_1^{},z_2^{},z_3^{}]_{LR} & = & [x,-x-y-1\mid z_1^{},z_3^{},z_2^{}]_{LR} \\ =\; [y,x\mid z_2^{},z_1^{},z_3^{}]_{LR} & = & [y,-x-y-1\mid z^{}_2,z^{}_3,z^{}_1]_{LR} \\ =\; [-x-y-1,x\mid z_3^{},z_1^{},z_2^{}]_{LR} & = & [-x-y-1,y\mid z_3^{},z_2^{},z_1^{}]_{LR} \, . \end{array} \end{equation} We evaluate one integration in (\ref{lozcoup}) explicitly for $x\leq -1$. The other values can be obtained by (\ref{sym}). This is a direct generalization of \cite{Ken1} to the case of arbitrary activities. One gets \begin{equation} [x,y|z_1^{},z_2^{},z_3^{}]_{LR} \; = \; \frac{i}{2 \pi} (-z_1^{})^x \int_{e^{i \varphi_0^{}}}^{e^{i(2\pi-\varphi_0^{})}} w^{-y-1} \left(z_2^{}+z_3^{}w\right)^{-x-1}dw, \end{equation} with $\varphi_0^{}=\arccos\frac{z_1^2-z_2^2-z_3^2}{2 z_2^{}z_3^{}}$. One easily finds the possible distance vectors of the scatterers. Away from the phase transition points, we have for the constant part of the autocorrelation measure for the scatterers (one per lozenge) \begin{equation} \label{lozauto} \begin{split} \left(\gamma_{\omega}^{}\right)_{const}^{} \; = \; \frac{2}{\sqrt{3}} \sum_{(x,y)\in \Gamma}\Bigl(&\left(\rho_1^2+\rho_2^2+\rho_3^2\right) \delta_{(x,y)}^{} \\ &+2\rho_1^{}\rho_2^{}\delta_{\left(\frac{2x+1}{2}, \frac{2y+1}{2}\right)}^{}+2\rho_1^{}\rho_3^{} \delta_{\left(\frac{2x+1}{2},y\right)}^{}+ 2\rho_2^{}\rho_3^{}\delta_{\left(x,\frac{2y+1}{2}\right)}^{}\Bigr). \end{split} \end{equation} The reciprocal lattice $\Gamma^*$ is spanned by the vectors $\left(1,-\frac{1}{\sqrt{3}}\right)^t$ and $\left(0,\frac{2}{\sqrt{3}}\right)^t$. Using (\ref{auto2d}) we can state \begin{theorem} \label{thm5} Under the above assumptions, the diffraction spectrum of the lozenge tiling exists with probabilistic certainty and consists of a pure point and an absolutely continuous part, i.e.\ $\hat{\gamma}^{}_{\omega} = (\hat{\gamma}^{}_{\omega})_{pp} + (\hat{\gamma}^{}_{\omega})_{ac}\,$, with \begin{equation} \left(\hat{\gamma}^{}_{\omega}\right)_{pp} \; = \; \frac{4}{3}\sum_{(h,k)\in \Gamma^*} \left((-1)^h\rho_1^{}+(-1)^k\rho_2^{}+\rho_3^{}\right)^2 \delta_{(h,k)}^{}. \end{equation} There is no singular continuous part, and $\hat{\gamma}^{}_{\omega}$ is periodic with lattice $2\Gamma^*$. \end{theorem} {\sc Proof}: The pure point part is simply the Fourier transform of (\ref{lozauto}), again calculated by means of Poisson's summation formula. \begin{figure}[ht] \centerline{\epsfysize=6cm \epsfbox{figs/hexpatch.eps} \hspace{1cm} \epsfysize=6cm \epsfbox{figs/hexfou.eps}} \caption{\label{hex} Typical tiling for $\rho_2^{}=0.24$ and $\rho_1^{}=\rho_3^{}=0.38$ (left) and its diffraction image (right). The scatterers are located at the centres of the tiles.} \end{figure} As before, we will now show that the remaining part of $\hat{\gamma}^{}_{\omega}$ converges to a periodic $L^1$-function and thus represents an absolutely continuous measure. Let us start with the case $x\leq -1$ (and $y$ arbitrary, but fixed) and show that $[x,y|z_1^{},z_2^{},z_3^{}]_{LR} = \mathcal{O}\left({|x|^{-1}}\right)$ as $x\to - \infty$. The necessary extension to the complete asymptotic behaviour will later follow from Eq.~(\ref{sym}). For now, and for fixed values of the activities in the admitted range, we have \begin{align} I \; = \; \left|\,[x,y|z_1^{},z_2^{},z_3^{}]_{LR}\,\right| &\;\leq\; \frac{z_1^x}{2\pi} \int_{e^{i\varphi_0^{}}}^{e^{i(2\pi-\varphi_0^{})}}|z_2^{}+z_3^{} w|^{-x-1}|dw| \nonumber\\ &\;=\; \frac{z_1^x}{\pi}\int_{\varphi_0^{}}^{\pi}\left(z_2^2+z_3^2+2z_2^{}z_3^{} \cos\vartheta\right)^{\frac{-x-1}{2}} d\vartheta. \end{align} We first bound $f(\vartheta)=(z_2^2+z_3^2+2 z_2^{} z_3^{}\cos\vartheta)/z_1^2$ by a straight line $\tilde{f}(\vartheta)$, i.e.\ we look for $f(\vartheta)\leq\tilde{f}(\vartheta)$ on the interval $[\varphi^{}_0,\pi]$. Note that $f(\vartheta)$ can only have one flex point at $\vartheta=\pi/2$ in $[0,\pi]$. One has to distinguish two cases: \begin{enumerate} \item $\cos\varphi_0^{}\geq(\pi-\varphi_0^{})\sin\varphi_0^{} - 1$: Choose $\tilde{f}_1(\vartheta) =(\vartheta-\varphi_0^{})f'(\varphi_0^{})+f(\varphi_0^{})$. The condition on $\varphi_0^{}$ implies $\tilde{f}_1(\pi)\geq f(\pi)$. Differentiating $f(\vartheta)-\tilde{f}_1(\vartheta)$ with respect to $\vartheta$ yields $f'(\vartheta)-\tilde{f}_1'(\vartheta) = f'(\vartheta) - f'(\varphi^{}_0)$ and this vanishes if $\vartheta = \varphi^{}_0$ (maximum) or if $\vartheta = \pi-\varphi_0^{}$ (minimum). Thus $f(\vartheta)\leq\tilde{f}_1(\vartheta)$. \item $\cos\varphi_0^{}<(\pi-\varphi_0^{})\sin\varphi_0^{} - 1$: Choose $\tilde{f}_2(\vartheta) =\frac{f(\pi) - 1}{\pi-\varphi_0^{}}\vartheta+ \frac{\pi-\varphi_0^{}f(\pi)} {\pi-\varphi_0^{}}$ (the line connecting $f(\varphi_0^{})=1$ and $f(\pi)$). Obviously $\tilde{f}'_2(\pi)\leq f'(\pi)$. Because of the angle condition, we further have $f'(\varphi_0^{})\leq \tilde{f}'_2(\varphi_0^{})$. As there is only one flex point, we conclude that $f(\vartheta)\leq\tilde{f}_2(\vartheta)$. \end{enumerate} Choose $f(\vartheta)$ as $f_1(\vartheta)$ or $f_2(\vartheta)$ according to the previous distinction. If $c=|\tilde{f}'(\vartheta)|$, we get $0<c\leq\frac{1}{\pi-\varphi^{}_0}\left(1-\frac{(z^{}_2-z^{}_3)^2}{z_1^2}\right) <\frac{1}{\pi-\varphi^{}_0}$. Consequently, we can estimate \begin{align} \pi z^{}_1 I&\;\leq\;\int_{\varphi_0^{}}^{\pi} \left(1-c(\vartheta-\varphi_0^{})\right)^{\frac{-x-1}{2}} d\vartheta \nonumber \\ &\;<\;\frac{1}{c}\int_0^1 (1-t)_{}^{\frac{-x-1}{2}} \, dt \\ &\;=\; \frac{2}{c(1-x)} \;=\;\mathcal{O}\left(\frac{1}{|x|}\right) . \nonumber \end{align} Now, we can use the symmetry relations of (\ref{sym}) and obtain the asymptotic behaviour $[x,y|z_1^{},z_2^{},z_3^{}]_{LR}=\mathcal{O}\left({(|x|+|y|)^{-1}}\right)$, whenever $|\bs r| \to\infty$, compare \cite{Ken1}. The rest of the argument is very similar to the domino case and need not be repeated here, the periodicity statement follows again constructively. \hfill $\square$ \smallskip Let us remark that an analogous scenario, with the same type of result, occurs for the more complicated dart-rhombus random tiling, see \cite{Hoeffe} for details. \subsection*{Addendum: The two-dimensional Ising model} For the sake of completeness, we add an application of the probably best analyzed model in statistical physics, the 2D Ising model without external field. It may be regarded as a lattice gas on $\mathbb Z^2$, compare \cite{Simon}, with scatterers of strength $s^{}_{(i,j)}\in\{1,0\}$. The partition function in the spin-formulation ($\sigma^{}_{(i,j)}\in\{+1,-1\}$) reads as follows \begin{equation} \mathcal{Z} \; = \; \sum_{\{\bs{\sigma}\}}\exp\left(\sum_{(i,j)} \sigma^{}_{(i,j)}\left(K^{}_1\sigma^{}_{(i+1,j)}+ K^{}_2\sigma^{}_{(i,j+1)}\right)\right) \, , \end{equation} where we sum over all configurations $\{\bs{\sigma}\}$. We consider the ferromagnetic case with coupling constants $K_i=J_i/(k_B T)>0$, temperature $T$ and Boltzmann's constant $k_B$. The model undergoes a phase transition at $k:=\left(\sinh(2 K_1)\sinh(2 K_2)\right)^{-1}=1$. It is common knowledge that in the regime with coupling constants smaller than the critical ones (corresponding to $T>T_c$) the ergodic equilibrium state with vanishing magnetization $m$ is unique, whereas above ($T<T_c$) there exist two extremal equilibrium states, which are thus ergodic \cite[Ch.~III.5]{Simon}. In this case, we assume to be in the extremal state with positive magnetization $m=(1-k^2)^{1/8}$. The diffraction properties of the Ising model can be extracted from the known asymptotic behaviour \cite{McCoy,Wu} of the autocorrelation coefficients. We first state the result for the isotropic case ($K_1=K_2=K$) and comment on the general case later. \begin{prop} \label{ising} Away from the critical point, the diffraction spectrum of the Ising lattice gas almost surely exists, is $\mathbb Z^2$-periodic and consists of a pure point and an absolutely continuous part with continuous density. The pure point part reads \begin{enumerate} \item $T>T_c$: $(\hat{\gamma}_{\omega})_{pp}=\frac{1}{4} \sum_{\bs{k}\in\mathbb Z^2}\delta_{\bs{k}}^{}$ \item $T<T_c$: $(\hat{\gamma}_{\omega})_{pp}=\rho^2 \sum_{\bs{k}\in\mathbb Z^2}\delta_{\bs{k}}^{}$, \end{enumerate} where the density $\rho$ is the ensemble average of the number of scatterers per unit volume. \end{prop} {\sc Proof}: First, note that $s^{}_{(i,j)}=(\sigma^{}_{(i,j)}+1)/2$ and thus $\langle\sigma^{}_{(i,j)}\rangle = m=2\rho-1$, so $\rho$ varies between $1$ and $1/2$. The asymptotic correlation function of two spins at distance $R=\sqrt{x^2+y^2}\,$ (as $R\to\infty$) is \cite{McCoy} \begin{equation} \label{asymp} \langle \sigma^{}_{(0,0)}\sigma^{}_{(x,y)}\rangle \; \simeq \; \begin{cases} c_1\frac{e^{-R/c_2}}{\sqrt{R}}, & T>T_c \\ m^2+c_3\frac{e^{-2R/c_2}}{R^2},\quad & T<T_c, \end{cases} \end{equation} with constants $c_1,c_2$ and $c_3$ depending only on $K$ and $T$, see also \cite[p.\ 51]{KS} and references given there for a summary. The pure point part $(\hat{\gamma}_{\omega}^{})_{pp}$ results directly from the Fourier transform of the constant part of $\gamma_{\omega}^{}$ as derived from the asymptotics of $\langle s_{(0,0)}^{} s_{(x,y)}^{} \rangle = (\langle\sigma_{(0,0)}^{}\sigma_{(x,y)}^{}\rangle + 2 m + 1)/4$. Here, already $\sum_{(x,y)\in\mathbb Z^2} e^{-R/c_2}/\sqrt{R}$ and $\sum_{(x,y)\in\mathbb Z^2} e^{-2R/c_2}/R^2$ converge absolutely, so we can view the corresponding correlation coefficients as functions in $L^1(\mathbb Z^2)$. Their Fourier transforms (which are uniformly converging Fourier series) are continuous functions on $\mathbb R^2/\mathbb Z^2$, see \cite[\S 1.2.3]{Rudin1}, which are then also in $L^1(\mathbb R^2/\mathbb Z^2)$. Applying the Radon-Nikodym theorem finishes the proof. \hfill $\square$ \smallskip {\sc Remark}: At the critical point, the correlation function $\langle \sigma_{(0,0)}^{}\sigma_{(x,y)}^{}\rangle$ is asymptotically proportional to $R^{-1/4}$ as $R\to\infty$ \cite{Wu,KS}. Again, taking out first the constant part of $\gamma_{\omega}^{}$, we get the same pure point part as in Prop.\ \ref{ising} for $T>T_c$. However, for the remaining part of $\gamma_{\omega}^{}$, both our previous arguments fail. Nevertheless, using a theorem of Hardy \cite[p.~97]{Bromwich}, we can show that the corresponding Fourier series still converges for $\bs{k}\not\in \mathbb Z^2$ (a natural order of summation is given by shells of increasing radius). {}For $\bs{k}\in\mathbb Z^2$, where the Bragg peaks reside, the series diverges. But this can neither result in further contributions to the Bragg peaks (the constant part of $\gamma_{\omega}^{}$ had already been taken care of) nor in singular continuous contributions (because the points of divergence form a uniformly discrete set). So, even though the series diverges for $\bs{k}\in\mathbb Z^2$, it still represents (we know that $\hat{\gamma}^{}_{\omega}$ exists) a function in $L^1(\mathbb R^2/\mathbb Z^2)$ and hence the Radon-Nikodym density of an absolutely continuous background. On the diffraction image, we thus can see, for any temperature, Bragg peaks on the square lattice and a $\mathbb Z^2$-periodic, absolutely continuous background concentrated around the peaks (the interaction is attractive). At the critical point, the intensity of the diffuse scattering diverges when approaching the lattice positions of the Bragg peaks. The same arguments hold in the anisotropic case, where the asymptotics still conforms to Eq.~(\ref{asymp}) and the above, if $R=R(x,y)$ is replaced by the formula given in \cite[Eq.\ 2.6]{Wu}. The pure point part is again that of Prop.~\ref{ising} with fourfold symmetry, while (as in the case of the domino tiling) the continuous background breaks this symmetry if $K_1\neq K_2$. Let us finally remark that a different choice of the scattering strengths (i.e.\ $\pm 1$ rather than $1$ and $0$) would result in the extinction of the Bragg peaks in the disordered phase ($T>T_c$), but no choice does so in the ordered phase ($T<T_c$). \subsection*{Introduction} The diffraction theory of crystals is a subject with a long history, and one can safely say that it is well understood \cite{Guinier,Cowley}. Even though the advent of quasicrystals, with their sharp diffraction images with perfect non-crystallographic symmetry, seemed to question the general understanding, the diffraction theory of perfect quasicrystals, in terms of the cut and project method, is also rather well understood by now, see \cite{Hof,Hof-Waterloo} and references therein. It should be noted though that this extension was by no means automatic, and required a good deal of mathematics to clear up the thicket. More recently, this has found a general extension to the setting of locally compact Abelian groups \cite{Martin,Martin98} which can be seen as a natural frame for mathematical diffraction theory and covers quite a number of interesting new cases \cite{BMS}. Another area with a wealth of knowledge is the diffraction theory of imperfect crystals and amorphous bodies \cite{Guinier,Welberry}, but the state of affairs here is a lot less rigorous, and many results and features seem to be more or less folklore. For example, the diffraction of simple stochastic systems, as soon as they are not bound to a lattice, is only in its infancy, see \cite{BM1} for some recent addition to its rigorous treatment. This does not mean that one would not know what to expect. However, one can often only find a qualitative argument in the literature, but no proof. May this be acceptable from a practical angle, it seems rather unsatisfactory from a more fundamental point of view. In other words, the answer to the question which distributions of matter diffract is a lot less known than one would like to believe, compare the discussion in \cite[Sec.\ 6]{Hof} and also \cite{R2,EM}. Note that this question contains several different aspects. On the one hand, one would like to know, in rigorous terms, under which circumstances the diffraction image is well defined in the sense that it has a unique infinite volume limit. This is certainly the case if one can refer to the ergodicity of the underlying distribution of scatterers \cite{Dworkin,Hof,Martin98}, in particular, if their positional arrangement is linearly repetitive \cite{LP}. However, this is often difficult to assess in situations {\em without} underlying ergodicity properties, see \cite{BMP} for an example. On the other hand, even if the image is uniquely defined, one still wants to know whether it contains Bragg peaks or not, or if there is any diffuse scattering present in it. This situation certainly did not improve with the more detailed investigation of quasicrystals, e.g.\ their less perfect versions, and in particular with the study of the so-called random tilings \cite{Elser,Henley,Richard}. Again, there is a good deal of folklore available, and a careful reasoning based upon scaling arguments (compare \cite{Jaric,Henley}) seems to give convincing and rather consistent results on their diffraction properties. However, various details, and in particular the exact nature of the diffraction spectrum, have always been the topic of ongoing discussion, so that a more rigorous treatment is desirable. It is the aim of this contribution to go one step into this direction, and to extend the analysis of \cite{BM1} on generalized lattice gases to the case of certain Markov type systems as they appear in the theory of random tilings. We will not be able to answer the real questions concerning those tilings relevant for quasicrystals, but we still think that the results derived below are a worth-while first step. Even this requires a bunch of methods and results which are scattered over rather different branches of mathematics and mathematical physics. It is thus also one of our aims to recollect the essential aspects and references, tailored for what we need here and for future work in this direction. Let us summarize how the article is organized. We start with a recapitulation of the measure theoretic setup needed for mathematical diffraction theory, where we essentially follow Hof \cite{Hof,Hof-Waterloo}, but adapt and extend it to our needs. We will be a little bit more explicit here than needed for an audience with background in mathematics or mathematical physics, because we hope that the article becomes more self-contained that way, and hence more readable for physicists and crystallographers who usually do not approach problems of diffraction theory in these more rigorous terms. We consider this as part of an attempt to penetrate the communication barrier. We then investigate several 1D systems, notably Markov systems and 1D random tilings, and derive their spectral properties. This is followed by an intermediate discussion of stochastic product tilings in arbitrary dimensions which already indicates that the appearance of mixed spectra with pure point, singular continuous and absolutely continuous parts is generic, though it also shows that its meaning in more than one dimension will have many facets. The next Section then deals with the main results of this article, the derivation of the diffraction spectrum of certain crystallographic random tilings in the plane, namely the domino and the rhombus (or lozenge) tiling. This requires some adaptation of results from the theory of Gibbs states to special hard-core lattice systems. Again, we explain that in slightly more detail than necessary from the point of view of mathematical physics in order to enhance self-containedness and readability. In both cases, the basic input of the explicit result has long been known in statistical mechanics, but the interest in the diffraction issue is rather recent. We also briefly comment on the diffraction of an interactive lattice gas based upon the classical 2D Ising model and its implications. The discussion addresses some open questions and what one should try to achieve next. \subsection*{Recollections from mathematical diffraction theory} Diffraction problems have many facets, but one important question certainly is which distributions of atoms lead to well-defined diffraction images, and if so, to what kind of images. This is a difficult problem, far from being solved. So, one often starts, as we will also do here, by looking at ``diffraction at infinity'' from single-scattering where it essentially reduces to questions of Fourier analysis \cite[Sec.\ 6]{Arsac}. This is also called kinematic diffraction in the Fraunhofer picture \cite{Cowley}, and we are looking into the more mathematical aspects of that now. Mathematical diffraction theory, in turn, is concerned with spectral properties of the Fourier transform of the autocorrelation measure of unbounded complex measures. Let us therefore first introduce and discuss the notions involved. Here, we start from the presentation in \cite{Hof,Hof-Waterloo} where the linear functional approach to measures is taken, compare \cite{Dieu} for details and background material. We also introduce our notation this way. Let ${\cal K}$ be the space of complex-valued continuous functions with compact support. A (complex) {\em measure} $\mu$ on $\mathbb R^n$ is a linear functional on ${\cal K}$ with the extra condition that for every compact subset $K$ of $\mathbb R^n$ there is a constant $a^{}_K$ such that \begin{equation} |\mu(f)| \; \leq \; a^{}_K \, \|f\| \end{equation} for all $f\in{\cal K}$ with support in $K$; here, $\|f\| = \sup_{x\in K} |f(x)|$ is the supremum norm of $f$. If $\mu$ is a measure, the {\em conjugate} of $\mu$ is defined by the mapping $f\to\overline{\mu(\bar{f})}$. It is again a measure and denoted by $\bar{\mu}$. A measure $\mu$ is called {\em real} (or signed), if $\bar{\mu}=\mu$, or, equivalently, if $\mu(f)$ is real for all real-valued $f\in{\cal K}$. A measure $\mu$ is called {\em positive} if $\mu(f)\geq 0$ for all $f\geq 0$. For every measure $\mu$, there is a smallest positive measure, denoted by $|\mu|$, such that $|\mu(f)|\leq |\mu|(f)$ for all non-negative $f\in{\cal K}$, and this is called the {\em absolute value} (or the total variation) of $\mu$. A measure $\mu$ is {\em bounded} if $|\mu|(\mathbb R^n)$ is finite (with obvious meaning, see below), otherwise it is called unbounded. Note that a measure $\mu$ is continuous on ${\cal K}$ with respect to the topology induced by the norm $\|.\|$ if and only if it is bounded \cite[Ch.\ XIII.20]{Dieu}. In view of this, the vector space of measures on $\mathbb R^n$, ${\cal M}(\mathbb R^n)$, is given the {\em vague topology}, i.e.\ a sequence of measures $\{\mu_n\}$ converges vaguely to $\mu$ if $\lim_{n\to\infty} \mu_n(f) = \mu(f)$ in $\mathbb C$ for all $f\in{\cal K}$. This is just the weak-* topology on ${\cal M}(\mathbb R^n)$, in which all the ``standard'' linear operations on measures are continuous, compare \cite[p.\ 114]{RS} for some consequences of this. The measures defined this way are, by proper decomposition \cite[Ch.\ XIII.2 and Ch.\ XIII.3]{Dieu} and an application of the Riesz-Markov representation theorem, see \cite[Thm.\ IV.18]{RS} or \cite[Thm.\ 69.1]{Berberian}, in one-to-one correspondence with the regular Borel measures on $\mathbb R^n$, wherefore we identify them. In particular, we write $\mu(A)$ (measure of a set) and $\mu(f)$ (measure of a function) for simplicity. {}For any function $f$, define $\tilde{f}$ by $\tilde{f}(x):=\overline{f(-x)}$. This is properly extended to measures via $\tilde{\mu}(f):=\overline{\mu(\tilde{f})}$. Recall that the convolution $\mu * \nu$ of two measures $\mu$ and $\nu$ is given by $\mu * \nu (f) := \int f(x+y) \mu(dx) \nu(dy) $ which is well-defined if at least one of the two measures has compact support. For $R>0$, let $B^{}_R$ denote the closed ball of radius $R$ with centre 0, and ${\rm vol}(B^{}_R)$ its volume. The characteristic function of a subset $A\subset\mathbb R^n$ is denoted by $1^{}_A$. Let $\mu^{}_R$ be the restriction of a measure $\mu$ to the ball $B^{}_R$. Since $\mu^{}_R$ then has compact support, \begin{equation} \gamma^R \; := \; \frac{1}{{\rm vol}(B^{}_R)} \, \mu^{}_R * \tilde{\mu}^{}_R \end{equation} is well defined. Every vague point of accumulation of $\gamma^R$, as $R\to\infty$, is called an {\em autocorrelation} of $\mu$, and as such it is, by definition, a measure. If only one point of accumulation exists, the autocorrelation is unique, and it is called the {\em natural autocorrelation}. It will be denoted by $\gamma$ or by $\gamma^{}_{\mu}$ to stress the dependence on $\mu$. One way to establish the existence of the limit is through the pointwise ergodic theorem, compare \cite{Dworkin}, if such methods apply. If not, explicit convergence proofs will be needed, as is apparent from known examples \cite{BMP} and counterexamples \cite{LP}. Note that Hof \cite{Hof} uses cubes rather than balls in his definition of $\gamma^R$. This simplifies some of his proofs technically, but they also work for balls which are more natural objects in a physical context. This is actually not important for our purposes here. One should keep in mind, however, that the autocorrelation will, in general, depend on the shape of the volume over which the average is taken --- with obvious meaning for the experimental situation where the shape corresponds to the aperture. To get rid of this problem, one often restricts the class of models to be considered and defines the limits over van Hove patches, thus demanding a stricter version of uniqueness \cite[Sec.\ 2.1]{Ruelle}. The space of complex measures is much too general for our aims, and we have to restrict ourselves to a natural class of objects now. A measure $\mu$ is called {\em translation bounded} \cite{AG} if for every compact set $K\subset\mathbb R^n$ there is a constant $b^{}_K$ such that \begin{equation} \sup_{x\in\mathbb R^n} |\mu|(K+x) \; \leq \; b^{}_K \, . \end{equation} For example, if $\Lambda$ is a point set of {\em finite local complexity}, i.e.\ if the set $\Delta=\Lambda-\Lambda$ of difference vectors is discrete and closed, the weighted Dirac comb \begin{equation} \label{comb1} \omega^{}_{\Lambda} \; := \; \sum_{x\in\Lambda} w(x) \delta_x \, , \end{equation} where $\delta_x$ is Dirac's measure at point $x$, is certainly translation bounded if the $w(x)$ are complex numbers with $\sup_{x\in\Lambda} |w(x)| < \infty$. This is so because $\Delta$ discrete and closed implies that $0\in\Delta$ is isolated and the points of $\Lambda$ are separated by a minimal distance, hence $\Lambda$ is uniformly discrete. Translation bounded measures $\mu$ have the property that all $\gamma^R$ are uniformly translation bounded, and if the natural autocorrelation exists, it is clearly also translation bounded \cite[Prop.\ 2.2]{Hof}. This is a very important property, upon which a fair bit of our later analysis rests. Note that such a restriction is neither necessary, nor even desirable (it would exclude the treatment of gases and liquids), but it is fulfilled in all our examples and puts us into a good setting in all cases where we cannot directly refer to pointwise ergodic theorems. Let us finally mention that different measures can lead to the same natural autocorrelation, namely if one adds to a given measure $\mu$ a sufficiently ``meager'' measure $\nu$, see \cite[Prop.\ 2.3]{Hof} for details. In particular, adding or removing finitely many points from $\Lambda$, or points of density 0, does not change $\gamma$, if it exists. Let us focus on the Dirac comb $\omega = \omega^{}_{\Lambda}$ from (\ref{comb1}), with $\Lambda$ of finite local complexity, and let us assume for the moment that its natural autocorrelation $\gamma^{}_{\omega}$ exists and is unique (Lagarias and Pleasants construct an example where this is not the case \cite{LP}). A short calculation shows that $\tilde{\omega}^{}_{\Lambda} = \sum_{x\in\Lambda} \overline{w(x)} \delta_{-x}$. Since $\delta_x * \delta_y = \delta_{x+y}$, we get \begin{equation} \label{comb2} \gamma^{}_{\omega} \; = \; \sum_{z\in\Delta} \nu(z) \delta_z \, , \end{equation} where the autocorrelation coefficient $\nu(z)$, for $z\in\Delta$, is given by the limit \begin{equation} \nu(z) \; = \; \lim_{R\to\infty} \frac{1}{{\rm vol}(B_R)} \sum_{\stackrel{\scriptstyle y \in \Lambda_R} {\scriptstyle z-y \in \Lambda}} w(y) \, \overline{w(z-y)} \, , \end{equation} where $\Lambda_R = \Lambda \cap B_R$. Conversely, if these limits exist for all $z\in\Delta$, the natural autocorrelation exists, too, because $\Delta$ is discrete and closed by assumption, and (\ref{comb2}) thus uniquely defines a translation bounded measure of positive type. This is one advantage of using sets of finite local complexity. We now have to turn our attention to the Fourier transform of unbounded measures on $\mathbb R^n$ which ties the previous together with the theory of tempered distributions \cite{Schwartz}, see \cite{AG,Martin98} for extensions to other locally compact Abelian groups. Let ${\cal S}(\mathbb R^n)$ be the space of rapidly decreasing functions \cite[Ch.\ VII.3]{Schwartz}, also called Schwartz functions. By the Fourier transform of a Schwartz function $\phi\in{\cal S}(\mathbb R^n)$ we mean \begin{equation} \label{fourierdef} ({\cal F}\phi) (k) \; = \; \hat{\phi} (k) \; := \; \int_{\mathbb R^n} e^{-2\pi i k\cdot x} \, \phi(x) dx \end{equation} which is again a Schwartz function \cite{Schwartz,RS}. Here, $k\cdot x$ is the Euclidean inner product of $\mathbb R^n$. The inverse operation is given by \begin{equation} \label{invfourierdef} \check{\psi} (x) \; = \; \int_{\mathbb R^n} e^{2\pi i x\cdot k} \, \psi(k) dk \, . \end{equation} The Fourier transform $\cal F$ is thus a linear bijection from ${\cal S}(\mathbb R^n)$ onto itself, and is bicontinuous \cite[Thm.\ IX.1]{RS}. Our definition (with the factor $2\pi$ in the exponent) results in the usual properties, such as $\check{\hat{\phi}\,}=\phi$ and $\hat{\check{\psi}\,}=\psi$. The convolution theorem takes the simple form $\widehat{\phi_1 * \phi_2} = \hat{\phi}_1 \cdot \hat{\phi}_2$ where convolution is defined by \begin{equation} \phi_1 * \phi_2 \, (x) \; := \; \int_{\mathbb R^n} \phi_1(x-y)\phi_2(y)dy \, . \end{equation} Let us also mention that $\cal F$ has a unique extension to the Hilbert space $L^2(\mathbb R^n)$, often called the Fourier-Plancherel transform, which turns out to be a unitary operator of fourth order, i.e.\ ${\cal F}^4=\mbox{Id}$. This is so because $({\cal F}^2\phi)(x)=\phi(-x)$, see \cite{Rudin} for details. {}Finally, the matching definition of the Fourier transform of a tempered distribution \cite{Schwartz} $T\in{\cal S}'(\mathbb R^n)$ is \begin{equation} \hat{T}(\phi ) \; := \; T( \hat{\phi} ) \end{equation} for all Schwartz functions $\phi$, as usual. The Fourier transform is then a linear bijection of ${\cal S}'(\mathbb R^n)$ onto itself which is the unique weakly continuous extension of the Fourier transform on ${\cal S}(\mathbb R^n)$ \cite[Thm.\ IX.2]{RS}. This is important, because it means that weak convergence of a sequence of tempered distributions, $T_n\to T$ as $n\to\infty$, implies weak convergence of their Fourier transforms, i.e.\ $\hat{T}_n\to \hat{T}$. Let us give three examples here, which will reappear later. First, the Fourier transform of Dirac's measure at $x$ is given by \begin{equation} \hat{\delta}_x \; = \; e^{-2\pi i k\cdot x} \end{equation} where the right hand side is actually the Radon-Nikodym density, and hence a function of the variable $k$, that represents the corresponding measure (we will not distinguish a measure from its density, if misunderstandings are unlikely). Second, consider the Dirac comb $\omega^{}_{\Gamma}=\sum_{x\in\Gamma}\delta_x$ of a lattice $\Gamma\subset\mathbb R^n$ (i.e.\ a discrete subgroup of $\mathbb R^n$ such that the factor group $\mathbb R^n/\Gamma$ is compact). Then, one has \begin{equation} \hat{\omega}^{}_{\Gamma} \; = \; {\rm dens}(\Gamma) \cdot \omega^{}_{\Gamma^*} \, , \end{equation} where ${\rm dens}(\Gamma)$ is the density of $\Gamma$, i.e.\ the number of lattice points per unit volume, and $\Gamma^*$ is the {\em dual} (or reciprocal) lattice, \begin{equation} \label{psf} \Gamma^* \; := \; \{ y\in\mathbb R^n\mid x\cdot y \in\mathbb Z \mbox{ for all } x\in\Gamma \} \, . \end{equation} This is {\em Poisson's summation formula} for distributions \cite[p.\ 254]{Schwartz} and will be central for the determination of the Bragg part of the diffraction spectrum. Finally, putting these two pieces together, we also get the formula \begin{equation} \label{exp-sum} \sum_{x\in\Gamma} e^{-2\pi i k\cdot x} \; = \; {\rm dens}(\Gamma) \cdot \sum_{y\in\Gamma^*} \delta_y \, , \end{equation} to be understood in the distribution sense. If a measure $\mu$ defines a tempered distribution $T_{\mu}$ by $T_{\mu}(\phi) = \mu(\phi)$ for all $\phi\in{\cal S}(\mathbb R^n)$, the measure is called a {\em tempered measure}. A sufficient condition for a measure to be tempered is that it increases only slowly, in the sense that $\int (1+|x|)^{-\ell} |\mu|(dx) < \infty$ for some $\ell\in\mathbb N$, see \cite[Thm.\ VII.VII]{Schwartz}. Consequently, every translation bounded measure is tempered -- and such measures form the right class for our purposes. We will usually not distinguish between a measure and the corresponding distribution, i.e.\ we will write $\hat{\mu}$ for $\hat{T}_{\mu}$. The Fourier transform of a tempered measure is a tempered distribution, but it need {\em not} be a measure. However, if $\mu$ is of {\em positive type} (also called positive definite) in the sense that $\mu(\phi*\tilde{\phi})\geq 0$ for all $\phi\in{\cal S}(\mathbb R^n)$, then $\hat{\mu}$ is a positive measure by the Bochner-Schwartz Theorem \cite[Thm.\ IX.10]{RS}. Every autocorrelation $\gamma$ is, by construction, a measure of positive type, so that $\hat{\gamma}$ is a positive measure. This explains why this is a natural approach to kinematic diffraction, because the observed intensity pattern is represented by a positive measure that tells us which amount of intensity is present in a given volume. Also, taking Lebesgue's measure as a reference, positive measures $\mu$ admit a unique decomposition into three parts, \begin{equation} \mu \; = \; \mu^{}_{pp} + \mu^{}_{sc} + \mu^{}_{ac} \, , \end{equation} where $pp$, $sc$ and $ac$ stand for pure point, singular continuous and absolutely continuous, see \cite[Sec.\ I.4]{RS} for background material. The set $P=\{x\mid\mu(\{x\})\neq 0\}$ is called the set of pure points of $\mu$, which supports the so-called Bragg part $\mu^{}_{pp}$ of $\mu$. Note that $P$ is at most a countable set. The rest, i.e.\ $\mu - \mu^{}_{pp}$, is the ``continuous background'' of $\mu$, and this is the unambiguous and mathematically precise formulation of what such terms are supposed to mean. Depending on the context, one also writes \begin{equation} \mu \; = \; \mu^{}_{pp} + \mu^{}_{cont} \; = \; \mu^{}_{sing} + \mu^{}_{ac} \, , \end{equation} where $\mu^{}_{cont}=\mu^{}_{sc}+\mu^{}_{ac}=\mu-\mu^{}_{pp}$ is the continuous part of $\mu$ (see above) and $\mu^{}_{sing}=\mu^{}_{pp}+\mu^{}_{sc}$ is the singular part, i.e.\ $\mu^{}_{sing}(S)=0$ for some set $S$ whose complement has vanishing Lebesgue measure (in other words, $\mu^{}_{sing}$ is concentrated to a set of vanishing Lebesgue measure). Finally, the absolutely continuous part, which is called diffuse scattering \cite{JF} in crystallography, can be represented by its Radon-Nikodym density \cite[Thm.\ I.19]{RS} which is often very handy. Examples for the various spectral types can easily be constructed by different substitution systems, see \cite{Queffelec} and references therein for details. Later on, we shall meet a simple example in the context of stochastic product tilings where all three spectral types are present, though their meaning will need a careful discussion. Hof discusses a number of properties of Fourier transforms of tempered measures \cite{Hof,Hof-Waterloo}. Important for us is the observation that temperedness of $\mu$ together with positivity of $\hat{\mu}$ implies translation boundedness of $\hat{\mu}$ \cite[Prop.\ 3.3]{Hof}. So, if $\mu$ is a translation bounded measure whose natural autocorrelation $\gamma^{}_{\mu}$ exists, then $\gamma^{}_{\mu}$ is also translation bounded (see above), hence tempered, and thus the positive measure $\hat{\gamma}^{}_{\mu}$ is also both translation bounded and tempered. This is the situation we shall meet throughout the article. In what follows, we shall restrict ourselves to the spectral analysis of measures $\mu$ that are concentrated on uniformly discrete point sets. They are seen as an idealization of pointlike scatterers at uniformly discrete positions, in the infinite volume limit. The rationale behind this is as follows. If one understands these cases well, one can always extend both to measures with extended local profiles (e.g.\ by convolution of $\omega$ with a smooth function of compact support or with a Schwartz function) and to measures that describe diffraction at positive temperatures (e.g.\ by using Hof's probabilistic treatment \cite{Hof95}). The treatment of gases or liquids might need some additional tools, but we focus on situation that stem from solids with long-range order and different types of disorder, because we feel that this is where the biggest gaps in our understanding are at present. \subsection*{Illustrative results in one dimension} The simplest cases to be understood are those in one dimension. We start with some examples obtained from stationary stochastic processes and then derive in detail the diffraction properties of 1D random tilings. The language and methods of this section closely follow those of classical ergodic theory because very good literature is available here \cite{Petersen}. \subsubsection*{Bernoulli and Markov systems} Let us start with a Bernoulli system, i.e.\ with a lattice gas without interaction. \begin{prop} \label{bernoulli} Consider the stochastic Dirac comb $\omega = \sum_{m\in\mathbb Z} \eta(m) \delta_m$ where $\eta(m)$ is a family of i.i.d.\ random variables that can take any of the $n$ complex numbers $h^{}_1, \ldots, h^{}_n$ (assumed pairwise different), with attached probabilities $p^{}_1, \ldots, p^{}_n$, $p^{}_i>0$. Then, the autocorrelation $\gamma^{}_{\omega}$ of $\omega$ exists with probabilistic certainty\footnote{Here and in the sequel, assertions of probabilistic certainty always refer to the invariant measure of the corresponding stochastic process.} and has the form $\gamma^{}_{\omega} = \sum_{m\in\mathbb Z} \nu(m) \delta_m$ with autocorrelation coefficients \begin{equation} \nu(m) \; = \; \begin{cases} \langle |\bs{h}|^2 \rangle \, , & \text{if $m=0$}, \\ |\langle \bs{h} \rangle|^2 \, , & \text{if $m\neq 0$}, \end{cases} \end{equation} where $\langle |\bs{h}|^2 \rangle = \sum_{i=1}^n p^{}_i |h^{}_i|^2$ and $\langle \bs{h} \rangle =\sum_{i=1}^n h^{}_i p^{}_i$. Consequently, the diffraction measure is, with probability one, $\mathbb Z$-periodic and given by \begin{equation} \hat{\gamma}^{}_{\omega} \; = \; \left(|\langle \bs{h} \rangle|^2 \sum_{m\in\mathbb Z} \delta_m \right) \; + \; \left(\,\langle |\bs{h}|^2 \rangle - |\langle \bs{h} \rangle|^2 \,\right) \, . \end{equation} \end{prop} {\sc Proof}: This is a straight-forward application of Birkhoff's pointwise ergodic theorem \cite[Thm.\ 2.3]{Petersen}, applied to the case of a Bernoulli system as described in \cite[Sec.\ 1.2 C]{Petersen}. One identifies the possible Dirac combs with the corresponding bi-infinite sequences $\bs{x}=(x^{}_i)^{}_{i\in\mathbb Z}$ of the Bernoulli process given above. Then, $\nu(m)$ is the orbit average of the function $f(\bs{x}) = \bar{x}^{}_0 x^{}_m$ under the action of the shift which almost surely exists and equals the average of $f$ over the invariant measure, because the process is ergodic. \hfill $\square$ \smallskip The diffraction thus consists of a pure point part (Bragg peaks) that is the one of the regular lattice $\mathbb Z$ multiplied by the absolute square of the average scattering strength and an absolutely continuous part (diffuse scattering) which is constant in this case (hence it is ``white noise''), as one would expect for a Bernoulli process. Note that the entropy density of this ensemble is given by $s=-\sum_{i=1}^n p_i\log(p_i)$, see \cite[Ch.\ 5.3, Ex.\ 3.4]{Petersen}. One could, alternatively, refer to the strong law of large numbers, under slightly different assumptions. This would then also give a generalization to higher dimensions, and to regular point sets beyond lattices, see \cite{BM1} for a detailed account of this. Let us now turn our attention to stochastic systems with interaction. Let $M$ be a Markov (or stochastic) matrix, i.e.\ $M=(M_{ij})^{}_{1\leq i,j\leq n}$ with $M_{ij}\geq 0$ and $\sum_{j=1}^{n}M_{ij}=1$. We assume that $M$ is {\em primitive}, so some power of $M$ has strictly positive entries only. As a consequence, the Perron-Frobenius (PF) eigenvalue $\lambda_1=1$ is unique, and all other eigenvalues $\lambda_i$ of $M$ have absolute value $|\lambda_i|<1$. The corresponding right eigenvector is $(1,1,\ldots,1)^t$, while the left eigenvector $\bs{p} = \bs{p} M$, $\bs{p}=(p^{}_1,\ldots,p^{}_n)$, defines the stationary state. Due to the primitivity of $M$, we can choose $p^{}_i>0$ and statistical normalization, $\sum_{i=1}^{n} p^{}_i=1$. Let $\Pi=\mbox{diag}(p^{}_1,\ldots,p^{}_n)$ which is thus invertible. In view of the applications we have in mind, we want to consider a Markov process that gives the same result for an operation in reverse direction. That is to say we restrict ourselves to {\em reversible} processes, i.e.\ to stochastic matrices $M$ with \begin{equation} \label{reversible} \Pi M \; = \; M^t \Pi \, . \end{equation} Since $M$ is also primitive, all $p^{}_i>0$ and $P=\Pi^{1/2}$ is well defined and non-singular, $P=\mbox{diag}(\sqrt{p^{}_1},\ldots,\sqrt{p^{}_n}\,)$. Now, $S=PMP^{-1}$ is a real symmetric matrix and can thus be diagonalized by an orthogonal matrix. The eigenvalues of $S$ and $M$ coincide and are real; we denote them by $\lambda_1=1, \lambda_2,\ldots,\lambda_n$, where $|\lambda_i|<1$ for all $i\geq 2$. If $\{\bs{b}_i\}_{1\leq i\leq n}$ is the corresponding orthonormal basis, the spectral decomposition of $S$ is \begin{equation}\label{decomp} S \; = \; |\bs{b}_1\rangle \langle\bs{b}_1| \oplus \sum_{i=2}^{n} |\bs{b}_i\rangle \lambda_i\langle\bs{b}_i| \; = \; S_0 \oplus S_1 \, , \end{equation} where we use Dirac's bra-ket notation for the standard Hermitian scalar product of $\mathbb C^n$. We embed $\mathbb R^n$ into $\mathbb C^n$ because we deal with complex scattering strengths later. Note that the decomposition on the right hand side of (\ref{decomp}) is into a projector (first term, $S_0$) and a contraction, i.e.\ $|S_1 \bs{x}| < |\bs{x}|$ for all $\bs{x}\in\mathbb C^n$. Also, we have $S_0 S_1 = S_1 S_0=0$ and, in the standard basis of $\mathbb C^n$, $S_0$ is explicitly given by $S_0=(\sqrt{p^{}_i p^{}_j}\,)^{}_{1\leq i,j \leq n}$. \begin{prop} \label{markov} Consider the stochastic Dirac comb $\omega = \sum_{m\in\mathbb Z} \eta(m) \delta_m$ where $\eta(m)$ is a family of random variables that take values out of the $n$ complex numbers $h^{}_1, \ldots, h^{}_n$ (assumed pairwise different), subject to a primitive, reversible Markov process defined by a matrix $M$ with left-PF-eigenvector $\bs{p}$ as described above. Then, the autocorrelation $\gamma^{}_{\omega}$ of $\omega$ exists with probabilistic certainty and has the form $\gamma^{}_{\omega} = \sum_{m\in\mathbb Z} \nu(m) \delta_m$ with non-negative autocorrelation coefficients \begin{equation} \nu(m) \; = \; \langle\bs{h}|\Pi M^{|m|}|\bs{h}\rangle \end{equation} for $m\in\mathbb Z$. In particular, $\nu(0)=\langle|\bs{h}|^2\rangle=\sum_{i=1}^{n} p^{}_i |h^{}_i|^2$. \end{prop} {\sc Proof}: This is another application of Birkhoff's pointwise ergodic theorem \cite[Thm.\ 2.3]{Petersen}, this time applied to the case of an ergodic Markov system as described in \cite[Sec.\ 1.2 D]{Petersen}. The setup is parallel to that of the Bernoulli system, only the invariant measure (defined via cylinder sets) is different, and this accounts for the different ensemble average. The latter is calculated as (for $m\geq 0$ say) $$ \nu(m) \; = \; \sum_{i^{}_0,i^{}_1,\ldots,i^{}_m} \bar{h}^{}_{i^{}_0} p^{}_{i^{}_0} M^{}_{i^{}_0 i^{}_1} M^{}_{i^{}_1 i^{}_2} \cdot \ldots \cdot M^{}_{i^{}_{m-1} i^{}_m} h^{}_{i^{}_m} \; = \; \langle \bs{h} | \Pi M^m | \bs{h} \rangle \, . $$ Due to (\ref{reversible}), we also have $$ \langle \bs{h} | \Pi M^m | \bs{h} \rangle \; = \; \langle \bs{h} | (M^t)^m \Pi | \bs{h} \rangle \; = \; \langle \bs{h} | (\Pi M^m)^{\dagger} | \bs{h} \rangle \; = \; \overline{\langle \bs{h} | \Pi M^m | \bs{h} \rangle} \, ,$$ which shows that $\overline{\nu(m)}=\nu(m)$, so $\nu(m)$ is real. The case $m<0$ is analogous. But we also have \begin{equation} \nu(m) \; = \; \langle \, \bs{h} \mid \frac{1}{2} (\Pi M^{|m|} + (M^t)^{|m|}\Pi) \mid \bs{h} \, \rangle \; \geq \; 0 \, , \end{equation} because it represents a quadratic form with a non-negative real symmetric (hence Hermitian) matrix. Finally, $m=0$ gives the result for $\nu(0)$. \hfill $\square$ \smallskip {}From the last Proposition, it is possible to derive the diffraction. Observe that, with $\bs{c} = P \bs{h}$ and $r > 0$, we have \begin{equation} \nu(r) \; = \; \langle \bs{c} | S^r | \bs{c} \rangle \; = \; \langle \bs{c} | S_0^r | \bs{c} \rangle + \langle \bs{c} | S_1^r | \bs{c} \rangle \, , \end{equation} and that $S_0$ is a projector, i.e.\ $S_0^r=S_0^{}$. With the explicit form of $S_0$ given above, we thus obtain ($r > 0$) \begin{equation} \nu(r) \; = \; \left|\sum_{i=1}^{n} p^{}_i h^{}_i\right|^2 + \langle \bs{c} | S_1^r | \bs{c} \rangle \; = \; |\langle\bs{h}\rangle|^2 + \langle \bs{c} | S_1^r | \bs{c} \rangle \, . \end{equation} With $\nu(0) = \langle|\bs{h}|^2\rangle = |\langle\bs{h}\rangle|^2 + (\langle|\bs{h}|^2\rangle-|\langle\bs{h}\rangle|^2)$, the autocorrelation gets the form \begin{equation}\label{markov-auto} \gamma^{}_{\omega} \; = \; |\langle\bs{h}\rangle|^2 \sum_{m\in\mathbb Z}\delta_m + (\langle|\bs{h}|^2\rangle-|\langle\bs{h}\rangle|^2)\delta_0 + \sum_{r=1}^{\infty} \langle\bs{c}|S_1^r|\bs{c}\rangle (\delta_r + \delta_{-r}) \, . \end{equation} Before we continue, let us point out that the Bernoulli case is a special case of this, namely $M_{ij}=1/n$, hence $p^{}_1=\ldots=p^{}_n=1/n$, $S_0=M$ and $S_1=0$. In this limit, (\ref{markov-auto}) gives back the corresponding result of Proposition \ref{bernoulli}. Also, one can treat more general cases of Markov chains, compare \cite[vol.\ 1, Ch.\ XV]{Feller1} for details, and Markov chains of higher order (or depth). This becomes technically more involved, but we think that the essential flavour is obvious from the situation discussed here. It should also be clear how to make the other examples of the crystallographic literature, see \cite{JF}, rigorous this way. Finally, let us remark that the similarity of the (binary) Markov chain to the 1D Ising model is anything but accidental, and that the form of the autocorrelation is closely related to the solution of the Ising model by means of transfer matrices, see \cite{Welberry} and \cite[Ch.\ 3.2]{Georgii}. The three terms on the right hand side of (\ref{markov-auto}) can now easily be Fourier transformed. The first, by means of Poisson's summation formula (\ref{psf}), gives the Bragg part. The second results in a constant continuous background, $\langle |\bs{h}|^2 \rangle - |\langle \bs{h} \rangle|^2$, as in our previous example. To calculate the Fourier transform of the third term, we can employ Neumann's series \cite[p.\ 191]{RS} twice, because with $S_1$ also $\exp(\pm2\pi i k) S_1$ is a contraction, for arbitrary $k\in\mathbb R$. This gives an absolutely continuous contribution which depends on the wave number $k$. Observing $\langle |\bs{h}|^2 \rangle = \langle \bs{h} | P^2 | \bs{h} \rangle$, one can combine the $k$-dependent part with the first term of the constant part. This finally gives \begin{theorem} The diffraction spectrum of the Markov system of Proposition \ref{markov} exists with probabilistic certainty and is given by the formula \begin{equation} \hat{\gamma}^{}_{\omega} \; = \; |\langle\bs{h}\rangle|^2 \cdot \omega^{}_{\mathbb Z} \; + \; (\hat{\gamma}^{}_{\omega})_{ac} \, , \end{equation} where $\omega^{}_{\mathbb Z}=\sum_{k\in\mathbb Z}\delta_k$ is the Dirac comb of the integer lattice. The absolutely continuous part $(\hat{\gamma}^{}_{\omega})_{ac}$ is represented by the $\mathbb Z$-periodic continuous (hence bounded) function \begin{equation}\label{cont2} f(k) \; = \; \langle \bs{h} \mid P \, \frac{1 - S_1^2}{1 - 2 \cos(2\pi k) S_1^{} + S_1^2} \; P \mid \bs{h} \rangle \; - \; |\langle\bs{h}\rangle|^2 \, , \end{equation} with obvious meaning of the quotient. \hfill $\square$ \end{theorem} Let us add that an explicit calculation of $f(k)$ is easily done by means of the orthonormal eigenbasis of $S$. If $P\bs{h}=\sum_{i=1}^{n} \beta_i \bs{b}_i$, one finds $|\beta_1|^2 = |\langle\bs{h}\rangle|^2$ and hence \begin{equation} \label{basis-sum} f(k) \; = \; \sum_{j=2}^{n} \frac{|\beta_j|^2 \, (1 - \lambda_j^2)} {1 - 2 \cos(2\pi k) \lambda_j^{} + \lambda_j^2} \; \geq \; 0 \, , \end{equation} which, for $n=2$, coincides with the result of \cite{JF,Welberry}. Note that the diffuse background corresponds to a positive entropy density which is given by $s=-\sum_{i,j} \, p_i M_{ij}\log(M_{ij})$, see \cite[Ch.\ 5.3, Ex.\ 3.5]{Petersen}. Let us also mention that (\ref{cont2}) can also be rewritten as \begin{equation} f(k) \; = \; \langle \bs{h} \mid P \, \frac{1 - S_1^2 - S^{}_0}{1 - 2 \cos(2\pi k) S_1^{} + S_1^2} \; P \mid \bs{h} \rangle \, , \end{equation} from which $f(k)\geq 0$ can easily be seen also in operator form. {}For a graphical illustration of the diffraction, we refer to \cite{Welberry}. The important observation is that, in the presence of an interaction, the (now structured) diffuse background is attracted or repelled by the Bragg peaks according to the interaction being attractive or repulsive. This is a general qualitative feature of diffuse scattering, and will reappear in our later examples. \subsubsection*{1D random tilings} Let us now change our point of view and consider a Bernoulli system not in the scattering strength $h$ but rather in the distances. To keep things simple, consider the case of placing two intervals, of length $u$ and $v$ (both $>0$), with probabilities $p$ and $q=1-p$ (hence, the entropy density per interval is $s=-p\log(p)-q\log(q)$), and assume that we have a unit point mass always at the left endpoint of each interval. Here, we have to distinguish the cases where $\alpha = u/v$ is rational or irrational. \begin{prop} \label{1drt} Consider the ensemble of binary random tilings of intervals of length $u$ and $v$, with probabilities $p$ and $q=1-p$, $pq>0$. Let, for each such random tiling, $\Lambda$ be the point set defined through the left endpoints of the intervals. Then, the natural density of $\Lambda$ exists with probabilistic certainty and is given by $d=(pu+qv)^{-1}$. If $\omega=\omega^{}_{\Lambda}=\sum_{x\in\Lambda}\delta_x$ denotes the corresponding stochastic Dirac comb, the autocorrelation $\gamma^{}_{\omega}$ of $\omega$ also exists with probabilistic certainty. It is a pure point measure that is supported on the set \begin{equation} \label{diff1} \Delta \; = \; \{ m u + n v \mid m,n\in\mathbb Z \mbox{ and } mn\geq 0 \} \, , \end{equation} and, with $z_{m,n}:=m u + n v$, it is given by \begin{equation} \label{corr3} \gamma^{}_{\omega} \; = \; d \; \sum_{N=-\infty}^{\infty} \sum_{\ell=0}^{|N|} \binom{|N|}{\ell} p^{\ell} q^{|N|-\ell} \, \delta_{{\rm sgn}(N) z^{}_{\ell,|N|-\ell}} \, . \end{equation} {}For $\alpha=u/v$ irrational, it is of the form $\gamma^{}_{\omega} = \sum_{z\in\Delta} \nu(z) \delta_z$, where $z\in\Delta$ has a unique representation $z=z_{m,n}=mu+nv$ with $m,n\in\mathbb Z$ and $mn\geq 0$. The corresponding autocorrelation coefficient is then given by \begin{equation} \label{coeff3} \nu(z_{m,n}) \; = \; d \, \binom{|n|\!+\!|m|}{|m|} \, p^{|m|} q^{|n|} \, . \end{equation} \end{prop} {\sc Proof}: Let $\Lambda$ be a random point set according to the assumptions. If $u=v$, $\omega^{}_{\Lambda}$ is the Dirac comb of a lattice and the statement is trivial. So, let us henceforth assume that $u\neq v$. If we view the random tiling ensemble as a Bernoulli system in the symbols $u,v$ with attached probabilities $p,q$, each tiling is a sequence $\bs{x}=(x^{}_i)^{}_{i\in\mathbb Z}$ with $x_i\in\{u,v\}$. Since $u,v$ also code the length of the intervals, the average distance between two consecutive points of the corresponding point set $\Lambda$ is the limit of $\frac{1}{2n+1}\sum_{i=-n}^{n} x_i$ as $n\to\infty$ which almost surely exists and, again by Birkhoff's pointwise ergodic theorem, is given by $(pu+qv)$. The density $d$ is then clearly the inverse of this, as stated. The possible differences between two points of $\Lambda$ are clearly given by $\Delta$ of Eq.~(\ref{diff1}), and this set is discrete and closed. To establish the existence of the autocorrelation, it is thus sufficient to show that its coefficients exist. Let $z=mu+nv$ be in $\Delta$. Although the representation of $z$ need not be unique (e.g.\ if $\alpha$ is rational), there is no other representation with $N=m+n$ intervals because we have excluded the case $u=v$. So, starting from an arbitrary point $x\in\Lambda$, $N+1$ different points can be reached by adding $N$ intervals (to the left or to the right, according to the sign of $z$), and the corresponding probabilities follow a binomial distribution, because the system is Bernoulli. If we pick a single sequence, and determine the average frequency to reach the point $x+z$ from $x$ in $N$ steps, the strong law of large numbers \cite[vol.\ 1, Ch.\ X.1]{Feller1} tells us that this frequency, almost surely, converges to the corresponding probability, i.e.\ we find the limiting frequency $$\binom{|n|+|m|}{|m|} p^{|m|} q^{|n|}\, .$$ Here, we have averaged over the number $M$ of starting points $x$ in a finite piece of the sequence $\Lambda$, and considered the limit $M\to\infty$. The corresponding contribution to the autocorrelation coefficient, however, is defined as a volume-averaged limit. This gives a prefactor that is the average number of points of $\Lambda$ per unit volume which is the density $d$. It exists with probability one as shown above. This establishes (\ref{corr3}). The full autocorrelation coefficient is now given by \begin{equation} \nu(z) \; = \; \lim_{r\to\infty} \frac{1}{2r} \sum_{\stackrel{\scriptstyle x\in\Lambda, \, |x|\leq r} {\scriptstyle x+z\in\Lambda}} 1 \; = \; d \sum_{\stackrel{\scriptstyle m,n\in\mathbb Z, \, mn\geq 0} {\scriptstyle mu+nv=z}} \binom{|n|+|m|}{|m|} p^{|m|} q^{|n|} \, , \end{equation} which clearly also exists with probability one. This shows the existence of $\gamma^{}_{\omega}$. Let us finally assume $\alpha\not\in\mathbb Q$. Then, $z_{m,n}=z_{m',n'}$ implies $m=m'$ and $n=n'$, so that the only possibility to fill this distance is by $m$ intervals of length $u$ and $n$ of length $v$, the remaining freedom just being the order in which this is done. So, the sum in the previous equation reduces to one term, the one given in (\ref{coeff3}). \hfill $\square$ \smallskip Let us add two remarks. First, the autocorrelation could also be worked out\footnote{We thank A.~Martin-L\"of for pointing this out to us.} by means of the renewal theorem \cite[vol.\ 2, Ch.\ 11]{Feller1}. This would have the advantage of also being applicable to gases. However, for our case, one has to pay attention to convergence questions wherefore the derivation is not shorter. Second, the argument given for the existence of the average distance between two consecutive points of $\Lambda$ can easily be modified to calculate that the average distance bridged by $N$ consecutive intervals is given by $N(pu+qv)$, because the system is Bernoulli. This average, however, can now also be calculated as the weighted sum over the possibilities to fill $N$ steps by $m$ intervals of type $u$ and $N-m$ of type $v$, i.e.\ we obtain the identity \begin{equation} \label{amaz1} \sum_{\ell=0}^{N} \binom{N}{\ell} p^{\ell} q^{N-\ell} \, (\ell u + (N-\ell) v) \; = \; N \, (pu+qv) \end{equation} which can also be checked explicitly by induction. It rests upon $p+q=1$, and the binomial formula for $(p+q)^N$. It can also be understood from the first moment of the binomial distribution, calculated as derivative of its generating function. To understand the diffraction, we have to determine the Fourier transform of $\gamma^{}_{\omega}$. To do so, it is advantageous to write the tempered measure $\gamma^{}_{\omega}$ as a weak limit of tempered measures with compact support, i.e.\ to write $\gamma^{}_{\omega} = \lim_{N\to\infty} \mu^{}_N$ where \begin{equation} \mu^{}_N \; = \; d \; \sum_{n=-N}^{N} \sum_{m=0}^{|n|} \binom{|n|}{m} p^{m} q^{|n|-m} \, \delta_{{\rm sgn}(n) z^{}_{m,|n|-m}} \, , \end{equation} with $z^{}_{m,n}=mu+nv$ as before. It is evident that this sequence of measures converges weakly to $\gamma^{}_{\omega}$ of (\ref{corr3}), and the support of $\mu^{}_N$ is certainly contained in the interval $[-w,w]$ where $w=N \max(u,v)$. So, due to the Paley-Wiener theorem \cite[Thm.\ IX.12]{RS}, the Fourier transform $\hat{\mu}^{}_N$ is naturally represented by an entire analytic function, $g^{}_N(k)$. Also, since the Fourier transform is continuous, the convergence of $\mu^{}_N\to\mu=\gamma^{}_{\omega}$ implies that of $\hat{\mu}^{}_N\to\hat{\mu}=\hat{\gamma}^{}_{\omega}$. Note, however, that the $\mu^{}_N$ are, in general, not measures of positive type, whence the $\hat{\mu}^{}_N$ are not positive measures. They are signed measures though, as we shall see shortly, and one could decompose them as $\hat{\mu}^{}_N=\hat{\mu}^{+}_N-\hat{\mu}^{-}_N$ with $\hat{\mu}^{\pm}_N=\frac{1}{2}(|\hat{\mu}^{}_N|\pm \hat{\mu}^{}_N)$. Then, $\hat{\mu}^{+}_N\to\hat{\gamma}^{}_{\omega}$ and $\hat{\mu}^{-}_N\to 0$ as $N\to\infty$, but later calculations would be more complicated wherefore we prefer to work with the signed measures $\hat{\mu}^{}_N$ rather than with the positive measures $\hat{\mu}^{+}_N$. \begin{theorem} \label{t1drt} Under the assumptions of Proposition \ref{1drt}, the diffraction spectrum consists, with probabilistic certainty, of a pure point (Bragg) part and an absolutely continuous part, so $\hat{\gamma}^{}_{\omega} = (\hat{\gamma}^{}_{\omega})_{pp} + (\hat{\gamma}^{}_{\omega})_{ac}$. If $\alpha=u/v$, the pure point part is \begin{equation} \label{point3} (\hat{\gamma}^{}_{\omega})_{pp} \; = \; d^2 \cdot \begin{cases} \delta_0 & \text{if $\alpha\not\in\mathbb Q$}, \\ \sum_{k\in\frac{1}{\xi}\mathbb Z} \delta_k & \text{if $\alpha\in\mathbb Q$}, \end{cases} \end{equation} where, if $\alpha\in\mathbb Q$, we set $\alpha=a/b$ with coprime $a,b \in\mathbb Z$ and define $\xi=u/a=v/b$. The absolutely continuous part $(\hat{\gamma}^{}_{\omega})_{ac}$ can be represented by the continuous function \begin{equation} \label{cont3} g(k) \; = \; \frac{d \cdot p q \sin^2(\pi k(u-v))} {p \sin^2(\pi ku) + q \sin^2(\pi kv) - pq \sin^2(\pi k(u-v))} \, , \end{equation} which is well defined for $k(u-v)\not\in\mathbb Z$. It has a smooth continuation to the excluded points. If $\alpha$ is irrational, this is $g(k)=0$ for $k(u-v)\in\mathbb Z$ with $k\neq 0$ and \begin{equation} \label{cont-irrational} g(0) \; = \; \frac{d\cdot pq (u-v)^2}{p u^2 + q v^2 - pq (u-v)^2} \; = \; d\, \frac{\, pq (u-v)^2}{(p u + q v)^2} \, . \end{equation} {}For $\alpha=a/b\in\mathbb Q$ as above, it is $g(k)=0$ for $k(u-v)\in\mathbb Z$, but $ku\not\in\mathbb Z$ (or, equivalently, $kv\not\in\mathbb Z$), and \begin{equation} \label{cont-rational} g(k) \; = \; d\, \frac{\, pq (a-b)^2}{(p a + q b)^2} \end{equation} for the case that also $ku\in\mathbb Z$. \end{theorem} {\sc Proof}: We employ the sequence of measures $\mu^{}_N$ introduced above which converges weakly to $\gamma^{}_{\omega}$. A direct calculation shows that the tempered measure $\hat{\mu}^{}_N$ is represented by the analytic function \begin{equation} \label{formal} g^{}_N(k) \; = \; d \, \sum_{n=-N}^{N} \left(p e^{-{\rm sgn}(n) 2\pi iku} + q e^{-{\rm sgn}(n) 2\pi ikv}\right)^{|n|}\, . \end{equation} If we define $r(k) = p e^{-2\pi iku} + q e^{-2\pi ikv}$, it is clear that this is a complex number inside the (closed) unit circle, i.e.\ $r(k)=R e^{i\phi}$ with $0\leq R\leq 1$ and $\phi\in[0,2\pi)$. This results in \begin{equation} g^{}_N(k) \; = \; d \,\left( 1 + 2 \,\sum_{m=1}^N R^m \cos(m\phi) \right)\, , \end{equation} which shows that $g^{}_N(k)$ represents a sequence of signed (or real), but not necessarily positive, measures. Their limit, however, is positive. Let us first check where the sequence of functions converges pointwise. We have $r(-k)=\overline{r(k)}$. Then, by the triangle inequality, $|r(k)|\leq p+q = 1$. Also, we have $|r(k)|^2 = 1 - 4pq\sin^2(\pi k(u-v))$, and thus $|r(k)|^2=1$ if and only if $p=0$, $p=1$, or $k(u-v)\in\mathbb Z$. We have excluded the trivial cases $p=0$ and $q=0$ by our assumptions (they correspond to periodic chains, up to defects of density zero). So, if $k(u-v)\not\in\mathbb Z$, the geometric series in (\ref{formal}) actually converges, with limit \begin{equation}\label{geom-limit} g(k) \; = \; d \; \frac{1-|r(k)|^2}{|1-r(k)|^2} \, , \end{equation} which immediately gives the expression in (\ref{cont3}). In particular, the denominator is always different from 0 for $k(u-v)\not\in\mathbb Z$. It is not difficult to check that $g(k)$ has a continuation to points $k$ with $k(u-v)\in\mathbb Z$. Consider first the case $\alpha$ irrational. Then, for $k\neq 0$, the denominator of (\ref{cont3}) is never 0, and $g(k)=0$ is the correct continuation. The case $k=0$ requires twice the application of de l'Hospital's rule, and gives the value of $g(0)$ of (\ref{cont-irrational}). Next, let $\alpha=a/b$ with coprime $a,b\in\mathbb Z$. If $k(u-v)\in\mathbb Z$, then $ku\in\mathbb Z$ if and only if $kv\in\mathbb Z$. So, if $ku\not\in\mathbb Z$, we are back to the case where $g(k)=0$ is the correct continuation, while $ku\in\mathbb Z$, again with de l'Hospital's rule, gives the extension stated in (\ref{cont-rational}). This in particular demonstrates that $g(k)$, with the appropriate continuation, is a continuous function. It is also a positive function, as can easily be checked, and thus represents an absolutely continuous positive measure. So, we have shown that $g^{}_N(k) - g(k)$ tends pointwise to 0, as $N\to\infty$, for all $k$ with $k(u-v)\not\in\mathbb Z$. The convergence is actually uniform on each compact interval that does not contain any of the exceptional points, as is clear from (\ref{geom-limit}). The latter form a 1D lattice of spacing $1/(u-v)$, and any singular part of $\hat{\gamma}^{}_{\omega}$ must thus be concentrated to this set. Since the latter is uniformly discrete, the singular part cannot be singular continuous, but at most consist of point measures. We now have to check what happens with $g^{}_N(k)$ for $k(u-v)\in\mathbb Z$, where the sequence of functions does {\em not} converge. First, let $\alpha$ be irrational, but $k\neq 0$. Then, it is impossible to have $r(k)=1$, because this would imply $ku\in\mathbb Z$ and hence $kv\in\mathbb Z$ -- a contradiction to $\alpha\not\in\mathbb Q$. But then, with $r(-k)=\overline{r(k)}$ and $|r(k)|\leq 1$, we get \begin{eqnarray} \label{converg} \frac{1}{d} \, |g^{}_N(k)| & \leq & 1 + 2 \, \left| \sum_{n=1}^{N} r(k)^n \right| \; = \; 1 + 2 \, \frac{|r(k)| \, |1-r(k)^{N}|}{|1-r(k)|} \nonumber \\ & \leq & 1 + 2 \, \frac{1 + |r(k)|^N}{|1-r(k)|} \; \leq \; 1 + \frac{4}{|1-r(k)|} \\ & = & 1 + \frac{2} {\sqrt{p \sin^2(\pi ku) + q \sin^2(\pi kv) - p q \sin^2(\pi k(u-v))} } \nonumber \end{eqnarray} which, for each fixed $k\neq 0$, has a denominator $\neq 0$. So, the sequence $g^{}_N(k)$, and hence (\ref{formal}), stays bounded in this case, even though it does not converge. With the previous result on $g$, this means that the sequence $g^{}_N$ is uniformly bounded on each closed interval that does not include $0$. Hence, $\hat{\gamma}^{}_{\omega}$ cannot be singular at the exceptional points $k$ with $k(u-v)\in\mathbb Z$ unless $k=0$. The analogous argument applies if $\alpha=a/b\in\mathbb Q$, as long as $ku\not\in\mathbb Z$: the sequence $g^{}_N$ is uniformly bounded on each closed interval that does not include any of the exceptional points with $ku\in\mathbb Z$, which we will call singular from now on. So, the remaining cases are $k=0$ for $\alpha$ irrational resp.\ $ku\in\mathbb Z$ for $\alpha$ rational. For such singular $k$, (\ref{formal}) gives \begin{equation} g^{}_N(k) \; = \; d (2N+1) \, , \end{equation} and this means that $g^{}_N(k)$ diverges for these $k$, always with the same rate, and the divergence is proportional to the system size. To make this precise, consider first $\alpha=a/b$ with coprime $a,b\in\mathbb Z$. The singular points are then $k=\ell a/u=\ell b/v$ for $\ell\in\mathbb Z$. Define $L(N)=\lfloor N(p a + q b)\rfloor$, where $\lfloor x\rfloor$ denotes the integer part of a positive $x$, and set \begin{equation} h^{}_N(k) \; = \; d^2 \xi \sum_{m=-L(N)}^{L(N)} e^{-2\pi i k m\xi} \end{equation} with $\xi=u/a=v/b$. Clearly, $h^{}_N(k) = d (2N+1) + {\cal O}(1)$ for all singular $k$. The form of $h^{}_N(k)$ is fixed by the requirement that both the height and the width of the finite approximations to the point measures at the singular points equals that of $g^{}_N(k)$, up to lower order terms. One can now show (though we will skip the details here) that $g^{}_N(k) - h^{}_N(k)$ is bounded on each compact interval even if it does contain singular values of $k$. On the other hand, we know from (\ref{exp-sum}) that \begin{equation} \lim_{N\to\infty} h^{}_N(k) \; = \; d^2 \xi \sum_{x\in\xi\mathbb Z} e^{-2\pi i k x} \; = \; d^2 \sum_{y\in\frac{1}{\xi}\mathbb Z} \delta_y \, . \end{equation} This then proves that $g^{}_N(k)$ converges to a point measure concentrated on $\frac{1}{\xi}\mathbb Z$ plus the $ac$ measure derived above. Similarly, one deals with the case $\alpha$ irrational, but $k=0$ (e.g.\ by taking a suitable sequence of rational cases while letting $\xi\to 0$). So, for all singular $k$, the structure factor converges to $d^2$, i.e.\ $(\hat{\gamma}^{}_{\omega})_{pp} (\{k\})\; = \; d^2$, and we obtain the result given in Eq.~(\ref{point3}). Consequently, the positive measure $\hat{\gamma}^{}_{\omega}$ has the decomposition claimed, and the absolutely continuous part can be represented by the continuous function $g(k)$ (with the appropriate continuation to all $k$). This is the Radon-Nikodym density \cite[Thm.\ I.19]{RS} which is uniquely determined almost everywhere. \hfill $\square$ \smallskip Let us mention that the function $g(k)$ appears simpler than it is --- if one tries a selection of different parameters and produces some plots, one quickly realizes that it actually shows some ``spiky'' structure (though it is smooth), and the way it does depends rather critically on the nature of $\alpha=u/v$. For example, if $\alpha=\tau$ (the golden ratio), a rather regular pattern emerges, and localized bell-shaped needles of increasing height appear at sequences of positions that scale with $\tau$. This is reminiscent of what happens in perfect Fibonacci model sets. If, however, $\alpha$ is transcendental (e.g.\ $\pi$), much more pronounced needles appear, but at rather irregular positions. This is a clear consequence of how these numbers can be approximated by rationals, and an analogous phenomenon is well known in the approximation of irrational numbers by finite continued fractions \cite[Ch.\ VII]{Cassels}. We have discussed the case of a binary random tiling in detail, to make the structure as transparent as possible. It is clear that one can treat, with the same methods, also the case of a random tiling with $n$ tiles of length $(u^{}_1,\ldots,u^{}_n)=\bs{u}$, $u^{}_i>0$, and attached frequencies $(p^{}_1,\ldots,p^{}_n)=\bs{p}$, $p^{}_i>0$, $\sum_{i=1}^{n}p^{}_i=1$. Let us state the results in an informal way, as they are extremely parallel to what we discussed above. Viewing this system again as a Bernoulli system in the symbols $u^{}_i$ reveals that the mean free path between two consecutive points of $\Lambda$ is, almost surely, given by $\bs{p}\cdot\bs{u}$, and the density is then $d=1/(\bs{p}\cdot\bs{u})$. To simplify the following formulas, it is advantageous to adopt standard multi-index notation. So, $\bs{m}=(m^{}_1,\ldots,m^{}_n)$ is a vector of non-negative integers, $|\bs{m}|^{}_1 = m^{}_1+\cdots +m^{}_n$ its 1-norm, and $\bs{p}^{\bs{m}}=p_1^{m^{}_1}\cdot\ldots\cdot p_n^{m^{}_n}$. Also, we shall need the multinomial coefficient \begin{equation} \binom{N}{\bs{m}} \; = \; \frac{N!}{m^{}_1! m^{}_2! \cdot\ldots\cdot m^{}_n!} \end{equation} where $N=|\bs{m}|^{}_1$. Let $\Lambda$ be the vertex set of a random tiling of this kind. The natural autocorrelation of $\omega=\omega^{}_{\Lambda}$ exists with probabilistic certainty, and has the form $\gamma^{}_{\omega} = \sum_{z\in\Delta} \nu(z)\delta_z$ where $\Delta = \{ \pm z \mid z = \bs{m}\cdot\bs{u} \}$ and the autocorrelation coefficient is given by \begin{equation} \nu(z) \; = \; d \sum_{\bs{m}\cdot\bs{u}=z} \binom{|\bs{m}|^{}_1}{\bs{m}} \bs{p}^{\bs{m}} \, . \end{equation} This is the previous result with the binomial structure replaced by a multinomial one. In particular, we also get an analogue of Eq.~(\ref{amaz1}), namely \begin{equation} \sum_{|\bs{m}|^{}_1=N} \binom{N}{\bs{m}} \bs{p}^{\bs{m}} \, (\bs{m}\cdot\bs{u}) \; = \; N \, (\bs{p}\cdot\bs{u}) \, . \end{equation} The autocorrelation measure $\gamma^{}_{\omega}$ can again be approximated by a weakly converging sequence of measures $\mu^{}_N$, and their Fourier transform now reads \begin{equation} \hat{\mu}^{}_N \; = \; d\cdot \sum_{m=-N}^{N} \left( \sum_{j=1}^{n} p^{}_j e^{-{\rm sgn}(m) 2\pi iku^{}_j} \right)^{|m|} \, . \end{equation} The analysis of pointwise convergence then reveals once again that the diffraction spectrum consists of a pure point part and an absolutely continuous part. The absolutely continuous part of the diffraction, $(\hat{\gamma}^{}_{\omega})_{ac}$, is represented by the continuous Radon-Nikodym density \begin{equation} g(k) \; = \; \frac{d \cdot \sum_{j<\ell} \, p^{}_{j} p^{}_{\ell} \sin^2(\pi k(u^{}_j - u^{}_{\ell}))} {\sum_j \, p^{}_j \sin^2(\pi k u^{}_j) - \sum_{j<\ell} \, p^{}_{j} p^{}_{\ell} \sin^2(\pi k(u^{}_j - u^{}_{\ell}))} \, , \end{equation} which is well defined as long as not all $k(u^{}_j - u^{}_{\ell})$ are integer. If they are, the continuation is to $g(k)=0$ if $k\bs{u}\not\in\mathbb Z^n$ and otherwise to \begin{equation} g(k) \; = \; \frac{d \cdot \sum_{j<\ell}\, p^{}_j p^{}_{\ell} (u^{}_j -u^{}_{\ell})^2} {\sum_j \, p^{}_j u^2_j \, - \, \sum_{j<\ell}\, p^{}_j p^{}_{\ell} (u^{}_j -u^{}_{\ell})^2} \; = \; \frac{d \cdot \sum_{j<\ell}\, p^{}_j p^{}_{\ell} (u^{}_j -u^{}_{\ell})^2} {(\bs{p}\cdot\bs{u})^2} \, . \end{equation} The Bragg part of the diffraction is determined by the condition that, whenever $k\bs{u}\in\mathbb Z^n$, $k$ is a pure point, and results in a contribution of $d^2 \delta_k$ to $(\hat{\gamma}^{}_{\omega})_{pp}$. Thus, if $\bs{u} = \xi (a^{}_1,\ldots,a^{}_n)$ with $a^{}_i\in\mathbb Z$ and $\gcd(a^{}_1,\ldots,a^{}_n)=1$, the condition $k\bs{u}\in\mathbb Z^n$ is equivalent to $k\xi\in\mathbb Z$, and we get a 1D lattice Dirac comb, \begin{equation} (\hat{\gamma}^{}_{\omega})_{pp} \; = \; d^2 \cdot \sum_{k\in\frac{1}{\xi}\mathbb Z} \delta_k \, . \end{equation} If, however, at least one quotient $u^{}_i/u^{}_j$ is irrational, we only get the trivial point part, $d^2 \delta_0$, while all other peaks are extinct. One reason for this rather detailed discussion will become apparent shortly when we use this to describe a class of very simple stochastic tilings in higher dimension. \subsection*{Intermezzo: Stochastic product tilings} With the use of 1D random tilings one can construct a particularly simple class of stochastic tilings in higher dimension, by simply taking one 1D random tiling per Cartesian direction and considering the space filling by cuboids obtained that way. The prototiles are thus cuboids whose edges in direction $j$ is any of the possible lengths of the $j$th 1D random tiling used. Since the diffraction theory of these objects is essentially an exercise in direct products, but nevertheless quite useful and instructive, we describe the result in an informal way. Note, however, that the entropy density of these tilings is zero if $D>1$, so that they are no ordinary random tilings in the sense of \cite{Henley} or \cite{Richard}. Consider $D$ different 1D random tilings, and the corresponding point sets $\Lambda_i$, $1\leq i\leq D$, characterized by vectors of possible tile lengths $\bs{u}^{(i)}$ and frequency vectors $\bs{p}^{(i)}$. The total number of tiles in each case may be different, and is given by $n_i$. Let us now consider the Cartesian product \begin{equation} \Lambda \; = \; \Lambda^{}_1\times\ldots\times\Lambda^{}_D \; = \; \{ (x^{}_1,\ldots,x^{}_D)\mid x^{}_i\in\Lambda_i \} \, , \end{equation} which is the vertex set of a stochastic tiling in $D$ dimensions whose prototiles are the $n^{}_1\cdot\ldots\cdot n^{}_D$ cuboids obtained as Cartesian products of the intervals $u^{(i)}_j$ with $1\leq i\leq D$ and $1\leq j\leq n_i$. The sets $\Lambda$ are thus all of finite local complexity, and we have \begin{equation} \Delta \; = \; \Lambda-\Lambda \; = \; \{ \bs{z}=(z^{}_1,\ldots,z^{}_D)\mid z^{}_i\in\Delta_i \} \end{equation} with probability one, where $\Delta_i=\Lambda_i - \Lambda_i$. Also, the density of $\Lambda$ exists again with probabilistic certainty, and is given by $d=d^{}_1\cdot\ldots\cdot d^{}_D$ where $d_i = (\bs{p}^{(i)}\cdot\bs{u}^{(i)})^{-1}$ as derived in the previous Section. The product structure of $\Lambda$ also implies that $\omega=\sum_{\bs{x}\in\Lambda}\delta_{\bs{x}}$ is a product measure (or, which is equivalent in this case, the tensor product of distributions, see \cite[Chap.\ IV]{Schwartz}), i.e.\ we have \begin{equation} \omega \; = \; \prod_{i=1}^D \omega^{(i)} \; = \; \prod_{i=1}^D \left( \sum_{x_i\in\Lambda_i}\delta^{(i)}_{x_i} \right) \, , \end{equation} where $\delta^{(i)}$ is meant as a 1D Dirac measure acting in the space along the $i$th coordinate. It also follows, with probabilistic certainty, that the autocorrelation $\gamma^{}_{\omega}$ exists and is a product measure, too. To prove the existence of the corresponding coefficients $\nu(\bs{z})$, it is easiest to take averages over cubes rather then balls, i.e.\ to use the definition of \cite{Hof}. Here, this gives the same limit as our definition of the natural autocorrelation. One obtains $\nu(\bs{z}) = \prod_{i=1}^D \nu^{(i)}(z_i)$, where $\nu^{(i)}(z_i)$ is the coefficient of the autocorrelation attached to $\Lambda_i$. So we have \begin{equation} \gamma^{}_{\omega} \; = \; \prod_{i=1}^D \left( \sum_{z_i\in\Lambda_i} \nu^{(i)}(z_i)\delta^{(i)}_{z_i} \right)\, . \end{equation} {}Finally, let us consider the diffraction spectrum $\hat{\gamma}^{}_{\omega}$. Since it is the Fourier transform of a product measure, it is a product measure itself \cite[Thm.\ XIV]{Schwartz}, and we thus obtain \begin{equation} \hat{\gamma}^{}_{\omega} \; = \; \prod_{i=1}^D \hat{\gamma}^{}_{\omega^{(i)}} \; = \; \prod_{i=1}^D \left( \, (\hat{\gamma}^{}_{\omega^{(i)}})^{}_{pp} + (\hat{\gamma}^{}_{\omega^{(i)}})^{}_{ac} \, \right) \, , \end{equation} where the $\hat{\gamma}^{}_{\omega^{(i)}}$ are determined through Theorem \ref{t1drt}. In particular, the absolutely continuous (pure point) part of $\hat{\gamma}^{}_{\omega}$ is precisely the product of the $ac$ ($pp$) parts of the $\hat{\gamma}^{}_{\omega^{(i)}}$, while all other combinations result in singular continuous components --- though the meaning of this will require some thought. The $sc$ property can be seen from the fact that terms in the expanded product which contain at least one component of each kind are concentrated to a support of vanishing Lebesgue measure, but contain no pure points themselves --- whence they must be singular continuous relative to Lebesgue measure. This also agrees with the common, intuitive scaling picture: a term with $m$ $ac$-components and $D-m$ $pp$-components would show intensities that stem from amplitudes (or Fourier-Bohr coefficients) which show a finite-size scaling with $L^{m/2} L^{D-m} = (L^D)^{\beta}$ where $L$ is the linear system extension and \begin{equation} \frac{1}{2} \; \leq \; \beta=1-\frac{m}{2D} \; \leq \; 1 \, . \end{equation} Here, $\beta=\frac{1}{2}$ and $\beta=1$ correspond to the cases of $ac$ and $pp$ part, respectively, compare the discussion in \cite[Sec.\ 6]{Hof} and \cite{Hof-scaling}. However, this notion of singular continuity is to be taken with a grain of salt. All we have constructed here are product measures, and such objects would perhaps not qualify to be singular continuous in a ``generic'' sense. They do show up in liquid crystals though, compare the discussion on the nature of their long-range order in \cite{HL}. In particular, they appear in Danzer's aperiodic prototile in 3-space \cite[Sec.\ 4]{Danzer}, which can be seen as a toy model of a smectic C$^*$ liquid crystal. Nevertheless, these simple product tilings show that one has to expect a larger variety of spectral types in higher dimensions, and that already in planar cases the appearance of singular continuous components should be typical. Also, one can easily construct examples with all three spectral types present. More specific and genuine random tilings, however, might bypass this, as we shall see in the next Section, although that should not be considered generic. \input{ran2d-6} \subsection*{Outlook} The diffraction of crystallographic and of perfect quasi-crystallographic structures is well understood. The main aim of this article was to begin to counterbalance this into the direction of certain stochastic arrangements of scatterers, and to random tiling arrangements in particular. This requires a careful investigation of the diffuse background and clear concepts about absolutely versus singular continuous contributions. Such problems are naturally studied via spectral properties of unbounded complex measures which we did for a number of simple, but relevant examples. While stochastic cuboid tilings would generically display a singular continuous contribution in the diffraction image, our results on the domino and the lozenge tilings show that no such contributions exist there. However, we do not think that this is a robust result. In fact, the natural next step would be an extension to planar random tilings with quasi-crystallographic symmetries such as 8-, 10- or 12-fold. Then, according to the folklore results, one should expect a purely continuous diffraction spectrum (except for the trivial Bragg peak at $\bs{k}=0$ which merely reflects the existing natural density of the scatterers). This spectrum should then split into an $ac$ and an $sc$ part. But is the replacement of Bragg peaks by $sc$ peaks significant? Scaling arguments \cite{Henley} and numerical calculations \cite{Dieter} indicate that the exponents of important $sc$ peaks can be extremely close to $1$ which means that their distinction from Bragg peaks is almost impossible in practice. Is it possible that this is one reason why structure refinement is so difficult for real decagonal quasicrystals? This certainly demands further thought, but it is not clear at the moment to what extent a rigorous treatment is possible. Finally, except for Bernoulli type systems \cite{BM1} and for extensions by means of products of measures, we have not touched the ``real'' diffraction issues in 3-space. One reason is that we presently do not know of any interesting model that can be solved exactly (let alone rigorously), another is that, for random tiling models relevant to real quasicrystals, one expects bounded fluctuations (again, due to heuristic scaling arguments, see \cite{Henley} and references therein). Consequently, the diffraction spectra should typically show Bragg peaks plus an absolutely continuous background. We hope to report on some progress soon. \subsection*{Acknowledgements} It is our pleasure to thank Joachim Hermisson, Anders Martin-L\"of, Robert V.\ Moody, Wolfram Prandl and Martin Schlott\-mann for several helpful discussions and comments. We are grateful to Aernout C.\ D.\ van Enter for helpful advice and for setting us right in the Addendum. This work was supported by the Cusanuswerk and by the German Research Council (DFG). \clearpage
\section{Introduction and summary} There is a growing interest in the magnetic properties of synthesized molecules \cite{TDP94,GCR94,FST96,LGB97} containing relatively small numbers of paramagnetic ions. With the ability to control the placement of magnetic moments of diverse species within stable molecular structures, one can test basic theories of magnetism and even begin to explore the design of novel systems that offer the prospect of useful applications. Most species of organic-based molecular magnets exhibit very weak intermolecular magnetic interactions, so that measurements performed on a bulk sample actually reflect intramolecular interactions only. The magnetic interaction appears to be well described by the Heisenberg model with isotropic, nearest-neighbor exchange. A key quantity is the time- and temperature-dependent correlation function for pairs of magnetic moments, as it serves as the basic ingredient for understanding diverse dynamical phenomena, such as inelastic neutron scattering \cite{BaL89} and spin lattice relaxation \cite{Mor56}. The present study is motivated by a desire to achieve a deeper understanding of spin dynamics in the Heisenberg model, especially concerning the trends that occur in arrays of $N$ interacting moments (individual spins $s$) for increasing values of both $N$ and $s$. The classical Heisenberg model turns out to provide accurate quantitative results for static properties, such as magnetic susceptibility, down to thermal energies of the order of the exchange coupling \cite{Sch99,LLB98}. It is quite easy to establish the connection of that model, for static properties, to the corresponding quantum model for arbitrary $s$. However, considerable care is required to successfully link up with classical Heisenberg spin dynamics starting from quantum Heisenberg spin dynamics. In this article we present the analytical form of the time-dependent equilibrium auto\-correlation function of the quantum mechanical dimer with general spin $s$. The trends for increasing $s$ are explored in some detail and in particular we compare the quantum results with the exact analytical result recently derived \cite{LBC,LuL99} for a classical Heisenberg spin dimer. The quantum results for arbitrary $s$ are obtained using Mathematica$^{\circledR}$ to evaluate the Clebsch-Gordan coefficients. Results have previously been obtained for $s={\frac{1}{2}}$ spin rings of length up to $N=16$ by complete diagonalization methods \cite{FaL97}, and in Ref.~\cite{MTP81} some aspects of the high spin limit were discussed. The present study of the equilibrium autocorrelation function for large values of $s$ is timely given the fact that NMR measurements have very recently been performed \cite{LTA99} on a dimer molecular magnet composed of Fe$^{3+}$ $(s=5/2)$ ions. Heretofore only $s={\frac{1}{2}}$ dimers have been available for NMR studies \cite{Kaw66,SvT71,FIK96}. For comparison between theory and experiment it will also be necessary to incorporate molecular and single-ion anisotropy terms in the Hamilton operator. This will be the subject of a forthcoming article. \section{The quantum dimer} The quantum dimer is specified by the Hamilton operator \begin{eqnarray} \label{E-2-1} \op{H} &=& \frac{J}{\hbar^2}\; \op{\vek{s}}_1 \cdot \op{\vek{s}}_2 = \frac{J}{2 \hbar^2}\; \left( \op{\vek{S}}^2 - \op{\vek{s}}_1^2 - \op{\vek{s}}_2^2 \right) \quad\ ;\quad \op{\vek{S}}=\op{\vek{s}}_1 + \op{\vek{s}}_2 \ , \end{eqnarray} where $J>0$ describes antiferromagnetic and $J<0$ ferromagnetic coupling. Throughout this article it is assumed that the spin quantum numbers of both sites of the dimer are identical, $s_1=s_2=s$. The eigenstates $\ket{S\,M}$ of total spin $\op{\vek{S}}^2$ \begin{eqnarray} \label{E-2-2} \op{\vek{S}}^2 \ket{S\,M} &=& \hbar^2 S (S+1) \ket{S\,M} \ ,\qquad \op{{S}}_z \ket{S\,M} = \hbar M \ket{S\,M} \end{eqnarray} are also eigenstates of the Hamilton operator with eigenvalues $E_S$, which, in the absence of a magnetic field, do not depend on the total magnetic quantum number $M$ \begin{eqnarray} \label{E-2-3} \op{H} \ket{S\,M} &=& \frac{J}{2}\; \left( S (S+1) - 2\,s (s+1) \right) \ket{S\,M} = E_S\,\ket{S\,M} \ . \end{eqnarray} Thus the partition function in the canonical ensemble reads \begin{eqnarray} \label{E-2-4} Z &=& \mbox{tr}\left\{ e^{-\beta\op{H}}\right\} = \sum_{S,M} \bra{S\,M} e^{-\beta\op{H}} \ket{S\,M} \\ &=& e^{\beta\,J\,s (s+1)} \;\sum_{S=0}^{2\,s} (2S+1)\;e^{-\frac{\beta\,J}{2}\,S\,(S+1)} \nonumber \end{eqnarray} and, considering that the Hamilton operator \fmref{E-2-1} is isotropic, one obtains for the unnormalized autocorrelation function \begin{eqnarray} \label{E-2-5} \Mean{\op{\vek{s}}_1(t) \cdot \op{\vek{s}}_1(0)} &=& \frac{3}{Z} \sum_{S,M} \bra{S\,M} \op{{s}}_{1z}(t) \cdot \op{{s}}_{1z}(0)\, e^{-\beta\op{H}} \ket{S\,M} \\ &=& \frac{3}{Z} \sum_{S,M,S^\prime,M^\prime} e^{\frac{i\,t}{\hbar}(E_S-E_{S^\prime})}\; e^{-\beta\,E_S}\; |\bra{S\,M}\op{{s}}_{1z}\ket{S^\prime\,M^\prime}|^2 \nonumber \ . \end{eqnarray} \begin{figure}[t] \begin{center} \epsfig{file=F-2.5.eps,height=150mm} $\qquad$ \epsfig{file=AF-2.5.eps,height=150mm} \caption{Normalized autocorrelation function $\Re{C(t)}$, see \eqref{E-A-2}, for a spin-$\frac{5}{2}$-dimer for four different temperatures (solid lines). The left panels display our results for the ferromagnetic dimer, the right panels the antiferromagnetic case \cite{MeS99}. The dashed lines show the classical result; for details see the next section.} \label{F-2-1} \end{center} \end{figure} The last expressions simplifies when we take into account that only matrix elements with $M=M^\prime$ and a difference in total spin not larger than one contribute \cite{CoS64}, i.e. \begin{eqnarray} \label{E-2-6} \bra{S\,M}\op{{s}}_{1z}\ket{S^\prime\,M^\prime} = 0 \qquad\mbox{if}\qquad |S-S^\prime| > 1 \qquad\mbox{or}\qquad M\ne M^\prime \ . \end{eqnarray} The resulting expression is shown in \eqref{E-A-1} in the appendix. Note that the angular frequency spectrum of the spin dimer is given by integer multiples of ${J}/{\hbar}$, which in turn means that the autocorrelation function is periodic in time and that the recurrence time $\tau$ only depends on the coupling $J$ but not on the spin quantum number $s$ \begin{eqnarray} \label{E-2-8} \omega \in \left\{ \frac{J}{\hbar} S \right\} \ , \quad S=0,\dots , 2\,s \quad\Rightarrow\quad \tau = \frac{2\,\pi\,\hbar}{J} \ . \end{eqnarray} This is of course true for all Hamilton operators that can be written like the term on the r.h.s. of \eqref{E-2-1}, namely for the spin trimer and the spin tetrahedron. Figure \xref{F-2-1} shows the autocorrelation function normalized to unity at $t=0$ for a spin-$\frac{5}{2}$-dimer, a system that has been synthesised (Fe dimer) and that is currently under investigation \cite{LTA99}. The analytical expression for this autocorrelation function is given in \eqref{E-A-2} in the appendix. One clearly sees that the autocorrelation function, which is a superposition of five harmonic oscillations and a constant, is dominated at low temperatures by the highest frequency in the ferromagnetic case and by the lowest frequency in the antiferromagnetic case. At higher temperatures other frequencies also contribute. One also notices that, independent of temperature, the autocorrelation function returns to its initial value after $\tau = \frac{2\,\pi\,\hbar}{J}$. \section{Comparison to the classical dimer} In order to compare the results of the quantum dimer for different spin quantum numbers $s$ with each other and with the classical dimer it is useful to introduce normalized spin operators \begin{eqnarray} \label{E-3-0} \op{\vek{\epsilon}}_n = \frac{\op{\vek{s}}_n}{\sqrt{\hbar^2\,s(s+1)}} \ ,\quad n=1,2 \quad \ , \end{eqnarray} which depend on $s$. Note, that \begin{eqnarray} \label{E-3-10} \left[ \op{\epsilon}_{n\,x} ,\op{\epsilon}_{n\,y} \right] = \frac{i}{\sqrt{s(s+1)}}\; \op{\epsilon}_{n\,z} \ , \end{eqnarray} and hence these become commuting operators for $s\rightarrow\infty$. \eqref{E-3-0} suggests that we define a classical Hamilton function $H_{c}$ \begin{eqnarray} \label{E-3-1} H_{c} = J_{c}\;\vek{e}_1 \cdot \vek{e}_2 \quad , \quad J_{c} = J\,s(s+1) \ , \end{eqnarray} where $\vek{e}_1$ and $\vek{e}_2$ are unit vectors (c-numbers). We expect that the thermal properties of this classical Heisenberg system will coincide with those of the quantum Heisenberg dimer if $s\gg 1$ except for very low temperatures. This is because the spectrum of eigenvalues of $\op{\epsilon}_{n\,z}$ is confined within $(-1,1)$ and becomes dense for $s\rightarrow\infty$ and thus coincides with the continuous range of ${e}_{n\,z}$. Similarly, if we substitute \eqref{E-3-0} in the quantum equations of motion for $\op{\vek{s}}_1$ and $\op{\vek{s}}_2$, we have \begin{eqnarray} \label{E-3-2} \dot{\op{\vek{\epsilon}}}_1 = -\Omega\; \op{\vek{\epsilon}}_1 \times \op{\vek{\epsilon}}_2 \quad , \quad \dot{\op{\vek{\epsilon}}}_2 = +\Omega\; \op{\vek{\epsilon}}_1 \times \op{\vek{\epsilon}}_2 \end{eqnarray} where \begin{eqnarray} \label{E-3-3} \Omega = \frac{J\,\sqrt{s(s+1)}}{\hbar} = \frac{J_{c}}{\hbar\,\sqrt{s(s+1)}} \ . \end{eqnarray} This suggests that we prescribe the following equations of motion for the classical unit vectors $\vek{e}_1$ and $\vek{e}_2$, \begin{eqnarray} \label{E-3-11} \dot{\vek{e}}_1 = -\Omega\; \vek{e}_1\times\vek{e}_2 \quad , \quad \dot{\vek{e}}_2 = +\Omega\; \vek{e}_1\times\vek{e}_2 \ . \end{eqnarray} We emphasize that in these equations $\Omega$ is given by \eqref{E-3-3}. It is expected that the autocorrelation function derived using \fmref{E-3-11} and the canonical ensemble average based on $H_c$ will coincide in the large $s$ limit with the normalized autocorrelation derived from \eqref{E-A-1}. This expectation is in fact confirmed as discussed below. Using the fact that the total spin is a constant of motion the classical partition function can be derived as \cite{Fis64,CLA99} \begin{eqnarray} \label{E-3-4} Z_{c} &=& {\frac{1}{2}} \int_0^2 dS \, S \, \exp\left\{-\frac{\beta\,J_{c}}{2} \, (S^2-2)\right\} \\ &=& \frac{1}{2\,J_{c}}\;\int_{-J_{c}}^{J_{c}}\,\mbox{d} E\; \exp\left\{-\beta\,E \right\} = \frac{\sinh(\beta\,J_{c})}{\beta\,J_{c}} \nonumber \ . \end{eqnarray} Note that the classical density of states turns out to be a constant in the energy interval $[-J_{c},J_{c}]$. This coincides nicely with the quantum density of states which can be obtained by counting the discrete eigenvalues per unit energy interval and normalizing the density so that its integral gives $1$. One can show, starting from \eqref{E-2-4}, that the quantity $Z/(4\,s(s+1))$ is in close numerical agreement with $Z_c$ for temperatures $k_B\,T\,>\,0.2\,J\,s(s+1)$. This serves to clearly define the classical regime for the thermal properties of the dimer. \begin{figure}[t] \begin{center} \epsfig{file=CF-T-0.2.eps,height=120mm} $\qquad$ \epsfig{file=CAF-T-0.2.eps,height=120mm} \caption{Normalized autocorrelation function for three different spins at the temperature $k_B\,T/(J\,s\,(s+1))=0.2$. The left panels displays the ferromagnetic dimer, the right panels the antiferromagnetic one. The solid lines show the quantum result, the dashed lines the classical. } \label{F-3-1} \end{center} \end{figure} For the classical autocorrelation function one finds that \cite{LBC,LuL99} \begin{eqnarray} \label{E-3-5} C_{c}(t) &=& {\frac{1}{2}} \left[ 1-\coth(\beta\,J_{c})+\frac{1}{\beta\,J_{c}} \right] + \frac{\beta\,J_{c}}{1-\exp(-2\,\beta\,J_{c})} \\ && \times\int_0^2\mbox{d} S\; S\left(1-\frac{S^2}{4}\right) \exp\left(-\frac{\beta\,J_{c}\,S^2}{2} \right) \cos\left(S\,\Omega\,t\right) \nonumber \ , \end{eqnarray} which can be integrated using error functions of complex arguments, see Ref.~\cite{MeS99}. In contrast to the quantum autocorrelation function \fmref{E-2-5}, the classical quantity is real. The reason is that $\op{\vek{s}}_1(t) \cdot \op{\vek{s}}_1(0)$ is not a hermitian operator for $t\ne 0$. If one would like to construct a hermitian operator, ${\frac{1}{2}}\left(\op{\vek{s}}_1(t) \cdot \op{\vek{s}}_1(0) +\op{\vek{s}}_1(0) \cdot \op{\vek{s}}_1(t)\right)$ would be appropriate. This coincides with the real part of our definition Eqs.~\fmref{E-2-5} and \fmref{E-A-1}. It is also interesting to note that the imaginary part of $\Mean{\op{\vek{s}}_1(t) \cdot \op{\vek{s}}_1(0)}$ does indeed vanish in the high temperature limit. In \figref{F-3-1} the dashed curves display the classical autocorrelation function obtained from \fmref{E-3-5} together with the quantum result (solid lines) for three different spin quantum numbers. In order to compare the different correlation functions all spectra have been mapped on the same energy interval $[-J_c,J_c]$. Thus the different figures show the autocorrelation functions for the same position of the mean excitation energy in the spectrum, i.e. the same \begin{eqnarray} \frac{k_B\,T}{J\,s\,(s+1)} = \frac{k_B\,T}{J_c} = 0.2 \ . \end{eqnarray} Based on our earlier remark concerning the close numerical agreement of the classical and quantum partition functions when $k_B\,T\,>\,0.2\,J_c$, we anticipate similar agreement for the autocorrelation function in this temperature range. This is confirmed on inspecting the various panels of \figref{F-3-1}, which demonstrate nicely that the quantum autocorrelation function approaches the classical result with increasing $s$. The most prominent difference between these results is that the classical autocorrelation function does not return to its initial value but approaches a unique non-zero limit, whereas the quantum autocorrelation function is recurrent with a recurrence time independent of spin and temperature. This is due to the fact that the classical system has a continuous spectrum of excitations in the angular frequency interval $[0, 2\Omega]$ whereas the quantum system possesses a discrete spectrum of excitations which are all integer multiples of the lowest one. \section{The quantum dimer} Using the matrix properties \fmref{E-2-6} the unnormalized autocorrelation function \fmref{E-2-5} can be simplified to \begin{eqnarray} \label{E-A-1} && \Mean{\op{\vek{s}}_1(t) \cdot \op{\vek{s}}_1(0)} = \frac{3}{Z}\; e^{\beta\,J\,s (s+1)} \\ &&\times \Bigg\{ \sum_{S,M} e^{-\frac{\beta\,J}{2}\,S (S+1)} |\bra{S\,M}\op{{s}}_{1z}\ket{S\,M}|^2 \nonumber \\ &&\quad + \sum_{S=1}^{2 s} \sum_{M=-S+1}^ {M=S-1} |\bra{S\,M}\op{{s}}_{1z}\ket{S-1\,M}|^2 \nonumber \\ &&\qquad\times \Big( \cos\left[\frac{t\,J}{\hbar} S\right] \left[ e^{-\frac{\beta\,J}{2} S (S+1)} +e^{-\frac{\beta\,J}{2}\,S (S-1)} \right] \nonumber \\ &&\qquad\quad + i \sin\left[\frac{t\,J}{\hbar} S\right] \left[ e^{-\frac{\beta\,J}{2} S (S+1)} -e^{-\frac{\beta\,J}{2}\,S (S-1)} \right] \Big) \Bigg\} \nonumber \ . \end{eqnarray} Taking as an example the case $s=5/2$ yields \begin{eqnarray} \label{E-A-2} && C(t) = \frac{\Mean{\op{\vek{s}}_1(t) \cdot \op{\vek{s}}_1(0)}} {\Mean{\op{\vek{s}}_1(0) \cdot \op{\vek{s}}_1(0)}} = \\&& \Bigg[ 330 + 180\,{e^{{{5\,J\,\beta}}}} + 84\,{e^{{{9\,J\,\beta}}}} + 30\,{e^{{{12\,J\,\beta}}}} + 6\,{e^{{{14\,J\,\beta}}}} \nonumber \\ && + 35\,{e^{{{J}\,\left( {\frac{i\,t}{\hbar}} + {{14}\beta}\right) }}} + 35\,{e^{{{J}\,\left( {-\frac{i\,t}{\hbar}} + {{15}\beta} \right) }}} + 64\,{e^{{{2\,J}\, \left( \frac{i\,t}{\hbar} + 6 \beta \right) }}} + 64\,{e^{{{2\,J}\,\left( {-\frac{i\,t}{\hbar}} + {{7}\beta} \right) }}} \nonumber \\ && + 81\,{e^{{{3\,J}\,\left( \frac{i\,t}{\hbar} + 3 \beta \right) }}} + 81\,{e^{{{3\,J}\, \left( -\frac{i\,t}{\hbar} + 4 \beta \right) }}} + 80\,{e^{{{J}\,\left( {\frac{4\,i\,t}{\hbar}} + {{5}\beta}\right) }}} + 80\,{e^{{{J}\,\left( -\frac{4\,i\,t}{\hbar} + 9 \beta \right) }}} \nonumber \\ && + 55 {e^{{\frac{{{5\,i}}\,J\, t}{\hbar}}}} + 55\,{e^{{{5\,J}\, \left( -\frac{i\,t}{\hbar} + \beta \right) }}} \Bigg] \nonumber\\&& \Bigg/ \Bigg[ {35\,\left( 11 + 9\,{e^{{{5\,J\,\beta}}}} + 7\,{e^{{{9\,J\,\beta}}}} + 5\,{e^{{{12\,J\,\beta}}}} + 3\,{e^{{{14\,J\,\beta}}}} + {e^{{{15\,J\,\beta}}}} \right) } \Bigg] \nonumber \ . \end{eqnarray} Autocorrelation functions for other spin quantum numbers can be evaluated using a Mathematica$^{\circledR}$ 3.0 script, that the reader is encouraged to download from our web site \cite{MeS99}.
\section{Introduction} One of the most intriguing features of the galaxy distribution on large scales (${\lower0.9ex\hbox{ $\buildrel > \over \sim$} } 10~h ^{-1}$Mpc.) is the organisation of matter into geometrically complex structures often described as being cellular, network-like, filamentary, sheet-like, honey-comb etc (\cite{zesh82,mel90,delgh91,sathya98}). The plethora of adjectives frequently used to qualify the large scale structure of the universe underscores the difficulties inherent in trying to quantify the morphology of the galaxy distribution. This has as much to do with with the need for robust statistical measures of the geometry and topology of large scale structure as it does with the virtual absence, until recently, of redshift surveys covering what may be regarded as a truly representative sample of the universe. This last gap in our knowledge has been partially filled by the Las Campanas Redshift Survey (LCRS), which for the first time appears to contain coherent structures whose size is significantly smaller than the survey size. The LCRS may therefore provide us with a statistically unbiased picture of the clustering pattern in the universe on very large scales. Evidence for large scale connectivity and the presence of filaments, pancakes and voids also comes from semi-analytic modelling of large scale structure and from direct N-body simulations. For instance N-body simulations describing clustering from Gaussian initial conditions reveal that as a distribution evolves into the non-linear regime it begins to develop non-Gaussian features characterised by the appearance, on the one hand of large empty regions -- voids, and on the other, of the concentration of matter into filaments and pancakes (\cite{sc95}). The visually prominent appearance of filaments in galaxy catalogues (CfA, SSRS etc.) and in N-body simulations led to the endeavour to characterise these morphological features mathematically in order that ``real filaments'' might be distinguished from ``apparent filaments'', the latter arising because the eye is prone to picking out structure when in fact there is none. In the process it has been shown that filamentarity is a very real feature of cosmological gravitational instability, and that the degree of filamentarity increases as gravitational clustering advances (\cite{sss96,sss98b}). Traditionally, the clustering of galaxies in groups and clusters has been probed with some success by the two point correlation function, $\xi(r)$, which provides an estimate of the probability (in excess of random), of finding a galaxy at a distance $r$ from another galaxy. A robust indicator of clustering, the two point correlation function nevertheless reveals very little about the morphology of the galaxy distribution, {{\em i.e.,~~}} the concentration of matter in sheets and/or filaments and the geometrical and topological properties of the supercluster-void distribution as a whole. The reason is simple, since $\xi(r)$ is the Fourier transform of the power spectrum $P(k) = \langle\vert\delta_k\vert^2\rangle$, it does not describe the build up of phase correlations which lead to the emergence of non-Gaussian features in gravitating systems arising as a result of clustering from Gaussian random initial conditions. Complete information about clustering is formally contained in the infinite hierarchy of correlation functions $\xi_N$, $N = 2,3,..\infty$. Although attempts have been made to calculate $\xi_3$ and $\xi_4$, the presence of a finite number of galaxies in any sample makes the measurement of $\xi_N$ on large scales difficult for large $N$. Lower order correlation functions must, therefore, be complemented by other statistical measures sensitive to geometry and topology, if we wish to probe the connectedness of large scale structure (and the associated non-Gaussianity) at a more fundamental level. The first statistical measures to probe the geometry of large scale structure were percolation analysis and the genus curve (\cite{zel82,sh83,gmd86}). When applied to N-body simulations and galaxy catalogues both methods pick out departures from Gaussianity and are helpful in discerning the presence of ``network-like'' and ``sponge-like'' features in systems undergoing gravitational clustering. The minimal spanning tree has also proved useful in quantifying geometrical features of large scale structure (\cite{bbs85}). More recently, Minkowski functionals have added to our understanding of morphology (\cite{mbw94,sb97}). As demonstrated in Sahni, Sathyaprakash \& Shandarin (1998), Shapefinders, a new shape diagnostic constructed out of ratios of Minkowski functionals, provides valuable information about the ``shape'' of a clustering pattern and can be used with considerable advantage (in conjunction with percolation analysis) to study issues relating to the morphology of large scale structure, such as the abundance of sheet-like, filament-like and ribbon-like objects. In the present paper we shall apply Shapefinders to assess the degree of filamentarity in the Las Camapanas Redshift Survey. The rest of the paper is organized as follows: Sec. 2 defines the statistics we use, Sec. 3 describes the data and the reference cataloges we deal with, finally in Sec. 4 we present our results and conclusions. \section{Minkowski Functionals and the Shapefinder Statistic} In three dimensions, the four Minkowski functionals characterising the morphology of a compact manifold are (\cite{mbw94}): (i) its volume, $V$, (ii) surface area, $S$, (iii) the integrated mean curvature, $C$, and (iv) the integrated Gaussian curvature, $G$, (equivalently the Euler characteristic or genus). Percolation analysis can be accomodated within this scheme if one studies the behavior of the volume of the largest cluster as a function of the filling factor for the full cluster distribution (for mass distributions clusters being defined as connected objects lying above a given density threshold). Likewise, the genus curve can be obtained if one plots $G$ as a function of the filling factor. The Shapefinder trio is constructed out of ratios of Minkowski functionals: $L = \frac{C}{4\pi}$, $W = \frac{S}{C}$, $T = \frac{3V}{S}$. $L,W$ \& $T$ have dimensions of length and provide an estimate of physical dimensions of an object such as its {\em length} $L$, {\em width} $W$ and {\em thickness} $T$ (\cite{sss98a}). Thus $L \simeq W \simeq T$ characterises a quasispherical cluster, $L \gg W \simeq T$ a filament, $L\simeq W \gg T$ a pancake and $L \gg W \gg T$ a ribbon. Based on $L,W,T$ a pair of dimensionless Shapefinders $\lbrace P, F\rbrace$ can be constructed which provide us with estimates of the planarity $P = \frac{W - T}{ W + T}$ and filamentarity $F = \frac{L - W}{L + W}$ of an object, $0 \leq P,F \leq 1$. ($F = P = 0$ for a sphere, $F \simeq 1, P \simeq 0$ for a filament, $F \simeq 0, P \simeq 1$ for a pancake and $F \sim P \sim 1$ for a ribbon.) In two dimensions the partial Minkowski functionals characterizing the morphology of a connected region are the area $S$, perimeter $L$ and the number of holes in the region or genus $G$. The ratio $T=S/L$ characterizes the thickness of the cluster and $L$ its extension. The ratio $L/G$ also has dimensions of length and becomes a meaningful parameter after the onset of percolation, characterizing the scale of large scale structure ($L$ itself obviously grows without limit after percolation with the growth of the survey size). It appears that the Shapefinder statistic is a much better diagnostic of shape than moment-based methods, particularly when applied to topologically complex bodies such as isodensity surfaces occuring at moderate density thresholds in N-body simulations and in galaxy catalogues (\cite{sss98b}). In two dimensions, the Shapefinder statistic simplifies to a {\em single number} \begin{equation} {\cal F} = \frac{L^2 - 4\pi S}{L^2 + 4\pi S} \label{eq:1} \end{equation} where $L$ is the perimeter and $S$ the area of a closed two dimensional contour. (This could be an isodensity contour in a two dimensional galaxy distribution, an isotherm of a hot/cold spot in a map of the Cosmic Microwave Background (\cite{nfsh99}), etc.) By definition $0 \leq {\cal F} \leq 1$; ${\cal F} \simeq 1$ for an ideal filament (having a finite length and zero width), and ${\cal F} \simeq 0$ for a circular disc. The Las Campanas Redshift Survey contains redshifts of about 25,000 galaxies making it the largest magnitude limited three-dimensional catalogue of galaxies to date. However, the six slices into which LCRS is divided are all quasi-two-dimensional, which makes the two-dimensional shapefinder ${\cal F}$ appropriate for its study. Before applying the Shapefinder statistic to discern filamentary features in the LCRS, we shall first demonstrate its effectiveness by determining ${\cal F}$ for certain eikonal shapes in two dimensions. The first object whose shape we study is an ellipse of semi-major axis $a$, semi-minor axis $b$ and eccentricity $\epsilon = \sqrt{1 - b^2/a^2}$. The area of such an ellipse is \begin{equation} S = \pi a b \label{eq:2} \end{equation} and its perimeter is given by \begin{equation} L = 4aE(\epsilon) \simeq \pi (a + b)\frac{64 - 3\lambda^4}{64 - 16\lambda^2} \label{eq:3} \end{equation} where $E(\epsilon) = \int_0^{\pi/2}\sqrt{1 - \epsilon^2\sin^2{\theta}} d\theta$ is the complete elliptic integral of the second kind, $\lambda = (a-b)/(a+b)$. The accuracy of (\ref{eq:3}) is better than $0.2\%$. In table 1 we show the value of ${\cal F}$ as the ellipse is gradually deformed from an initially circular shape to a highly filamentary final shape. As pointed out in Sahni, Sathyaprakash \& Shandarin (1998) other Shapefinders can also be constructed out of the Minkowski functionals $S$ and $L$, a good example being provided by the transformations ${\cal F}'$ = ${\cal F}^p$ and ${\cal F}'' = \sin{(\pi{\cal F}/2)}$ which define new Shapefinders from the canonical form (\ref{eq:1}). Transformations such as these can be made to define statistics that are in conformity with the visual impression of morphology of archetypal structures. In table 1, we show results for ${\cal F}$, ${\cal F}'$ (with $p=1/2$) and ${\cal F}''$. >From the results presented in this Table, we see that the Shapefinder family ${\cal F}$, ${\cal F}'$ and ${\cal F}''$ acquires continuous values between 0 and 1 as a circle is deformed into a filament, and that filamentarity is more accentuated in ${\cal F}'$ and ${\cal F}''$ than in ${\cal F}$. It is interesting to contrast the behaviour of ${\cal F}$ with that of the eccentricity, $\epsilon$, another parameter characterising shape but defined {\em only} for an ellipse. While the value of ${\cal F}$ grows in proportion to the deformation of the ellipse (and is therefore reflective of the latter's visual shape), the change in $\epsilon$ is sudden, with the result that $\epsilon$ rapidly approaches unity even for relatively small deformations of the ellipse. Although the eccentricity is extremely good for measuring the small deviations from a circle it is clearly not the kind of behaviour one wishes to see in a ``well behaved'' shape-statistic which should also be sensitive to large deviations from a circle. Next, we examine a {\em topologically non-trivial} object -- the region between two concentric circles with radii $R_1$ and $R_2$, $R_2 \leq R_1$ (circular disc-with-a-hole). In this case: \begin{equation} S = \pi(R_1^2 - R_2^2), ~~~~ L = 2\pi (R_1 + R_2), \end{equation} and the expression for ${\cal F}$ turns out to be very simple \begin{equation} {\cal F} = \frac{R_2}{R_1}. \end{equation} As the radius of the hole $R_2$ shrinks to zero, the object becomes a circular disc with ${\cal F} \simeq 0$, in the other extreme case when $R_1 \sim R_2$ the disc-with-a-hole reduces to a circular filament having ${\cal F} \simeq 1$ (rather like the mythological serpent eating its tail). One should note that the presence of circular symmetry would lead a moment-based statistic to wrongly declare such an object as being circular or homogeneous (\cite{sss98b}). Results of increasing the hole size by increasing $R_2$ are shown in Table 2 for ${\cal F}$, ${\cal F}'$ and ${\cal F}''$. They are in broad agreement with those obtained earlier for the ellipse. We therefore find that the two dimensional Shapefinder statistic gives sensible results when applied to both simple and topologically complicated eikonal shapes. This conclusion is supported by results obtained for three dimensional Shapefinders by Sahni et al. (1998) and Sathyaprakash, Sahni \& Shandarin (1998). In this paper we shall apply ${\cal F}$ to study filamentarity in the Las Camapanas Redshift Survey (the results of applying ${\cal F}'$ and ${\cal F}''$ are qualitatively similar and will not be discussed separately). \section{Analysis.} \subsection{The Las Camapanas Redshift Survey.} The Las Camapanas Redshift Survey (LCRS) contains approximately 25,000 galaxies with known redshifts. The survey region is divided into six slices, three each in the Northern (N$_1$, N$_2$, N$_3$) and Southern (S$_1$, S$_2$, S$_3$) Galactic hemispheres. The slices are strips of the sky $1.5^\circ$ thick and $80^\circ$ wide which are separated by $3^\circ$ and probe a distance of up to $600~h^{-1}$Mpc. (h is the value of the Hubble parameter in units of 100 km~sec$^{-1}$Mpc$^{-1}$). The centers of the three northern slices are at declinations $-3^\circ$ (N$_1$), $-6^\circ$ (N$_2$) and $-12^\circ$ (N$_3$), whereas the three southern slices are centered at $-39^\circ$ (S$_1$), $-42^\circ$ (S$_2$) and $-45^\circ$ (S$_3$). The survey is complete to limiting magnitude $m = 17.75$ (for more details of the LCRS survey see \cite{shectman96}). In order to minimise selection and projection effects we apply the Shapefinder statistic to a volume limited subsample of the LCRS derived from the dense central region $200 \leq R \leq 400~h^{-1}$ Mpc. (The LCRS selection function peaks at $R \simeq 200 h^{-1}$Mpc. and the region $200 \leq R \leq 400~h^{-1}$ Mpc corresponds to the densest part of the survey.) In order to ascertain the statistical significance of our results we compare them with a statistically homogeneous two-dimensional Poisson distribution with identical selection function and geometry as the survey and corrected for projection effects (for more details, see \cite{sy98}). Each slice is embedded in a $560 \times 260$ grid with resolution $1~h^{-1} {\rm Mpc}$. This enables us to define galaxy locations on a two dimensional lattice. Lattice cells containing a galaxy are said to be occupied or {\em filled}, and lattice cells with no galaxies are referred to as {\em empty}. The cells which lie outside the boundaries of the survey are eliminated from the lattice. Next, we proceed to ``grow'' structure (or coarse-grain) by means of the following iterative procedure: At each step filled cells are made to grow isotropically ({\em i.e.,~~} in all directions) by a single unit of the mesh size. Thus any cell neighboring a filled cell (one containing a galaxy) in any of eight possible directions is also designated ``filled''. This algorithm leads to the growth of filled cells at each successive iteration. The procedure of sequentially coarse-graining a LCRS slice is shown in Figure 1 for the N1 slice. Incidentally, it also demonstrates what happens when the size of the dots in the wedge diagram plots are arbitrarily chosen. Selecting the size of the dots or the linking length one can emphasize ``desired'' features of the distribution such as filamentarity, connectivity, the size of the greatest supercluster or something else. So far no reason has been suggested in the literature for a particular choice of the linking length (except one corresponding to percolation transition) and therefore we use the linking length or the actual size of the dots for the purpose of parameterization. The associated {\em filling factor} (FF), defined as the fraction of filled cells in the total slice, increases from a small initial value to almost unity when all the cells in the LCRS slice have inter-connected into one big all-pervading cluster. At any given value of FF ({\em i.e.,} at any given iterative step) we determine clusters using a friends-of-friends algorithm: any filled cell sharing a common side with another filled cell is its friend. All such ``friends'' define a cluster. The shape of individual clusters defined in this manner can be analysed using the Shapefinder statistic at different values of the filling factor. Large scale connectedness of galaxies in LCRS is also revealed by studying the Largest Cluster Statistics (LCS) characterizing the percolation transition as a function of the filling factor (LCS is defined as the ratio of the area of the largest cluster to the total area of all clusters, see \cite{sy98}). It is also necessary to point out that measurements of structure in galaxy catalogues can in principle depend upon the observational strategy, particularly on the proximity of fibres mounted on a CCD. Ones intrinsic inability to pack fibres sufficiently close together in order to measure redshifts of all galaxies in a dense field might lead to the undersampling of galaxies in clusters. However since filaments are much lower density objects than clusters their detection will not be as sensitive to the spacing between fibres with the result that most of the results presented in the present paper are expected to be sufficiently robust. (For the rare case of a filament aligned along the line of sight and therefore viewed ``head on'' this effect could lead to undercounting of galaxies along the core of the filament and therefore weaken the signal for filamentarity. Our present estimates of filamentarity in LCRS should therefore be treated as a lower bound to the extent of filamentarity in the ``real universe''.) \subsection{Shapefinders on a grid} The Shapefinders (\ref{eq:1}) were defined for continuous contours. For contours on a grid this definition must be modified. It is easy to see that the lattice version of (\ref{eq:1}) is \begin{equation} {\cal F}_1 = \frac{L^2 - 16 S}{L^2 + 16 S} \label{eq:4} \end{equation} (Note that the factor $4\pi$ in (\ref{eq:1}) has been replaced by 16 in (\ref{eq:4}).) Consider a rectangle defined on a grid with sides $n\times l$ and $m\times l$, respectively, where $l$ is the inter-grid spacing. It is easy to see that in this case ${\cal F}_1$ reduces to \begin{equation} {\cal F}_1 = \frac{(n - m)^2}{(n + m)^2 + 4nm} \label{eq:5} \end{equation} as a result ${\cal F}_1 = 0$ for $n = m$ (a square), and ${\cal F}_1 \rightarrow 1$ for $n \gg m$ (a filament). It is interesting that one can define a second Shapefinder statistic which is also suitable for determining shapes of contour lines defined on a grid. Again, if $l$ is the inter-grid spacing we have \begin{equation} {\cal F}_2 = \frac{L^2 - 16S}{(L - 4l)^2} \label{eq:6} \end{equation} so that, for the rectangle $(n,m)$ \begin{equation} {\cal F}_2 = \frac{(n - m)^2}{(n + m - 2)^2}. \label{eq:7} \end{equation} We see that once more ${\cal F}_2 = 0$ for $n = m$ (a square), and ${\cal F}_2 \rightarrow 1$ for $n \gg m$ (a filament). A specific feature of ${\cal F}_2$ is that a rectangular contour of unit width $(n,1)$ is always delared to be a filament regardless of its length $n$, since, substituting $m = 1$ in (\ref{eq:7}) we get ${\cal F}_2 = (n-1)^2/(n-1)^2 = 1$. One might question the need for introducing two separate statistics to study shapes on a lattice. The reason for this is related to the following question: Consider an ``ideal'' one-dimensional filament having a finite length but zero breadth. How would one go about representing this object on a grid? Depending upon how one answers this question one arrives at either ${\cal F}_1$ or ${\cal F}_2$. One might argue that the discrete analogue of an ideal filament would be a rectangular object with width $\sim l$, ({\em i.e.,~~} width = mesh size) in which case ${\cal F}_2$ gives the correct continuum limit: ${\cal F}_2 = {\cal F} = 1$. On the other hand one might equally argue that a filament having width $< l$ is impossible to correctly define on a grid for which the Nyquist wavelength determines an effective small scale ``resolution cutoff''. Therefore, if one is confined to a grid one must necessarily differentiate between objects having identical widths ($\sim l)$ but varying in length, in this case the correct statistic to use is ${\cal F}_1$. In this paper, we report on work carried out using both ${\cal F}_1$ and ${\cal F}_2$. As we shall show, ${\cal F}_1$ and ${\cal F}_2$ contain complementary information about filamentarity, as a result both prove to be very useful shape-diagnostics for large scale structure. \cite{nfsh99} used yet another approach to measuring the Minkowski functionals on a grid. Studying the 4-year COBE maps they assumed a smooth underlying field and, therefore, used a linear interpolation scheme for measuring the perimeters and areas of the clusters. \section{Results and Conclusions.} A general statistical quantity that can be constructed out of the Shapefinder ${\cal F}$ is the functional $n({\cal F}|A,FF)$ which describes the number density of disjoint regions of area $A$ having given values of the filamentarity ${\cal F}$ and located above a given density threshold characterised by the filling factor $FF$. However a good simple indicator of filamentarity is provided by the averaged quantity \begin{equation} \langle {\cal F}^{(d)}(FF)\rangle = \frac{\int\int n({\cal F}|A,FF){\cal F} A^d~dA~ d{\cal F}} { \int\int n({\cal F}|A,FF) A^d~dA~ d{\cal F}} \label{eq:7a} \end{equation} describing particular moments of filamentarity in the distribution. (Each cluster of area $A$ is identified using a nearest neighbors algorithm at a given value of the coarse-graining.) Clearly the extent of filamentarity in a survey is strongly influenced by the morphology of its largest and most massive members: a supercluster contributes more to the overall texture of large scale structure than an individual galaxy. We therefore examine moments with $d > 1$ thereby giving greater weight to larger objects. Since our results are qualitatively similar for $d = 2,3,4...$ we show only results for $d = 2$ and refer to the statistic $\langle {\cal F}^{(2)}(FF)\rangle$ simply as $F$. For the discrete sample under consideration (\ref{eq:7a}) reduces to \begin{equation} F \equiv F_{1,2} = \frac{\sum_i A_i^2{\cal F}_{1,2}^i}{\sum_i A_i^2}, ~~~~~~ 0 \leq F_{1,2} \leq 1 \label{eq:8} \end{equation} where the sum is evaluated at a given value of the filling factor. $A_i$ is the area of a given cluster and ${\cal F}_{1,2}^i$ is the filamentarity of the $i^{th}$ cluster obtained by using (\ref{eq:4}) for ${\cal F}_1$ or (\ref{eq:6}) for ${\cal F}_2$. In Figures 2 and 3, we show the extent of filamentarity as a function of the filling factor for three northern and three southern slices of LCRS (the volume limited slice contain from 800 to 1,600 galaxies). To assess the statistical significance of our results we compare $F$ with the corresponding quantity derived from four independent Poisson randomizations of the given LCRS slice. From Figures 2 and 3 it is clear that for moderate values of the filling factor the extent of filamentarity in {\em all} six slices of LCRS is {\em significantly greater} than in the Poisson sample. This is true for $F$ constructed from both ${\cal F}_1$ and ${\cal F}_2$. Comparing Fig. 2 (obtained using ${\cal F}_1$) with Fig. 3 (obtained using ${\cal F}_2$) we find that the difference between LCRS and the Poisson sample shows up at much lower filling factors for ${\cal F}_2$. The reason for this is simple, as we demonstrated earlier, ${\cal F}_2$ is designed to be more successful in picking out smaller filaments than ${\cal F}_1$. At low FF, small filaments appear to be much more abundant in LCRS than they are in the Poisson sample, as a result ${\cal F}_2$ easily discerns filamentarity in LCRS at small FF. Together ${\cal F}_1$ and ${\cal F}_2$ probe the extent of filamentarity over a wide range of FF. Complementary information emphasising the ``connectivity'' of a distribution is revealed by percolation analysis (\cite{sy98}). At small FF each slice contains a large number of distinct clusters whose shape must be determined individually. As the FF increases, neighboring clusters begin to merge leading to a decrease in the total number of clusters in the sample and to the emergence of a single dominant ``supercluster''. With increased FF we find that the largest cluster shows a very rapid increase in size as a result of which it soon spans the entire survey region. This corresponds to the onset of percolation. We determine ${\cal F}$ for individual clusters and the area of the largest cluster at each level of coarse graining. As $FF \rightarrow 1$ the percolating supercluster fills the entire slice, the value of ${\cal F}$ then drops to a small value which describes the shape of the survey region. (If the slice were an exact square the value ${\cal F}$ would approach 0 as $FF \rightarrow 1$.) In Fig. 4 we plot the Largest Cluster Statistic (LCS is the fractional size of the largest cluster relative to all clusters at a given value of coarse graining) against FF (the fraction of filled cells in the total area of the slice). Comparing these results with those of the Poisson distribution we clearly see that the growth in the largest cluster is more rapid in LCRS than it is in the randomized catalogue (constructed with the same selection function and number of galaxies and therefore having identical geometrical properties to LCRS) confirming the earlier results of Shandarin and Yess (1998). To summarise we have shown that clustering of galaxies in LCRS is significantly non-Gaussian: it is dominated by filaments and its global geometry is network-like (in the parlance of \cite{sy98}). Our analysis also indicates that Shapefinders recognize non-Gaussian features in LCRS at lower filling factors than the percolation curve and that Shapefinders and percolation analysis supplement each other in discerning the geometrical properties of the galaxy distribution on large scales. Finally we note that at $FF \sim 1$ the extent of filamentarity in LCRS declines (from a value close to unity) and becomes smaller than in the Poisson catalogue. The decrease in $F$ in both LCRS and the Poisson catalogue is simply due to the fact that all clusters in the sample have now merged into a single ``super-duper'' cluster which resembles a quasi-homogeneous object with several holes (empty regions). Such an object when probed using the Shapefinder statistic appears very filamentary both in LCRS and the Poisson sample. However as $FF \rightarrow 1$ the holes gradually get filled and $F$ drops to a small value which describes the shape of the survey region. We do not attach much significance to results shown in Fig.2 and Fig.3 at $FF > 0.6$ because at such large values of the filling factor the percolating supercluster spans the entire sample region. As a result at large $FF$ some geometrical properties of the supercluster begin to be progressively determined by the size and geometry of the sample volume and cease to be reflective of the (formally infinite) galaxy distribution from which the finite sample is drawn. The largest cluster in two dimensions can be characterized by two parameters: the thickness of the cluster wall separating nearby voids $T=S/L$ and the length of the cluster boundary per void $L/|G|$. The cluster perimeter $L$ itself is not a good measure because for the percolating cluster it diverges as the area of the survey grows without bound. On the other hand the parameter $L/|G|$ defined for the percolating cluster proves to be remarkably stable over a wide range of linking lengths. Fig. 5 shows the length of the cluster boundary per void ($L/|G|$) as a function of wall thickness ($S/L$) for both the LCRS and reference catalogues. The stability of the statistic $L/|G|$ is reflected by the fact that while the thickness of the supercluster grows by a factor of three from $\approx 5~h^{-1}$Mpc to $\approx 16~h^{-1}$Mpc the mean value of $L/|G|$ decreases only marginally: from $\approx 110~h^{-1}$Mpc to $\approx 70~h^{-1}$ Mpc in the LCRS catalogue. The structure in the reference Poisson catalogues is of considerably smaller size than in the LCRS. This is reflective of the existence of large scale structure in the galaxy distribution marked by a percolating network of filaments separating large voids. Although Fig. 5 gives the scale of the structure, we wish to stress that it is not the diameter of the empty cells seen in Fig. 1. The parameter $L/|G|$ has a meaning of ``length of the supercluster contour line per void''. In our case the mean number of voids is large: about 30 in the LCRS and about 130 in the randomized catalogues (at wall thickness of $\approx 10~h^{-1}$Mpc). This allows us to say that the sample is sufficiently large and the average diameter of the voids (at thickness $\approx 10~h^{-1}$Mpc) is about $80/\pi \approx 25~h^{-1}$Mpc. (Large values of $|G|$ for a given cluster indicate a large number of holes and consequently a greater porosity of the cluster. This applies particularly to the percolating supercluster for which $G$ can be very large and negative.) It is interesting to note that an analysis of LCRS using the genus curve has shown that at large values of the smoothing radius, LCRS reveals statistical properties which are consistent with those of a Gaussian random field (\cite{colley97}). Our results on the other hand (and those of \cite{sy98}) show that the {\em unsmoothed} LCRS catalogue displays strongly non-Gaussian features. In this context, it is worthwhile to reiterate that N-body gravitating systems clustering from Gaussian initial conditions rapidly develop non-Gaussian features reflected both in a network-like structure and by a growth of filamentarity in the morphology of clusters and superclusters (\cite{sss96,sss97,sss98b}). The effects of smoothing such a distribution on large scales would be to add greater weight to the clustering of long range modes still in the linear regime, and hence statistically distributed in the manner of a Gaussian random field (\cite{ds92,sc95}). When viewed in this context the results of this paper (and those of \cite{sy98}), together with the results of Colley (1997), appear to provide strong support for the scenario in which the large scale structure of the universe evolved via gravitational clustering from Gaussian initial conditions predicted (for instance) by Inflationary Models of the very early universe. Acknowledgements. S. Shandarin acknowledges the support of NSF-EPSCoR grant, GRF grant at the University of Kansas and from TAC Copenhagen. We wish to thank an anonymous referee for comments that led to an improvement in the manuscript.
\section*{ABSTRACT} We describe {{\it Beppo}SAX}\ \ignorespaces observations of the black hole candidates LMC X--1 and LMC X--3 performed in Oct. 1997. Both sources can be modelled by a multicolor accretion disk spectrum, with temperature $\sim 1$ keV. However, there is some evidence that a thin emitting component coexists with the thick disk at these temperatures. In the direction of LMC X--1, we detected a significant emission above 10 keV, which we suspect originates from the nearby source PSR 0540-69. For LMC X--1, we estimate an absorbing column density of $\simeq 6\times 10^{21}$ cm$^{-2}$, which is almost ten times larger than that found for LMC X--3. In both sources, we find no indication of emission or absorption features whatsoever. \section{INTRODUCTION} LMC X--3 and LMC X--1 are two luminous persistent X--ray binaries. LMC X--3 has an orbital period of 1.7 days and is one of the most secure black hole candidates ($M_{\rm X} \simeq 10M_{\odot}$). For LMC X--1 there is still some uncertainty regarding the optical counterpart. The most probable optical candidate has a period of 4.23 d, which implies a companion (hole) mass of $\simeq 5 M_{\odot}$ (Cowley et al. 1995). The two objects have soft X--ray spectra reminiscent of Cyg X--1 in the high/soft state. A high energy tail extending above 10 keV is detected in both systems (Ebisawa, Mitsuda \& Inoue 1989; Treves et al. 1990). We observed the two sources with {{\it Beppo}SAX}\ \ignorespaces in Oct. 1997, which allowed us to construct simultaneous spectra in a broad energy band (0.1--100 keV). All the four narrow field instruments (LECS 0.1--4 keV, MECS 1.8--10 keV, HPGSPC 7--70 keV, PDS 12--150 keV) performed nominally; the MECS functioning with two units. Here we are presenting some preliminary results on the X--ray spectral distribution. HPGSPC data analysis is not discussed in this paper. Our observations correspond to a low intensity state of LMC X--3, while LMC X--1 appears stable on month/year time scale, as apparent from the the XTE--ASM light curves. \begin{table*} \begin{tabular}{cccccccc} \multicolumn{8}{l}{{TABLE 1: Observation Log}}\\ &&&&&&&\\ \hline \hline &&&&&&&\\ \multicolumn{1}{c}{Source Name} &\multicolumn{3}{c}{Exposure Time (ks)}& &\multicolumn{3}{c}{Count rate (cts/s)}\\ & \multicolumn{1}{c} {LECS} & \multicolumn{1}{c} {MECS} & \multicolumn{1}{c} {PDS} & & \multicolumn{1}{c} {LECS} & \multicolumn{1}{c} {MECS} & \multicolumn{1}{c} {PDS}\\ &&&&&&&\\ \hline &&&&&&&\\ {LMC X--1} & {14.4} & {38.2} & {43.3}& & {3.26$\pm{0.02}$} & {2.67$\pm{0.01}$} & {0.21$\pm{0.05}$}\\ &&&&&&&\\ {LMC X--3} & {17.4} & {39.5} & {38.0}& & {3.37$\pm{0.01}$} & {2.19$\pm{0.01}$} & {0.03$\pm{0.05}$}\\ &&&&&&&\\ \hline &&&&&&&\\ &&&&&&&\\ \multicolumn{8}{l}{Note: MECS counts are for one unit. PDS counts are for four units.} \end{tabular} \end{table*} \begin{table*} \begin{tabular}{cccccccc} \multicolumn{8}{l}{{TABLE 2: Spectral Fits: Disk Spectrum$^{a}$ + Free--Free}}\\ &&&&&&&\\ \hline \hline &&&&&&&\\ \multicolumn{1}{c}{Source Name} &\multicolumn{1}{c}{$N_{\rm H}$} &\multicolumn{1}{c}{$kT_{\rm in}$} &\multicolumn{1}{c}{N$_{\rm diskbb}^{b}$} &\multicolumn{1}{c}{$kT_{\rm brems}$} &\multicolumn{1}{c}{N$_{\rm bremss}^{c}$} &\multicolumn{1}{c}{F$^{d}_{[2-10]{\rm keV}}$} &\multicolumn{1}{c}{$\chi^2/dof$}\\ &{($10^{22}$ cm$^{-2}$)} & {(keV)} & &{(keV)} & & & \\ &&&&&&&\\ \hline &&&&&&&\\ {LMC X-1} & {$0.61^{+0.03}_{-0.03}$} & {$0.85^{+0.01}_{-0.02}$} & {$45.1^{+7.0}_{-8.5}$} & {$2.39^{+0.25}_{-0.19}$} & {$0.15^{+0.04}_{-0.04}$}& {$3.25$} & {$190/214$}\\ &&&&&&&\\ {LMC X-3} & {$0.056^{+0.005}_{-0.004}$} & {$1.05^{+0.01}_{-0.02}$} & {$19.4^{+1.7}_{-1.2}$} & {$2.01^{+0.42}_{-0.90}$} & {$0.036^{+0.008}_{-0.009}$}& {$2.71$} & {$217/214$}\\ &&&&&&&\\ \hline &&&&&&&\\ &&&&&&&\\ \multicolumn{8}{l}{Note: errors are 90\% confidence level for one parameter.}\\ &&&&&&&\\ \multicolumn{8}{l}{ $^a$~ ${dN \over dE}= {8\pi \over 3}{R_{\rm in}\over D^2} \cos{\theta} \, \int_{T_{\rm out}}^{T_{\rm in}} (T/T_{\rm in})^{-11/3} B(E,T) dT/T_{\rm in}$}\\ &&&&&&&\\ \multicolumn{8}{l}{ $^b$~ $[(R_{\rm in}/{\rm km})/ (D/10{\rm kpc})]^2 \cos{\theta}$ } \\ &&&&&&&\\ \multicolumn{8}{l}{ $^c$~ ${3.02\times 10^{-15} \over 4\pi D^2} \int n_e n_I\, dV$}\\ &&&&&&&\\ \multicolumn{8}{l}{ $^d$~Flux unit: (10$^{-10}$ erg/cm$^2$/s)} \end{tabular} \end{table*} \section{{{\it Beppo}SAX}\ \ignorespaces SPECTRA} Data reduction followed the standard procedure. Total exposure times and count rates are reported in Table 1. Results of the spectral analysis described below are reported in Table 2. \begin{itemize} \item[LMC X--1:] In the field of view of the MECS and PDS there is the source PSR 0540-69. In the $[2-10]$ keV band, the pulsar is fainter than LMC X--1 by a factor $\sim$ 10. The pulsar spectrum is rather hard, and a power--law fit of the MECS data yields an energy index $\Gamma=2.01\pm{0.07}$. In principle, the pulsar emission could account of (part of) the signal detected in the PDS. To test this hypothesis, we first fitted separately the LECS+MECS data of LMC X--1 with an absorbed disk blackbody, obtaining a totally unacceptable fit ($\chi^2/dof > 3$). The inclusion of an optically thin free--free component greatly improves the fit ($\chi^2/dof=0.89$; see Table 2 and Fig. 1). We did not find any narrow emission or absorption feature. The column density is quite large, $N_{\rm H} \simeq 6.5 \times 10^{21}$ cm$^{-2}$. Then, we fitted the PDS data alone. They are well represented by a hard power--law, which is completely consistent, in terms of flux and spectral index, with the extrapolation in the PDS energy range of the MECS spectrum of the pulsar. We therefore suspect that the pulsar is responsible of the high energy emission detected in the PDS. \begin{figure} \centerline{\psfig{figure=treves-fig1.ps,height=9.0cm,angle=-90}} \caption{ LMC X--1: count spectrum and contribution to $\chi^2$ when data are fitted with a {\tt wabs[diskbb+bremss]} model.} \end{figure} \item[LMC X--3:] In this source, we did not find any positive detection in the PDS. A variable hard X--ray tail was observed in the PDS in only one of the two one--month spaced {{\it Beppo}SAX}\ \ignorespaces science verification phase observations performed in 1996, while in the other it had an upper limit comparable to ours (Siddiqui et al. 1998). As in the case of LMC X--1, a fit with an absorbed multicolor disk spectrum showed evidence of features in the 0.8--2 keV range ($\chi^2$ null probability $\lta 0.1$\%). The inclusion of a free--free component reduced the $\chi^2$ to an acceptable value (see Table 2 and Fig. 2). Again, we did not find any narrow emission or absorption feature above 2 keV. The absorbing column lies in the range reported by previous X--ray observations ($N_{\rm H}\simeq 5.7 \times 10^{20}$ cm$^{-2}$, Treves et al. 1988). \end{itemize} \begin{figure} \centerline{\psfig{figure=treves-fig2.ps,height=9.0cm,angle=-90}} \caption{LMC X--3: count spectrum and contribution to $\chi^2$ when data are fitted with a {\tt wabs[diskbb+bremss]} model.} \end{figure} \section{CONCLUSIONS} Our analysis shows that in LMC X--1 and LMC X--3 an optically thick accretion disk coexists with an optically thin, X--ray emitting gas, with comparable temperatures. This may indicate that the thermal emission from the innermost region of the accretion disk is modified by electron scattering. In LMC X--1 direction there is also indication of a component at much higher energy, responsible for the power--law emission in the PDS, which however we suspect arises from a nearby pulsar. Previous claims of the presence of a hard tail in LMC X--1 were based on non imaging instruments, where confusion with the nearby pulsar can not be excluded. LMC X--1 is found to be severely absorbed at low energies. Theoretical interpretations of the results are in progress and will be presented in a forthcoming paper. \section{REFERENCES} \vspace{-5mm} \begin{itemize} \setlength{\itemindent}{-8mm} \setlength{\itemsep}{-1mm} \item[] Cowley, A.P., Schmidtke, P.C., Anderson, A.L., and McGrath, T.K., 1995, PASP, 107, 145. \item[] Ebisawa, K., Mitsuda, K., and Inoue, H., 1989, PASJ, 41, 519. \item[] Siddiqui, H., et al., 1998, in prep. \item[] Treves, A., et al., 1988, ApJ, 325, 119. \item[] Treves, A., et al., 1990, ApJ, 364, 266. \end{itemize} \end{document}
\section{#1}} \input epsf \newcounter{fignum} \newcommand{\arabic{fignum}}{\arabic{fignum}} \newcommand{\figur}[2]{ \addtocounter{fignum}{1} \addcontentsline{lof}{figure}{\protect \numberline{\arabic{section}.\arabic{fignum}}{#2}} \hspace{-3mm}{\it fig.}\ \arabic{fignum}. \begin{figure}[t]\begin{center} \leavevmode\hbox{\epsffile{#1.eps}}\\[3mm] \parbox{10cm}{\small \bf Fig.\ \arabic{fignum} : \it #2} \end{center} \end{figure}\hspace{-1.5mm}} \newcommand{\figurplus}[3]{ \addtocounter{fignum}{1} \addcontentsline{lof}{figure}{\protect \numberline{\arabic{section}.\arabic{fignum}}{#3}} \hspace{-3mm}{\it fig.}\ \arabic{fignum}. \begin{figure}[t]\begin{center} \leavevmode\hbox{\epsfxsize=#2 \epsffile{#1.eps}}\\[5mm] \parbox{10cm}{\small \bf Fig.\ \arabic{fignum}: \it #3} \end{center} \end{figure}\hspace{-1.5mm}} \newcommand{{\it fig.}\ }{{\it fig.}\ } \begin{document} \begin{titlepage} \rightline{UCB-PTH-99/11, LBNL-43079, NSF-ITP-99-22} \rightline{hep-th/9904191} \begin{center} \vskip 1.5 cm {\Large \bf Wilson Loops and Minimal Surfaces} \vskip 1 cm {\large Nadav Drukker$^{1,2}$, David J. Gross$^1$ and Hirosi Ooguri$^{3,4}$}\\ \vskip 1cm $^1$Institute for Theoretical Physics, University of California,\\ Santa Barbara, CA 93106 \smallskip $^2$Department of Physics, Princeton University,\\ Princeton, NJ 08544 \smallskip $^3$Department of Physics, University of California,\\ Berkeley, CA 94720-7300 \smallskip $^4$Lawrence Berkeley National Laboratory, \\ Mail Stop 50A-5101, Berkeley, CA 94720 \bigskip {\tt <EMAIL>, <EMAIL>, <EMAIL>} \end{center} \vskip 0.5 cm \begin{abstract} The $AdS$/CFT correspondence suggests that the Wilson loop of the large $N$ gauge theory with ${\cal N}=4$ supersymmetry in 4 dimensions is described by a minimal surface in $AdS_5 \times S^5$. We examine various aspects of this proposal, comparing gauge theory expectations with computations of minimal surfaces. There is a distinguished class of loops, which we call BPS loops, whose expectation values are free from ultra-violet divergence. We formulate the loop equation for such loops. To the extent that we have checked, the minimal surface in $AdS_5 \times S^5$ gives a solution of the equation. We also discuss the zig-zag symmetry of the loop operator. In the ${\cal N}=4$ gauge theory, we expect the zig-zag symmetry to hold when the loop does not couple the scalar fields in the supermultiplet. We will show how this is realized for the minimal surface. \end{abstract} \end{titlepage} \mysection{Introduction} The remarkable duality between 4-dimensional supersymmetric gauge theories and type IIB string theory on $AdS_5\times S^5$ background \cite{Maldacena:1997re} has been studied extensively over the past year and a half. This conjecture is difficult to test. As with many dualities, it relates a weakly coupled string theory to a strongly coupled gauge theory. Weakly coupled string theory is well defined, even though there are technical problems in doing calculations with Ramond-Ramond backgrounds. But how can one compare the results to the gauge theory, which is strongly coupled? Even if there is no phase transition in going from weak to strong coupling in the gauge theory, there is little that can be said about the strongly coupled gauge theory. By virtue of non-renormalization theorems, it is possible to calculate some quantities in perturbation theory and extrapolate to strong coupling. Such techniques, however, raise the question of whether these comparisons can be regarded as strong evidence for the conjecture or whether the result is dictated by symmetry alone. Gauge theory without fermions has a non-perturbative formulation on the lattice. This allows one to define, if not compute, quantities at arbitrarily large bare couplings. The lattice formulation of gauge theory enables one to derive a rigorous form of the loop equation \cite{Makeenko:1979pb}, for the large $N$ limit of the theory. These equations are satisfied on the lattice and are solved by the master field of the theory. The only case where the loop equation has been explicitly solved is 2 dimensions, where the theory is soluble \cite{Kazakov:1980zi}. The loop equation can also be derived formally in the continuum field theory. It has been shown that the perturbative expansion of the theory yields a solution to the loop equation. This is also the case for supersymmetric theories. Thus, although there is no formulation of supersymmetric theories on the lattice, we assume that those theories still satisfy a large $N$ loop equation. Since this equation holds for all couplings we can use it for strong coupling as well. One of the goals of this paper is to check if the $AdS_5$ ansatz for the expectation value of the Wilson loop operator satisfies the loop equation. To the extent that we were able to reliably estimate properties of string in $AdS_5$, the loop equation is satisfied. However we were unable to test them in all interesting cases. In the course of our investigation we will also learn new facts about Wilson loops and strings in Anti de-Sitter space. We discuss the best understood and most studied case of the $AdS$/CFT correspondence between type IIB superstring on $AdS_5\times S^5$ and ${\cal N}=4$ super Yang-Mills theory with gauge group $SU(N)$ in 4 dimensions. We will concentrate on the case with Euclidean signature metric. Let us review some basic facts about this duality\footnote{For more complete reviews see \cite{Schwarz:1998fd,Douglas:1999ww}.}. The near horizon geometry of $N$ D3-branes is given by the metric \begin{equation} {ds^2\over \alpha'} ={U^2\over\sqrt{4\pi g_sN}}\sum_{\mu=0}^3dX^\mu dX^\mu +\sqrt{4\pi g_sN}{dU^2\over U^2} +\sqrt{4\pi g_sN}d\Omega_5^2 , \end{equation} where $g_s$ is the string coupling constant and the string tension is $(2\pi \alpha')^{-1}$. The background contains $N$ units of Ramond-Ramond flux. The $X$ and $U$ are coordinates on $AdS_5$, and $d\Omega_5^2$ is the metric on $S^5$ with unit radius. The curvature radii of both $AdS_5$ and $S^5$ are given by $(4\pi g_sN)^{{1\over 4}} l_s$ where $\alpha' = l_s^2$. We will find it more convenient to rescale the coordinates $X^\mu$ by $1/\sqrt{4\pi g_sN}$ and introduce new coordinates $Y^i=\theta^i/ U$ ($i=1,\cdots,6$), where $\theta^i$ are the coordinates on $S^5$ and $\theta^2 = 1$. The metric in this coordinate system is \begin{equation} {ds^2 \over \alpha'} =\sqrt{4\pi g_s N} Y^{-2} \left(\sum_{\mu=0}^3dX^\mu dX^\mu+\sum_{i=1}^6dY^i dY^i\right). \label{metric} \end{equation} It is interesting to note that $AdS_5 \times S^5$ is conformal to flat ${\mathbb R}^{10}$ if the radii of $AdS_5$ and $S^5$ are the same. In this coordinate system, the boundary of $AdS_5$ is mapped to the origin $Y^i=0$ of ${\mathbb R}^6$. The gauge theory coupling $g_{YM}$ and the string coupling $g_s$ are related by $g_{YM}^2=4\pi g_s$. We are interested in the limit of $N \rightarrow \infty$ while keeping the 't Hooft coupling $\lambda = g_{YM}^2 N$ finite \cite{'tHooft:1974jz}. After taking the large $N$ limit, we will consider the region $\lambda \gg 1$, where the curvature is small compared to the string scale and stringy excitations are negligible. In this case, the supergravity approximation is reliable. According to the $AdS$/CFT correspondence, every supergravity field has a corresponding local operator in the gauge theory. Correlators of local operators are given by the supergravity action for fields with point sources on the boundary of $AdS_5$ \cite{Gubser:1998bc,Witten:1998qj}. In the classical limit one just solves the equations of motion with such sources. An interesting set of non-local operators in a gauge theory are Wilson loops. It was proposed in \cite{Maldacena:1998im,Rey:1998ik} that the Wilson loop is defined by an open string ending on the loop at the boundary of $AdS_5$. In the classical limit, the string is described by a minimal surface. Due to the curvature of $AdS_5$, the minimal surface does not stay near the boundary, but goes deep into the interior of space, where the area element can be made smaller. Because of this the behavior of the Wilson loop, for large area, is that of a conformal theory, and the area law does not produce confinement. The gauge theory under discussion does not contain quarks or other fields in the fundamental representation of the gauge group. To construct the Wilson loop describing the phase associated with moving a particle in the fundamental representation around a closed curve, we place one of the D-branes very far away from the others. The ground states of the string stretched from the distant D-brane to the others consist of the $W$-bosons and their superpartners in the fundamental representation of the gauge group of the remaining branes. Thus, for large $\lambda$, the expectation value of the Wilson loop is related to the classical action of the string, with appropriate boundary conditions. To the leading order in $\lambda$, we can ignore the effect of the Ramond-Ramond flux and use the Nambu-Goto action, namely the area of the minimal surface \begin{equation} A=\int {d\sigma_1d\sigma_2 \over2\pi\alpha'\sqrt{\lambda}} \sqrt{g} =\int {d\sigma_1d\sigma_2 \over 2 \pi Y^2}\sqrt{\det(\partial_\alpha X^\mu\partial_\beta X^\mu+ \partial_\alpha Y^i\partial_\beta Y^i)}. \label{area} \end{equation} Because of the $Y^{-2}$ factor, this area is infinite. After regularizing the divergence, the infinite part was identified as due to the mass of the $W$-boson and subtracted \cite{Maldacena:1998im}. Taking 2 parallel lines (with opposite orientation) as a quark-anti quark pair, the remaining finite part defines the quark-anti quark potential. Such calculations were used to study the phases of the ${\cal N}=4$ super Yang-Mills theory and to demonstrate confinement in non-supersymmetric generalizations \cite{Witten:1998zw,Brandhuber:1998bs}. We will argue below that the correct action of the Wilson loop is not the area of the minimal surface, but the Legendre transform of it with respect to some of the loop variables. The reason is that some of the string coordinates satisfy Neumann conditions rather than Dirichlet conditions. For a certain class of loops, this Legendre transform exactly removes the divergent piece from the area. As the result, the expectation values of such loops are finite. The appropriate Wilson loop for ${\cal N}=4$ super Yang-Mills theory is an operator of the form (suppressing all fermion fields for the moment) \begin{equation} W[C] ={1\over N}{\rm Tr}\,{\cal P}\, \exp\left(\oint (iA_\mu \dot x^\mu+\Phi_i\dot y^i)ds\right), \end{equation} where $A_\mu$ are the gauge fields and $\Phi_i$ are the six scalars in the adjoint representation, and $C$ represents the loop variables $(x^\mu(s), y^i(s))$. $(x^\mu(s))$ determines the actual loop in four dimensions, $(y^i(s))$ can be thought of as the extra six coordinates of the ten-dimensional ${\cal N}=1$ super Yang-Mills theory, of which our theory is the dimensionally reduced version. It turns out that minimal surfaces terminating at the boundary of $AdS_5$ correspond only to loops that satisfy the constraint $\dot x^2=\dot y^2$. This constraint was derived before, and we study in greater depth its origin and meaning. In \cite{Maldacena:1998im}, the constraint was introduced as a consequence of the fact that the mass of the open string and the Higgs VEV are proportional to each other. We will show that the constraint also has a geometric interpretation in terms of a minimal surface in $AdS_5 \times S^5$. Another interpretation of the constraint has to do with the ${\cal N}=4$ supersymmetry; the loops obeying the constraint are BPS-type objects in loop-space. After discussing various aspects of loops obeying the constraint, we present some idea on how to extend the calculation to a more general class of loops. The loop equation is a differential equation on the loop space. We evaluate, using string theory on $AdS_5$, the action of the loop differential operator $\hat{L}$ on a certain class of Wilson loops. On a smooth loop $C$, we find the differential operator annihilates the vacuum expectation of the loop $\langle W \rangle$, in accord with the loop equation as derived in the gauge theory. On the other hand, for a loop with a self-intersection point, the gauge theory predicts that $\hat{L} \langle W \rangle$ is non-zero and proportional to $g_{YM}^2 N$. We point out the gauge theory also predicts that a cusp (a sharp turning point) in a loop gives a non-zero contribution to the loop equation, proportional to $g_{YM}^2 N$. We will show that $\hat{L} \langle W \rangle$ for a loop with a cusp evaluated by the minimal surface in $AdS_5 \times S^5$ is indeed non-vanishing and proportional to $g_{YM}^2 N$. We have not been able to reproduce the precise dependence on the angle at the cusp due to our lack of detailed understanding of loops not obeying the constraint $\dot x^2 = \dot y^2$. For the same reason we were unable to reproduce the expected result at an intersection. \medskip The paper is organized as follows. \nopagebreak \nopagebreak \medskip \nopagebreak \noindent In Section 2, we start with a brief review of the Wilson loop operator in the pure Yang-Mills theory. We then point out an important subtlety in performing the Wick rotation in the supersymmetric theory. We will present some results from the perturbation theory where the subtlety in the Wick rotation plays an interesting role. \medskip \noindent In Section 3, we turn to string theory in $AdS_5 \times S^5$. We will give a precise specification of boundary conditions on the string worldsheet and the geometric origin of the constraint $\dot x^2 = \dot y^2$. For some cases, we can compute the area of minimal surfaces explicitly. These include loops with intersections or cusps. For such loops, the areas have logarithmic divergences. After calculating those areas, we explain the need for the Legendre transform and show that it removes the linear divergence. The absence of a linear divergence fits well with what we expect for the supersymmetric gauge theory. We will clarify the issue of zig-zag symmetry, and end the section with a discussion of loops that do not satisfy the constraint. \medskip \noindent Section 4, we give a review of the loop equation in the pure Yang-Mills theory and derive its generalization to the case of ${\cal N} = 4$ super Yang-Mills theory in 4 dimensions. \medskip \noindent In Section 5, we will discuss to what extent the minimal surface calculation in $AdS_5$ is consistent with the loop equation. \medskip \noindent To make the body of the paper more readable, some details are presented in appendices. In Appendix \ref{appendix-phase} we derive the Wilson loop as the first quantized action of the $W$-boson. In Appendix \ref{appendix-cusp} we calculate the area of a minimal surface near a cusp. In Appendix \ref{appendix-N=4}, we present some more details on the loop equation of the ${\cal N}=4$ theory. \mysection{Wilson Loops in N=4 Gauge Theory} We define the Wilson loop operator in the supersymmetric gauge theory, and review some of its basic properties. We pay particular attention to its coupling to the scalar fields in the supermultiplet. \subsection{Definition} \label{generalities} One of the most interesting observables in gauge theories is the Wilson loop, the path-ordered exponential of the gauge field, \begin{equation} W={1\over N}{\rm Tr}\,{\cal P}\,\exp\left(i\oint A_\mu dx^\mu\right), \label{wilsonloop} \end{equation} with the trace in the fundamental representation. The Wilson loop can be defined for any closed path in space, providing a large class of gauge invariant observables. In fact, these operators, and their products, form a complete basis of gauge invariant operators for pure Yang-Mills theory. An appropriate definition of the loop operator for the ${\cal N}=4$ super Yang-Mills theory in 4 dimensions will be given below. One of physical applications of Wilson loops stems from the fact that an infinitely massive quark in the fundamental representation moving along the loop will be transformed by the phase factor in (\ref{wilsonloop}). Thus the dynamical effects of the gauge dynamics on external quark sources is measured by the Wilson loop. In particular for a parallel quark anti-quark pair, the Wilson loop is the exponent of the effective potential between the quarks and serves as an order parameter for confinement \cite{Wilson:1974sk}. The Maldacena conjecture states that type IIB string theory on $AdS_5\times S^5$ is dual to ${\cal N}=4$ super Yang-Mills theory in 4 dimensions. This gauge theory does not contain quarks in the fundamental representation. To construct the Wilson loop, we separate a single D-brane from the $N$ D-branes and take it very far away. For large $N$, we can ignore the fields on the distant D-brane, except for open strings stretching between it and the other $N$. The ground states of the open string are the $W$-bosons and their superpartners of the broken, $SU(N)$, gauge group. Their trajectories should give the same effect as that of an infinitely massive particle in the fundamental representation. The correlation functions of the $W$-boson can be written in the first quantized formalism as an integral over paths. This description is studied in detail in Appendix \ref{appendix-phase}. When the 4-dimensional space has the Lorentzian signature metric, the phase factor associated to the loop is given by the vacuum expectation value of the operator \begin{equation} W={1\over N}{\rm Tr}\,{\cal P}\, \exp\left(i\oint (A_\mu \dot x^\mu+|\dot x|\Phi_i\theta^i)ds\right). \end{equation} When the metric is Euclidean, there is an important modification to this formula as \begin{equation} W={1\over N}{\rm Tr}\,{\cal P}\, \exp\left(i\oint (A_\mu \dot x^\mu-i|\dot x|\Phi_i\theta^i)ds\right). \label{euclid-phase} \end{equation} Notice the presence of $i$ in the second term in the exponent. The ``phase-factor'' in the Euclidean theory is not really a phase, but contains a real part. In the above, $\theta^i$ are angular coordinates of magnitude $1$ and can be regarded as coordinates on $S^5$. In the gauge theory, we may consider a more general class of Wilson loops of the form \begin{equation} W={1\over N}{\rm Tr}\,{\cal P}\, \exp\left(\oint (iA_\mu \dot x^\mu+\Phi_i\dot y^i)ds\right). \label{mostgeneral} \end{equation} with an arbitrary function $y^i(s)$. This is the general loop we would get by dimensional reduction from the 10-dimensional gauge theory, where $\Phi_i$ would be the extra six components of the gauge field. Equation (\ref{euclid-phase}) restricts us to the case of $\dot x^2-\dot y^2=0$. This suggests that the metric on the loop variables $(x^\mu(s), y^i(s))$ has the signature $(4,6)$. It is important to stress that this is not the signature of $AdS_5\times S^5$ but of the space where the loops are defined\footnote{One may regard the extra factor of $i$ in the Euclidean case (\ref{euclid-phase}) as a Wick-rotation of the 6 $y$-coordinates so that we can express the constraint as $\dot x^\mu\dot x_\mu+\dot y^i\dot y_i=0$, both in the Lorentzian and the Euclidean cases. To avoid confusions, we will not use this convention and write the $i$ explicitly in all our expressions in the Euclidean case.}. As we will show later, the signature of the loop space metric is related to the fact that the 6 loop variables $y^i(s)$ correspond to T-dual coordinates on the string worldsheet. The constraint $\dot x^2-\dot y^2=0$ is also related to supersymmetry. Gauge invariance in 4 dimensions requires that the Wilson loop close in 4 dimensions, $i.e.$ the loop variables $x^\mu(s)$ are continuous and periodic around the loop. This is not the case for the other 6 variables $y^i(s)$, and the loop may have a jump in these 6 directions. \subsection{Perturbation Theory} \label{pertsection} As a warm-up, we study properties of the Wilson loops in perturbation theory. To first order in $g_{YM}^2 N$, the expectation value of the loop $\langle W \rangle$ is given by \begin{eqnarray} \langle W[C] \rangle &=&1-g_{YM}^2N\oint ds\oint ds' \Big[ \dot x^\mu(s)\dot x^\nu(s') G_{\mu\nu}\left(x(s)-x(s')\right) \phantom{G_{ij}^2} \nonumber\\ &&\hskip1.5in -\dot y^i(s)\dot y^j(s') G_{ij}\left(x(s)-x(s')\right) \Big], \end{eqnarray} where $G_{\mu\nu}$ and $G_{ij}$ are the gauge field and scalar propagators. The relative minus sign comes from the extra $i$ in front of the scalar piece in the exponent in (\ref{euclid-phase}). This integral is linearly divergent. With a regularization of the propagator with cutoff $\epsilon$ ({\it i.e.} replacing $1/x^2$ with $1/(x^2+\epsilon^2)$), the divergent piece coming from the exchange of the gauge field $A_\mu$ is evaluated as \begin{equation} -{\lambda\over8\pi^2} \oint ds\int_{-{\epsilon\over|\dot x|}}^{\epsilon\over|\dot x|}ds' \dot x^\mu(s)\dot x^\nu(s') {\delta_{\mu\nu}\over\epsilon^2} =-{\lambda\over(2\pi)^2\epsilon}\oint ds|\dot x| =-\lambda{L\over(2\pi)^2\epsilon}, \label{pert-A} \end{equation} where $L$ is the circumference of the loop. The divergent contribution from the exchange of the scalars $\Phi_i$ is \begin{equation} {\lambda\over8\pi^2} \oint ds\int_{-{\epsilon\over|\dot x|}}^{\epsilon\over|\dot x|}ds' \dot y^i(s)\dot y^j(s') {\delta_{ij}\over\epsilon^2} ={\lambda\over(2\pi)^2\epsilon}\oint ds |\dot x| {\dot y^2\over \dot x^2}. \label{pert-Phi} \end{equation} Combining these terms together, we find \begin{equation} W = 1 + {\lambda \over (2\pi)^2 \epsilon} \oint ds |\dot x| \left(1 - {\dot y^2 \over \dot x^2} \right) + {\rm finite}. \label{pert-div} \end{equation} We note that the linear divergence cancels when the constraint $\dot x^2=\dot y^2$ is satisfied. At $n$-th order in the $\lambda = g_{YM}^2 N$ expansion, one finds a linear divergence of the form \begin{equation} {\lambda^n\over\epsilon} \oint ds|\dot x|G_n \left({\dot y^2 \over \dot x^2}\right), \end{equation} for some polynomial $G_n(z)$. We now argue that $G_n(1)=0$, namely the linear divergence cancels when $\dot x^2 = \dot y^2$, to all order in the perturbative expansion. The $n$-th order term is calculated by connected Feynman diagrams with external legs attached to the loop. The linear divergence appears when all the external legs come together in 4 dimensions. Since the Feynman rule of the ${\cal N}=4$ gauge theory is obtained by the dimensional reduction of the 10-dimensional theory, the 10-dimensional rotational invariance of the Feynman rule is recovered in the coincidence limit. Therefore the contractions of the external indices by the Feynman rule produce only rotational invariant combinations of $(\dot x^\mu, i \dot y^i)$, namely a polynomial of $(\dot x^2 - \dot y^2)$. The polynomial does not have a constant term since a connected Feynman diagram for $\langle W \rangle$ needs to have at least 2 external lines attached to the loop. Therefore the polynomial vanishes when $\dot x^2 - \dot y^2=0$. When the loop has a cusp, there is an extra logarithmic divergence from graphs as shown in \figur{ads5} {(a) At one loop, there is a linear divergence from the propagator connecting coincident points. The divergence is proportional to the circumference of the loop. (b) At cusps and intersections, an additional logarithmic divergence appears when the 2 external legs approach the singular point.} Let us denote the angle at the cusp by $\Omega$. We choose the angle so that $\Omega = \pi$ at a regular point of the loop. A one-loop computation with the gauge field gives \begin{equation} {\lambda\over(2\pi)^2}((\pi- \Omega)\cot\Omega+1)\log{L\over\epsilon}. \label{oneloop1} \end{equation} A cusp is a discontinuity of $\dot x^\mu$. There may also be a discontinuity in $\dot y^i$, which we measure by an angle $\Theta$. We choose $\Theta$ so that $\Theta=0$ when $\dot y^i$ is continuous. A one-loop computation with the scalar fields gives \begin{equation} -{\lambda\over (2\pi)^2}\left(-{\pi- \Omega\over\sin\Omega}\cos\Theta+1\right) \log{L\over\epsilon}. \label{oneloop2} \end{equation} Combining (\ref{oneloop1}) and (\ref{oneloop2}) together, we obtain \begin{equation} {\lambda\over (2\pi)^2}{\pi- \Omega\over\sin\Omega}(\cos\Omega+\cos\Theta) \log{L\over\epsilon}. \label{pert-log} \end{equation} A similar computation at an intersection gives \begin{equation} {\lambda\over2\pi}{1\over\sin\Omega}(\cos\Omega+\cos\Theta) \log{L\over\epsilon}. \label{pert-int} \end{equation} \mysection{Minimal surfaces in Anti de-Sitter Space} \label{section-minimal} According to the Maldacena conjecture, the expectation value of the Wilson loop is given by the action of a string bounded by the curve at the boundary of space: \begin{equation} \langle W[C]\rangle =\int_{\partial X=C}{\cal D} X\, \exp( -\sqrt{\lambda} S[X]), \label{pathintegral} \end{equation} for some string action $S[X]$. Here $X$ represents both the bosonic and the fermionic coordinates of the string. For large $\lambda$, we can estimate the path integral by the steepest descent method. Consequently the expectation value of the Wilson loop is related to the area $A$ of the minimal surface bounded by $C$ as \begin{equation} \langle W \rangle \simeq \exp ( - \sqrt{\lambda} A ) . \label{firstansatz} \end{equation} The motivation for this ansatz is that the $W$-boson considered in section \ref{generalities} is described in the D-brane language by an open string going between the single separated D-brane and the other $N$ D-branes. In the near horizon limit, the $N$ D-branes are replaced by the $AdS_5$ geometry and the open string is stretched from the boundary to the interior of $AdS_5$. To be precise, this argument only tells us that the Wilson loop and the string in $AdS_5$ are related to each other. The expression (\ref{pathintegral}) is schematic at best, and there may be an additional loop-dependent factor in (\ref{firstansatz}). A similar problem exists in computation of correlation functions of local operators; there is no known way to fix the relative normalization of local operators in the gauge theory and supergravity fields in $AdS_5$. To determine the normalization factor, one has to compute the 2 point functions \cite{Freedman:1998tz,Lee:1998bx}. In our case, the normalization factor in (\ref{firstansatz}) may depend on the loop variables $C=(x^\mu(s), y^i(s))$. In fact, we will argue below that the correct action to be used in (\ref{firstansatz}) is not the area $A$ of the surface, but the Legendre transform of it. This modification does not change the equations of motion, and the solutions are still minimal surfaces. However the values of the classical action for these surfaces are different than their areas. We will assume that, to the leading order in $\lambda$, there is no further $C$ dependent factor. Otherwise the conjecture would be meaningless as it would produce no falsifiable predictions. On the other hand, one expects a $C$ dependent factor in the subleading order, such as the fluctuation determinant of the surface in $AdS_5$. There can also be a factor in the relation between the $W$-boson propagation amplitude and the Wilson loop computed in Appendix A. Such a factor would be kinematic in nature and independent of $\lambda$, and therefore negligible in our analysis. \subsection{Boundary Conditions and BPS Loop} The Wilson loop discussed in \cite{Maldacena:1998im} obeys the constraint \begin{equation} \dot x^2=\dot y^2. \label{constraint} \end{equation} This constraint was originally derived by using the coupling of the fundamental string to the gauge fields and to the scalars. In our derivation of the loop operator from the phase factor for the $W$-boson amplitude in Appendix \ref{appendix-phase}, the constraint arises from the saddle point in integrating over different reparametrizations of the same loop; essentially for the same reason as in \cite{Maldacena:1998im}. In this section, we will give another interpretation of the constraint (\ref{constraint}), in terms of the string theory in $AdS_5 \times S^5$. For this interpretation we need to give a precise specification of the boundary condition on the string in $AdS_5 \times S^5$. We begin with super Yang-Mills theory in 10 dimensions, which is realized on space-filling D$9$-branes. We ignore the fact that this theory is anomalous since we will reduce it to the anomaly free theory in 4 dimensions. Moreover, we are only interested in the boundary conditions on bosonic variables\footnote{Boundary conditions for fermionic variables are not relevant in our analysis of the loop for large $\lambda$.}. The Wilson loop in 10 dimensions corresponds to an open string worldsheet bounded by the loop, $i.e.$ we should impose full Dirichlet boundary conditions on the string worldsheet. This is natural since, without the Wilson loop operator, the string end-point obeys fully Neumann boundary conditions along the D$9$-brane. The conditions imposed by the Wilson loop are complementary to the boundary conditions on the D$9$-brane. To reduce the theory to 4 dimensions, we perform T-duality along 6 directions. An open string ending on the D$3$-brane obeys 4 Neumann and 6 Dirichlet boundary conditions. Consequently, the Wilson loop operator in the 4-dimensional gauge theory imposes complementary boundary conditions; namely 4 Dirichlet and 6 Neumann boundary conditions. If the Wilson loop is parametrized by the loop variables $(x^\mu(s), y^i(s))$, where $\dot y^i(s)$ couples to the 6 scalar fields, then the 6 loop variables $\dot y^i(s)$ are to be identified with the 6 Neumann boundary conditions on the string worldsheet. We are ready to specify the boundary condition on the string worldsheet living in $AdS_5 \times S^5$, with line element, \begin{equation} {ds^2 \over \alpha'} = \sqrt{\lambda} Y^{-2} \left(\sum_{\mu=0}^3dX^\mu dX^\mu+\sum_{i=1}^6 dY^i dY^i\right). \end{equation} Choose the string world-sheet coordinates to be $(\sigma^1, \sigma^2)$ such that the boundary is located at $\sigma^2=0$. Since $X^\mu$ is identified with the 4 dimensional coordinates where the gauge theory lives, it is natural to impose Dirichlet conditions on $X^\mu$, so that \begin{equation} X^\mu(\sigma_1,0)=x^\mu(\sigma_1). \label{dirichlet} \end{equation} The remaining 6 string coordinates $Y^i(\sigma^1, \sigma^2)$ obey Neumann boundary conditions. We propose that these boundary conditions are \begin{equation} J_1^{~\alpha} \partial_\alpha Y^i(\sigma^1,0) = \dot y^i(\sigma^1), \label{neumann} \end{equation} where $J_\alpha^{~\beta}$ ($\alpha,\beta = 1,2$) is the complex structure on the string worldsheet given in terms of the induced metric $g_{\alpha\beta}$, \begin{equation} J_\alpha^{~\beta} = {1 \over \sqrt{g}} g_{\alpha\gamma} \epsilon^{\gamma\beta}. \end{equation} Although we do not have a derivation of the boundary condition (\ref{neumann}) from first principles, it can be motivated as follows. Because of the identification of the $SO(6)$ symmetries in the $AdS$/CFT correspondence, it is clear that Neumann boundary conditions must set $\dot y^i$ equal to $J_1^{~\alpha} \partial_\alpha Y^i$ up to a relative normalization of the two. The use of the induced complex structure $J_\alpha^{~\beta}$ in the Neumann boundary condition is required by the reparametrization invariance on the worldsheet. The fact that the condition $\dot x^2 = \dot y^2$ has a natural interpretation in terms of the minimal surface, as we will explain below, suggests that the normalization factor is $1$, as in (\ref{neumann}). For a generic choice of the loop variables $(x^\mu(s), y^i(s))$, there is a unique minimal surface in Euclidean space obeying the 10 boundary conditions, (\ref{dirichlet}) and (\ref{neumann}). However the resulting minimal surface does not necessarily terminate at the boundary $Y^i=0$ of $AdS_5$. The condition $Y^i=0$ would be additional Dirichlet conditions, which may or may not be compatible with (\ref{neumann}). In fact, one can show that, for a smooth loop, the additional condition $Y^i(\sigma^1,0)=0$ is satisfied by the minimal surface if and only if the loop variables obey the constraint $\dot x^2 = \dot y^2$. To see this consider the Hamilton-Jacobi equation\footnote{ In general, the Hamilton-Jacobi equation for the area of a minimal surface on a Riemannian manifold with a metric $G_{IJ}$ takes the form, $$G^{IJ} (\delta A/\delta X^I) (\delta A/\delta X^J) = G_{IJ} \partial_1 X^I \partial_1 X^J.$$} for the area $A$ of a minimal surface bounded by a loop $(X^\mu(s), Y^i(s))$ in $AdS_5 \times S^5$: \begin{equation} \left( {\delta A \over \delta X^\mu} \right)^2 + \left( {\delta A \over \delta Y^i} \right)^2 = {1 \over (2 \pi )^2Y^4} \left( (\partial_1 X^\mu)^2 + (\partial_1 Y^i)^2 \right). \end{equation} Since the momenta conjugate to the $X^\mu$'s and the $Y^i$'s are given by \begin{equation} {\delta A \over \delta X^\mu} = {1 \over 2\pi Y^2} J_1^{~\alpha} \partial_\alpha X^\mu, \qquad {\delta A \over \delta Y^i} = {1 \over 2 \pi Y^2} J_1^{~\alpha} \partial_\alpha Y^i, \end{equation} we obtain \begin{equation} (J_1^{~\alpha} \partial_\alpha X^\mu)^2 + (J_1^{~\alpha} \partial_\alpha Y^i)^2 = (\partial_1 X^\mu)^2 + (\partial_1 Y^i)^2. \end{equation} If the minimal surface obeys the boundary conditions (\ref{dirichlet}) and (\ref{neumann}), this becomes \begin{equation} \dot x^2 - \dot y^2 = (J_1^{~\alpha} \partial_\alpha X^\mu)^2 - (\partial_1 Y^i)^2. \label{hamjac} \end{equation} Now impose the additional condition that the string worldsheet terminates at the boundary of $AdS_5$, $i.e.$ $Y^i(\sigma^1,0)=0$. Obviously $\partial_1 Y^i(\sigma^1,0)=0$. This alone tells us that $\dot x^2 - \dot y^2 \geq 0$. Moreover, if the boundary is smooth, it costs a large area to keep $J_1^{~\alpha} \partial_\alpha X^\mu$ non-zero near the boundary of $AdS_5$, so it has to vanish at the boundary $Y=0$ \cite{Maldacena:1998im}. Therefore, the condition that the minimal surface terminates at the boundary of $AdS_5$ requires $\dot x^2 = \dot y^2$. When the constraint $\dot x^2 = \dot y^2$ is satisfied, one can reinterpret the 6 Neumann condition (\ref{neumann}) as Dirichlet conditions on $S^5$. To see this, it is useful to decompose the 6 coordinates $Y^i$ as \begin{equation} Y^i = Y \theta^i \end{equation} where $\theta^i$ are coordinates on $S^5$ and $Y = U^{-1}$ is one of the coordinates on $AdS_5$. Since for a smooth loop the classical solution has $\partial_\alpha Y^i = (\partial_\alpha Y) \theta^i$ at the boundary $Y=0$ of $AdS_5$, the Neumann conditions (\ref{neumann}) turn into the Dirichlet conditions on $S^5$ as \begin{equation} \theta^i(\sigma^1,0) = {\dot y^i \over |\dot y|}. \label{fixedpt} \end{equation} This justifies the boundary conditions used in \cite{Maldacena:1998im}. There is yet another interpretation of the constraint $\dot x^2 = \dot y^2$, and it has to do with supersymmetry. The loops we have considered so far couple only to bosonic fields: the gauge field $A_\mu$ and scalars $\Phi^i$. We also need to allow coupling to the fermionic fields in the exponent. Fermionic variables $\zeta(s)$ along the loop couple to the gauginos $\Psi$ as \begin{equation} \bar\zeta(\dot x^\mu\Gamma_\mu-i\dot y^i\Gamma_i)\Psi, \end{equation} where we are using 10-dimensional gamma matrices $\Gamma_\mu$ and $\Gamma_i$ with signature (10,0). This is derived in Appendix \ref{appendix-N=4}. Exactly when the constraint is satisfied this combination of gamma matrices becomes nilpotent. Consequently only half the components of $\zeta$ couple to $\Psi$, putting the loop in a short representation of local supersymmetry in super loop-space. The simplest example is when the Wilson loop is a straight line, when $\dot x$ and $\dot y$ are independent of $s$. If $\zeta$ is also constant, this loop is the phase factor associated with the a trajectory of a free BPS particle. \subsection{Calculating the Area} \label{section-area} The computation of the Wilson loop in $AdS_5$ requires an infrared regularization, since the area of the minimal surface terminating at the boundary of $AdS_5$ is infinite due to the factor $Y^{-2}$ in the metric. In order to make sense of the ansatz (\ref{firstansatz}), we need to regularize the area. One natural way to do so is to impose the boundary conditions (\ref{dirichlet}) and (\ref{neumann}) at $Y=0$, but integrate the area element only over the part of the surface with $Y \geq \epsilon$. On the gauge theory side, the Wilson loop requires regularization in the ultraviolet. According to the UV/IR relation in the $AdS$/CFT correspondence \cite{Susskind:1998dq}, the IR cutoff $\epsilon$ in $AdS_5$ should be identified with the UV cutoff in the gauge theory. \medskip There are a few cases when minimal surfaces can be studied analytically. \smallskip \noindent (1) Parallel Lines: \noindent The minimal surface for parallel lines, each of length $L$ and separated by a distance $R$, was obtained in \cite{Maldacena:1998im,Rey:1998ik}. The area of the loop is \begin{equation} A={2L\over2 \pi \epsilon} -{4\pi\sqrt{2}\over\Gamma(1/4)^4}{L\over R}. \label{parallel} \end{equation} \smallskip \noindent (2) Circular Loop: \noindent The minimal surface in $AdS_5$ bounded by a circle of radius $R$ is found in \cite{dgo,Berenstein:1998ij} as \begin{equation} Y(r,\varphi)=\sqrt{R^2-r^2}, \label{circlesol} \end{equation} where $r$ and $\varphi$ are radial coordinates on a plane in the 4 dimensions, and we use them as coordinates on the string worldsheet also. The area of the surface with the cutoff $\epsilon$ is \begin{equation} A={1\over2\pi}\int dr \,r d\varphi\,Y^{-2}\sqrt{1+Y^{\prime2}} =R\int_0^{\sqrt{R^2-\epsilon^2}}{r\,dr\over(R^2-r^2)^{3\over2}} ={2\pi R\over 2 \pi \epsilon}-1. \label{circle} \end{equation} \smallskip \noindent \nopagebreak{ (3) Cusp: \noindent } Another family of minimal surfaces we can solve analytically is a surface near a cusp on ${\mathbb R}^4$ and its generalization including a jump on $S^5$. We can find analytical solutions in this case since the boundary conditions are scale invariant. Using radial coordinates in the vicinity of the cusp, $r$ and $\varphi$, as world sheet coordinates, the scale invariant ansatz, \begin{equation} Y(r,\varphi)={r\over f(\varphi)}, \end{equation} reduces the determination of the minimal surface to a one-dimensional problem. The resulting surface is depicted in \figur{ads4}{A minimal surface for a Wilson loop with a cusp. The regularized area is evaluated over the shaded region.} When there is also a jump on $S^5$, one needs to introduce another variable. An analytical solution in this case is found in a similar way. These solutions are presented in Appendix \ref{appendix-cusp}. The result is that the area of the surface has a logarithmic divergence as well as a linear divergence. It behaves as \begin{equation} A={L\over2 \pi \epsilon} -{1\over2\pi}F(\Omega,\Theta)\log{L\over\epsilon} + \cdots, \label{cusp} \end{equation} where $\Omega$ and $\Theta$ are the cusp angles in $R^4$ and $S^5$ respectively. When either $\Theta$ or $\Omega$ vanishes, we can express $F(\Omega,\Theta)/2\pi$ in terms of elliptic integrals. In \figur{ads6}{The solid curve shows the function $F(\Omega,0)/2\pi$, which appears in the logarithmic divergence of the minimal surface with the cusp of angle $\Omega$. This is compared with the perturbative result (\ref{pert-log}) at the cusp shown in the dashed curve. The dotted curve is half of the perturbative result (\ref{pert-int}) at an intersection.} we show the numerical evaluation of the function $F(\Omega, 0)$ in the solid curve. This is to be compared with the perturbative expression (\ref{pert-log}) shown in the dashed curve. The function $F(\Omega, 0)$ is zero at $\Omega=\pi$ and has a pole at $\Omega=0$. As the angle $\Omega \rightarrow 0$ at the cusp, the loop goes back along it's original path, or backtracks. Regularizing the extra divergence from the pole turns it into a linear divergence which cancels part of the linear divergence from the length of the loop. This is related to issues discussed in the section on the zig-zag symmetry. Away from the cusp, the surface approaches the boundary along the $Y$-direction without a momentum in the $X$-direction. Right at the cusp, however, the surface has momentum in both the $Y$ and $r$ direction. This means that, although the constraint $\dot x^2 = \dot y^2$ is obeyed almost everywhere, it is modified at the cusp as \begin{equation} \dot x^2= \big( 1 +f_0^2 \big) \dot y^2, \label{bad-cusp} \end{equation} where $f_0=f(\varphi=\Omega/2)$ is the minimal value of $f(\varphi)$. \medskip \noindent (4) Intersection: \noindent The minimal surface for a self-intersecting loop is just the sum of 2 cusps. The only difference is that, by the exchange symmetry of the 2 components of the loop, the intersection forces \begin{equation} {\dot y\over|\dot x|} = 0 \label{bad-intersection} \end{equation} instead of (\ref{bad-cusp}). \medskip In all the examples above, there is a linear divergence $(2\pi \epsilon)^{-1}$ in the regularized area. This is true for any loop. As explained in \cite{Maldacena:1998im}, this leading divergence in the area of the minimal surface in $AdS_5$ is proportional to the circumference of the loop\footnote{ We are using the coordinates $X^\mu$ in (\ref{metric}) to describe the configurations of the Wilson loops. With these coordinates, there is no factor of $\lambda$ in the relation between the IR cutoff $\epsilon$ in $AdS_5$ and the UV cutoff of the gauge theory \cite{Susskind:1998dq}. These coordinates are different from the coordinates on the D$3$-brane probe, by a factor of $\sqrt{\lambda}$ \cite{Peet:1999wn}.}. The linear divergence arises from the leading behavior of the surface at small $Y$, $i.e.$ near the boundary of $AdS_5$. In this section, we have computed the regularized area by imposing the boundary condition at the boundary $Y=0$ of $AdS_5$ and integrating the area element over the part of the surface $Y \geq \epsilon$. This is not the unique way to regularize the area. Another reasonable way to compute the minimal surface is to impose the boundary conditions, not at $Y=0$, but at $Y = \epsilon$. The area bounded by the loop on $Y=\epsilon$ is then by itself finite. A comparison of the two regularization prescriptions are illustrated in \figur{ads1}{The comparison of the two regularization prescriptions. The boundary conditions are imposed at $Y=0$ in (a) and at $Y=\epsilon$ in (b). The shaded regions represent the regularized areas.} These two regularizations give the same values for the area, up to terms which vanish as $\epsilon \rightarrow 0$. For example, consider the circular loop. The solution (\ref{circlesol}) can also be regarded as a minimal surface with the boundary condition on $Y=\epsilon$, except that the radius of the circle on $Y=\epsilon$ is now $R_0=\sqrt{R^2-\epsilon^2}$. The area computed in this new regularization is then \begin{equation} A ={1\over\epsilon}\sqrt{R_0^2+\epsilon^2}-1 ={2\pi R_0\over2 \pi\epsilon}-1+{\epsilon\over 2R_0}+\ldots. \end{equation} Thus the results of the two regularizations are the same up to terms which vanish as $\epsilon \rightarrow 0$. It is straightforward to show that this is also the case for the parallel lines. We have also verified that when the loop has a cusp or an intersection, the two regularizations give the same area modulo terms which are finite as $\epsilon \rightarrow 0$, which are subleading compared to the logarithmic divergence. When we impose the boundary condition at $Y=\epsilon$, the constraint on the loop variables is not exactly $\dot x^2 = \dot y^2$, but it is modified. If the loop is smooth, the modification is only by $O(\epsilon)$ terms\footnote{ If the loop has a cusp or an intersection, as we saw earlier, the boundary conditions imposed at $Y=0$ imply the constraint $\dot x^2=\dot y^2$ holds almost everywhere along the loop, except at a cusp or an intersection point. When we impose the boundary conditions at $Y=\epsilon$, the constraint is modified in regions of size $\epsilon$ near the cusp and the intersection point.}. Therefore most of the results in this paper are independent of the choice between the two ways of imposing the boundary conditions. The only exception to this rule is the discussion of the zig-zag symmetry. The zig-zag symmetry of the string worldsheet on $AdS_5$ seems to fit well with our expectations about the gauge theory when we use the boundary conditions at $Y=\epsilon$ rather than at $Y=0$. \subsection{Legendre Transformation} \label{corrected-section} The Maldacena conjecture implies that the Wilson loop is related to a string ending along the loop on the boundary of space. In the classical limit, we expect that the string worldsheet is described by a minimal surface. This argument, however, does not completely determine the value of $\langle W \rangle$ for large $\lambda$ since there are many actions whose equations of motion are solved by minimal surfaces. They differ by total derivatives, or boundary terms. Since the surface has boundaries, such terms can be important. In \cite{Maldacena:1998im,Rey:1998ik} it was assumed that one should use the Nambu-Goto action, so the Wilson loop was given in terms of the area $A$ of the minimal surface. This is what we have studied so far. In this section, we argue that $\langle W \rangle$ is in fact given not by $A$ but by an appropriate Legendre transform. We have shown that the loop variables $\dot y^i$ impose Neumann boundary conditions (\ref{neumann}) on the coordinates $Y^i$. Therefore $\langle W \rangle$ should be regarded as a functional of the coordinates $X^\mu$ and the momenta $P_i$ conjugate to $Y^i$, defined by, \begin{equation} P_i ={\delta A\over\delta\partial_2Y^i} ={1\over2\pi\sqrt{\lambda} \alpha'} \sqrt{g}g^{2\alpha}\partial_\alpha Y^j G_{ij}. \end{equation} The Nambu-Goto action is a natural functional of $X^\mu(s)$ and $Y^i(s)$ and is more appropriate for the full Dirichlet boundary conditions. To replace it with a functional of $X^\mu(s)$ and $P^i(s)$, we need to perform the Legendre transform \begin{equation} \tilde{\cal L} ={\cal L}- \partial_2\left(P_i Y^i\right), \end{equation} or \begin{equation} \tilde A =A-\oint d\sigma_1 P_i Y^i. \label{new-action} \end{equation} To show that $\tilde A$ is a natural functional of $(X^\mu,P^i)$, we use Hamilton-Jacobi theory. Under a general variation of the $Y$ coordinates, the variation of the area $A$ of the minimal surface is given by \begin{eqnarray} \delta A &=& \int d\sigma_1d\sigma_2 \left({\delta A\over\delta Y^i} -\partial_\alpha{\delta A\over\delta\partial_\alpha Y^i} \right) \delta Y^i(\sigma_1,\sigma_2) +\oint d\sigma_1{\delta A\over\delta\partial_2 Y^i}\delta Y^i(\sigma_1,0) \nonumber\\ &=& \oint d\sigma_1P_i(\sigma_1,0)\delta Y^i(\sigma_1,0). \end{eqnarray} Here we used the equations of motion. Therefore, after performing the Legendre transformation, we obtain \begin{equation} \delta\tilde A =-\oint d\sigma_1Y^i(\sigma_1,0)\delta P_i(\sigma_1,0). \end{equation} Thus $\tilde A$ is a functional of the momenta $P^i$ at the boundary, not the coordinates $Y^i$. The Neumann boundary conditions (\ref{neumann}) are conditions on the momenta $P^i$, \begin{equation} {\dot y^i\over2\pi}=P^i= Y^2 P_i. \end{equation} In fact, if the loop variables $\dot y^i(s)$ are continuous, the coordinates $Y^i$ are parallel to the momenta $P_i$, as we saw in (\ref{fixedpt}). In this case, the Legendre transform gives \begin{equation} \tilde A =A-{1\over2\pi}\oint d\sigma_1 {\dot y^i \over Y^2} Y^i =A-{1\over2\pi}\oint d\sigma_1 {|\dot y|\over Y} =A-{1\over2\pi\epsilon}\oint ds|\dot y|, \label{legendre2} \end{equation} where $\epsilon$ is the regulator. In the last step, we have set $Y=\epsilon$ since the regularized action is evaluated for $Y \geq \epsilon$. In the previous section, we saw that the area $A$ of minimal surface has a linear divergence proportional to the circumference of the boundary. By combining it with (\ref{legendre2}), we find \begin{equation} \tilde A ={1\over2\pi\epsilon}\oint ds\left(|\dot x|-|\dot y|\right) +\hbox{finite} \end{equation} for a smooth loop. Therefore the linear divergence cancels when the constraint $\dot x^2 = \dot y^2$ is satisfied. The minimal surface in $AdS_5$ is supposed to describe the Wilson loop for large coupling $\lambda$. We saw in section (\ref{pertsection}) that the cancellation of the divergence also takes place to all order in the perturbative expansion $\lambda$. This suggests that the cancellation of the linear divergence is exact, and a smooth loop obeying $\dot x^2 = \dot y^2$ does not require regularization. We suspect that this is a consequence of the BPS property of the loop. When the loop is a straight line, it preserves a global supersymmetry, not only the local one. In that case the lowest order perturbation calculation is exact. The modified action is zero, the expectation value of the Wilson loop is 1. We were not able to find an explicit expression for $\tilde{\cal L}$ as a function of $X^\mu$, $P^i$ and their derivatives. We only know how to evaluate it for classical solutions in terms of the old variables. By definition, the area $A$ of the minimal surface is positive. On the other hand, its Legendre transform $\tilde A$ may be negative and the expectation value of the loop $\langle W \rangle = \exp(-\sqrt{\lambda} \tilde{A})$ may be larger than $1$. In the pure Yang-Mills theory, the Wilson loop is a trace of a unitary operator (divided by the rank $N$ of the gauge group), and its expectation value has to obey the inequality $\langle W \rangle \leq 1$. This is not the case in the supersymmetric theory in the Euclidean signature space since $W$ in (\ref{euclid-phase}) is not a pure phase, and there is no unitarity bound on its expectation value. We have shown that the expectation value of a smooth Wilson loop obeying $\dot x^2 = \dot y^2$ is finite. If the loop has a cusp or an intersection, the cancellation is not exact and we are left with the logarithmic divergence.\footnote{ If $\Theta\neq0$, the function $F(\Omega,\Theta)$ gets a contribution from the Legendre transformation.} \begin{equation} \tilde A = - {1\over2\pi} F(\Omega,\Theta) \log{L\over\epsilon} +\hbox{finite}. \end{equation} It is interesting to note that the constraint $\dot x^2 = \dot y^2$ is not satisfied either at a cusp \begin{equation} \dot x^2 = \big(1 + f_0^2 \big) \dot y^2, \end{equation} or at an intersection point \begin{equation} {\dot y^i\over|\dot x|} = 0 \end{equation} We suspect that the logarithmic divergences at the cusp and the intersection are caused by the failure of the loop to satisfy the BPS condition at these points. \subsection{Zig-Zag Symmetry} A Wilson loop of the form \begin{equation} W={1\over N}{\rm Tr}\,{\cal P}\,\exp( i\oint ds A_\mu \dot x^\mu) \end{equation} is reparametrization invariant, in $s$, namely unchanged by $s\to f(s)$. Formally it is even invariant under reparametrizations which backtrack (namely when $\dot f(s)$ is not always positive) since the phase factor going forward and then backwards will cancel. Polyakov has argued in \cite{Polyakov:1997tj} that this \lq\lq zig-zag symmetry" is one of the basic properties of the QCD string. One must however be careful, even in pure Yang-Mills theory, since the loop requires regularization. Zig-zag symmetry, in fact, is only true perturbatively for regularized loops, where the backtracking paths are closer than the ultraviolet cutoff. It was pointed out in \cite{Maldacena:1998im} that the Wilson loop in the supersymmetric theory (\ref{euclid-phase}), with the constraint $\dot x^2 = \dot y^2$, does not have this symmetry. This is because the couplings of the Wilson loop to the scalar fields $\Phi^i$ is proportional to $|\dot x|$, which does not change the sign when the loop backtracks. Thus if the loop stays at the same point $\theta^i$ on $S^5$, there is no cancellation of the coupling to the scalar fields. In perturbation theory, one can easily prove that the zig-zag symmetry holds for the Wilson loop (\ref{mostgeneral}) when $\dot y^i = 0$. Suppose we have a segment $C_1$ of a loop which goes in one direction and another segment $C_2$ which comes back parallel to $C_1$ but in the opposite direction, as shown in \figur{ads7}{The zig-zag loop; the loop goes in one direction along $C_1$ and comes back along $C_2$. The two segments $C_1$ and $C_2$ are parallel and their distance $\eta$ is less that the gauge theory UV cutoff $\epsilon$.} If the distance $\eta$ between $C_1$ and $C_2$ is much less than the UV regularization $\epsilon$ of the gauge theory, there is one-to-one cancellation between a Feynman diagram $\Gamma$ which has one of its external leg ending on $C_1$ and another diagram $\Gamma'$ which is identical to $\Gamma$ except that the corresponding leg ends on $C_2$. Therefore, to all order in the perturbative expansion, the segments $C_1$ and $C_2$ do not contribute to the expectation value of the Wilson loop. On the other hand, if $\dot y^i = |\dot x| \theta^i$ and $\theta^i$ is fixed at a point on $S^5$, a diagram with a leg coupled to $\dot y^i$ on $C_1$ and one with the corresponding leg coupled to $\dot y^i$ on $C_2$ add up, rather than cancel each other. The perturbative computation therefore shows no zig-zag symmetry in this case. When the coupling $\lambda$ is large, we expect that $\langle W \rangle$ is related to the minimal surface. The area functional, and as a matter of fact any other functional which is an integral over a minimal surface, has zig-zag symmetry. The proof is simple. If we look at the region $Y \geq \epsilon$, the minimal surface bounded by a backtracking loop is almost identical to the surface bound by the curve without backtracking if the separation $\eta$ between $C_1$ and $C_2$ is much less than the cutoff $\epsilon$. This is illustrated in \figur{ads3}{The area of a loop with a zig-zag (a) is roughly the same as the loop without it (b).} Therefore an action on the surface given by an integral over the part of the surface in $Y \geq \epsilon$ is the same with or without the backtracking. At first sight, the zig-zag symmetry of the minimal surface appears in contradiction with the gauge theory expectation since we know the minimal surface ending along a smooth loop on the boundary of $AdS_5$ obeys the constraint $\dot x^2 = \dot y^2$ and therefore $\dot y^i \neq 0$. In the gauge theory, we do {\it not} expect zig-zag symmetry when $\dot y^i$ is non-zero and constant. A close examination of the boundary condition, however, reveals that the situation is more subtle. It is true that, if we impose the boundary conditions at $Y=0$, the part of the surface connecting $C_1$ and $C_2$ do not reach $Y=\epsilon$ and does not contribute to the regularized area for $Y \geq \epsilon$. Therefore zig-zag symmetry holds for $\langle W \rangle$. This is also the case when we impose the boundary condition at $Y=\epsilon$. In this case, if $\epsilon \gg \eta$, the minimal surface goes from $C_1$ to $C_2$ along the $Y=\epsilon$ surface. Therefore the contribution of the segments to the regularized area is proportional to $\eta/\epsilon^2$ times the length of the segment and vanish in the limit $\eta \rightarrow 0$. However the physical interpretation of the two computations are quite different. If the boundary conditions are imposed at $Y=0$, the constraint $\dot x^2 = \dot y^2$ holds provided the segments $C_1$ and $C_2$ are smooth. On the other hand, if the conditions are imposed on the $Y=\epsilon$ hypersurface, the minimal surface bounded by $C_1$ and $C_2$ stays within $\eta$ from $Y=\epsilon$, and $\dot y^2$ vanishes as $\eta/\epsilon \rightarrow 0$. If we take the latter point of view, the apparent contradiction with the gauge theory expectation disappears since the minimal surface in question is related to the Wilson loop which does not couple to the scalar fields in the segments $C_1$ and $C_2$. This is exactly the situation in which zig-zag symmetry arises in the gauge theory. One may argue that the boundary condition at $Y=\epsilon$ gives a more precise definition of the Wilson loop $\langle W \rangle$ as a functional of the loop variables $(x^\mu(s), y^i(s))$. The Legendre transformation of the area $A$ in section (\ref{corrected-section}), for example, is a way to define a functional of the momenta $P^i$ evaluated at $Y=\epsilon$ and not at $Y=0$. It does not make sense to perform this procedure at $Y=0$ since the factor $1/\epsilon$ in the right-hand side of (\ref{new-action}) needs to be replaced by $\infty$. In most of the cases discussed in this paper, whether we impose the boundary conditions at $Y=0$ or $Y=\epsilon$ does not make much difference since the value of the momenta $P^i$ stays almost the same in the region $0 \leq Y \leq \epsilon$. The analysis of zig-zag symmetry, however, seems to be an exception to this rule. If we use the boundary condition at $Y=\epsilon$, the existence of the minimal surface requires the constraint $\dot y^i(s) = 0$ rather than $\dot x^2 = \dot y^2$ for the backtracking loop, and the result fits well with the gauge theory expectation. Clearly the regularization dependent nature of zig-zag symmetry needs to be clarified further. An analysis similar to the one given above leads to the following observations about the Wilson loop, which we find interesting. Consider a self-intersecting loop as in \figur{ads2}{(a) A self-intersecting loop which corresponds to a single trace operator and (b) A pair of loops obtained by reconnecting the loop at the intersection.} The area calculated on the minimal surface bound by the loop (a) is the same as the sum of the two areas bounded by the separated loops (b). In the gauge theory, these loops are very different objects. One is a single trace operator and the other a multi-trace operator. We can even connect two distant closed loops by a long neck without changing the value of the loop since the minimal surface spanning the neck region does not contribute to the area. Graphically this can be written as \begin{equation} \vev{\matrix{\scriptstyle i\cr \scriptstyle j} \!\!\!=\!=\!=\!=\!=\!=\!=\!\!\! \matrix{\scriptstyle k\cr \scriptstyle l}} ={1\over N}\delta_{ij}\delta_{kl} \end{equation} This suggests that the parallel transport $U = {\cal P} \exp (i\int A_\mu dx^\mu )$ along an open curve behaves as a random matrix. As in the case of the zig-zag symmetry, if we impose the boundary condition at $Y=\epsilon$, the minimal surface exists only when $\dot y^i(s)=0$, and we are considering a loop which does not couple to the scalar fields in the neck region. \subsection{Removing the Constraint} \label{section-lifting} So far we considered loops of the form (\ref{euclid-phase}) which satisfy the constraint $\dot x^2-\dot y^2=0$. When the loop has a cusp or an intersection, this constraint is modified as in (\ref{bad-cusp}) and (\ref{bad-intersection}). In the gauge theory, we can define the loop operator for any $(x^\mu(s), y^i(s))$, not necessarily obeying the constraint. Consequently, we need to find a way to calculate an expectation value of such a loop in $AdS_5$ so that the relation between the gauge theory and string theory is complete. The reason given by Maldacena for the constraint (and also in Appendix \ref{appendix-phase}) is that the $W$-bosons are BPS particles and their charges and masses are related. To break the constraint, one needs a non-BPS object with an arbitrary mass. Fortunately string theory contains many such objects. Instead of considering the ground state of the open string corresponding to the $W$-boson, one may use excited string states, which have extra mass from the string oscillations. As shown in the Appendix A, an excited string indeed generates a loop obeying the modified constraint, \begin{equation} \dot y^2=\dot x^2 {M^2\over M^2+m^2}, \label{excited} \end{equation} where $M=\epsilon^{-1}$ is the original $W$-boson mass and $m$ is the mass of the excitations. This makes it possible to relax the constraint, at least for $\dot x^2 \geq \do y^2$. For the loop obeying the original constraint $\dot x^2 = \dot y^2$, the regularized area has the linear divergence of the form \begin{equation} A = \frac{1}{2\pi\epsilon} \oint ds |\dot x| + \cdots = \frac{1}{2\pi} \oint M |\dot x| + \cdots. \end{equation} We expect that the corresponding computation using the string excitation replaces $M$ by $\sqrt{M^2 + m^2}$ as \begin{equation} A = \frac{1}{2\pi} \oint ds \sqrt{M^2 + m^2} |\dot x| + \cdots = \frac{1}{2\pi \epsilon} \oint ds \frac{\dot x^2}{|\dot y|} + \cdots. \end{equation} The Legendre transformation turns this into \begin{eqnarray} \tilde{A} &=& A - \frac{1}{2\pi \epsilon} \oint ds |\dot y| \nonumber \\ &=& \frac{1}{2\pi\epsilon} \oint ds \left( \frac{\dot x^2}{|\dot y|} - |\dot y| \right) + \cdots. \end{eqnarray} This shows that the linear divergence is not completely canceled for $|\dot x| \neq |\dot y|$. Since a highly excited string state may be sensitive to stringy corrections, we can trust this estimate of the linear divergence only for small deviation from the constraint. In the following, we will use an approximate expression for $|\dot x| \sim |\dot y|$ as \begin{equation} \tilde{A} = \frac{1}{\pi \epsilon} \oint ds (|\dot x| - |\dot y|) + \cdots. \label{largeNdiv} \end{equation} \mysection{The Loop Equation} Since the expectation value of the Wilson loop is a measure of confinement, much attention has been given to calculating them. In particular, in the large $N$ limit of gauge theory, they satisfy a closed set of equations \cite{Makeenko:1979pb}. In this section, we first give a review of the loop equation for pure Yang-Mills theory (for more details see \cite{Migdal:1984gj,Polyakov:1987ez}). The equation is easy to write down and is formally satisfied, order by order, in the perturbative expansion of the gauge theory. The lattice version of the loop equations are also satisfied in the non-perturbative lattice formulation of the theory. However, the only case where one can solve explicitly for Wilson loops is in 2 dimensions. There indeed they do satisfy the loop equation. We will then formulate the loop equation for the ${\cal N}=4$ super Yang-Mills theory in 4 dimensions. As far as we know, the loop equation in this case has not been derived before. We will find that the BPS condition (\ref{constraint}) will play a crucial role. We will discuss details of the construction in Appendix \ref{appendix-N=4} and present only the general ideas here. \subsection{Bosonic Theories} The action of pure gauge theory in any number of dimensions is\footnote{The complete action contains a gauge fixing term and ghosts. Those appear also in the equations of motion, but can be dropped by a Ward identity \cite{Brandt:1982gz}.} \begin{equation} {\cal S}={1\over4g_{YM}^2}\int dx\,{\rm Tr} F_{\mu\nu}F^{\mu\nu}, \end{equation} and the Wilson loop is given by \begin{equation} W={1\over N}{\rm Tr}\,{\cal P}\, \exp\left(i\oint A_\mu dx^\mu\right), \end{equation} where the integral is over a path parametrized by~$x^\mu$. The main observation is that there is a differential operator on loop space which brings down the variation of the action $D^\nu F_{\mu\nu}$ as \begin{equation} \hat L\,\vev{W}=-i\oint ds\,\dot x^\mu \vev{(D^\nu F_{\mu\nu})^a(s) \,{1\over N}{\rm Tr}{\cal P}\,T^a(s)\exp\left(i\oint A_\mu dx^\mu\right)} \end{equation} where $T^a(s)$ is the generator of the gauge group inserted at the point $s$ along the loop. There are a few equivalent definitions of~$\hat L$. We will use \begin{equation} \hat L=\lim_{\eta\rightarrow 0} \oint ds \int_{s-\eta}^{s+\eta} ds' {\delta^2\over\delta x^\mu(s')\delta x_\mu(s)}. \label{boson-deriv} \end{equation} As we will explain below, $\eta$ has to be taken much shorter than the UV cutoff scale $\epsilon$ in order to extract the term $D^\nu F_{\mu\nu}$. The insertion of $D^\nu F_{\mu\nu}$ into the loop would be zero if we use the classical equation of motion, but quantum corrections produce contact terms. To see that, one can write the equations of motion as the functional derivative of the action ${\cal S}$ and use the Schwinger-Dyson equations, $i.e.$ integration by parts in the functional integral, \begin{eqnarray} \hat L\,\vev{W} \hskip-.1in&=&\hskip-.1in ig_{YM}^2\int{\cal D} A\oint ds\,{1\over N}{\rm Tr}{\cal P}\,T^a(s) \exp\left(i\oint A_\mu dx^\mu\right) \dot x^\mu(s){\delta e^{-{\cal S}} \over\delta A^{\mu a}(x(s))} \nonumber\\ \hskip-.1in&=&\hskip-.1in -ig_{YM}^2 \vev{\oint ds\,\dot x^\mu(s){\delta\over\delta A^{\mu a}(x(s))} {1\over N}{\rm Tr}{\cal P}\,T^a(s)\exp\left(i\oint A_\mu dx^\mu\right)}. \phantom{(4.5)} \end{eqnarray} The functional derivative $\delta/\delta A_\mu(x(s))$ in this equation is formally evaluated as \begin{eqnarray} \hat L\,\vev{W} &=&{\lambda\over N^2}\oint ds\oint ds'\, \delta(x^\mu(s')-x^\mu(s)) \dot x_\mu(s)\dot x^\mu(s') \times \nonumber \\ &&~~~~~~~~~~~~~\times \vev{{\rm Tr}\,{\cal P}\, T^a(s)T^a(s') \exp\left(i\oint A_\mu dx^\mu\right)}. \end{eqnarray} We then use the relation between the generators of $SU(N)$, \begin{equation} T^a_{nm}T^a_{kl} =\delta_{nk}\delta_{ml}-{\delta_{nm}\delta_{kl}\over N}. \end{equation} Ignoring the $1/N$ term, the trace is broken into two. This gives the correlation function of two loops. In the large $N$ limit, the correlator factorizes and we obtain, \begin{equation} \hat L\,\vev{W} =\lambda\oint ds\oint ds'\, \delta(x^\mu(s')-x^\mu(s)) \dot x_\mu(s)\dot x^\mu(s') \vev{W_{ss'}}\vev{W_{s's}}. \label{boson-eqn} \end{equation} Here $W_{ss'}$ is a Wilson-loop that start at~$s$ and goes to~$s'$ and~$W_{s's}$ goes from~$s'$ to~$s$. They are closed due to the delta function\footnote{The delta-function is not sharp, but is regularized by the cutoff $\epsilon$. That means that the loops $W_{ss'}$ and $W_{s's}$ are not exactly closed loops, and the two ends may be separated by a distance $\epsilon$. This does not contradict gauge invariance since one may consider only gauge transformations which do not vary much over that scale, so the ``almost'' closed loop are ``almost'' gauge invariant. We expect those loops to be equal to the closed loops up to $O(\epsilon)$ corrections.}. Equation (\ref{boson-eqn}) shows that $\hat{L} \langle W \rangle$ receives contributions from self-in\-ter\-sec\-tions of the loop. Since the derivation of the equation is rather formal, it is not clear whether we need to count the trivial case of $s=s'$, in which case $W_{ss'}=1$ and $W_{s's}=W$. In most of the literature on the loop equation, this trivial self-intersection is ignored. In any case, it can be taken care of by multiplicative renormalization of the loop operator. In the supersymmetric gauge theory, the leading contribution from the trivial self-intersection cancels when $\dot x^2 = \dot y^2$. In the definition of the loop derivative $\hat L$, it is important to take the limit $\eta\rightarrow0$. This procedure isolates the term $D^\nu F_{\nu\mu}$, which is a contact term of the double functional derivative. If $\eta$ is of the order of the UV cutoff $\epsilon$, there will be other contributions to the loop equation such as $F_{\mu\nu} F^{\nu\rho} \dot x_\rho$. When calculating the loop equation in perturbation theory, we can take $\eta$ to be arbitrarily small, and in particular $\eta \ll \epsilon$. This is how we view the loop equation in the continuum theory. In fact, it was shown that the perturbative expansion of the Wilson loop solves the loop equation \cite{Brandt:1982gz}. When we study the loop equation the string in $AdS_5$, we will consider the same limit $\eta \rightarrow 0$. In the lattice regularization, it is not possible to calibrate the variation of the loop in distance shorter than the lattice spacing $\epsilon$. In this case, a different definition of $\hat L$ is used which does not require taking such a limit. It is possible to define a loop derivative localized at a point on the loop, instead of the integrated version considered above. The entire derivation goes through by simply dropping one $\oint ds$. \subsection{Supersymmetric Case} We briefly summarize how to derive the loop equation in the supersymmetric theory, leaving the details in Appendix C. We derive them only for variations from constrained loops $\dot x^2 = \dot y^2$. One important modification is due to the extra factor of $i$ in front of the scalars in the Wilson loop operator in the Euclidean theory, \begin{equation} W={1\over N}{\rm Tr}\,{\cal P} \exp\left(\oint (iA_\mu \dot x^\mu+\Phi_i\dot y^i)\,ds\right). \end{equation} Another novelty is the need to include the fermions. The fermions are important even when the loop equation is evaluated at the body part $\zeta(s)=0$ of super loop-space since the fermions appear as source terms in the equations of motion for the gauge fields and the scalars. Here we will explain the effect of the extra $i$. In Appendix \ref{appendix-N=4}, we will discuss how to deal with the fermions. If we define loop derivative \begin{equation} \hat L=\lim_{\eta\rightarrow 0} \oint ds \int_{s-\eta}^{s+\eta} ds' \left({\delta^2\over\delta x^\mu(s')\delta x_\mu(s)} -{\delta^2\over\delta y^i(s')\delta y_i(s)}\right), \label{susy-deriv} \end{equation} then the relative minus sign combines with the extra $i$ to give \begin{eqnarray} \hskip-.1in&&\hskip-.1in\hskip-.25in\hat L\,\vev{W} \nonumber\\ \hskip-.1in&=&\hskip-.1in -i {g_{YM}^2\over N}\oint ds\,\vev{ \left(\dot x^\mu{\delta\over\delta A^{\mu a}} -i\dot y^i{\delta\over\delta\Phi^{ia}}\right) {\rm Tr}{\cal P}\,T^a\exp\left(\oint (iA_\mu \dot x^\mu+\Phi_i\dot y^i)\,ds\right)} \nonumber\\ \hskip-.1in&=&\hskip-.1in \lambda\oint ds\oint ds' \left(\dot x^\mu(s)\dot x_\mu(s')-\dot y^i(s)\dot y_i(s')\right) \delta^4(x(s)-x(s'))\vev{W_1}\vev{W_2}. \phantom{(3.11)} \label{susy-eqn} \end{eqnarray} A simple way to obtain this is by considering the extra $i$ as the Wick rotation of the $y^i$ coordinates and repeat the derivation from (\ref{boson-deriv}) to (\ref{boson-eqn}). The right-hand side of the bosonic loop equation contains a cubic divergence proportional to the circumference of the loop. In the supersymmetric case this ``zero-point energy'' cancels for a smooth loop by the constraint $\dot x^2 = \dot y^2$. \subsection{Predictions} In this subsection, we evaluate the right-hand side of the loop equation (\ref{susy-eqn}) for various types of loops. In the next section, we will compare it with computations of the loop using the minimal surface spanned by the loop in $AdS_5$. In the supersymmetric theory, the trivial self-intersection at $s=s'$ does not contribute to the right-hand side if the loop is smooth and obeys the constraint $\dot x^2 = \dot y^2$. This is related to the fact that such a loop does not require regularization. To be precise, the constraint only cancels the leading divergence proportional to $\epsilon^{-3}$. Since the delta-function in (\ref{susy-eqn}) has a width $\epsilon$, the Taylor expansion of $x(s')$ at $s'=s$ gives subleading terms in $\epsilon$ such as \begin{equation} -{\lambda\over3\epsilon}\oint ds\,(\ddot x^2-\ddot y^2). \label{ambiguity} \end{equation} However this expression is highly regularization dependent. Moreover there are other contributions of the same order due to the fact that the loops $W_{ss'}$ and $W_{s's}$ are not precisely closed, as explained in the last footnote. At any rate, these terms are negligible (by a factor $\epsilon$) compared to the terms we will find at cusps and intersections, and we will ignore them for the rest of the paper. For a loop with an intersection, the integral over the regularized delta-function in the right-hand side of the loop equation gives \begin{eqnarray} && \lambda (\cos \Omega + \cos \Theta) \oint ds\oint ds'\, |\dot x(s)||\dot x(s')| \delta^4_\epsilon\left(x^\mu(s)-x^\mu(s')\right) \nonumber\\ &=& \lambda (\cos \Omega + \cos \Theta) \int_{-\infty}^\infty dx \int_{-\infty}^\infty dx'\, \delta^4_\epsilon\left(\sin\Omega(x-x')\right) \nonumber \\ &=&\lambda {\cos \Omega + \cos \Theta \over2\pi\epsilon^2\sin\Omega}. \end{eqnarray} It is important to note that the result depends explicitly on the UV cutoff $\epsilon^{-2}$. Here we have evaluated the leading term in the $\epsilon^{-1}$ expansion only. There are subleading terms in the expansion which are comparable to (\ref{ambiguity}) at the trivial self-intersection. A cusp also gives an interesting contribution to the loop equation. This may be regarded as a special case of the trivial self-intersection. In fact, in the literature, this effect is ignored together with that of the trivial self-intersection\footnote{ In the lattice formulation, the effect of the cusp to the loop equation is not seen since there is no local definition of a cusp.}. In the supersymmetric theory, the contribution from the trivial self-intersection at a smooth point on the loop is canceled by the constraint $\dot x^2 = \dot y^2$. The situation is more interesting at the cusp since the tangent vector $\dot x^\mu(s)$ is discontinuous there. If there is a jump on $S^5$, $\dot y^i(s)$ is also discontinuous. A simple calculation (identical to (\ref{pert-log}), where we found the log divergence in perturbation theory) shows that the cusp contribute to the right-hand side of the loop equation as \begin{eqnarray} &&2\lambda (\cos \Omega + \cos \Theta) \int_{-\infty}^0 dx\int_0^\infty dx'\, \delta^4_\epsilon\left(\sin\Omega\right(x-x'))\nonumber \\ &=& \lambda {(\pi-\Omega)(\cos \Omega + \cos \Theta) \over(2\pi\epsilon)^2\sin\Omega}. \label{predict-cusp} \end{eqnarray} To summarize, we can express the loop equation as \begin{eqnarray} &&\hat L\,\vev{W} ={\lambda\over2\pi\epsilon^2}\left( \sum_{n:{\rm cusps}} {(\pi- \Omega_n)(\cos\Omega_n+\cos\Theta_n)\over2\pi\sin\Omega_n} \vev{W}+ \right. \nonumber\\ &&\hskip.2in\left. +\sum_{m:{\rm intersections}} {\cos\Omega_m+\cos\Theta_m\over\sin\Omega_m} \vev{W_m}\vev{\widetilde{W}_m}\right) +O\left({\lambda\over\epsilon}\right), \label{predict} \end{eqnarray} where $W_m$ and $\widetilde{W}_m$ are Wilson loops one obtains by detaching the original loop into two at the intersection point $m$. \mysection{Loop Equation in $AdS_5 \times S^5$} \subsection{General Case} In this section, we will examine whether the computation of the loop using string theory in $AdS_5$ agrees with the predictions of the loop equation. A general form of the loop expectation value is \begin{equation} \vev{W}=\Delta \exp \left(-\sqrt{\lambda} \tilde{A}\right). \label{generalform} \end{equation} We assume that the dependence of the prefactor $\Delta$ on the loop variables is subleading for large $\lambda$. Since the loop derivative $\hat{L}$ does not commute with the constraint $\dot x^2 = \dot y^2$, we need an expression for $\tilde{A}$ when the constraint is not satisfied. As we saw in section 3.5, the exponent $\tilde{A}$ has a linear divergence of the form \begin{equation} \tilde{A}(x,y) = \frac{1}{\pi \epsilon} \oint ds (|\dot x| - |\dot y|) + \cdots \label{exponentofw} \end{equation} to the leading order in $(|\dot x| - |\dot y|)$. The loop derivative is a second order differential operator. When the derivatives act on the exponent and bring it down twice, the result is proportional to $\lambda$. On the other hand, when they act on $\Delta$ or on the same $\tilde{A}$ twice, we get things only of order $\sqrt\lambda$ or less. In the following, we will pay attention to the leading term in $\lambda$ only. The exact expression we have to evaluate is therefore, \begin{equation} \lambda \lim_{\eta\rightarrow 0} \oint ds \int_{s-\eta}^{s+\eta} ds' \left({\delta\tilde A\over\delta x^\mu(s')} {\delta\tilde A\over\delta x_\mu(s)} -{\delta\tilde A\over\delta y^i(s')} {\delta\tilde A\over\delta y_i(s)}\right). \label{twoderivative} \end{equation} We do not have to include the fermionic derivative. When it acts once on a bosonic loop, it gives a fermion whose expectation value is zero. There are also non-zero contributions when it acts twice on $\tilde A$, but they are subleading in $\lambda$. Let us evaluate (\ref{twoderivative}). Although the linear divergence ${1 \over2\pi\epsilon}\oint ds(|\dot x|-|\dot y|)$ in $\tilde{A}(x,y)$ vanishes for the loop obeying the constraint, the variation $\hat{L}$ does not commute with the constraint. Thus the linear divergence term gives an important contribution to (\ref{twoderivative}). Since the variation of the length functional \begin{equation} L=\oint ds\sqrt{\dot x^2} \end{equation} gives the acceleration $\ddot x^\mu$ (in the parametrization where $|\dot x|=1$) and the same for $y$, we obtain \begin{eqnarray} &&\lambda \left( {\delta\tilde A\over\delta x^\mu(s')} {\delta\tilde A\over\delta x_\mu(s)} -{\delta\tilde A\over\delta y^i(s')} {\delta\tilde A\over\delta y_i(s)} \right)\nonumber \\ &=&{\lambda \over\pi^2\epsilon^2} \left(\ddot x_\mu(s)\ddot x^\mu(s')-\ddot y_i(s)\ddot y^i(s')\right) + \cdots. \label{ads-deriv} \end{eqnarray} Note that it has the same divergence, $\epsilon^{-2}$, as the right-hand side of the loop equation. Moreover the powers of $\lambda$ match up in the loop equation and in (\ref{ads-deriv}). The $\cdots$ in the right-hand side represents variations of the remaining terms in $\tilde{A}$, which are finite for a smooth loop. To compute $\hat{L} \langle W \rangle$, we integrate (\ref{ads-deriv}) over $s-\eta \leq s' \leq s+ \eta$. When the loop is smooth, the acceleration $(\ddot x^\mu, \ddot y^i)$ itself is finite. Therefore, by taking $\eta \rightarrow 0$, one finds that $\hat{L} \vev{W}=0$ in this case. This is consistent with the loop equation. Therefore we reach the first conclusion that a minimal surface in $AdS_5$ bounded by a smooth loop solves the loop equation. \subsection{Loops with Cusps} If the loop has a cusp of angle $\Omega$, the tangent vector is discontinuous and $\ddot x$ has a delta-function pointing along the unit vector bisector $\hat e$ \begin{equation} \ddot x^\mu=2\cos{\Omega\over2}\delta(s)\hat e^\mu. \end{equation} A similar thing happens when $\dot y$ is discontinuous, with the angle $\Theta$ replacing $\Omega$ in the above. This delta-function is regularized by $\eta$, not $\epsilon$, since it is related to the shortest length scale on which the loop is defined. Thus the integral of (\ref{ads-deriv}) over $s$ and $s'$ gives a non-zero result as \begin{eqnarray} && \frac{\lambda}{\pi^2 \epsilon^2} \oint ds\int_{s-\eta}^{s+\eta}ds' \left(\ddot x_\mu(s)\ddot x^\mu(s')-\ddot y_i(s)\ddot y^i(s')\right) \nonumber \\ &=&\frac{4\lambda}{\pi^2 \epsilon^2} \left(\cos^2{\Omega\over2}-\sin^2{\Theta\over2}\right) \nonumber\\ &=&\frac{2\lambda}{\pi^2 \epsilon^2} \left(\cos\Omega+\cos\Theta\right) \label{ads-cusp} \end{eqnarray} In comparison with the prediction of (\ref{predict-cusp}) of the loop equation, we are missing the factor of $(\pi-\Omega)/\sin\Omega$. This, however, is not a contradiction. The expression for the linear divergence term in (\ref{exponentofw}) is an approximation for small $(|\dot x| - |\dot y|)$. Since $\dot x^2 = (1 + f_0) \dot y^2$ with $f_0 = f(\Omega/2)$ at the cusp, this approximation is valid only when $f_0$ is small. Apart from this factor, (\ref{ads-cusp}) agrees with the prediction of the loop equation that the cusp gives a non-zero contribution to the loop equation proportional to $\lambda = g_{YM}^2 N$ times $\epsilon^{-2}$. When $(|\dot x| - |\dot y|)$ is not small, the expression (\ref{exponentofw}) needs to be modified as \begin{equation} \tilde A(x,y) = {1 \over\pi\epsilon} \oint ds|\dot x|G\left({\dot y^2\over\dot x^2}\right) + \cdots \label{modifydivergence} \end{equation} for some function $G(z)$. By repeating the computation that lead to (\ref{ads-cusp}), we find that the contribution of the cusp takes the form \begin{equation} \hat{L} \exp(-\sqrt{\lambda} \tilde{A}) = \lambda {\cal G}(f_0) (\cos \Omega + \cos \Theta) \exp(-\sqrt{\lambda} \tilde{A}) + \cdots, \label{cuspfinal} \end{equation} where ${\cal G}(f_0)$ is a function related to $G(z)$. The agreement with (\ref{predict-cusp}) requires \begin{equation} {\cal G}\big(f(\Omega/2)\big) = \frac{\pi - \Omega}{8\sin \Omega}. \end{equation} Proving this would be a very strong evidence for the conjecture. Loops with cusps have also logarithmic divergences, which could contribute to the loop equations. To see that, one may write the logarithmically divergent term as \begin{equation} {1\over2\pi}F(\Omega)\log{L\over\epsilon} ={1\over2\pi}\int ds\int ds' |\dot x(s)||\dot x(s')| {\sin\varphi\over\pi-\varphi}F(\varphi){1\over(x-x')^2+\epsilon^2} \eeql{L-zig} where $\pi-\varphi$ is the angle between $\dot x(s)$ and $\dot x(s')$. To check this equation one should integrate over two straight lines meeting at a point. Differentiating (\ref{L-zig}) gives a few terms, among them \begin{equation} \ddot x(s){1\over\epsilon}{\sin\Omega\over\pi-\Omega}F(\Omega) \eeql{ads-cusp2} which has the same divergence as the piece that gave (\ref{ads-cusp}). \subsection{Self-Intersecting Loops} The situation at a self-intersection is more mysterious since $\dot x$ and $\dot y$ are both continuous at the intersection point. However we have problems in our ability to test the loop equation in this case. First of all, $\dot y^i = 0$ at the intersection, and the function $G(z)$ which appears in the linear divergence term in (\ref{modifydivergence}) may be singular at $z=|\dot y|/|\dot x| =0$. Since we do not know about the function $G(z)$ except for its behavior near $z = 1$, it is difficult to tell whether there is a contribution from the intersection. The presence of the unknown factor $\Delta$ in (\ref{generalform}) makes the situation worse. As we explained before, the Wilson loop is \begin{equation} \vev{W}=\Delta \exp\left(-\sqrt{\lambda}\tilde A\right). \end{equation} For a self-intersecting loop we expect \begin{equation} \hat L\,\vev{W_{1+2}}=\lambda {\cos\Omega+\cos\Theta\over\sin\Omega} \vev{W_1}\vev{W_2}, \end{equation} where $W_{1+2}$ is the self-intersecting loop and $W_1$ and $W_2$ its two pieces. In order for this to be consistent with the $AdS_5$ computation, we need to find \begin{equation} \hat L\,\exp\left(-\sqrt{\lambda}\tilde A_{1+2}\right) =\lambda {\cos\Omega+\cos\Theta\over\sin\Omega} {\Delta_1\Delta_2\over\Delta_{1+2}}\exp\left( -\sqrt{\lambda}(\tilde A_1+\tilde A_2)\right). \end{equation} Since we do not know the relation between the factors $\Delta_1$, $\Delta_2$ and $\Delta_{1+2}$, a quantitative test is difficult in this case. Though it seems unlikely that the ration would be zero. It would be very interesting to determine the function $G(z)$ what appears in the linear divergence as it would settle the question as to whether the intersection gives the contribution to $\hat{L} \exp(-\sqrt{\lambda}\tilde{A})$ predicted by the loop equation. \mysection{Discussion} The $AdS$/CFT correspondence allows us to calculate certain Wilson loops in terms of minimal surfaces in anti de-Sitter space. We presented a few reasons why only loops satisfying the constraint $\dot x^2=\dot y^2$ (generically) are given in terms of minimal surfaces. For more general loops we run into the problem of inconsistent boundary conditions. The constrained loops are invariant under half of the local supersymmetry in super loop-space. As such they are BPS objects and are free from divergences. The area of the minimal surface is divergent, so it is not the correct functional that yields the Wilson loop. Since the minimal surface satisfies Neumann boundary conditions, it's natural to take for the action the Legendre transform of the area. We showed this yields a finite result. In other examples of the $AdS$/CFT correspondence the action has to be modified as well. In non-supersymmetric cases, such as the near extremal D$3$-brane, the effect of adding the boundary term is to subtract $L/(2\pi\epsilon)$. The result is finite, but contains a piece proportional to the circumference times the radius of the horizon. This may be considered a mass renormalization of the $W$-boson. The scale of the renormalization is not the UV cutoff, but rather the scale of supersymmetry breaking. In addition, if $\dot x^2 \neq \dot y^2$, the Wilson loop will contain a linear divergence proportional to the UV cutoff. The surface observables on the $M5$ brane theory, as calculated in $AdS_7\times S^4$ have quadratic and logarithmic divergences \cite{Maldacena:1998im,Berenstein:1998ij,Graham:1999pm}. Taking the Legendre transformation will eliminate the quadratic divergence, but we are not sure whether it will also remove the log divergence. Recently there were some attempts to go beyond the classical calculation and include fluctuations of the minimal surfaces \cite{Forste:1999qn,Greensite:1999jw,Naik}. One of the goals was to find the ``L\"uscher term,'' the Coulomb like correction to the linear potential in a confining phase \cite{Luscher:1980fr}. Any attempt to perform such a calculation will require using the correct Neumann boundary conditions on the spherical coordinates, and including the appropriate boundary terms. Finally we formulated the loop equations for those loops, and checked if the $AdS$ ansatz satisfies them. For smooth loops, due to the supersymmetry, the loop equations should give zero. This is indeed the result we find also from the variation of the minimal surface. This calculation actually requires extending the prescription to loops that do not satisfy the constraint. We propose that the natural extension for small deviation from the constraint gives a linear divergence proportional to $\sqrt\lambda L$. This term is particularly important when we consider the loop equations for loops with cusps. The expected result is finite and proportional to $\lambda$. This is in fact what we find, but we do not have enough control over the calculation to compare the coefficients. The situation with self-intersecting loops is more mysterious, we expect a non-zero answer, but cannot reproduce that. There are, however, some reasons why this test is more difficult than the other cases. In particular, the constraint is broken by a large amount at the intersection. Classical string theory tells us only how to calculate loops satisfying the constraint. These are BPS objects in loop space, and therefore easier to control. As we argued, non-BPS Wilson loops are related to excited open strings, but we are unable to evaluate them reliably. A similar statement is true for local operators, one has control only over the chiral operators. Non-chiral operators should be given by excited closed string states. Despite the large effort devoted to testing the Maldacena conjecture, there is still no good understanding of non-BPS objects. \section*{Acknowledgements} We thank Korkut Bardakci, Sunny Itzhaki, Juan Maldacena, Joe Polchinski, Bruno Zumino. N.D. thanks Theory Group of UC Berkeley and LBNL for their hospitality. H.O. thanks the Institute for Theoretical Physics at Santa Barbara, Department of Mathematics and Physics at University of Amsterdam, and Theory Group at CERN, for their hospitality and for providing excellent working environments. The work of H.O. was supported in part by the NSF grant PHY-95-14797 and the DOE grant DE-AC03-76SF00098. This work was supported in part by the NSF under grant No. PHY94-07194.
\section{Introduction} The physics of columnar crystals is relevant to the Abrikosov lattice of flux lines in Type-II superconductors and liquid crystalline materials like concentrated phases of long polymers or discotics. The stability of the columnar crystal has been investigated, and various mechanisms proposed for its melting. Conventional melting, which arises when phonon displacements reach a fixed fraction of the lattice constant, can easily be located via the Lindemann criterion~\cite{Nelson:directed,Jain:ice}. Melting destroys the two-dimensional crystalline order perpendicular to the columns leading to a nematic liquid of lines or columns, which is entangled at sufficiently high densities. Crystal defects play an important role above the melting transition. If edge dislocations in the crystal proliferate, they drive the shear modulus to zero, leading to a liquid-like shear viscosity. However, dislocations alone cannot destroy the six-fold orientational order of the triangular lattice in a two-dimensional cross-section. Thus, provided disclination lines do not also proliferate, the resulting liquid of lines is hexatic, not isotropic~\cite{MarchNel:hexatic}. The screw component of the unbound dislocations leads to entanglement. A finite concentration of unbound disclinations superimposed on the hexatic liquid leads to isotropic in-plane order. Another kind of transition is brought about by vacancy/interstitial line defects in columnar crystals composed of long, continuous lines. As discussed in Ref.~\cite{Frey:defect}, under suitable conditions (such as high field and small interlayer coupling in layered superconductors), it can become favourable for these line defects to proliferate. If this happens at a temperature $T_d$ below the melting temperature $T_m$, then the phase that exists between $T_d$ and $T_m$ will be simultaneously crystalline and highly entangled. In the boson analogy of an aligned system of lines, where the lines represent two-dimensional bosons traveling in the ``time-like'' axial ($\hat{\mathbf z}$) direction~\cite{Nelson:directed}, such a phase is analogous to the supersolid phase of the bosonic system which incorporates vacancies and interstitials in its ground state. This entangled solid melts into an entangled liquid or an entangled hexatic at even higher temperatures. The proliferation of vacancy or interstitial strings could also affect a crystal-to-hexatic transition mediated by dislocations. Dislocations in the columnar crystalline geometry are normally constrained to lie in the vertical plane formed by their Burger's vector and the $\hat{\mathbf z}$-axis, because a dislocation in a two-dimensional cross-section can move along the columnar axis only through glide parallel to its Burger's vector. Transverse motion (climb) would require it to absorb or emit vacancies or interstitials. This becomes possible in the supersolid phase, thus allowing dislocation loops to take on arbitrary non-planar configurations which would have to be included in the treatment of Ref.~\cite{MarchNel:hexatic} to study melting out of a supersolid phase~\cite{MR:supersolid}. Vacancy/interstitial strings in a columnar crystal tend to be lines themselves because of the continuity of the columns. If the columns are constrained to be continuous across the entire sample (as is the case for vortex lines in Type II superconductors), these defects must either thread the entire sample (Fig.~\ref{string}) or appear in vacancy/interstitial pairs forming loops (Fig.~\ref{loop})~\cite{Frey:defect}. The situation is different, however, for finite-length polymers, or columns of discotic liquid crystal molecules which can break and reform freely. As illustrated in Fig.~\ref{polymer}a, a slice through a low temperature configuration in a polymer columnar crystal (with translational order perpendicular to the column axis but not parallel to it) would consist of tightly bound polymer ``heads and tails''. At higher temperatures, however, the heads and tails will separate, either moving apart to form a vacancy string or sliding past each other to form a line of interstitials (Fig.~\ref{polymer}b)~\cite{polymer-defects}. In columnar discotic crystals with similar translational order, ``heads'' and ``tails'' are absent at low temperatures, but appear spontaneously when vacancy and interstitial strings are excited (Fig.~\ref{discotic}). (Head and tail defects appear superficially like dislocations in the cross sections shown in Figs.~\ref{polymer} and~\ref{discotic}. A three-dimensional analysis of lines and columns in neighbouring sheets like that shown in Figs.~\ref{string} and~\ref{loop} is necessary to clearly reveal that these are strings of vacancies and interstitials.) \noindent \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=3.2in \epsfbox{string.eps} \caption{Vacancy string ${\mathbf r}_d(z)$ (thick dashed curve) meandering through a columnar crystal. Dashed lines represent columns just above or below the plane of the figure. (Taken from Ref.~\protect\cite{Frey:defect}.)} \label{string} \end{figure} \end{minipage} \hfill \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=2.8in \epsfbox{loop.eps} \smallskip \caption{Vacancy-interstitial loop in a columnar crystal. Dashed lines represent columns just above or below the plane of the figure. (Taken from Ref.~\protect\cite{Frey:defect}.)} \label{loop} \end{figure} \end{minipage} \medskip \noindent \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=3in \epsfbox{poly.eps} \caption{Formation of vacancy/interstitial strings by sliding of polymers within columns in a columnar crystal of finite-length polymers.} \label{polymer} \end{figure} \end{minipage} \hfill \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=3in \epsfbox{discotic.eps} \bigskip \caption{Formation of vacancy/interstitial strings by sliding of polymers within columns in a columnar crystal of finite-length polymers.} \label{discotic} \end{figure} \end{minipage} \medskip Unlike dislocation lines, these strings (and loops) are not constrained to be planar: the lines can jump to any neighbouring lattice site as they traverse the crystal. Several horizontal jumps connecting a head to a tail are shown in Fig.~\ref{jumps}. Note that \underline{left}ward deflections of the interstitial segment connecting a head to a tail are accompanied by \underline{right}ward deflections of the lines or columns themselves. A typical string can be approximated by an alternating sequence of straight segments and kinks joining the head of one column or polymer chain to the tail of another (see Fig.~\ref{walk}). Vacancy/interstitial strings are suppressed at low temperatures because they have a finite line tension, and hence an energy proportional to their length. At higher temperatures, heads and tails can move apart, forming variable-length strings that wander or ``diffuse'' perpendicular to their length by forming kinks. These strings thus resemble living polymers~\cite{Safran}, except that they are directed, on average, along the $\hat{\mathbf z}$-axis. In polymer crystals, the number of such strings is determined by the fixed concentration of heads and tails. In columnar discotic crystals, heads and tails can be created freely, and it is appropriate to treat their statistical mechanics in a grand canonical ensemble by introducing a head/tail fugacity, similar to the fugacity which controls defect concentrations in theories of vortex or dislocation unbinding transitions~\cite{Nelson:trans}. We assume here that we can treat polymer crystals using the same formalism provided we tune the head/tail fugacity to achieve the fixed concentration determined by the mean polymer length. Long polymers imply a dilute distribution of heads and tails. We exclude, for simplicity, the possibility of hairpin excitations in polymer systems, which can be regarded as doubly quantized interstitial excitations leading to a higher energy. As we shall see, the sharp defect proliferation transition discussed in Ref.~\cite{Frey:defect} is blurred when there is a finite concentration of heads and tails in equilibrium. \noindent \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=2in \epsfbox{jumps.eps} \caption{Illustration of a vacancy string (thick dashed curve) joining a column head to another column's tail in a columnar crystal composed of long-chain polymers.} \label{jumps} \end{figure} \end{minipage} \hfill \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=2.8in \epsfbox{walk.eps} \caption{Schematic of a defect string (composed of straight segments and kinks) wandering through the columnar crystal.} \label{walk} \end{figure} \end{minipage} \medskip Given an appropriate combination of parameters, namely, low line tension combined with head/tail and kink energies comparable to the temperature, the entropy of diffusion of the strings can overcome the line tension and lead to string proliferation, allowing heads and tails to separate to arbitrarily large distances. As in its bosonic counterpart, there exists off-diagonal long-range order in this phase, represented by \begin{equation} \label{entangle} \lim_{|{{\mathbf r}_\perp}'-{\mathbf r}_\perp|\rightarrow \infty} \langle\psi({\mathbf r}_\perp,z)\psi^*({{\mathbf r}_\perp}',z')\rangle \neq 0 \end{equation} where $\psi$ and $\psi^*$ are head and tail ``destruction'' and ``creation'' operators~\cite{Nelson:directed}, implying entanglement of lines on a macroscopic scale. If defects are absent or appear only in closed loops, the expression above would vanish as $|{{\mathbf r}_\perp}' - {\mathbf r}_\perp| \rightarrow \infty$. Once defects proliferate, a line can wander to any other column and Eq.~(\ref{entangle}) has a finite limit. A crystal with proliferating vacancies and interstitials is an incommensurate phase --- the magnitude of the smallest reciprocal vector $G = 4\pi/\sqrt{3}a_0$ is no longer related to the areal density in the obvious way as $\rho = \sqrt{3}G^2/8\pi^2$ because the density differs from its defect-free value $\rho_0 = 2/\sqrt{3}a_0^2$ ($a_0$ being the lattice constant of the triangular lattice in cross-section). All crystals of pointlike atoms or molecules are trivially ``incommensurate'' in this sense --- the corresponding pointlike vacancies and interstitials proliferate at any finite temperature. It is the anomalous suppression of vacancies and interstitials and their organization into lines at low temperatures in columnar crystals which makes these materials unusual. In this paper we apply the physics of directed lines to vacancy/interstitial strings. With this in mind, we briefly review the elasticity theory of these systems in the next Section. In Sec. \ref{single} we model a single string and estimate its transverse wandering. The form of this wandering is unchanged by coupling to phonon distortions of the lattice, as shown in Appendix \ref{bend}. So is its magnitude, as calculated in Appendix \ref{DR}. In Sec. \ref{many} we apply the statistical mechanics of living polymers to an ensemble of directed strings and calculate their volume fraction, average length, etc. in the non-interacting limit. A simple quadratic-interaction model is presented in Section \ref{int}, similar to the one discussed via the boson mapping in Ref.~\cite{Nelson:directed}, and we reproduce the results therein. Numerical calculations of the line tensions of various species of defects are presented in Sec. \ref{num}. The interaction potentials considered are repulsive and monotonic; we study simple power laws as well as a screened Debye-H\"{u}ckel interaction. We find many metastable species of vacancies. However, the lowest energy defect is always found to be the one with the highest symmetry in its category. For very short range interactions, this is the symmetric vacancy ($V_6$), whereas for most interactions the centered interstitial ($I_3$) is most favoured. Appendix \ref{Ewald} contains details of the Ewald summation calculations for the potentials considered here. \section{Review of elasticity theory} \label{theory} Before discussing defects in a columnar crystal, we review the aspects of elasticity theory common to all the systems mentioned in the Introduction. We consider lines or columns aligned along a common direction ($\hat{\mathbf z}$) up to thermal fluctuations, with crystalline order in any cross-section perpendicular to the columnar axis. In the case of flux lines, the average direction of alignment is imposed by an external field (${\mathbf H}=H \hat{\mathbf z}$) and local deviations from this direction cost energy. With columnar crystals of long-chain molecules composed of covalently bonded nematogens or disk-shaped molecules cylindrically stacked via hydrogen bonds, or amphiphilic molecules in cylindrical micellar aggregates, the columnar axis represents spontaneously broken rotational symmetry. Therefore local deviations from the alignment direction are not penalized, but undulations of the column are. The rotational symmetry can, however, be broken by imposing an external field. In addition, the two-dimensional crystalline order resists shear and areal deformations perpendicular to the $\hat{\mathbf z}$-axis. Low-energy fluctuations of the system can be described by a ``continuum'' model that works for small amplitude, long-wavelength deformations~\cite{SelB,Nelson:directed,deGennes}. The important fluctuations in this limit can be characterized by a two-dimensional displacement field ${\mathbf u}({\mathbf r}_\perp,z)$, representing the average deviation of lines in the ($x,y$) plane in a small region centered at $({\mathbf r}_\perp,z)$. With it can be associated a local areal density change $\delta \rho/\rho_0 = - {\mathbf \nabla}_\perp \cdot {\mathbf u}$ ($\rho_0 = 2/\sqrt{3} a_0^2$) and a local nematic director ${\hat{\mathbf n}} = {\hat{\mathbf z}} + {\mathbf t}$, with ${\mathbf t} \equiv \partial{\mathbf u}/\partial z$. The free energy of the system is a sum of nematic and crystalline contributions: \begin{equation} \label{Ftot} {\mathcal F} = {\mathcal F}_{nematic} + {\mathcal F}_{crystal}, \end{equation} To the lowest order in the fluctuations, these are given by \begin{equation} {\mathcal F}_{nematic} = \frac{1}{2} \int\!\! d^{3}r \left[K_1 ({\mathbf \nabla}_\perp \cdot {\mathbf t})^2 + K_2 ({\mathbf \nabla}_\perp \times {\mathbf t})^2 + K_3 (\partial_z {\mathbf t})^2 \right] \end{equation} and \begin{equation} \label{Fxtal} {\mathcal F}_{crystal} = \int\!\! dz\! \int\!\! d^2{\mathbf r}_\perp \left[\mu\,{u_{i j}}^2 + \frac{1}{2} \lambda\,\left(\frac{\delta\rho}{\rho_0}\right)^2 \right] \end{equation} where $K_1, K_2, K_3$ are the Frank constants for splay, twist and bend respectively, and $\lambda$ and $\mu$ are the Lam\'{e} coefficients. The matrix $u_{ij} = (\partial_i u_j + \partial_j u_i)/2$ is the linearized 2D strain field. In the presence of an external field $H {\hat{\mathbf z}}$, one should add to ${\mathcal F}$: \begin{equation} \label{ext} {\mathcal F}_{ext} = \frac{1}{2} \chi_a H^2 \int\!\! dz \int\!\! d^2r_\perp |{\mathbf t}|^2 , \end{equation} where $\chi_a$ is the anisotropic part of the susceptibility~\cite{deGennes}. The last two contributions to ${\mathcal F}$ are quadratic in the derivatives, and can be rewritten as \begin{eqnarray} {\mathcal F}_{crystal} + {\mathcal F}_{ext} = & \frac{1}{2} \int\!\! d^{3}r \left[c_{11} ({\mathbf \nabla}_\perp \cdot {\mathbf u})^2 + c_{66} ({\mathbf \nabla}_\perp \times {\mathbf u})^2 + c_{44} (\partial_z{\mathbf u})^2 \right] \nonumber\\ & \mbox{ } + \mu \;(\mbox{surface terms}) \end{eqnarray} where $c_{11} \equiv \lambda + 2 \mu$, $c_{66} \equiv \mu$, and $c_{44} \equiv \chi_a H^2 \rho$. The surface terms become important when there are defects within the bulk of the crystal, like vacancy/interstitial strings, represented by cuts joining column-end singularities in the field ${\mathbf u}({\mathbf r}_\perp,z)$. Evaluating these terms over a cylindrical surface enclosing such a string yields the energy cost of the defect string: a line tension $\tau_z \approx \mu a^2$ due to the elastic distortion around the string, in addition to a core energy $E_c$ per unit length (of the same order of magnitude) within the cylindrical core. ${\mathcal F}_{nematic}$ can be further simplified if, as is often the case with nematic polymers, the splay and twist constants are small in comparison to the bend constant. Specifically, if $K_1$ and $K_2$ satisfy $K_{1,2} a_0^{-1}/\sqrt{K_3 c_{11}} \ll 1$~\cite{Jain:ice}, then they can be neglected. For long-wavelength distortions along the columnar axis, the dominant free energy contribution is then $K_3 (\partial_z^2 {\mathbf u})^2$ in the absence of an external field. $K_3$ can be simply related to the persistence length $l_P$ of the polymer as $K_3 = k_B T l_P \rho$. \noindent \begin{minipage}{2in} \begin{figure} \centering \leavevmode \epsfxsize=2in \epsfbox{distort.eps} \caption{Distortion induced by a column end in the neighbouring columnar crystalline matrix. The distortion is confined to a vertical extent $|z| < \sqrt{\lambda_L r_\perp}$ (shaded region) around the column end.} \label{distort} \end{figure} \end{minipage} \hfill \newlength{\textw} \setlength{\textw}{\textwidth} \addtolength{\textw}{-2.3in} \begin{minipage}{\textw} The statistical mechanics of defects in polymer liquid crystals has been discussd in detail by Selinger \& Bruinsma~\cite{SelB:defect,Meyer:defect}. The presence of defects imposes a deformation on the $T=0$ equilibrium configuration. In the case of a semi-infinite vacancy/interstitial string with a head or tail at the origin, this distortion follows from minimization of the free energy above with respect to ${\mathbf u}({\mathbf r}_\perp,z)$ under the constraint \begin{equation} \label{constraint} {\mathbf \nabla}_\perp \cdot {\mathbf u} = \pm \rho_0^{-1} \delta({\mathbf r}_\perp) \theta(z) + (\mbox{non-singular terms}) \end{equation} where the $\pm$ sign refers to a column tail/head located at the origin. Since the planar distortion about a string has azimuthal symmetry in the continuum approximation, ${\mathbf \nabla}_\perp \times {\mathbf u} = 0$. Hence, the only relevant terms in the free energy are the bend and bulk distortion terms (neglecting splay). The resulting distortion around the column end spans a parabolic region about the radial direction (see Fig.~\ref{distort}) defined by \begin{equation} \label{lambda} z^2 \lesssim \lambda_L r_\perp \end{equation} where $\lambda_L = \sqrt{K_3/c_{11}}$ is the length scale relating the distortions parallel and perpendicular to ${\hat{\mathbf z}}$. \end{minipage} \medskip Selinger \& Bruinsma also calculate the interaction energy between two column ends by superimposing the distortion created by each. They find the interesting result that a head and tail in a \textit{nematic} medium attract weakly if they fall within each other's region of influence, as just described, but repel otherwise. However, in a columnar crystal (with non-zero shear modulus), the interaction is always a strong attractive linear potential due to the finite line tension associated with the string of distortions joining a head to a tail. \section{Wandering of a single string} \label{single} Consider a single vacancy/interstitial string in a hexagonal columnar crystal of, say, polymer strands with lattice constant $a_0$ and monomer spacing $c$ along the columnar axis $\hat{\mathbf z}$. For a discotic columnar liquid crystal, $c$ is the spacing between oblate molecules along the column axis. For a flux line in a layered Type-II superconductor with magnetic field perpendicular to the layers, $c$ is the layer spacing. If the string is vertical, the energy per unit length $\tau_z$ is of the order of $\mu a_0^2$ (see Section \ref{theory}) where $\mu$ is the in-plane shear modulus of the crystal. For a horizontal string, $\tau_\perp = \varepsilon_k/a_0$ where the kink energy $\varepsilon_k \sim \kappa^{1/4} \mu^{3/4} a_0^2$~\cite{Nelson:directed}, $\kappa \equiv K_3/\rho$ being the bending rigidity. The ratio is $\tau_\perp/\tau_z \sim (\kappa/\mu)^{1/4}/a_0 \sim l^*/a$ where $l^*$ is the kink size. Typically $l^* \gg a_0$, so that the strings are predominantly vertical, with few kinks. For flux lines on the other hand, the kink energy is $g^{1/2} \mu^{1/2} a_0$ with $g \equiv c_{44}/\rho$, where $c_{44}$ is the tilt modulus and $\rho$ is the areal line density. The ratio is then $(g/\mu)^{1/2}/a_0$. In highly anisotropic layered superconductors, this ratio can be small, favouring large, nearly horizontal defect excursions. We will for now work with nearly vertical strings, allowing for a gas of kinks sufficiently dilute so that the interaction between kinks can be ignored (see Fig.~\ref{walk}). We thus assign to a string of vertical extent $l$ and $n_k$ kinks an energy $l \tau + n_k \varepsilon_k + 2 \varepsilon_0$ where $\tau \equiv \tau_z$ and $\varepsilon_0$ is the energy of a polymer end. We expect that the results for defects with a high density of kinks would be qualitatively similar. In units such that $k_B=1$, the partition function of a string of length $l$ is \begin{equation} {\mathcal Z}_1 = (1 + q e^{-\varepsilon_k/T})^{l/l^*} e^{-l \tau/T} \end{equation} where $T$ is the temperature, and $q$ is the two-dimensional co-ordination number of the lattice on which the defect string lives --- for a symmetric vacancy this is the same as that of the original triangular lattice, $q = 6$, whereas for a symmetric interstitial it is that of the dual honeycomb lattice, $q = 3$ (see Section \ref{num}). The above expression represents the freedom of the string to jump to any of the neighbouring lattice sites anywhere along its length. These transverse meanderings cause an entropic lowering of the free energy per unit length of the string: \begin{eqnarray} f_1 &=& \lim_{l \rightarrow \infty} -T \ln{{\mathcal Z}_1} /l \nonumber \\ &=& \tau - \frac{T}{l^*} \ln{\left(1 + q e^{-\varepsilon_k/T}\right)} \nonumber \\ &\simeq& \tau - \frac{T q}{l^*} e^{-\varepsilon_k/T} \quad\mbox{for}\quad e^{-\varepsilon_k/T} \ll 1 \end{eqnarray} If $N_k$ is the total number of kinks, the average kink density is \begin{eqnarray} n_k \equiv \frac{\langle{N_k}\rangle}{l} &=& \frac{1}{l^*} \frac{q e^{\varepsilon_k/T}}{1+q e^{\varepsilon_k/T}} \nonumber \\ &\simeq& \frac{q}{l^*} e^{-\varepsilon_k/T}, \quad\mbox{for}\quad e^{-\varepsilon_k/T} \ll 1. \end{eqnarray} Thus, kinks are on the average $l_k = l^* e^{\varepsilon_k/T}/q$ monomers apart. The assumption of dilute kinks then translates into the condition $l^* n_k \ll 1$, or, $\varepsilon_k \gg T$, which can be rephrased as $\langle |{\mathbf u}|^2 \rangle/a_0^2 \ll 1$~\cite{Nelson:directed,Jain:ice}, a condition clearly satisfied by a crystal below its Lindemann melting point. The above is a ``diffusive'' model for the string --- if ${\mathbf d}$ denotes the horizontal end-to-end displacement, the mean square wandering is $\langle|{\mathbf d}|^2\rangle = 2 D l$, where the ``diffusion constant'' $D$ is given by $2 D = a_0^2 n_k$. Consider a continuum description of the string in terms of a function ${\mathbf r}_d(z)$, ${\mathbf r}_d(z)$ being the transverse displacement. Provided the average slope $|d{\mathbf r}_d/dz|$ is small, this ``diffusive'' wandering would correspond to an effective Hamiltonian of the form \begin{equation} \label{defEg} H_1 = \int_0^l dz \left[\frac{g}{2} \left|\frac{d{\mathbf r}_d}{dz}\right|^2 + \tau\right], \quad g=\frac{T}{D} \end{equation} Here we have assumed that the string is wandering within a frozen crystal. However, the lattice around the vacancy/interstitial string responds to its presence by collapsing or expanding around it. For a straight string at ${\mathbf r}_d = {\mathbf 0}$, the deformation ${\mathbf u}({\mathbf r}_\perp,z)$ is given by \begin{equation} {\mathbf u}_d({\mathbf r}_\perp,z) = \pm \frac{\Omega}{2 \pi} \frac{{\mathbf r}_\perp}{r_\perp^2} \end{equation} in the continuum description of the crystal, that is, away from the defect where the deformations are small. $\Omega$ is the area change due to the vacancy/interstitial, $\Omega \simeq a_0^2$. The energy of this deformation has to be included in the energy cost of the defect string. Again invoking the continuum approximation, we assume that for a defect string with small average slope, the resulting deformation away from the string in any plane perpendicular to $\hat{\mathbf z}$ would be approximately that resulting from a straight string at the location of the defect in that plane: \begin{equation} {\mathbf u}({\mathbf r}_\perp,z) \simeq {\mathbf u}_d({\mathbf r}_\perp-{\mathbf r}_d(z),z). \end{equation} (In general ${\mathbf u}({\mathbf r}_\perp,z)$ would depend on the derivatives of ${\mathbf r}_d(z)$ as well.) Within this approximation, the distortion energy of the crystal with bending Frank's constant $K_3 \equiv T l_P \rho$ is, keeping terms up to fourth-order in the derivatives (see Appendix \ref{bend}): \begin{equation} \label{defE4} \frac{\Delta H_1}{T} \sim l_P \int dz \left[ \left|\frac{d^2 {\mathbf r}_d}{dz^2}\right|^2, a_0^{-2} \left|\frac{d{\mathbf r}_d}{dz}\right|^4 \right] \end{equation} These impart an effective stiffness to the defect string and suppress transverse fluctuations over a length scale $\sim a_0\sqrt{D K_3/T} \sim a_0\sqrt{l_P n_k}$. However, they do not change the long scale diffusive nature of the string. The lattice distortions renormalize the diffusion constant of the string when the symmetry direction of the crystal is externally imposed, as in the case of flux lines, or in a polymer crystal with an external field along the $\hat{\mathbf z}$-direction. The tilt modulus $c_{44}$ is then non-zero (Eq.~(\ref{ext})), and D is renormalized to $D_R$, where (see Appendix \ref{DR}) \begin{equation} \frac{1}{D_R} \simeq \frac{1}{D} + {\mathcal O}\left(\frac{c_{44}}{T \rho}\right) \end{equation} For a dense vortex \textit{liquid} this effect has been analyzed in detail by Marchetti~\cite{Marchetti:D} and $D$ is found to be renormalized to a value independent of its bare value in the long-wavelength limit. The correction comes from convection of a tagged flux line along the local tangent-field direction. If a similar calculation is carried out for a \textit{crystal} of spontaneously aligned long semi-flexible polymers (see Appendix \ref{DR}), one finds a qualitatively different renormalization of $D$ --- the correction in the long-wavelength limit is proportional to its bare value, and $\delta D/D \sim 1.45 \langle|{\mathbf u}|^2\rangle/a_0^2 \lesssim 3~\%$ using $c_L^2 \simeq 1/50$~\cite{cl} ($c_L$ is the Lindemann constant for melting of a columnar crystal). The correction is negligible. It can be ignored for another reason --- the idea of convection of a line by the mean local field, although appropriate for a dense fluid, would not be applicable in a crystalline environment where diffusion can only occur through discrete jumps from column to column. Although thermal fluctuations are already implicit in the exponential factor in $D = a_0^2 n_k/2$ coming from $n_k$, defects in this case move only on a discrete lattice, without phonon fluctuations. To summarize this section, we characterize the statistical mechanics of a defect string with a head/tail energy $\varepsilon_0$, a line tension $\tau$, and a diffusion constant $D$. The latter two can be combined in an effective chemical potential $\overline{\mu} \equiv T \mu_d$ per kink size ($l^*$) of the string: \begin{equation} \mu_d = l^* (-\tau/T + n_k) = q e^{-\varepsilon_k/T} - \varepsilon_k/T, \end{equation} with $n_k$ related to $D$ through $D = a_0^2 n_k/2$. Because $n_k$ is exponentially small, $\mu_d \approx -l^* \tau/T \approx -l^* \mu a_0^2/T$ and is usually negative, which suppresses long vacancy \& interstitial strings. Turning it positive would require raising the temperature and lowering the kink energy $\varepsilon_k$, and is favored by a larger co-ordination number $q$. Although we have assumed a constant shear modulus, the presence of the defects themselves can drive it down exponentially with the defect concentration, as discussed by Carruzzo \& Yu~\cite{CarYu:shear}. Thus, positive $\mu_d$ becomes possible when softening of the bare elastic constants with increasing defect concentration is taken into account. \section{Statistical Mechanics of non-interacting strings} \label{many} At any finite temperature, a crystal with a negative string line-chemical potential will contain a distribution of thermally excited vacancy and interstitial strings. Since the string energy is proportional to length in the non-interacting-kinks approximation, the equilibrium probability distribution would be an exponentially decaying function of length with mean determined by the line chemical potential, in the dilute string-gas limit where inter-string interactions can also be neglected~\cite{Safran}. In discotic crystals string heads and tails can be created as necessary. In a crystal of long polymers, the number of heads and tails is fixed by the mean polymer length. Let N be the total number of possible kink sites in the lattice, $N = \mbox{volume} \times \rho/l^*$, and ${\mathcal P}_l$ be $1/N \times$ the number of defect strings $l$-links long. Assuming that only one kind of defect string is present --- those with the lowest line tension --- we can write the defect free energy in terms of $\{{\mathcal P}_l\}$ as~\cite{Safran} \begin{equation} \label{F} {\mathcal F}_d(\{{\mathcal P}_l\}) = \sum_l N {\mathcal P}_l (2 \varepsilon_0 - l T \mu_d) + T \sum_l N {\mathcal P}_l (\ln{{\mathcal P}_l} - 1) \end{equation} Minimizing with respect to the $\{{\mathcal P}_l\}$ yields the expected exponential distribution: \begin{equation} {\mathcal P}_l = h^2 z^l \end{equation} where $z = e^{\mu_d}$, and the head/tail fugacity $h = e^{-\varepsilon_0/T}$ is expected to be small. For hexagonal columnar crystals of \textit{polymers}, we work in a grand canonical ensemble and adjust $\varepsilon_0$ so that the average head/tail concentration agrees with the fixed value determined by the mean polymer length. The head/tail concentration will be small if the polymers are long. For \textit{discotic} crystals, the grand canonical ensemble is the natural one and the head/tail concentration fluctuates, with an average value determined by the fixed value of $h = e^{-\varepsilon_0/T}$, and the monomer fugacity $z = e^{\mu_d} < 1$. The net defect volume fraction $\phi$ is \begin{equation} \phi \equiv \sum_l l {\mathcal P}_l = h^2 \frac{z}{(1-z)^2} . \end{equation} The total number of strings $N_d \equiv N n_s$ is given by the string density \begin{equation} n_s \equiv \sum_l {\mathcal P}_l = \frac{h^2}{1-z} \end{equation} A defect monomer is most likely to be found in a string of mean length (in units of the kink size) \begin{equation} l_m = \frac{1}{|\mu_d|} \end{equation} The length distribution has an average at $2 l_m$, and a spread also of $\sqrt{2} l_m$. The form (\ref{F}) of the energy, linear in $l$, is really applicable only when $l \gg 1$, so that end effects can be parametrized by the $l$-independent constant $\varepsilon_0$. Then, $\mu_d$ is close to $0$, and the relation $\phi \simeq n_s l_{mp}$ holds. The asymptotic behaviours in the dilute and dense limits are as follows: \begin{eqnarray} \phi &=& \left\{\begin{array}{ll} h^2 e^{\mu_d}, & z \ll 1 \\ \frac{h^2}{|\mu_d|^2}, & z \lesssim 1 \end{array}\right. \\ n_s &=& \left\{\begin{array}{ll} h^2, & z \ll 1 \\ \frac{h^2}{|\mu_d|}, & z \lesssim 1 \end{array}\right. \end{eqnarray} A string proliferation transition thus occurs at $\mu_d = 0$ in this model, corresponding to a temperature $T_d = \tau l_k$. In the limit $\varepsilon_0 \rightarrow \infty$, it corresponds to the appearance of a supersolid phase~\cite{Frey:defect} which is simultaneously crystalline and entangled, where infinitely long vacancy/interstitial strings facilitate the wandering and entanglement of lines in the crystalline phase. If the melting temperature $T_m > T_d$, this supersolid/incommensurate solid phase will exist between $T_d$ and $T_m$. The non-interacting approximation breaks down in the vicinity of $T_d$ as calculated here, and its estimate will have to be refined by including interactions. For finite $\varepsilon_0$, the sharp transition discussed in Ref. ~\cite{Frey:defect} will be blurred, as discussed in Sec. \ref{int}. \section{$\phi^2$-interaction model} \label{int} Interactions between polymer ends in a columnar crystal have been calculated by Selinger \& Bruinsma~\cite{SelB:defect} within the continuum approximation. Because of the uniaxial anisotropy, the interaction has a rather complicated form. The distortion due to an isolated head or tail placed at the origin at in-plane distance $r_\perp$ extends over a vertical extent $|z| \sim \sqrt{\lambda_L r_\perp}$ where $\lambda_L = \sqrt{K_3/c_{11}}$ (see Eq.~(\ref{lambda})). The resulting interaction between heads and tails falls as $1/|z|^3$ for predominantly vertical separations $z$ ($|z| \gg \sqrt{\lambda_L r_\perp}$), and as $-1/(\lambda_L r_\perp)^{3/2}$ for predominantly horizontal separations $r_\perp$. In polymer crystals, these contributions must be superimposed on the linear energy cost of the vacancy or interstitial string joining them. At low defect densities where the string length is much smaller than the average separation of string centers of mass, we have $1/|\mu_d| \ll 1/\phi^{1/3}$, i.e., $|\mu_d| \gg h^{2/3}$, and a string interacts with other strings as a head-tail dipole. The effective interaction between dipoles then falls off very rapidly, becoming short-ranged not only in the axial, but also in the radial direction. At the other extreme, the strings are long, which would happen in the vicinity of the head-tail unbinding transition and in the supersolid phase itself. End-interactions can then be neglected and the remaining interaction between effectively infinite strings becomes predominantly ``radial'' (i.e., perpendicular to $\hat{\mathbf z}$) provided the root mean square tilt with respect to the $\hat{\mathbf z}$ axis is small. The defects are then non-interacting in the continuum model unless their anisotropy is taken into account. The interaction between defects with n-fold symmetry (n = 2, 3 or 6) falls off at least as fast as $1/r^n$ (see Appendix \ref{n-def}). This interaction has an azimuthal dependence of the form $\cos{n \theta}$ or higher harmonics. The angular average vanishes, leading to an effective interaction which vanishes as an even higher power which is effectively short-ranged. As mentioned in the Introduction, the lowest-energy vacancy or interstitial defects for simple repulsive pair potentials in the radial direction are in fact of high (three-fold or six-fold) symmetry. We discuss here the simplest model for a short-ranged interaction --- a repulsive $\phi^2$ model that has been treated earlier in Ref.~\cite{Nelson:directed} using a coherent state path integral representation which exploits an analogy with the quantum mechanics of two-dimensional bosons. The defect volume fraction $\phi$ corresponds to the mean square boson field amplitude $\langle|\psi|^2\rangle$ in that description. Here, we reproduce the essential results without resorting to the sophisticated boson formalism. Upon adding a term $u \phi^2 /2$ to the free energy $f \equiv F/N T$ in Eq.~(\ref{F}) of the previous section, we find after minimization, \begin{equation} {\mathcal P}_l = h^2 e^{l (\mu_d - u \phi)} . \end{equation} As discussed in Ref.~\cite{Nelson:directed}, the coupling $u$ is an excluded volume parameter describing defect line repulsion. Thus $\phi$ and $N_d$ have the same form as before, but with $z$ replaced by an effective fugacity $\zeta$: \begin{equation} \label{zeta} z \rightarrow \zeta(z,\phi) \equiv z e^{-u \phi} , \end{equation} so that \begin{equation} \label{phi-zeta} \phi(h,\zeta) = h^2 \frac{\zeta}{(1-\zeta)^2} . \end{equation} The volume fraction $\phi(h,z)$ now has to be solved for self-consistently from Eq.~(\ref{phi-zeta}). Note that the effective chemical potential has been reduced by $u \phi$ due to the repulsive interaction: \begin{equation} \mu_{eff} \equiv \ln{\zeta} = \mu_d - u \phi \end{equation} Accordingly, the mean string length $l_m$ changes to \begin{equation} l_m = -\frac{1}{\ln{\zeta}} \equiv \frac{1}{u \phi - \mu_d} . \end{equation} The free energy of the distribution is $f \approx -u \phi^2 /2$. The behaviour of the string volume fraction for $h = 0$ and $h \neq 0$ is illustrated schematically in Fig.~\ref{phimu}. Four distinct regimes emerge, with the following asymptotic behaviours: \begin{figure} \centering \leavevmode \epsfxsize=6.1in \epsfbox{phimu.eps} \caption{The volume fraction $\phi$ is plotted against the effective defect chemical potential $\mu_d$ for the $\phi^2$-interaction model of a gas of defect strings. The strings are short and dilute in regime A, but long, dense and entangled in regime B. (Taken from Ref.~\protect\cite{Nelson:directed}.)} \label{phimu} \end{figure} \begin{enumerate} \item $\mu_d \ll -1$ (point A in Fig.~\ref{phimu}): \begin{equation} \phi \simeq h^2 e^{\mu_d},\quad n_s \simeq h^2,\quad l_m = \frac{1}{|\mu_d|} . \end{equation} This is again the dilute limit where heads and tails are tightly bound. \item $-1 \ll \mu_d \ll -(u h^2)^{1/3}$: \begin{equation} \label{noni} \phi \simeq \frac{h^2}{|\mu_d|^2},\quad n_s \simeq \frac{h^2}{|\mu_d|},\quad l_m = \frac{1}{|\mu_d|} . \end{equation} These results are again identical to those for non-interacting strings. This correspondence is expected, because $|\mu_d| > (u h^2)^{1/3} > u \phi$, therefore the effective chemical potential is still approximately $\mu_d$. The relation $\mu_d \sim -(u h^2)^{1/3}$ marks the limit of validity of the non-interacting approximation, as we argued in the beginning of this section. As we approach this limit, we find for $h \rightarrow 0$: $\phi, n_s \rightarrow 0$, whereas $l_m \rightarrow \infty$. Thus, the strings are still dilute, although lengthening. Note that the results in this regime coincide with those of Ref.~\cite{Nelson:directed} in the short and dilute strings limit. \item $|\mu_d| \ll (u h^2)^{1/3} \equiv \mu_c$ ($\mu_d$ around the transition which occurs for $h = 0$): \begin{equation} \phi \simeq \frac{h^2}{|\mu_c|^2} \left[ 1 + \frac{2}{3}\frac{\mu_d}{\mu_c}\right],\quad n_s \simeq \frac{h^2}{|\mu_c|} \left[ 1 + \frac{1}{3}\frac{\mu_d}{\mu_c}\right],\quad l_m \simeq \frac{1}{|\mu_c|} \left[ 1 + \frac{1}{3}\frac{\mu_d}{\mu_c}\right] . \end{equation} These results can be matched onto those in the non-interacting regime above by replacing $\mu_d$ with \begin{equation} \mu_{eff} = -\mu_c + \mu_d/3 = -\mu_c \left(1-\frac{\mu_d}{3 \mu_c}\right) , \end{equation} which is now dominated by the repulsive interaction: $\mu_{eff} \approx - u \phi$. The unphysical divergences of the non-interacting model have been suppressed and we find at the transition point: \begin{equation} \phi = \frac{h^{2/3}}{u^{4/3}},\quad n_s = \frac{h^{4/3}}{u^{1/3}},\quad l_m = \frac{1}{u^{1/3} h^{2/3}} . \end{equation} Note that all quantities have interesting singularities in the limit $h \rightarrow 0$. If the head/tail fugacity $h$ is small, the defect volume fraction remains negligible at the transition, but the average string length grows large so that it could become greater than the inter-string separation, now given by $1/\phi^{1/2}$. Indeed, $1/\phi^{1/2} \ll l_m$ if $h \ll 1/u^2$ which would be true if polymer ends are highly unfavourable. This long \& dilute regime interpolates between the short \& dilute and the long \& dense limits described in Ref.~\cite{Nelson:directed}. \item $\mu_d \gg \mu_c$ (Point B in Fig.~\ref{phimu}):\\ In this limit, we have \begin{equation} \mu_{eff} = -\mu_c \sqrt{\frac{\mu_c}{\mu_d}} . \end{equation} The repulsion now keeps in check the string proliferation, and $\mu_{eff}$ approaches $0$ as $1/\sqrt{\mu_d}$. Thus, \begin{equation} \phi \simeq \frac{\mu_d}{u},\quad n_s \simeq h \sqrt{\frac{\mu_d}{u}},\quad l_m \simeq \frac{1}{|\mu_c|} \sqrt{\frac{\mu_d}{\mu_c}} . \end{equation} This is the phase where strings are dense and entangled --- $\phi$ is ${\mathcal O}(1)$. These results also agree with Ref.~\cite{Nelson:directed}. \end{enumerate} As the head/tail fugacity $h \rightarrow 0$, the intermediate regime~3 above (around $\mu = 0$) shrinks to zero. At $h = 0$, heads/tails are completely expelled, and we have a second-order phase transition at $\mu_d = 0$ with $\phi = 0$ for $\mu_d < 0$, and growing as $\mu_d$ for $\mu_d > 0$, as in Ref.~\cite{Nelson:directed}. This limit corresponds to the situation in thermally excited vortex lattices~\cite{Frey:defect} because flux lines cannot start or stop within the sample. In the boson picture, $h$ acts like an external field coupled to the order parameter, injecting magnetic monopoles into the superconductor. We have neglected vacancy/interstitial loops, which exist even in the limit $h \rightarrow 0$. For finite $h$, their contribution can be neglected near the transition because for long loops, the energy of a loop exceeds the energy of a string of the same vertical extent: Whereas a string of length $l$ has energy $l \tau_{interstitial} + 2 \varepsilon_0$ (we expect interstitials to be the preferred defect at the transition in most cases), the energy of a vacancy-interstitial loop of the same length would be approximately $l\: (\tau_{vacancy} + \tau_{interstitial})$. For large $l$, the difference $l \tau_{vacancy} - 2 \varepsilon_0$ will strongly suppress vacancy/interstitial loops. Because of this energetic barrier, loops cannot become arbitrarily large, and cannot cause entanglement over macroscopic scales. For $h = 0$, as is the case for vortex matter, fluctuations in the low temperature phase are entirely in the form of loops~\cite{Frey:defect}, and similar to vortex ring fluctuations in the Meissner phase. For systems with a finite axial length, the balance may be tilted in favour of long strings because the end penalty is removed if the ends move to the surface and the string threads the sample. For threading strings the expression for entropy in Eq.~(\ref{F}) is no longer valid because the freedom in the z-direction is lost. The remaining two-dimensional entropy can be ignored in a three-dimensional system, and we are left with \begin{equation} f \simeq -\mu_d \phi + u \phi^2 /2 \end{equation} where $\phi$ now is also the areal fraction of defects; and one finds $\phi \simeq \mu_d/u$, similar to region~4 discussed above. \section{Numerical calculation of defect line tensions} \label{num} Line tension calculations require that we find the lowest energy lattice deformation associated with a vacancy or interstitial. These line tensions depend on the \textit{type} of vacancy or interstitial, e.g., whether the defect sits in an environment which is two-, three- or six-fold symmetric. If thermal fluctuations out of this configuration are small enough to be described within a quadratic approximation, they decouple from the equilibrium configuration. Since these $T=0$ equilibrium defect configurations are composed of straight columns, the 3-dimensional deformation energy can be reduced to an effective 2-dimensional interaction energy $V(r)$ per unit length between columns separated by distance $r$. The calculations can then be performed on a two-dimensional triangular lattice of points interacting with potential $V(r)$. Thus, the defect energies in a two-dimensional Wigner crystal of electrons~\cite{Morf:defect} would correspond to the \textit{line tensions} of the corresponding string defects in a hexagonal columnar crystal of lines interacting with an effective radial $1/r$-potential per unit length. Such calculations have been carried out by several authors~\cite{Frey:defect,Morf:defect,CockEl:defect}. Whereas Refs.~\cite{Morf:defect} and~\cite{CockEl:defect} have considered defects in a Wigner crystal of electrons ($V_p(r)=1/r$), Frey \textit{et al.}~\cite{Frey:defect} have studied a modified Bessel-function potential $V_{\kappa}(r) = u_0 K_0(\kappa r)$ in the $\kappa \rightarrow 0$ limit. Here $\kappa \equiv \lambda^{-1}$, where $\lambda$ is the Debye screening length in the case of long polyelectrolytes in an ionic solution, and the London penetration depth in the case of vortex lines in a type-II superconductor. The limit $\kappa \rightarrow 0$ corresponds to a long-range logarithmic interaction, whereas in the short-range limit $\kappa a_0 \gg 1$ the interaction is exponentially decaying. Both Refs.~\cite{Frey:defect} and~\cite{CockEl:defect} dealt with long-range interactions ($\ln{r}$ and $1/r$ repectively), and found that the centered interstitial (see Fig.~\ref{defects}) has the lowest line tension. We denote the centered interstitial by $CI$, or by $I_3$ when we want to stress its three-fold symmetry. The edge interstitial (denoted $EI$ or $I_2$) was found to be a saddle-point and buckled into a $CI$. The three-fold symmetric centered interstitial $CI$ is the lowest energy interstitial defect over the entire range of interactions we studied. Among the vacancies, the two-fold symmetric crushed vacancy (denoted $V_2$ or $V_{2a}$ --- see Fig.~\ref{defects}) is the only stable one, the symmetric six-fold vacancy ($V_6$) being unstable to it. The long-range interactions between the energetically preferred types of interstitials and vacancies were found to be attractive for interstitials and repulsive for vacancies. To determine the correct type of microscopic defect to insert into the phenomenological considerations of Secs. \ref{single}--\ref{int}, we have extended the work of Frey \textit{et al.} to the short-ranged regime of the $K_0(\kappa r)$-interaction, to which end we studied values of $\kappa a_0$ from $0$ to $7$ ($7$ being large enough to represent the short-range $\kappa a_0 \rightarrow \infty$ limit) (Fig.~\ref{bes:ek}). The aim was to determine the point of cross-over from centered interstitials to vacancies as the lowest-energy defect, since it is known from simulations of short-range interactions (for a review, see Ref.~\cite{point-defects}) that vacancies are preferred in this limit. In the same spirit, we have also extended the Coulomb interaction to power-law interactions $1/r^p$ with exponent values ranging from $p = 0$ ($\sim \ln{r}$) to $p = 12$ (Fig.~\ref{gam:ep}). We checked our minimization procedure by first reproducing the results of Refs.~\cite{Frey:defect} and~\cite{CockEl:defect} for $\ln{r}$ and $1/r$ potentials respectively. As we move away from the long-range interaction limit $\kappa a = 0$, the metastable crushed vacancy ($V_{2a}$) exchanges stability with the metastable split vacancy ($SV$), also of two-fold symmetry. Two metastable species, a three-fold symmetric vacancy ($V_3$) and a two-fold symmetric vacancy ($V_{2b}$) crushed along the basis vector of a triangular unit cell, also exist, but are of higher energy. The differences in energy can be as small as one part in a few thousand. As the interaction gets shorter-ranged, $V_{2b}$ loses stability to $V_3$ at $\kappa a_0 \simeq 5.2$, and the 3-fold deformation of $V_3$ gets smaller so that it transforms continuously into $V_6$ at $\kappa a_0 \simeq 5.9$. When $V_6$ appears, the $SV$ also loses stability to it. By the time $I_3$ and $V_6$ finally cross in energy, $V_6$ is the only stable vacancy left. The crossing happens at surprisingly large parameter values, $\kappa a_0 \simeq 6.9$ for $V_{\kappa a}$ (Fig.~\ref{bes:diff}), and $p \simeq 5.9$ for $V_p$ (Fig.~\ref{gam:diff}), each very close to the short-range limit. We thus find that the interstitial has a very wide range of stability, extending well into the short-ranged regime. \begin{figure} \centering \leavevmode \epsfxsize=6.3in \epsfbox{def.eps} \vfill \caption{Various defects obtained in a two-dimensional triangular lattice. The centered interstitial is the only stable interstitial defect.} \label{defects} \end{figure} \noindent \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=3.2in \epsfbox{besek.eps} \caption{Defect energy as a function of the screening $\kappa a$ for $V(r) = K_0(\kappa r)$ at system size $n = 4$ ($N = 480$). Only the centered interstitial is shown, because the edge interstitial is always unstable to it. Various species of vacancies exist, within limited parameter ranges, very close in energy. Lines joining the data points are only an aid to the eye.} \label{bes:ek} \end{figure} \end{minipage} \hfill \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=3.2in \epsfbox{gamep.eps} \caption{Defect energy as a function of the screening $\kappa a$ for $V(r) = 1/r^p$ at system size $n = 5$ ($N = 750$). The apparent increase in energy with $p$ (interaction getting shorter-ranged) would go away with proper normalization of the potential. Lines joining the data points are only an aid to the eye.} \label{gam:ep} \end{figure} \end{minipage} \medskip \noindent \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=3.2in \epsfbox{besdiff.eps} \caption{Defect energies for $V(r) = K_0(\kappa r)$, $n = 4$, on the log-scale, with respect to $V_3$/$V_6$, in order to illustrate the detailed structure of the energy diagram. The $CI$ can be seen crossing $V_6$ at $\kappa a \approx 6.9$. Lines joining the data points are only an aid to the eye.} \label{bes:diff} \end{figure} \end{minipage} \hfill \begin{minipage}{3.4in} \begin{figure} \centering \leavevmode \epsfxsize=3.2in \epsfbox{gamdiff.eps} \smallskip \caption{Defect energies for $V(r) = 1/r^p$, $n = 5$, on the log-scale, with respect to $V_3$/$V_6$. The $CI$ and $V_6$ cross at $p \approx 5.9$. Lines joining the data points are only an aid to the eye.} \label{gam:diff} \end{figure} \end{minipage} \medskip Following previous authors, the simulations were performed in an almost square (length-to-width ratio $5 : 3\sqrt{3}$) cell containing $N = 5n \times 6n = 30 n^2$ lattice points with $n =$ 1 -- 5 (rather than a more nearly square but bigger rectangle of, say, $7n \times 8n$ ($7 : 4\sqrt{3}$) which would allow us to sample fewer number of system sizes $n$ with a given computational limit on $N$). Fig.~\ref{defects} corresponds to $n=3$. A defect is introduced by adding or removing a particle, and then allowing the resulting configuration to relax. The difference between the energies of the relaxed defect configuration and the perfect lattice configuration gives the energy of the defect. There are two modifications to this simple calculation. We want the defect energy corresponding to the physical conditions of constant chemical potential or line density, so we rescale the cell dimensions (by changing the lattice constant $a_0$) after inserting the defect to restore the system to its original density (following Ref.~\cite{CockEl:defect}). Moreover, since we would ideally like to study an infinite system, the large, but finite cell containing $30 n^2$ particles is assumed to be repeated in all directions, so that we are effectively dealing with a periodic array of defects, or, an infinite lattice in the absence of a defect. The periodic boundary conditions maintain the average line density during the relaxation process. However, now the energy per cell also includes the energy of interaction of a defect with all its periodic images. As discussed earlier, this energy is finite, and by extrapolating its dependence on cell size $n$, i.e., inter-defect separation ($\approx 5n$), to large $n$, the energy of an isolated defect can be extracted~\cite{Frey:defect,CockEl:defect}. For short-ranged interactions, the energy calculation can be simplified. We introduce a cut-off interaction radius $r_c$ where the interaction falls to a small fraction of its nearest-neighbour value. The interaction with the particles outside can be approximately accounted for by assuming a uniform density outside and integrating over it. The radius $r_c$ is chosen to make this correction small compared to the total energy, say, less than $10^{-3}$ of it. Interactions within the shell are calculated explicitly. As long as $r_c < L/2$, $L$ being the cell width, this short-range method should be very accurate. For long-ranged interactions such as $\ln{r}$, $1/r$, or $1/r^2$, the above method breaks down, and we must resort to the Ewald summation technique~\cite{Rosenfeld:Ewald,Heyes:Ewald} which yields an effective two-particle interaction that includes the interaction of one particle with all the periodic images of the other. This effective potential consists of a real space sum (corresponding to a screened interaction) and a reciprocal space sum (corresponding to the screening charge). The division between the two is controlled by an Ewald parameter, and by a judicious choice of its value, the interaction can be made sufficiently short-ranged for both sums. We then employ cut-offs in both spaces, with values determined by the desired precision (see Appendix C for details). \noindent \begin{minipage}{\textwidth} \begin{table} \begin{tabular}{|l||c|c|c|c|c|c|} $\kappa a$ & $I_3$ & $SV$ & $V_{2a}$ & $V_3$ & $V_{2b}$ & $V_6$\\ \hline\hline 0 & .073016802 & $V_{2a}$ & .107018876 & .108206944 & .109320135 & $V_3$\\ 1 & .066331581 & .096728537 & .096661116 & .097578530 & .099169907 & $V_3$\\ 2 & .050588818 & .072306827 & .072341149 & .072594220 & .073852944 & $V_3$\\ 3 & .033575192 & .046095915 & $SV$ & .046131759 & .047174061 & $V_3$\\ 4 & .020037313 & .025980648 & $SV$ & .025962421 & .026641900 & $V_3$\\ \hline\hline 4 & .020036\hspace*{\fill} & .025980\hspace*{\fill} & $SV$ & .025961\hspace*{\fill} & .026641\hspace*{\fill} & $V_3$\\ 5 & .0110170\hspace*{\fill} & .0133112\hspace*{\fill} & $SV$ & .0133146\hspace*{\fill} & .0136217\hspace*{\fill} & $V_3$\\ 5.1 & .010338333 & .012397139 & $SV$ & .012400742 & .012674362 & $V_3$\\ 5.2 & .009695442 & .011537972 & $SV$ & .011541059 & $V_3$ & $V_3$\\ 5.3 & .009087036 & .010731274 & $SV$ & .010733113 & $V_3$ & $V_3$\\ 5.4 & .008511788 & .009974612 & $SV$ & .009974441 & $V_3$ & $V_3$\\ 5.5 & .007968369 & .009265581 & $SV$ & .009262603 & $V_3$ & $V_3$\\ 5.6 & .007455456 & .008601808 & $SV$ & .008595187 & $V_3$ & $V_3$\\ 5.7 & .006971737 & .007980968 & $SV$ & .007969812 & $V_3$ & $V_3$\\ 5.8 & .006515917 & .007400791 & $V_3$ & .007384121 & $V_3$ & $V_3$\\ 5.9 & .006086722 & $V_6$ & $V_6$ & $V_6$ & $V_6$ & .006835768\\ 6 & .005682901 & $V_6$ & $V_6$ & $V_6$ & $V_6$ & .006322377\\ 7 & .002788486 & $V_6$ & $V_6$ & $V_6$ & $V_6$ & .002771295\\ \end{tabular} \caption{Defect energies for $V(r) = K_0(\kappa r)$; $a_0=1$; system size $n = 4$ ($N = 480$). The upper part corresponds to the Ewald Sum method for long-range interactions, the lower part to a simple cut-off method for short-range interactions. The centered interstitial and the symmetric vacancy cross at $\kappa a \approx 6.9$. Entries such as ``$V_{2a}$'', ``SV'', ``$V_3$'', ``$V_6$'' indicate an instability to a lower energy defect.} \label{tbl:bes} \end{table} \end{minipage} \noindent \begin{minipage}{\textwidth} \begin{table} \begin{tabular}{|l||c|c|c|c|c|c|} \ $p$& $I_3$& $SV$& $V_{2a}$& $V_3$& $V_{2b}$& $V_6$\\ \hline\hline \ 0& \ 0.073061685& $V_{2a}$& 0.106775085& 0.108253779& 0.108994418& $V_3$\\ \ 1& \ 0.146421440& $V_{2a}$& 0.209046876& 0.209331872& 0.213568209& $V_3$\\ \ 2& \ 0.487928019& 0.677444176& $SV$& 0.672359275& 0.694143882& $V_3$\\ \ 3& \ 1.08543992\hspace*{\fill}& 1.39071722\hspace*{\fill}& $SV$& 1.38704618\hspace*{\fill}& 1.42628053\hspace*{\fill}& $V_3$\\ \ 4& \ 1.99663790\hspace*{\fill}& 2.37494467\hspace*{\fill}& $SV$& 2.37649196\hspace*{\fill}& 2.43341170\hspace*{\fill}& $V_3$\\ \ 5& \ 3.2620983\hspace*{\fill}& 3.5889518\hspace*{\fill}& $SV$& 3.5851010\hspace*{\fill}& $V_3$& $V_3$\\ \ 5.8& \ 4.5498400\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& \ 4.6053332\\ \ 5.9& \ 4.7286554\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& \ 4.7341340\\ \ 6& \ 4.9114956\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& \ 4..8637723\\ \ 7& \ 6.9642383\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& \ 6.1999848\\ \ 8& \ 9.4317462\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& \ 7.5920876\\ \ 9& 12.319586\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& \ 9.0220754\\ 10& 15.629229\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& 10.477581\hspace*{\fill}\\ 11& 19.359421\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& 11.950259\hspace*{\fill}\\ 12& 23.495660\hspace*{\fill}& $V_6$& $V_6$& $V_6$& $V_6$& 13.434556\hspace*{\fill}\\ \end{tabular} \caption{Defect energies for $V(r) = 1/r^p$; $a_0=1$; system size $n = 5$ ($N = 750$). The Ewald sum technique was used to calculate the energies. The centered interstitial and the symmetric vacancy cross at $p \approx 5.9$. Entries such as ``$V_3$'' and ``$V_6$'' indicate an instability to a lower energy defect.} \label{tbl:gam} \end{table} \end{minipage} To find the minimum of the interaction energy as a function of the configuration of N particles, we use the conjugate-gradient method~\cite{Num-Rec}. The forces are also needed for this method, and are easily derived from the energy and conveniently calculated along with it. The results for $n = 4$ ($480$ particles) for $V_{\kappa a}$ and for $n = 5$ for $V_p$ ($750$ particles) are shown in Tables~\ref{tbl:bes} and~\ref{tbl:gam} and Figs.~\ref{bes:ek} and~\ref{gam:ep}. ($n = 5$ was computationally prohibitive for the long-ranged regime with $\kappa a_0 > 0$). Note that, for the screened Bessel-function interaction, we find that calculations optimized for the long- and short-ranged regimes agree to within $1$ part in $20,000$ at $\kappa a_0 = 4$. Moreover, we find that the interaction of a defect with all its periodic images is repulsive for defects with (even) two- and six-fold symmetry, and attractive for (odd) three-fold symmetry, consistent with Ref.~\cite{Frey:defect}. As discussed in Refs.~\cite{Frey:defect} and~\cite{CockEl:defect}, the true asymptotic form of the power law defect interaction probably isn't reached for the distance scales $r \sim 20-30$ lattice spacings studied here. \section{Conclusions} We have studied factors contributing to the wandering of a vacancy or interstitial string defect in a hexagonal columnar crystal. A gas of such strings in the crystalline phase, interacting via short-range potentials, can proliferate via continuous or first-order transitions when the corresponding defect chemical potential changes sign, leading to a supersolid phase. The transition can be modified by the presence of vacancy-interstitial loops, especially in a system of finite thickness. We have also numerically calculated defect line tensions for two families of line interactions which interpolate between long- and short-ranged interaction potentials. In each case, we determine the point where interstitial and vacancy defects exchange stability. A complete accounting requires consideration of a variety of nearly degenerate vacancy configurations. At finite temperatures, the small energy differences between diffferent species will further lower the free energy of the vacancy through a gain in fluctuation entropy. The interstitial itself can fluctuate between the centered and edge configurations. The point where vacancies and interstitials exchange stability will shift at finite temperatures due to entropic effects of this kind. In the context of long-range potential calculations, we show in an Appendix how to extend the Ewald summation to the modified Bessel function potential $K_0(x)$. \acknowledgements This research was supported by the National Science Foundation, in part by the MRSEC Program through Grant No. DMR-9400396 and through Grant No. DMR-9714725.
\subsection*{Acknowledgements} I would like to thank Dorit Aharonov for suggesting the possibility of enhanced models of quantum finite automata, which motivated much of this work, Andris Ambainis for useful comments on the paper, Umesh Vazirani for many discussions that lead to crucial insights and for help with the presentation of the results, and the referees for their feedback on the paper.
\section*{\normalsize{\bf 1. Introduction}} In this work we consider five different string constructions, which lead to the same low-energy spectrum. This consists of a four-dimensional supergravity theory with two space-time supersymmetries whose spectrum, besides the gravity multiplet, contains (when the gauge group is broken to its Cartan subgroup) 19 vector multiplets and 20 hypermultiplets. This massless spectrum is obtained via compactification and orbifold projections from all the known perturbative string constructions, namely the types IIA/B, the heterotic, and the type I. The gauge group being a broken phase of $SO(16) \times SO(16)$\footnote{When we speak of gauge group we mean, on the heterotic string, the part that comes from the currents. For the sake of simplicity we always omit to mention the left and right gauge groups which come from the compact space.}, this spectrum can be obtained from both the heterotic strings, starting either from $E_8 \times E_8$ or from $SO(32)$. In all these constructions the sypersymmetry and/or the gauge group are broken by $Z_2$ projections. The type II orbifolds are obtained either with projections which act symmetrically on the right and left movers \cite{vm,lc,gkr}, giving rise to type IIA or IIB models, or with an asymmetric projection, in which all the supersymmetries come only from the left movers. These orbifolds are dual to one another: the type IIA and IIB are trivially related by the inversion of an odd number of radii, the compactification being self-mirror; the type II asymmetric orbifold is instead related to these by a $U$-duality that exchanges perturbative and non-perturbative moduli (see Ref.\cite{gkr}). The type I orbifolds were recently constructed in Ref.\cite{adds} as orientifolds of certain type IIB orbifolds, with a spontaneous breaking of the $N=8$ supersymmetry. In Ref.\cite{adds} they are indicated respectively as the ``Scherk--Schwarz breaking'' and the ``M-theory breaking'' models. The heterotic orbifold, which we present here as a new construction, is an interesting example of heterotic compactification in which, although the gauge group has maximal rank, there are always, even away from the Abelian point, an equal number of vector and hypermultiplets, leading to vanishing gauge beta-functions. It therefore behaves as the higher-level, reduced rank models considered in Refs.\cite{fhsv,gkp}. We provide evidence of duality between such heterotic orbifold and the above two type I models, which correspond to two different, but continuously related phases of the heterotic theory. This allows us to see the connection between the two models: their relation is non-perturbative from the point of view of type I. Thanks to the duality between these constructions, we are able to determine at least part of the non-perturbative correction to the effective coupling constant of a special, gravitational ($R^2$) amplitude. As a byproduct, we determine also the $U$-duality group. In particular, $S$-duality appears to be broken by the action of a freely acting projection applied to the eleventh coordinate of M-theory. The type IIA and heterotic constructions, on the other hand, are not dual. This is related to the fact that the type IIA orbifold cannot be regarded as a singular limit of a K3 fibration \cite{lc,gkr}. However, even though these models cannot be compared for finite values of the moduli, we argue that they correspond to two different phases, or regions in the moduli space, of M-theory. These two regions are connected in the moduli space. The above issues are discussed according to the following order: In Section 2 we review the type II orbifolds, which were discussed in detail in Ref.\cite{gkr}. We discuss also the corrections to the $R^2$ term, which, being a function of the moduli of the vector manifold, will provide us with a quantity on which to test the duality relations. In Section 3 we recall, from Ref.\cite{adds}, the type I orientifolds, commenting on the addition of discrete Wilson lines which break the gauge group to the Cartan subgroup and discussing the gravitational corrections. In Section 4 we then discuss in detail the heterotic construction. As the previous sections, also this ends with a discussion of the $R^2$ corrections. Through the analysis of this term, we discuss also the duality relation between this model and the type I of Section 3. The duality relations are used to obtain (at least part of) the non-perturbative gravitational corrections, as well as an insight into other non-perturbative properties. Finally, in Section 5 we comment on the connections between the various phases of these $N=2$, M-theory compactifications. Our conclusions are given in Section 6. \noindent \vskip 0.3cm \setcounter{section}{2} \setcounter{equation}{0} \section*{\normalsize{\bf 2. The type II constructions}} \subsection*{\normalsize{\sl 2.1. The type IIA construction}} We start by reviewing the construction of the type IIA, which is obtained by compactification of the ten-dimensional superstring on a Calabi--Yau manifold with Hodge numbers $h^{1,1}=h^{2,1}=19$ \cite{vm,lc}. We recall that this manifold has an orbifold limit in which the $N=8$ supersymmetry of the type IIA string compactified on $T^6$ is reduced to $N=2$ by two $Z_2$ projections, $Z_2^{(1)}$ and $Z_2^{(2)}$ \cite{lc}. The action of these on $T^6=T^2_{(1)} \times T^2_{(2)} \times T^2_{(3)}$ is the following: $Z_2^{(1)}$ acts as a rotation in $T^2_{(1)} \times T^2_{(3)}$, while $Z_2^{(2)}$ acts as a rotation on $T^4=T^2_{(2)} \times T^2_{(3)}$ and as a translation in $T^2_{(1)}$. The partition function of this orbifold reads (see Ref.\cite{gkr}): \begin{eqnarray} Z_{\rm II}^{(1,1)} & = & {1 \over \,{\rm Im}\, \tau \vert \eta \vert^{24} } {1 \over 4} \sum_{H^1,G^1} \sum_{H^2,G^2} \Gamma_{6,6} \ar{H^1,H^2}{G^1,G^2} \nonumber \\ &&\nonumber \\ && \times {1 \over 2} \sum_{a,b} (-)^{a+b+ab} \vartheta \ar{a}{b} \vartheta \ar{a+H^1}{b+G^1} \vartheta \ar{a+H^2}{b+G^2} \vartheta \ar{a-H^1-H^2}{b-G^1-G^2} \nonumber \\ &&\nonumber \\ && \times {1 \over 2} \sum_{\bar{a},\bar{b}} (-)^{\bar{a}+\bar{b}+\bar{a}\bar{b}} \bar{\vartheta} \ar{\bar{a}}{\bar{b}} \bar{\vartheta} \ar{\bar{a}+H^1}{\bar{b}+G^1} \bar{\vartheta} \ar{\bar{a}+H^2}{\bar{b}+G^2} \bar{\vartheta} \ar{\bar{a}-H^1-H^2}{\bar{b}-G^1-G^2}\, , \label{zII} \end{eqnarray} where the contribution of the compactified bosons, $X^I, \bar X^I,I=1,\ldots,6$ is contained in the factor $\Gamma_{6,6} \ar{H^1,H^2}{G^1,G^2} \Big / \vert \eta \vert^{12}$, while the other factors contain the contribution of their fermionic superpartners, $\Psi^I$, ${\bar\Psi}^I$, of the left- and right-moving non-compact supercoordinates $X^{\mu}$, $\Psi^{\mu}$, ${\bar X}^{\mu}$, ${\bar \Psi}^{\mu}$ and of the super-reparametrization ghosts $b,c,\beta,\gamma$ and ${\bar b},{\bar c},{\bar \beta},{\bar \gamma}$. $(H^{1},G^{1})$ refer to the boundary conditions introduced by the projection $Z_2^{(1)}$, and $(H^{2},G^{2})$ to the projection $Z_2^{(2)}$. $\Gamma_{6,6} \ar{H^1,H^2}{G^1,G^2}$ factorizes into the contributions corresponding to the three tori of $T^6$: \begin{equation} \Gamma_{6,6} \ar{H^1,H^2}{G^1,G^2}= \Gamma_{2,2}^{(1)} \ar{H^1 \vert H^2}{G^1 \vert G^2}\, \Gamma_{2,2}^{(2)} \ar{H^2 \vert 0}{G^2 \vert 0}\, \Gamma_{2,2}^{(3)} \ar{H^1+H^2 \vert 0} {G^1+G^2 \vert 0}\, , \end{equation} which are expressed in terms of the twisted and shifted characters of a $c=(2,2)$ block, $\Gamma_{2,2} \ar{h\vert h'}{g\vert g'}$; the first column refers to the twist, the second to the shift. The non-vanishing components are the following: \begin{eqnarray} \Gamma_{2,2} \ar{h\vert h'}{g\vert g'} &=& {4\, \vert\eta \vert^6\over \left\vert \vartheta{1+h\atopwithdelims[] 1+g}\, \vartheta{1-h\atopwithdelims[] 1-g} \right\vert} \ , \ \ {\rm for\ } (h',g')=(0,0) \ {\rm or\ } (h',g')=(h,g) \nonumber \\ &=& \Gamma_{2,2} \ar{h'}{g'} \ , \ \ ~~~~~~~~~ {\rm for\ } (h,g)=(0,0)\, , \label{g22ts} \end{eqnarray} where $\Gamma_{2,2} \ar{h'}{g'}$ is the $Z_2$-shifted $(2,2)$ lattice sum. In this case, the only shift is that due to the $Z_2^{(2)}$ translation on $T^2_{(1)}$, $(-)^{m_2 G^2}$. For a detailed analysis of the spectrum of this model, we refer to Ref.\cite{gkr}. Here we simply recall that the $Z_2^{(1)}$ twisted sector has sixteen fixed points, which give rise to eight vector and eight hypermultiplets. The twist of $Z_2^{(2)}$ on the other hand is accompanied by a lattice shift, so there are no massless states from this twisted sector. The other eight vector and eight hypermultiplets come from the sector twisted by $Z_2^{(1)} \times Z_2^{(2)}$ (the $H^{(1)}+H^{(2)}$-twisted sector). We now consider the corrections to the $R^2$ term. They receive a non-zero contribution at one loop, and are related to the infrared-regularized integral of the fourth helicity supertrace, $B_4$ (for the definition and the details we refer, for instance, to Ref.\cite{gkr}). The one-loop correction to the coupling constant is: \begin{eqnarray} {16 \, \pi^2 \over g^2_{\rm grav} \left( \mu^{(\rm IIA)} \right)} & = & -2 \log \,{\rm Im}\, T^1 \vert \vartheta_4 \left( T^1 \right) \vert^4 - 6 \log \,{\rm Im}\, T^2 \vert \eta \left( T^2 \right) \vert^4 - 6 \log \,{\rm Im}\, T^3 \vert \eta \left( T^3 \right) \vert^4 \nonumber \\ && + 14 \log {M^{(\rm IIA)} \over \mu^{(\rm IIA)}} \, , \label{IIthr} \end{eqnarray} where $T^1,T^2,T^3$ are the K\"{a}hler class moduli of the three tori of the compact space. The last term in Eq.(\ref{IIthr}), which encodes the infrared running due to the massless contributions, is expressed in terms of the type IIA string scale $M^{(\rm IIA)} \equiv {1 / \sqrt{\alpha'_{\rm IIA}}}$ and infrared cut-off $\mu^{(\rm IIA)}$. The coefficient, 14, is actually the massless contribution to $(B_4-B_2) \big/ 3$, as is computed in field theory (see Ref.\cite{gkp}). Observe that there is no limit in the space of moduli $T^1$, $T^2$, $T^3$, at which this correction reproduces the behaviour of an $N=4$ orbifold, for which it is expected to depend on the K\"{a}hler class modulus of only one torus (see for instance Refs.\cite{hmn=4,6auth}). This implies that, in the space of these three moduli, there is no region in which the $N=4$ supersymmetry is restored, as happens instead in orbifold constructions with a spontaneous breaking of supersymmetry, such as those considered in Refs.\cite{gkp,gkp2}. There is therefore no perturbative connection to an $N=4$ theory, and therefore this orbifold cannot be seen as a singular limit in the moduli space of a K3 fibration \cite{gkr}. As explained in Ref.\cite{gkr}, the type IIB dual compactification is trivially obtained by changing the chirality of the right-moving spinors. This is obtained by changing the phase $(-)^{\bar{a}+\bar{b}+\bar{a}\bar{b}}$ in Eq.(\ref{zII}) to $(-)^{\bar{a}+\bar{b}}$. The analysis is similar and we obtain analogous results, with the role of the fields $T^i$, $i=1,2,3$, associated to the K\"{a}hler classes of the three tori, interchanged with that of the fields $U^i$, associated to the complex structures. \subsection*{\normalsize{\sl 2.2. The type II asymmetric dual}} The model above described possesses a type II dual constructed as an asymmetric orbifold \cite{gkr}, obtained by combining the projection $Z_2^{(1)}$, acting in the same way as before, with $Z_2^{\rm F}$, which projects out all the right-moving supersymmetries by relating the action of the right fermion number operator, $(-)^{\rm F_R}$, to a translation in the compact space. The partition function reads \cite{gkr}: \begin{eqnarray} Z_{\rm II}^{(2,0)} & = & {1 \over \,{\rm Im}\, \tau |\eta|^{24} } {1 \over 4} \sum_{H^{\rm F},G^{\rm F}} \sum_{H^{\rm o},G^{\rm o}} \Gamma_{6,6} \ar{H^{\rm F}, H^{\rm o}}{G^{\rm F}, G^{\rm o}} \nonumber \\ && \times {1 \over 2} \sum_{a,b}(-)^{a+b+ab} \vartheta^2 \ar{a}{b} \vartheta \ar{a+H^{\rm o}}{b+G^{\rm o}}\vartheta \ar{a-H^{\rm o}}{b-G^{\rm o}} \nonumber \\ && \times {1 \over 2} \sum_{\bar{a},\bar{b}}(-)^{\bar{a}+\bar{b}+\bar{a}\bar{b}} (-)^{\bar{a}G^{\rm F}+\bar{b}H^{\rm F}+H^{\rm F}G^{\rm F}} \bar{\vartheta}^2 \ar{\bar{a}}{\bar{b}} \bar{\vartheta} \ar{\bar{a}+H^{\rm o}}{\bar{b}+G^{\rm o}} \bar{\vartheta} \ar{\bar{a}-H^{\rm o}}{\bar{b}-G^{\rm o}}~, \label{z2as} \end{eqnarray} with \begin{equation} \Gamma_{6,6} \ar{H^{\rm F}, H^{\rm o}}{G^{\rm F}, G^{\rm o}} = \Gamma_{2,2}^{(1)} \ar{H^{\rm o}~|~H^{\rm F}}{G^{\rm o}~|~G^{\rm F}} \Gamma_{2,2}^{(2)} \ar{H^{\rm o}~|~0}{G^{\rm o}~|~0} \Gamma_{2,2}^{(3)} \ar{0~|~0}{0~|~0}~. \label{z20} \end{equation} As was discussed in Ref.\cite{gkr}, the $R^2$ gravitational correction, which receives contribution only from one loop, is a function of the moduli $T^{\rm As}$, $U^{\rm As}$, associated respectively to the K\"{a}hler class and the complex structure of the third complex plane; it reads \begin{eqnarray} {16 \, \pi ^2 \over g^2_{\rm grav} \left( \mu^{(\rm As)} \right)} & = & -6 \log \,{\rm Im}\, T^{\rm As} \left\vert \eta \left( T^{\rm As} \right) \right\vert^4\, - 6 \log \,{\rm Im}\, U^{\rm As} \left\vert \eta \left( U^{\rm As} \right) \right\vert^4 \, + \nonumber \\ && + 14 \log {M^{(\rm As )} \over \mu^{(\rm As )}}~, \label{dAs1} \end{eqnarray} where we introduced the type II asymmetric mass scale and infrared cut-off, $M^{(\rm As )}$ and $\mu^{(\rm As )}$ respectively. A comparison of Eq.(\ref{dAs1}) and Eq.(\ref{IIthr}) leads directly to the identifications $T^2 = T^{\rm As}$, $T^3 = U^{\rm As}$ and $T^1=\tau^{\rm As}_S$, where $\tau^{\rm As}_S \equiv 4 \pi S^{\rm As}$, $S^{\rm As}$ being the dilaton--axion field of the type II asymmetric models\footnote{We recall that this field belongs to a vector multiplet, as in the heterotic constructions \cite{gkr}.}. For later use, we note here that the corrections Eq.(\ref{IIthr}) and Eq.(\ref{dAs1}) remain the same if we correct the $R^2$ term by adding gauge amplitudes $F^2$: owing to the absence, in the perturbative type II strings, of gauge charges, these amplitudes are identically vanishing. \noindent \vskip 0.3cm \setcounter{section}{3} \setcounter{equation}{0} \section*{\normalsize{\bf 3. The type I models}} There are two type I orbifolds that possess the desired massless spectrum. They were constructed in Ref.\cite{adds}, as orientifolds of type IIB orbifolds, with the $N=8$ supersymmetry spontaneously broken to $N=4$ by a freely acting projection, $Z_2^{(\rm f)}$. The latter acts as a twist on $T^4$, and, in the first case, as a translation in the momenta, produced by the projection $(-)^{m G^{\rm f}}$, on a circle of $T^2$. In Ref.\cite{adds} this construction is called the ``Scherk--Schwarz breaking'' model; in the following we will refer to it as the model {\bf A}. Model {\bf B}, referred to in Ref.\cite{adds} as the ``M-theory breaking'' model, is obtained when the translation is performed on the windings instead. After the orientifold projection, the resulting type I models possess an $N=4$ supersymmetry spontaneously broken to $N=2$. In model {\bf A} the $N=4$ is restored in the limit of decompactification (large radius) of the circle translated by $Z_2^{(\rm f)}$; in model {\bf B}, instead, it is restored when the radius goes to zero. Because of the spontaneous nature of the breaking of the $N=4$ supersymmetry, the massless spectrum contains an equal number of vector and hypermultiplets\footnote{As usual, we don't count here the three vector multiplets and the four hypermultiplets originating from the compact space, which are common to all the $N=2$ orbifolds we consider in this paper.} in the two models. However, while in model {\bf A} the gauge group consists only of factors arising from the contribution of 99-brane sector states, in model {\bf B} the consistency conditions of the construction, as derived by imposing tadpole cancellation, require the presence of both D9- and D5-branes, and the gauge group is the product of the 99-brane sector and the 55-brane sector contribution. Model {\bf A}, in which the gauge group is a broken phase of $SO(32)$, is expected to possess a heterotic dual, and indeed we will construct such a dual in the next section. As explained in \cite{adds}, there is a wide choice of Wilson lines compatible with this Scherk--Schwarz projection, leading to different breakings of the gauge group. In order to compare the type I and the heterotic constructions, we introduce a further set of discrete Wilson lines, which break the gauge group to its Cartan subgroup, $U(1)^{16,}$ \footnote{Here also we omit the contribution coming from the compact space.}. In this way we obtain $3+N_V$ vector multiplets and $4+N_H$ hypermultiplets, with $N_V=N_H=16$. It is possible to choose the Wilson lines to act as $Z_2$ shifts in the directions twisted by $Z_2^{(\rm f)}$. This choice corresponds to a breaking of the gauge group at the $N=2$, six-dimensional level, before the further torus compactification and supersymmetry breaking via orbifold projection; these Wilson lines therefore do not enter the $\Gamma_{2,18}$ lattice explicitly. The only vector multiplets moduli that appear are those of the $\Gamma_{2,2}$ lattice associated with the two-torus of the compactification from six to four dimensions. We will test the duality with the heterotic construction of the next section, through a comparison of the ``holomorphic'' gravitational corrections. These are defined as the corrections to the effective coupling constant of a special combination of $R^2$ and $F^2$ terms \cite{gkp,gkp2}, namely \begin{equation} \langle R^{\prime 2}_{\rm grav} \rangle \equiv \langle R^{2}_{\rm grav} \rangle + {1\over 12} \langle F_{\mu \nu} F^{\mu \nu} \rangle_{T^2} + {5 \over 48} \langle F_{\mu \nu} F^{\mu \nu} \rangle_{\rm gauge} \, . \label{r2f2} \end{equation} This combination was introduced in the above cited works in the context of heterotic orbifolds; it was shown to possess an amplitude smooth in the moduli space, in which the non-harmonic contributions to the $R^2$ and the $F^2$ amplitudes cancel each other, leaving only the $\Gamma_{2,2}$ ($Z_2$-shifted) lattice sum. This is due to the presence of the full bunch of states, contained in the part of the heterotic spectrum which has, as massless excitations, the states of the $c=(0,16)$ currents. On the type I side, the pure $R^2$ amplitude, although smooth in the moduli space, contains, besides the lattice sum provided by the torus ${\cal T}$, the contributions of the Klein bottle ${\cal K}$, the annulus $\cal{A}$ and the M\"{o}bius strip $\cal{M}$ \cite{ms}--\cite{apt}. However, only $N=2$ BPS multiples contribute to such amplitudes, which therefore are all proportional to a ``supersymmetric index'' \cite{ms}; for the combination of gravitational and gauge amplitudes Eq.(\ref{r2f2}), the $\cal{K}$, $\cal{A}$ and $\cal{M}$ contributions cancel (notice that in the case of type I, the gauge amplitude $\langle F_{\mu \nu} F^{\mu \nu} \rangle_{T^2}$ vanishes, because these states come from the RR sector of the type IIB string). The only contribution therefore comes from the torus ${\cal T}$. Since in this model there are no D5-branes, we expect the tree-level effective coupling ${1 \over g^2}$ to be given by the imaginary part of only one complex field, $S=S_1+iS_2$, whose real part $S_1$ is the scalar dual to $B_{\mu \nu}$, while the imaginary part is \begin{equation} S_2={\rm e}^{-\phi_4} G^{1/4} \omega^2~, \end{equation} where $\phi_4$ is the dilaton of the four-dimensional compactification, $\sqrt{G} \sim R_4 R_5$ is the volume of the two-torus, and $\omega^4$ is the volume of the K3, which in the case at hand is in his $T^4 \big/ Z_2$ limit. This situation has to be contrasted with the most general one \cite{abfpt,sgn}, in which the coupling is given by a combination \begin{equation} vS_2+v'S_2' \label{vsvsp} \end{equation} in which $S_2'={\rm e}^{-\phi_4} G^{1/4} \omega^{-2}$ is part of a complex field $S'$ whose real part $S_1'$ is the dual of $B_{45}$. The part of the coupling proportional to the inverse of the K3 volume is due to the presence of D5-branes. The one-loop contribution is given by the complex structure-dependent part of the integral, over the fundamental domain, of the $\Gamma_{2,2}$, $Z_2$-shifted lattice sum appearing in the ${\cal T}$ amplitude. This can be computed in an infrared-regularized background, as in Ref.\cite{infra}. This would lead to the introduction of a curvature in the space-time, providing a cut-off $\mu$ that, once the flat limit is taken, appears in the running of the effective coupling \cite{infra}--\cite{gauge}. There is no need to go into the details of such a procedure: such a running is in fact fixed by a regularization prescription, which imposes the matching of field theory and string computations in the infrared, and can therefore be determined simply by field-theory arguments. On the other hand, the full dependence of the corrections on the modulus $U$, the complex structure of the torus $T^2$ and the only non-trivial modulus that appears in such terms, the Wilson lines being frozen to fixed values, can be easily derived by knowing the action of $Z_2^{(\rm f)}$ on the torus \cite{kkprn}. The total result, including tree-level and one-loop contributions, is therefore given by \begin{equation} {16 \, \pi^2 \over g^2_{\rm grav}(\mu^{(\rm I)})} = 16 \, \pi^2 \,{\rm Im}\, S -2 \log \,{\rm Im}\, U \left \vert \vartheta_4 \left( U \right) \right\vert^4 +14 \log {M^{(\rm I )} \over \mu^{(\rm I )}}~, \label{Ithr} \end{equation} where the Jacobi function $\vartheta_4$ corresponds to a translation $(-)^{m_2 G^{\rm f}}$, in the second circle of $T^2$. The $SL(2,Z)_U$ duality group is broken to a $\Gamma(2)$ subgroup (see Refs.\cite{gkp,6auth,gkp2,kkprn}). In the last term we collect the dependence on the infrared cut-off $\mu^{(\rm I)}$ and the type I string mass scale $M^{(\rm I)}$. We now consider the ``M-theory breaking'', model {\bf B}. We specialize to the case in which, with reference to the notation of Ref.\cite{adds}, $n_2=d_2=0$, i.e. the symplectic factors do not appear in the gauge group, which is therefore given by $SO(16)_{99} \times SO(16)_{55}$ (the subscripts indicate the origin of these factors). We now introduce Wilson lines, as we did in the previous case, to break $SO(16)_{99}$ to $U(1)^8$. We also move the D5-branes a bit far from each other, in order to break also the second factor to the Cartan subgroup. We can now repeat the same arguments as before and compute the analogous ``gravitational'' corrections. In this case, due to the presence of the D5-branes, we expect a dependence of the tree level effective coupling also on the field $S'$. We therefore obtain: \begin{equation} {16 \, \pi^2 \over g^2_{\rm grav}(\mu^{(\rm I)})} = 16 \, \pi^2 v \,{\rm Im}\, S +16 \, \pi^2 v' \,{\rm Im}\, S' -2 \log \,{\rm Im}\, U \left \vert \vartheta_4 \left( U \right) \right\vert^4 +14 \log {M^{(\rm I )} \over \mu^{(\rm I )}}~, \label{Ithrp} \end{equation} where, as in Refs.\cite{abfpt,apt}, we allow for the presence of two independent contributions to the effective coupling constant ($v v' \neq 0$). Actually, since the model is symmetric under the exchange of the D9- and the D5-branes sectors, we deduce that $v=v'=1$. The $\vartheta_4(U)$ is obtained for a translation $(-)^{n_1 G^{\rm f}}$ on the windings of the first circle of $T^2$. \noindent \vskip 0.3cm \setcounter{section}{3} \setcounter{equation}{0} \section*{\normalsize{\bf 3. The heterotic construction}} We now discuss in detail the heterotic dual. This is constructed as a $Z_2$ freely acting orbifold of the heterotic string compactified on $T^6$, with the gauge group broken to $SO(16)\times SO(16)$ by a $Z_2$, discrete Wilson line. The type I construction corresponds to a special region in the moduli space of the theory, and in order to be able to compare heterotic and type I, we must choose on the heterotic side a special set of Wilson lines. What we need is a set of three other discrete Wilson lines, which act on the compact space as $Z_2$ translations, and further break $SO(16) \times SO(16)$ to $U(1)^{8} \times SO(4)^4$. The $N=4$ supersymmetry is then reduced to $N=2$ by a further $Z_2$ freely acting projection, $Z_2^{(\rm f)}$, which moreover acts as a further Wilson line, which breaks the gauge group, leaving massless only the bosons in the Cartan subgroup. The partition function, $Z_{\rm Het}$, can be easily written in terms of the usual fermionic and bosonic characters for the compact space and the $c=(0,2)$, $SO(4)$ twisted characters introduced in Ref.\cite{gkp}, $F_1$ and $F_2$, for the $c=(0,16)$ currents. We recall that: \begin{equation} F_{1}\ar{\gamma,h}{\delta, g} \equiv {1\over {\eta}^2}\, {\vartheta}^{1/2} \ar{\gamma+h_1}{\delta+g_1}\, {\vartheta}^{1/2} \ar{\gamma+h_2}{\delta+g_2}\, {\vartheta}^{1/2} \ar{\gamma+h_3}{\delta+g_3}\, {\vartheta}^{1/2} \ar{\gamma-h_1-h_2-h_3}{\delta-g_1-g_2-g_3} \end{equation} and \begin{equation} F_{2}\ar{\gamma, h}{\delta, g} \equiv {1\over {\eta}^2}\, {\vartheta}^{1/2} \ar{\gamma}{\delta}\, {\vartheta}^{1/2} \ar{\gamma+h_1-h_2}{\delta+g_1-g_2}\, {\vartheta}^{1/2} \ar{\gamma+h_2-h_3}{\delta+g_2-g_3}\, {\vartheta}^{1/2} \ar{\gamma+h_3-h_1}{\delta+g_3-g_1}\, , \end{equation} where $h\equiv (h_1,h_2,h_3)$ and similarly for $g$. Under $\tau\to \tau+1$, $F_{I}$ transforms as: \begin{eqnarray} F_{1}\ar{\gamma, h}{\delta, g} & \to & F_{1}\ar{\gamma, h}{\gamma+\delta +1, h + g} \nonumber \\ && \times \exp -{i\pi\over 4} \left({2\over 3} - 4\gamma +2 \gamma^2+h_1^2+h_2^2 +h_3^2 + h_1 h_2+h_2 h_3+h_3 h_1 \right) \, , \\ F_{2}\ar{\gamma, h}{\delta, g} & \to & F_{2}\ar{\gamma, h}{\gamma+\delta +1, h + g} \nonumber \\ && \times \exp -{i\pi\over 4} \left({2\over 3} - 4\gamma +2 \gamma^2+h_1^2+h_2^2 +h_3^2 - h_1 h_2-h_2 h_3-h_3 h_1 \right)\, . \end{eqnarray} In terms of these, we have \begin{eqnarray} Z_{\rm Het} & = & {1 \over \,{\rm Im}\, \tau | \eta|^4 } {1 \over 2} \sum_{H^{\rm f},G^{\rm f}} Z_{6,22} \ar{H^{\rm f}}{G^{\rm f}} \nonumber \\ && \times {1 \over 2} \sum_{a,b}{1\over \eta^4}~ (-)^{a+b+ab} \vartheta \ar{a}{b}^2 \vartheta \ar{a+H^{\rm f}}{b+G^{\rm f}} \vartheta \ar{a-H^{\rm f}}{b-G^{\rm f}} \, , \label{hH} \end{eqnarray} where the second line stands for the contribution of the 10 left-moving world-sheet fermions $\psi^{\mu},\Psi^I$ and the ghosts $\beta,\gamma$ of the super-reparametrization; $Z_{6,22}\ar{H^{\rm f}}{G^{\rm f}}$ accounts for the $(6,6)$ compactified coordinates and the $c=(0,16 )$ conformal system, which is described by the 32 right-moving fermions $\Psi_A$, $A=1,\ldots,32$: \begin{equation} Z_{6,22}\ar{H^{\rm f}}{G^{\rm f}}= {1\over 2^{5}}\, \sum_{\vec h, \vec g}~{1\over \eta^6 {\bar \eta}^6 } \Gamma_{2,2} \ar{H^{\rm f},h_1}{G^{\rm f},g_1} \, \Gamma_{4,4} \ar{H^{\rm f}\vert \vec h}{G^{\rm f}\vert \vec g}\, {\overline \Phi}\ar{H^{\rm f}, \vec h}{G^{\rm f}, \vec g}\, ,\label{hH622} \end{equation} and \begin{eqnarray} \Phi\ar{H^{\rm f}, \vec h}{G^{\rm f}, \vec g} & = & {1 \over 2} \sum_{\gamma,\delta} F^2_1 \ar{\gamma,h_1,h_2,h_3}{\delta,g_1,g_2,g_3} _{(H^{\rm f},G^{\rm f})} F^2_2 \ar{\gamma,h_1,h_2,h_3}{\delta,g_1,g_2,g_3} _{(H^{\rm f},G^{\rm f})} \nonumber \\ && \times \, F^2_1 \ar{\gamma+h_4,H^{\rm f},h_2,h_3}{\delta+g_4,G^{\rm f},g_2,g_3} F^2_2 \ar{\gamma+h_4,H^{\rm f},h_2,h_3}{\delta+g_4,G^{\rm f},g_2,g_3}~. \label{phi} \end{eqnarray} In Eq.(\ref{hH622}) we used the twisted-shifted bosonic character for $\Gamma_{4,4} \ar{h|h'}{g|g'}$, in which the first column indicates the twist $(h,g)$, the second the shift $(h',g')$, and the doubly-shifted character of the two-torus $\Gamma_{2,2} \ar{h,h'}{g,g'}$. In Eq.(\ref{phi}) the subscripts $(H^{\rm f},G^{\rm f})$ indicate the embedding of the spin connection in the gauge group. This is realized explicitly through a modification of the arguments in the first Ising character, both in $F_1$ and in $F_2$: \[ F_{1}\ar{\gamma,h}{\delta, g}_{(H^{\rm f},G^{\rm f})} \equiv {1\over {\eta}^2}\, {\vartheta}^{1/2} \ar{\gamma+h_1+H^{\rm f}}{\delta+g_1+G^{\rm f}}\, {\vartheta}^{1/2} \ar{\gamma+h_2}{\delta+g_2}\, {\vartheta}^{1/2} \ar{\gamma+h_3}{\delta+g_3}\, {\vartheta}^{1/2} \ar{\gamma-h_1-h_2-h_3}{\delta-g_1-g_2-g_3}~; \] \begin{equation} F_{2}\ar{\gamma, h}{\delta, g}_{(H^{\rm f},G^{\rm f})} \equiv {1\over {\eta}^2}\, {\vartheta}^{1/2} \ar{\gamma+H^{\rm f}}{\delta+G^{\rm f}}\, {\vartheta}^{1/2} \ar{\gamma+h_1-h_2}{\delta+g_1-g_2}\, {\vartheta}^{1/2} \ar{\gamma+h_2-h_3}{\delta+g_2-g_3}\, {\vartheta}^{1/2} \ar{\gamma+h_3-h_1}{\delta+g_3-g_1}\, , \end{equation} With this embedding, the shift in $\Gamma_{2,2}$ is produced by the projection $(-)^{m_2G^{\rm f}+n_2g_1}$ \footnote{As in Refs.\cite{gkp,gkp2}, there exists also an alternative heterotic construction with the same massless spectrum, in which the spin connection is embedded in the right $U(1)^2$ factor from the untwisted two-torus. In this case, the shift in $\Gamma_{2,2}$ is produced by an asymmetric projection, e.g. $(-)^{(m_1+n_1)G^{\rm f}}$. We will not consider this alternative construction, because it is not dual to the type I.}. The conformal blocks in the second line of Eq.(\ref{phi}) provide the right-moving part of eight vector multiplets, corresponding to a factor $U(1)^8$ in the gauge group, and sixteen hypermultiplets. From the blocks on the r.h.s. of the first line we get the other eight vectors, to make $U(1)^{16}$, and no further hypermultiplets. In addition to these, the massless spectrum of the model contains the usual three vector multiplets and four hypermultiplets of an ordinary $T^2 \times T^4 / Z_2$ compactification of the heterotic string. Thanks to the free action of all the projections, there are no additional massless states coming from the twisted sectors. It is worth remarking that, when some of the Wilson lines are absent, the gauge group is enlarged but still $N_V=N_H$. Notice that, because of the embedding of the spin connection into the gauge group, it is not possible to construct this model at the point $SO(16) \times SO(16)$, but only at a broken phase of it. It is easy to recognize that the $N=2$ sector of this orbifold, specified by $(H^{\rm f},G^{\rm f}) \neq (0,0)$, belongs to the same universality class as the $N=2$ heterotic constructions with $N_V=N_H$, considered in Refs.\cite{gkp,gkp2}. In fact, the modular transformation properties of the untwisted $\Gamma_{2,2}$ lattice, toghether with the condition $N_V=N_H$, as was already pointed out in Ref.\cite{gkp}, are sufficient to fix this orbifold sector uniquely, the difference between the various models residing in the $N=4$ sector. A consequence of this is that not only the $N=2$ singularities, but also all the threshold corrections that receive contribution only from this sector, are the same as for the other models of this class; we therefore skip the details about the analysis of this model and go directly to the discussion of the gravitational corrections. As we anticipated in Section 2, it was pointed out in Refs.\cite{gkp,gkp2} that the ``pure'' gravitational amplitude, $\langle R^2 \rangle$, must be corrected with a term proportional to the gauge amplitude, $\langle F_{\mu \nu} F^{\mu \nu} \rangle$, in order to make it holomorphic and non-singular. We recall here that the precise combination is: \begin{equation} \langle R^{\prime 2}_{\rm grav} \rangle \equiv \langle R^{2}_{\rm grav} \rangle + {1\over 12} \langle F_{\mu \nu} F^{\mu \nu} \rangle_{T^2} + {5 \over 48} \langle F_{\mu \nu} F^{\mu \nu} \rangle_{\rm gauge} \, . \label{r2f2h} \end{equation} The tree-level plus one-loop contribution reads: \begin{eqnarray} {16 \, \pi^2 \over g^2_{\rm grav}(\mu^{(\rm Het)})} & = & 16 \, \pi^2 \,{\rm Im}\, S^{(\rm Het)} -2 \log \,{\rm Im}\, T \left \vert \vartheta_4 \left( T \right) \right\vert^4 -2 \log \,{\rm Im}\, U \left \vert \vartheta_4 \left( U \right) \right\vert^4 \nonumber \\ && +14 \log {M^{(\rm Het)} \over \mu^{(\rm Het)} } + \ {\rm const.} \, , \label{htr} \end{eqnarray} where $S^{(\rm Het)}$ is the heterotic axion--dilaton field, \begin{equation} \,{\rm Im}\, S^{(\rm Het)} ={1 \over g^2_{\rm Het} } \end{equation} and we used the string scale $M^{(\rm Het)} \equiv {1 / \sqrt{\alpha'_{\rm Het}}}$ and the infrared cut-off $\mu^{(\rm Het)}$ of the heterotic string. Owing to the free action of the projection $Z_2^{(\rm f)}$, also in this model the $N=4$ supersymmetry is spontaneously broken perturbatively \cite{kk}; it is restored when $T \to \infty$, $U \to \infty$. This limit corresponds, as in the type I construction, to a decompactification to five dimensions \footnote{When $\,{\rm Re}\, T=\,{\rm Re}\, U=0$, we have $\,{\rm Im}\, T \sim R_1 R_2$ and $\,{\rm Im}\, U \sim R_2/R_1$, where $R_1$, $R_2$ are the radii of the two circles of the torus. In this case this limit corresponds to $R_2 \to \infty$, with $R_1$ fixed.}. In order to compare this model with the type I, ``Scherk--Schwarz breaking'' model {\bf A} of the previous section, we consider the limit in which $T$ is large while $U$ is kept finite (we stress that in this limit the supersymmetry remains broken to $N=2$). In this limit, the correction Eq.(\ref{htr}) depends on $T$ only logarithmically: \begin{equation} {16 \, \pi^2 \over g^2_{\rm grav}(\mu^{(\rm Het)})} \approx 16 \, \pi^2 \,{\rm Im}\, S^{(\rm Het)} -2 \log \,{\rm Im}\, U \left \vert \vartheta_4 \left( U \right) \right\vert^4 + {\cal O}(\log \,{\rm Im}\, T)~. \label{htr1} \end{equation} If we discard this logarithmic dependence for the moment, we see that Eq.(\ref{htr1}) exactly reproduces the type I correction given in Eq.(\ref{Ithr}), with the identifications $S^{(\rm Het)} \equiv S$ and $U^{(\rm Het)} \equiv U^{(\rm I)}$. We can interpret the logarithmic dependence as due to effects that are non-perturbative from the type I point of view\footnote{This is the same phenomenon as that appearing in the examples of type IIA/type II asymmetric orbifolds dualities considered in Refs.\cite{gkr,6auth,gkp2}, where the absence of tree level dilaton dependence in the analogous corrections on the asymmetric orbifolds indeed corresponds to a logarithmic dependence due to non-perturbative phenomena.}. The limit of restoration of the $N=4$ supersymmetry, in both the theories, corresponds to the decompactification of one radius ($R \to \infty$). In this limit, also $U \to \infty$, and the effective coupling constant, in the $N=4$ phase, depends only on the dilaton $S=S^{(\rm Het)}$, as expected in both the theories\footnote{Still, there is the presence of the logarithmic terms, both in $T$ and $U$. These terms can be lifted by switching on an appropriate cut-off, as is discussed in Refs.\cite{gkr,gkp,gkp2,solving}.}. We consider the coincidence of the massless spectrum and the rather non-trivial correspondence of these threshold corrections as compelling evidence of the duality of the heterotic and the type I, model {\bf A} constructions\footnote{In the construction with the asymmetric shift in $\Gamma_{2,2}$ referred to in footnote 5, there are lines, $T=f(U)$, in the $(T,U)$ space, along which the ``smooth gravitational amplitude'' we are considering indeed becomes singular. This is due to the appearance in the massless spectrum of new hypermultiplets, which are uncharged under the gauge group of the torus, and therefore lead to a jump in the $\beta$-function of the $R^2$ term, which is not compensated by an opposite jump in the $\beta$-function of $F^2$ (see Refs.\cite{gkp,kkprn}). By duality, these singularities should appear also on the type I side, where new massless states should appear for large or small values of the modulus $U$. The absence of these rules out the alternative heterotic construction.}. We now consider the limit $T \to 0$, with $U$ fixed. In this limit, the theory is better described in terms of the inverse modulus $\tilde{T} \equiv -1/T$. By performing an $SL(2,Z)$ inversion, the second term in Eq.(\ref{htr}) becomes \begin{equation} -2 \log \,{\rm Im}\, \tilde{T} \left\vert \vartheta_2 \left( \tilde{T} \right) \right\vert^4~, \end{equation} which diverges linearly when $\,{\rm Im}\, \tilde{T}$ is large: \begin{equation} \sim 2 \pi \,{\rm Im}\, \tilde{T} ~. \end{equation} Therefore, for large $\,{\rm Im}\, S^{(\rm Het)}$ and small $T$, Eq.(\ref{htr}) becomes \begin{equation} {16 \, \pi^2 \over g^2_{\rm grav}(\mu^{(\rm Het)})} \approx \, 16 \, \pi^2 \,{\rm Im}\, S^{(\rm Het)} +2 \pi \,{\rm Im}\, \tilde{T} -2 \,{\rm Im}\, U \vert \vartheta_4 \left( U \right)\vert^4 \, . \label{hthr2} \end{equation} It is tantalizing to interpret the linear divergence in the field $\tilde{T}$ as corresponding to the appearance of a D5-brane sector in the dual type I theory. Indeed, in this region of the moduli space, the effective coupling constant of the heterotic theory behaves like that of the ``M-theory breaking'' type I model {\bf B}, Eq.(\ref{Ithrp}), provided we identify the field $S^{(\rm Het)}$ with $S$, as before\footnote{Here, by an abuse of language, we are using the same notation, $S$, for both the type I constructions.} and $\tilde{T}$ with $8 \pi S'=2\tau_S'$ (we have introduced here as usual the field $\tau_S'$, in analogy with the field $\tau_S=4 \pi S$, the actual dilaton--axion field that enters the $SL(2,Z)$, Montonen--Olive duality transformations). As before, we can identify the complex structure moduli $U$ of the heterotic and type I constructions. It therefore seems that, for large $\tilde{T}$, the heterotic theory is perturbatively dual to the type I construction {\bf B}. This correspondence is better understood in terms of the ``T-dual'' heterotic theory, obtained by exchanging the two radii of the torus, $R_1 \leftrightarrow R_2$, and then inverting them: $R_1 \to \tilde{R}_1= 1 / R_1$, $R_2 \to \tilde{R}_2=1 / R_2$. With these inversions, we remain in a spontaneously broken phase of the $SO(16) \times SO(16)$ string, but with the K\"{a}hler class modulus of the torus given by $\tilde{T}$ instead of $T$. The modulus $\tilde{T}$ is coupled to the windings of the torus of this T-dual theory, and in the limit of large $\,{\rm Im}\, \tilde{T}$ all the string states with a non-zero winding number decouple from the spectrum, leaving only the Kaluza--Klein states, of the type I dual string\footnote{This limit can also be viewed as the infinite-tension limit of the heterotic string. We have in fact $\,{\rm Im}\, \tilde{T} \sim \tilde{R}_1 \tilde{R}_2 / \alpha'$, $\,{\rm Im}\, U \sim \tilde{R}_2 / \tilde{R}_1$, and the limit $\,{\rm Im}\, \tilde{T} \to \infty$ with $U$ fixed is equivalent to the limit $\alpha' \to 0$ with fixed radii.}. The above duality implies in this example that the massless states of the 55-branes sector appear on the heterotic side as perturbative states associated to the gauge currents, on the same footing as the states of the 99-branes sector. The situation is therefore rather different from that of the models of Gimon and Polchinski \cite{gp}, in which the states of the 55-branes sector are non-perturbative on the heterotic side \cite{w,blpssw}. A solution to this (apparent) puzzle comes from considering the T-dual, type I$^{\prime}$ picture \cite{adds}, in which the two gauge factors are provided by the two ``Ho\v{r}ava--Witten'' walls of the M-theory, $S^1 \big/ Z_2$ orbifold, where the D9-branes on one wall are wrapped on a T-dualized four-torus, and effectively appear as D5-branes. Nevertheless, because of their origin, the states corresponding to open strings ending on these D-branes are expected to appear as perturbative states on the heterotic theory\footnote{We thank E. Dudas for a clarification of this point.}. We can now try to see what happens from the type I side point of view when the heterotic theory passes from small to large $T$ (or equivalently from large to small $\tilde{T}$). Because of the identifications \begin{equation} S^{(\rm Het)}=S={\rm e}^{-\phi_4}G^{1/4}\omega^{2}~,~~~~~~~~~~ \tilde{T}=8 \pi S'=8 \pi {\rm e}^{-\phi_4}G^{1/4}\omega^{-2}~, \label{ssp} \end{equation} we have $8 \pi S^{(\rm Het)} \big/ \tilde{T} = S \big/ S'= \omega^4$. We therefore see that such a motion, performed while keeping the field $S^{(\rm Het)}$ fixed, corresponds on the type I side to increasing the volume of the $K3$, $\omega^4 \to \infty$. In order to keep the field $S$ fixed, we have also to shrink the volume of the two-torus and/or adjust the value of the dilaton $\phi_4$. In order to remain in the phase of broken $N=4$ supersymmetry, we must shrink the second circle, leaving the first one fixed. Notice that, according to Eq.(\ref{vsvsp}), this motion is non-perturbative from the type I point of view, involving a change of the tree-level coupling constant. In this limit, which corresponds to an effective decompactification of the theory to eight dimensions (or nine, if a circle of the two-torus is shrinked), indeed the D5-branes look like D9-branes wrapped around the $T^2$ torus. The $U(1)$ gauge bosons, which were provided by open strings ending on the same D5-brane, are still there; they contribute for a $U(1)^8$ factor to the gauge group, although they must now be reinterpreted as due to strings ending on D9-branes. What essentially distinguishes the behaviour of this model with respect to the type I constructions, in which supersymmetry is not spontaneously broken, as in Refs.\cite{gp,gj}, is that there is an unbroken $S$-duality in those cases; combined with the symmetry under exchange of the fields $S$ and $S'$, i.e. of the 99- and 55-branes sectors, this duality implies that, along such a motion, non-perturbative phenomena enter heavily in the game and, at the limit we are considering, the theory is perturbatively described by the $S$-, $S'$-dual, identical theory. In our case, instead, we expect $S$-duality to be broken, because, as explained in Ref.\cite{adds}, the type I model {\bf B} indeed corresponds to a Scherk--Schwarz mechanism applied to the 11-th dimension of M-theory\footnote{We will come back to this point, which contradicts the results of Ref.\cite{dg}.}. It is then reasonable to find that in this limit, since model {\bf B} is not falling back into itself, it ends up to coincide with model {\bf A}, in which all the gauge bosons of $U(1)^{16}$ are provided by the D9-branes\footnote{In order to understand how in this limit the D5-branes can look like D9-branes, consider the T-dual situation in which the four circles of $T^4 \big/ Z_2$ are T-dualized. In this case, for finite values of the radii, the D9-branes become D5-branes and vice versa. However, when the radii are shrunk to zero, the theory lives effectively in four extended and two compact dimensions, where there are no 9-branes but only 5-branes wrapped around the compact torus. T-dualizing again the four circles brings us back to the original limit, in which all the D5-branes become D9-branes.}. We now consider the restoration of the $N=4$ supersymmetry. The higher amount of supersymmetry can be restored essentially in two ways, which correspond on the type I side to the two decompactifications: $R_2 \to \infty$ in model {\bf A} and $R_1 \to 0$ in model {\bf B}. In the first case, the restoration is perturbative in both the type I and the heterotic side (the field $S \to \infty$). In the second case, the restoration is non-perturbative from both the type I and heterotic points of view ($S$, $S' \to 0$). There are, however, also intermediate possibilities, which involve a change also of $\omega$. In these cases, the restoration, although non-perturbative on the type I side, can look perturbative on the heterotic side. This happens when the product $G^{1/4} \omega^2$ is kept fixed. \subsection*{\sl Non-perturbative corrections} We have seen that through the duality between heterotic and type I constructions, we gained insight into the non-perturbative behaviour of both of them, at least regarding the restoration of the $N=4$ supersymmetry. We now try to go further: indeed, through the heterotic dual, we learned that the two type I constructions are actually two realizations of the same theory. The type I/heterotic duality can be used to get insight into the non-perturbative corrections to the effective coupling constant of the combination of gravitational and gauge amplitudes given in Eq.(\ref{r2f2}). From the heterotic dual we know that the $SL(2,Z)_{S'}$ duality group is broken to a $\Gamma(2)$ subgroup. On the heterotic side the $\Gamma(2)_T \times \Gamma(2)_U$ group is by construction a symmetry that remains valid at any value of the coupling. On the other hand, we know that the type I ``M-theory breaking'' model {\bf B}, is perturbatively symmetric under the exchange of the fields $S$ and $S'$: this is a consequence of the symmetry under the exchange of the D9- and D5-branes sectors. We claim that this implies that also the $SL(2,Z)_{S}$ duality group is indeed broken in the same way as the $SL(2,Z)_{S'}$ group. This statement, which promotes a perturbative symmetry to a non-perturbative one, is supported by the observation that, as is discussed in Ref.\cite{adds}, in the T-dual, type I$^{\prime}$ picture, the two contributions come from the two ``Ho\v{r}ava--Witten'' walls of the M-theory on $S^1 \big/ Z_2$, and the symmetry of the problem under exchange of the two remains true at any value of the 11-th coordinate. Once observed that the symmetry of the theory is $\Gamma(2)_{2S} \times \Gamma(2)_{2S'}\times \Gamma(2)_U$ times the permutations of the three factors, we can write the full, non-perturbative correction, which reduces to Eq.(\ref{htr}) in the large-$S$ limit: \begin{eqnarray} {16 \, \pi^2 \over g^2_{\rm grav}(\mu)} & = & -2 \log \,{\rm Im}\, \tau_S \left \vert \vartheta_2 \left( 2 \tau_S \right) \right\vert^4 -2 \log \,{\rm Im}\, \tau_S' \left \vert \vartheta_2 \left( 2 \tau_S' \right) \right\vert^4 -2 \log \,{\rm Im}\, U \left \vert \vartheta_4 \left( U \right) \right\vert^4 \nonumber \\ && +{\cal E}(2\tau_S,2\tau_S',U)+ 14 \log {M \over \mu }~. \label{tr} \end{eqnarray} In this expression we used the type I fields $\tau_S=4 \pi S$, $\tau_S'=4 \pi S'$ and $U$, which by now we know to be the same for both the type I models and equivalent to the heterotic fields $\tau_S^{\rm Het}=4 \pi S^{\rm Het}$, $\tilde{T}/2$ and $U$. The infrared cut-off $\mu$ and the mass scale $M$ can be indifferently those of the type I or of the heterotic string, their relation being \begin{equation} { M^{(\rm Het)} \over \mu^{(\rm Het)}} = { M^{(\rm I)} \over \mu^{(\rm I)}}~. \end{equation} In Eq.(\ref{tr}) we allow for the presence of a series, ${\cal E}(2\tau_S,2\tau_S',U)$, of exponentials symmetric in $2\tau_S, 2\tau_S', -1/U$. Such a term, always suppressed in the perturbative limit, cannot be excluded by the symmetries of the theory. In each of the limits in which the $N=4$ supersymmetry is perturbatively restored (either on the heterotic or on both sides), the contribution of two moduli drops out (it reduces to the already mentioned logarithmic dependence) and the correction Eq.(\ref{tr}) diverges linearly as a function of only one modulus (the field $S$, as expected for $N=4$ corrections). The term ${\cal E}(2\tau_S,2\tau_S',U)$ is suppressed in the limits of a restoration of supersymmetry. By using its symmetry properties it is easy to see that it is suppressed also in the non-perturbative limit $S \to 0$. The form of Eq.(\ref{tr}) therefore tells us that there exists a limit, which is non-perturbative on both the heterotic and type I sides, in which there is an effective restoration of an $N=8$ supersymmetry. This takes place when the three moduli contributions drop out, namely when $S \to 0$, $S' \to 0$ and $U \to \infty$. This for instance happens when, in model {\bf B}, we send $R_1$ to zero while keeping $\omega$, $R_2$ and the field $\phi_4$ fixed. In this case the effective coupling Eq.(\ref{tr}) vanishes (modulo the usual logarithmic divergences). Through the above analysis we have learned that in this theory, which can be interpreted as obtained by a freely acting projection of M-theory, the $U$-duality group $SL(2,Z)_S \times SL(2,Z)_T \times SL(2,Z)_U$ is indeed broken. In particular, the $S$-duality is broken to a $\Gamma(2)_{2\tau_S}$ subgroup. This is in contrast with the results of Ref.\cite{dg}, in which, even in the presence of a Scherk--Schwarz compactification of the 11-th coordinate of M-theory, $S$-duality remains unbroken. We believe that, once properly treated, a freely acting projection applied on the 11-th coordinate must instead necessarily break the $S$-duality. The result of Ref.\cite{dg} is obtained by using as a starting point the effective Ho\v{r}ava--Witten action, in which the space-time metric appearing in the gauge terms, which live on the two ten-dimensional walls, is a ``coordinate-independent'' restriction of the eleven-dimensional metric to ten dimensions. In this way, the heterotic dilaton has no dependence on the 11-th radius. This scenario is correct for a genuine Ho\v{r}ava--Witten orbifold of M-theory, in which the original amount of supersymmetry can never be restored by a decompactification of the eleventh dimension. This is, however, not the case of a Scherk--Schwarz compactification of M-theory, in which we expect the supersymmetry to be restored in the large-radius limit (as happens in our present case, where the radius of M-theory is T-dual to the $R_1$ of model {\bf B}). In this limit, the gauge bosons of the terms introduced to cancel the ten-dimensional anomaly should decouple. This is indeed what happens in our case, and is consistent with general arguments that fix the relation between the four-dimensional couplings and the eleventh radius $\rho$ to be \cite{hw,aq}: \begin{equation} \,{\rm Im}\, S,\, \,{\rm Im}\, S' \, \sim \, \rho^{-2 / 3}~, \end{equation} where we have omitted the factors that contain all the other parameters. As a last point, we remark that the model is symmetric under permutations of the fields $X=8 \pi S$, $Y=8 \pi S'$ and $-1/ U$. By proceeding as in Ref.\cite{gkp2}, it is therefore possible to calculate (at least a part of) the non-perturbative prepotential. This is obtained by a proper symmetrization of the perturbative prepotential, computed, as a function of the moduli $T$ and $U$, on the heterotic side. The result is the same as in Ref.\cite{gkp2}. \noindent \vskip 0.3cm \setcounter{section}{5} \setcounter{equation}{0} \section*{\normalsize{\bf 5. Discussion}} In the previous section we were able to determine part of the non-perturbative behaviour of the gravitational corrections. In order to determine also the term ${\cal E}(2\tau_S,2\tau_S',U)$ in Eq.(\ref{tr}), we would need to identify a type IIA dual, in which these three fields would belong to the perturbative moduli. This would allow us to repeat the analysis of Refs.\cite{gkp,hmn=4,gkp2,hmn=2}. In order for the duality to work, such a type IIA dual would have to be constructed as a compactification on (an orbifold limit of) a K3 fibration \cite{kv}--\cite{al}. A necessary condition is therefore a symmetric breaking of the supersymmetry; this requirement leads directly to the type IIA orbifold of Section 2. However, the heterotic and the type IIA orbifolds under consideration cannot be compared for finite values of the heterotic moduli $T$ and $U$. This is related to the fact that the compact space of the type IIA orbifold, a $T^6/( Z_2 \times Z_2)$ limit of $CY^{19,19}$, cannot be seen as a singular limit of a K3 fibration (see Refs.\cite{vm}--\cite{gkr}). Indeed, the spaces of the vector moduli of the two models do not correspond. A signal of the non-coincidence of these subspaces is provided by the absence of a perturbative super-Higgs phenomenon on the type IIA construction: unlike in the heterotic model, on the type IIA side it is not possible to restore the $N=4$ supersymmetry in a corner of the space of moduli $T^i$, $U^i$, $i=1,2,3$ \footnote{As is discussed in Ref.\cite{gkr}, compactification on a K3 fibration and spontaneous breaking of the $N=4$ supersymmetry are directly related.}. However, these models can be viewed as different phases of a single theory, corresponding to different regions of a wider moduli space\footnote{We consider here only the space of moduli belonging to the vector multiplets.}. These two regions are connected at the limits $\,{\rm Im}\, T^2 \to 0,\infty$, $\,{\rm Im}\, T^3 \to 0,\infty$ of the type IIA moduli space, for any value of the modulus $T^1$. In these limits, the type II correction, Eq.(\ref{IIthr}), reproduces the heterotic/type I, Eq.(\ref{tr}), at the appropriate corners in the space of moduli $T$ and $U$ ($\,{\rm Im}\, T, \,{\rm Im}\, U \to 0$ and/or $\infty$). For any value of $T^1$, we have the identifications $T^1=\tau_S^{\rm As}=-1 \big/ 2\tau_S$. Indeed, for any value of this modulus, there are two branches of the theory: a branch corresponds to a heterotic/type I phase, with moduli $T,U$; crossing the borders at $\,{\rm Im}\, T,\,{\rm Im}\, U \to 0,\infty$, one passes to another branch, which corresponds to the type II phase, described by the type IIA or type II asymmetric orbifolds, with moduli $T^2$, $T^3$. We can gain a deeper understanding of the reason why the heterotic and the type IIA constructions cannot be dual, if we consider the relation between these heterotic/type I constructions and the heterotic/type IIA, $S,T,U$ models with $N_V=H_H=0$ of Ref.\cite{gkp2}, which also possess a spontaneously broken $N=8$ supersymmetry. In that case, the spectrum is simply the truncation of the $N=8$ supergravity, and the restoration of the $N=8$ supersymmetry, which is non-perturbative from the heterotic point of view, being related to a motion in the dilaton field $S$, necessarily appears as perturbative on the type IIA side, where the field $S$ is the volume form of the K3 fibration. Indeed, on the type IIA side, the spontaneous breaking of the $N=8$ supersymmetry is due to a perturbative Scherk--Schwarz mechanism realized through freely acting projections, and is therefore directly related to the absence of extra massless states coming from orbifold twisted sectors. On the other hand, the heterotic/type I theory considered in this paper is an interesting example of (non-perturbative) spontaneous breaking of the $N=8$ supersymmetry, which, possessing an enlarged gauge group, cannot correspond to a truncation of the type IIA, $N=8$ spectrum, but is directly related to another phase of the M-theory. The two phases are interpolated by switching on and turning off ``Wilson lines'', which act as freely acting $Z_2$ projections twisting all the gauge bosons of the two ``Ho\v{r}ava--Witten'' planes. As is clear from the partition function given in Eqs. (4.2)--(4.8) of Ref.\cite{gkp2}, they appear on the heterotic side as ``$N=4$'' Wilson lines $Y$, which act on the twisted coordinates, corresponding to the hypermultiplets. Switching on/off these Wilson lines involves a blowing up of the moduli frozen at the fixed points of $T^4 \big/ Z_2$ and a motion to another $T^4 \big/ Z_2$ singularity\footnote{This necessarily involves passing through a $N=4$ phase of the model, so that the results of Ref.\cite{gkp2} for the $R^2$ correction cannot be used to rule out the term ${\cal E}(2\tau_S,2\tau_S',U)$ in Eq.(\ref{tr}).}. Such a motion is on the other hand non-perturbative from the type IIA point of view, the moduli associated to $Y$ including the dual of the type II dilaton field. The type II constructions of Section 2, on the other hand, correspond to a limit in the moduli space in which the $N=8$ supersymmetry is not spontaneously broken; in this limit, according to Ref.\cite{adds}, on the type I$^{\prime}$ picture, the combined action of the $Z_2$ which acts as a translation on the 11-th coordinate of M-theory, breaking $N=8$ to $N=4$, and that of the $Z_2$ which further breaks to $N=2$, is no longer free. {}From a perturbative point of view, the region of the moduli space corresponding to this limit is achieved at the above specified corners in the space of moduli $T^2, T^3$ (or $T^{(\rm Het)},U^{(\rm Het)}$), where the two theories can be connected. The above discussion is sketched in Fig. 1.\\ \begin{figure}[htb] \begin{center} \makebox[8cm]{ \epsfxsize=14cm \epsfysize=10cm \epsfbox{type1.eps} } \end{center} \vspace{1.4cm} \refstepcounter{figure} \parbox{15cm}{\hspace{1.55cm} \parbox{13cm}{ Figure \thefigure : The connections between the $N=2$ models and M-theory. ({\bf N}){\bf F} indicates a (non-)freely acting orbifold projection. With $(N_V,N_H)$ we indicate the number of vector and hypermultiplets of the ``twisted sector''. } } \label{figuretype1} \end{figure} \vskip 0.3cm \setcounter{section}{6} \setcounter{equation}{0} \section*{\normalsize{\bf 6. Conclusions}} In this paper we investigated the connections between several four-dimensional string constructions with the same massless spectrum, namely an $N=2$ supergravity with $3+N_V$ vector multiplets and $4+N_H$ hypermultiplets, and $N_V=N_H=16$. This spectrum is obtained via type IIA/B, heterotic and type I orbifold compactifications. We found evidence that the two type I constructions with spontaneous breaking of supersymmetry, presented in Ref.\cite{adds}, namely the $N=2$ ``Scherk--Schwarz'' and ``M-theory'' breaking models, indeed constitute two phases of the same theory, and are non-perturbatively related by a motion in the field $S'$, which parametrizes the coupling constant of the gauge fields of the D5-branes sector. This relation appears as perturbative on the heterotic dual construction. Collecting the knowledge coming from the heterotic model and the type I duals, we got some insight into the non-perturbative aspects of this theory. In particular, we discovered the existence of a non-perturbative super-Higgs phenomenon responsible for the spontaneous breaking of the $N=8$ supersymmetry. This is consistent with the interpretation of the theory as due to a ``Scherk--Schwarz'' mechanism applied to the 11-th dimension of the M-theory. This mechanism is also responsible for the breaking of the $SL(2,Z)$, Montonen--Olive $S$-duality, to a $\Gamma(2)$ subgroup, and it reflects on the dilaton dependence of string corrections to effective coupling constants, as the gravitational ones we considered. On the other hand, these constructions do not possess type II duals. Indeed we show that the type IIA, type II asymmetric orbifolds with the same massless spectrum actually correspond to a different phase of the M-theory. The two phases are connected at certain corners in the moduli space. \vskip 1.cm \centerline{\bf Acknowledgements} \noindent We thank R. Blumenhagen, E. Dudas, B. K\"{o}rs, A. Miemiec, H. Partouche, P.M. Petropoulos, A. Sagnotti and D. Smith for valuable discussions. This work was partially supported by the EEC under the contract TMR-ERBFMRX-CT96-0045. \noindent
\section{A MODEL INDEPENDENT DISCUSSION} \subsection{Introduction} CP violation arises naturally in the three generation Standard Model. The CP violation that has been measured in neutral $K$-meson decays ($\varepsilon_K$ and $\varepsilon^\prime_K$) is accommodated in the Standard Model in simple way \cite{KoMa}. Yet, CP violation is one of the least tested aspects of the Standard Model. The value of the $\varepsilon_K$ parameter \cite{CCFT}\ as well as bounds on other CP violating parameters (most noticeably, the electric dipole moments of the neutron, $d_N$, and of the electron, $d_e$) can be accounted for in models where CP violation has features that are very different from the Standard Model ones. It is unlikely that the Standard Model provides the complete description of CP violation in nature. First, it is quite clear that there exists New Physics beyond the Standard Model. Almost any extension of the Standard Model has additional sources of CP violating effects. In addition there is a great puzzle in cosmology that relates to CP violation, and that is the baryon asymmetry of the universe \cite{Sakh}. Theories that explain the observed asymmetry must include new sources of CP violation \cite{CKN}: the Standard Model cannot generate a large enough matter-antimatter imbalance to produce the baryon number to entropy ratio observed in the universe today \cite{FaSh,Gave,HuSa}. In the near future, significant new information on CP violation will be provided by various experiments. The main source of information will be measurements of CP violation in various $B$ decays, particularly neutral $B$ decays into final CP eigenstates \cite{BCP,BiSa,DuRo}. Another piece of valuable information might come from a measurement of the $K_L\rightarrow\pi^0\nu\bar\nu$ decay \cite{Litt,BuBuB,BuBun,GrNi}. For the first time, the pattern of CP violation that is predicted by the Standard Model will be tested. Basic questions such as whether CP is an approximate symmetry in nature will be answered. It could be that the scale where new CP violating sources appear is too high above the Standard Model scale ({\it e.g.} the GUT scale) to give any observable deviations from the Standard Model predictions. In such a case, the outcome of the experiments will be a (frustratingly) successful test of the Standard Model and a significant improvement in our knowledge of the CKM matrix. A much more interesting situation will arise if the new sources of CP violation appear at a scale that is not too high above the electroweak scale. Then they might be discovered in the forthcoming experiments. Once enough independent observations of CP violating effects are made, we will find that there is no single choice of CKM parameters that is consistent with all measurements. There may even be enough information in the pattern of the inconsistencies to tell us something about the nature of the new physics contributions \cite{NiSia,DLN,NiQuR,GrLoR}. The aim of this review is to explain the theoretical tools with which we will analyze new information about CP violation. In this chapter, we give a brief, model-independent discussion of CP violating observables. In the next chapter, we discuss CP violation in the Standard Model. In the third chapter we briefly explain why CP violation is a powerful probe of new physics. In the last chapter, we describe CP violation in Supersymmetric models. This discussion enables us to elucidate the uniqueness of the Standard Model description of CP violation and how little it has been tested so far. It further demonstrates how the information from CP violation can help us probe in detail models of New Physics. \subsection{Neutral Meson Mixing} Much of the exciting CP violation in meson decays is related to neutral meson mixing. Before we focus on CP violation, we briefly discuss then the physics and formalism of neutral meson mixing. We refer specifically to the neutral $B$ meson system, but most of our discussion applies equally well to the neutral $K$, $B_s$ and $D$ meson systems. Our phase convention for the CP transformation law of the neutral $B$ mesons is defined by \begin{equation}\label{phacon} {\rm CP}|{B^0}\rangle=\omega_B|{\bar B^0}\rangle,\ \ \ {\rm CP}|{\bar B^0}\rangle=\omega_B^*|{B^0}\rangle,\ \ \ (|\omega_B|=1). \end{equation} Physical observables do not depend on the phase factor $\omega_B$. An arbitrary linear combination of the neutral $B$-meson flavor eigenstates, \begin{equation}\label{defab} a|{B^0}\rangle+b|{\bar B^0}\rangle, \end{equation} is governed by a time-dependent Schr\"odinger equation, \begin{equation}\label{Schro} i{d\over dt}\pmatrix{a\cr b\cr}=H\pmatrix{a\cr b\cr} \equiv\left(M-{i\over2}\Gamma\right)\pmatrix{a\cr b\cr}, \end{equation} for which $M$ and $\Gamma$ are $2\times2$ Hermitian matrices. The off-diagonal terms in these matrices, $M_{12}$ and $\Gamma_{12}$, are particularly important in the discussion of mixing and CP violation. $M_{12}$ is the dispersive part of the transition amplitude from $B^0$ to $\bar B^0$. In the Standard Model it arises only at order $g^4$. In the language of quark diagrams, the leading contribution is from box diagrams. At sufficiently high loop momentum, $k\gg\Lambda_{\rm QCD}$, these diagrams are a very good approximation to the Standard Model contribution to $M_{12}$. This, or any other contribution from heavy intermediate states from new physics, is the {\it short distance} contribution. For small loop momenta, $k\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\; 1$ GeV, we do not expect quark hadron duality to hold. The box diagram is a poor approximation to the contribution from light intermediate states, namely to {\it long distance} contributions. Fortunately, in the $B$ and $B_s$ systems, the long distance contributions are expected to be negligible. (This is not the case for $K$ and $D$ mesons. Consequently, it is difficult to extract useful information from the measurement of $\Delta m_K$ and from the bound on $\Delta m_D$.) $\Gamma_{12}$ is the absorptive part of the transition amplitude. Since the cut of a diagram always involves on-shell particles and thus long distance physics, the cut of the quark box diagram is a poor approximation to $\Gamma_{12}$. However, it does correctly give the suppression from small electroweak parameters such as the weak coupling. In other words, though the hadronic uncertainties are large and could change the result by order $50\%$, the cut in the box diagram is expected to give a reasonable order of magnitude estimate of $\Gamma_{12}$. (For $\Gamma_{12}(B_s)$ it has been shown that local quark-hadron duality holds exactly in the simultaneous limit of small velocity and large number of colors. We thus expect an uncertainty of ${\cal O}(1/N_C)\sim30\%$ \cite{Alek,Bene}. For $\Gamma_{12}(B_d)$ the small velocity limit is not as good an approximation but an uncertainty of order 50\% still seems a reasonable estimate.) New physics is not expected to affect $\Gamma_{12}$ significantly because it usually takes place at a high energy scale and is relevant to the short distance part only. The light $B_L$ and heavy $B_H$ mass eigenstates are given by \begin{equation}\label{defqp} |{B_{L,H}}\rangle=p|{B^0}\rangle\pm q|{\bar B^0}\rangle. \end{equation} The complex coefficients $q$ and $p$ obey the normalization condition $|q|^2+|p|^2=1$. Note that $\arg(q/p^*)$ is just an overall common phase for $|B_L\rangle$ and $|B_H\rangle$ and has no physical significance. The mass difference and the width difference between the physical states are given by \begin{equation}\label{DelmG} \Delta m\equiv M_H-M_L,\ \ \ \Delta\Gamma\equiv\Gamma_H-\Gamma_L. \end{equation} Solving the eigenvalue equation gives \begin{equation}\label{eveq} (\Delta m)^2-{1\over4}(\Delta\Gamma)^2= (4|M_{12}|^2-|\Gamma_{12}|^2),\ \ \ \ \ \Delta m\Delta\Gamma=4{\cal R}e(M_{12}\Gamma_{12}^*), \end{equation} \begin{equation}\label{solveqp} {q\over p}=-{2M_{12}^*-i\Gamma_{12}^*\over \Delta m-{i\over2}\Delta\Gamma}=-{\Delta m-{i\over2}\Delta\Gamma \over 2M_{12}-i\Gamma_{12}}. \end{equation} In the $B$ system, $|\Gamma_{12}|\ll|M_{12}|$ (see discussion below), and then, to leading order in $|\Gamma_{12}/M_{12}|$, (\ref{eveq}) and (\ref{solveqp}) can be written as \begin{equation}\label{eveqB} \Delta m_B=2|M_{12}|,\ \ \ \Delta\Gamma_B=\ 2{\cal R}e(M_{12}\Gamma_{12}^*)/|M_{12}|, \end{equation} \begin{equation}\label{solveqpB} {q\over p}=-{M_{12}^*\over|M_{12}|}. \end{equation} \subsection{CP Violation in Neutral Meson Mixing} To discuss CP violation in mixing (see below), it is useful to write (\ref{solveqpB}) to first order in $|\Gamma_{12}/M_{12}|$: \begin{equation}\label{solveqpC} {q\over p}=-{M_{12}^*\over|M_{12}|}\left[1-{1\over2} {\cal I}m\left({\Gamma_{12}\over M_{12}}\right)\right]. \end{equation} To discuss CP violation in decay (see below), we need to consider decay amplitudes. The CP transformation law for a final state $f$ is \begin{equation}\label{phaconf} {\rm CP}|{f}\rangle=\omega_f|{\bar f}\rangle,\ \ \ {\rm CP}|{\bar f}\rangle=\omega_f^*|{f}\rangle,\ \ \ (|\omega_f|)=1. \end{equation} For a final CP eigenstate $f=\bar f=f_{{\rm CP}}$, the phase factor $\omega_f$ is replaced by $\eta_{f_{{\rm CP}}}=\pm1$, the CP eigenvalue of the final state. We define the decay amplitudes $A_f$ and $\bar A_f$ according to \begin{equation}\label{defAf} A_f=\vev{f|{\cal H}_d|B^0},\ \ \ \bar A_f=\vev{f|{\cal H}_d|\bar B^0}, \end{equation} where ${\cal H}_d$ is the decay Hamiltonian. To discuss CP violation in the interference of decays with and without mixing (see below), we introduce a complex quantity $\lambda_f$ defined by \begin{equation}\label{deflam} \lambda_f\ =\ {q\over p}\ {\bar A_f\over A_f}. \end{equation} We further define the CP transformation law for the quark fields in the Hamiltonian (a careful treatment of CP conventions can be found in \cite{BLS}): \begin{equation}\label{CPofq} q\ \rightarrow\ \omega_q\bar q,\ \ \ \bar q\ \rightarrow\ \omega_q^*q,\ \ \ (|\omega_q|=1). \end{equation} The effective Hamiltonian that is relevant to $M_{12}$ is of the form \begin{equation}\label{Hbtwo} H^{\Delta b=2}_{\rm eff}\propto e^{+2i\phi_B}\left[\bar d\gamma^\mu(1-\gamma_5)b\right]^2 +e^{-2i\phi_B}\left[\bar b\gamma^\mu(1-\gamma_5)d\right]^2, \end{equation} where $2\phi_B$ is a CP violating (weak) phase. (We use the Standard Model $V-A$ amplitude, but the results can be generalized to any Dirac structure.) For the $B$ system, where $|\Gamma_{12}|\ll |M_{12}|$, this leads to \begin{equation}\label{qpforB} q/p=\omega_B\omega_b^*\omega_d e^{-2i\phi_B}. \end{equation} (We implicitly assumed that the vacuum insertion approximation gives the correct sign for $M_{12}$. In general, there is a sign($B_B$) factor on the right hand side of (\ref{qpforB}) \cite{GKN}.) To understand the phase structure of decay amplitudes, we take as an example the $b\rightarrow q\bar qd$ decay ($q=u$ or $c$). The decay Hamiltonian is of the form \begin{equation}\label{Hdecay} H_d\propto e^{+i\phi_f}\left[\bar q\gamma^\mu(1-\gamma_5)d\right] \left[\bar b\gamma_\mu(1-\gamma_5)q\right] +e^{-i\phi_f}\left[\bar q\gamma^\mu(1-\gamma_5)b\right] \left[\bar d\gamma_\mu(1-\gamma_5)q\right], \end{equation} where $\phi_f$ is the appropriate weak phase. (Again, for simplicity we use a $V-A$ structure, but the results hold for any Dirac structure.) Then \begin{equation}\label{AbarA} \bar A_{\bar f}/A_f=\omega_f\omega_B^*\omega_b\omega_d^* e^{-2i\phi_f}. \end{equation} Eqs. (\ref{qpforB}) and (\ref{AbarA}) together imply that for a final CP eigenstate, \begin{equation}\label{lamfCP} \lambda_{f_{{\rm CP}}}=\eta_{f_{{\rm CP}}}e^{-2i(\phi_B+\phi_f)}. \end{equation} \subsection{The Three Types of CP Violation in Meson Decays} There are three different types of CP violation in meson decays: \begin{itemize} \item[(i)] CP violation in mixing, which occurs when the two neutral mass eigenstate admixtures cannot be chosen to be CP-eigenstates; \item[(ii)] CP violation in decay, which occurs in both charged and neutral decays, when the amplitude for a decay and its CP-conjugate process have different magnitudes; \item[(iii)] CP violation in the interference of decays with and without mixing, which occurs in decays into final states that are common to $B^0$ and $\bar B^0$. (It often occurs in combination with the other two types but there are cases when, to an excellent approximation, it is the only effect.) \end{itemize} (In cascade decays \cite{Azim,AzDu,KaSt,KaCA}, there appears a fourth type of CP violation \cite{MeSi,ASS}. We do not discuss this type of CP violation here.) {\bf (i) CP violation in mixing:} \begin{equation}\label{inmixin} |q/p|\neq1. \end{equation} This results from the mass eigenstates being different from the CP eigenstates, and requires a relative phase between $M_{12}$ and $\Gamma_{12}$. For the neutral $B$ system, this effect could be observed through the asymmetries in semileptonic decays: \begin{equation}\label{mixexa} a_{\rm SL}={\Gamma(\bar B^0_{\rm phys}(t)\rightarrow\ell^+\nu X)- \Gamma(B^0_{\rm phys}(t)\rightarrow\ell^-\nu X)\over \Gamma(\bar B^0_{\rm phys}(t)\rightarrow\ell^+\nu X)+ \Gamma(B^0_{\rm phys}(t)\rightarrow\ell^-\nu X)}. \end{equation} In terms of $q$ and $p$, \begin{equation}\label{mixter} a_{\rm SL}={1-|q/p|^4\over1+|q/p|^4}. \end{equation} CP violation in mixing has been observed in the neutral $K$ system (${\cal R}e\ \varepsilon_K\neq0$). In the neutral $B$ system, the effect is expected to be small, $\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\;{\cal O}(10^{-2})$. The reason is that, model independently, the effect cannot be larger than ${\cal O}(\Delta\Gamma_B/\Delta m_B)$. The difference in width is produced by decay channels common to $B^0$ and $\bar B^0$. The branching ratios for such channels are at or below the level of $10^{-3}$. Since various channels contribute with differing signs, one expects that their sum does not exceed the individual level. Hence, we can safely assume that $\Delta\Gamma_B/\Gamma_B={\cal O}(10^{-2})$. On the other hand, it is experimentaly known that $\Delta m_B/\Gamma_B\approx0.7$. To calculate the deviation of $|q/p|$ from a pure phase (see (\ref{solveqpC})), \begin{equation}\label{Absqp} 1-\left|{q\over p}\right|={1\over2}{\cal I}m{\Gamma_{12}\over M_{12}}, \end{equation} one needs to calculate $M_{12}$ and $\Gamma_{12}$. This involves large hadronic uncertainties, in particular in the hadronization models for $\Gamma_{12}$. {\bf (ii) CP violation in decay:} \begin{equation}\label{indecay} |\bar A_{\bar f}/A_f|\neq1. \end{equation} This appears as a result of interference among various terms in the decay amplitude, and will not occur unless at least two terms have different weak phases and different strong phases. CP asymmetries in charged $B$ decays, \begin{equation}\label{decexa} a_{f}={\Gamma(B^+\rightarrow f^+)-\Gamma(B^-\rightarrow f^-)\over \Gamma(B^+\rightarrow f^+)+\Gamma(B^-\rightarrow f^-)}, \end{equation} are purely an effect of CP violation in decay. In terms of the decay amplitudes, \begin{equation}\label{decter} a_{f^\pm}={1-|\bar A_{f^-}/A_{f^+}|^2\over1+|\bar A_{f^-}/A_{f^+}|^2}. \end{equation} CP violation in decay has been observed in the neutral $K$ system (${\cal R}e\ \varepsilon_K^\prime\neq0$ \cite{NAto,Esto,KTeVep}). There are two types of phases that may appear in $A_f$ and $\bar A_{\bar f}$. Complex parameters in any Lagrangian term that contributes to the amplitude will appear in complex conjugate form in the CP-conjugate amplitude. Thus their phases appear in $A_f$ and $\bar A_{\bar f}$ with opposite signs. In the Standard Model these phases occur only in the CKM matrix which is part of the electroweak sector of the theory, hence these are often called ``weak phases''. The weak phase of any single term is convention dependent. However the difference between the weak phases in two different terms in $A_f$ is convention independent because the phase rotations of the initial and final states are the same for every term. A second type of phase can appear in scattering or decay amplitudes even when the Lagrangian is real. Such phases do not violate CP, since they appear in $A_f$ and $\bar A_{\bar f}$ with the same sign. Their origin is the possible contribution from intermediate on-shell states in the decay process, that is an absorptive part of an amplitude that has contributions from coupled channels. Usually the dominant rescattering is due to strong interactions and hence the designation ``strong phases'' for the phase shifts so induced. Again only the relative strong phases of different terms in a scattering amplitude have physical content, an overall phase rotation of the entire amplitude has no physical consequences. Thus it is useful to write each contribution to $A$ in three parts: its magnitude $A_i$; its weak phase term $e^{i\phi_i}$; and its strong phase term $e^{i\delta_i}$. Then, if several amplitudes contribute to $B\rightarrow f$, we have \begin{equation}\label{defAtoA} \left|{\bar A_{\bar f}\over A_f}\right|=\left|{\sum_i A_i e^{i(\delta_i-\phi_i)}\over \sum_i A_i e^{i(\delta_i+\phi_i)}}\right|. \end{equation} The magnitude and strong phase of any amplitude involve long distance strong interaction physics, and our ability to calculate these from first principles is limited. Thus quantities that depend only on the weak phases are much cleaner than those that require knowledge of the relative magnitudes or strong phases of various amplitude contributions, such as CP violation in decay. There is however a large literature and considerable theoretical effort that goes into the calculation of amplitudes and strong phases . In many cases we can only relate experiment to Standard Model parameters through such calculations. The techniques that are used are expected to be more accurate for $B$ decays than for $K$ decays, because of the larger $B$ mass, but theoretical uncertainty remains significant. The calculations generally contain two parts. First the operator product expansion and QCD perturbation theory are used to write any underlying quark process as a sum of local quark operators with well-determined coefficients. Then the matrix elements of the operators between the initial and final hadron states must be calculated. This is where theory is weakest and the results are most model dependent. Ideally lattice calculations should be able to provide accurate determinations for the matrix elements, and in certain cases this is already true, but much remains to be done. {\bf (iii) CP violation in the interference between decays with and without mixing:} \begin{equation}\label{ininter} |\lambda_{f_{{\rm CP}}}|=1,\ \ {\cal I}m\ \lambda_{f_{{\rm CP}}}\neq0. \end{equation} Any $\lambda_{f_{{\rm CP}}}\neq\pm1$ is a manifestation of CP violation. The special case (\ref{ininter}) isolates the effects of interest since both CP violation in decay (\ref{indecay}) and in mixing (\ref{inmixin}) lead to $|\lambda_{f_{{\rm CP}}}|\neq1$. For the neutral $B$ system, this effect can be observed by comparing decays into final CP eigenstates of a time-evolving neutral $B$ state that begins at time zero as $B^0$ to those of the state that begins as $\bar B^0$: \begin{equation}\label{intexa} a_{f_{{\rm CP}}}={\Gamma(\bar B^0_{\rm phys}(t)\rightarrow f_{{\rm CP}})- \Gamma(B^0_{\rm phys}(t)\rightarrow f_{{\rm CP}})\over \Gamma(\bar B^0_{\rm phys}(t)\rightarrow f_{{\rm CP}})+ \Gamma(B^0_{\rm phys}(t)\rightarrow f_{{\rm CP}})}. \end{equation} This time dependent asymmetry is given (for $|\lambda_{f_{{\rm CP}}}|=1$) by \begin{equation}\label{intters} a_{f_{{\rm CP}}}=-{\cal I}m\lambda_{f_{{\rm CP}}}\sin(\Delta m_B t). \end{equation} CP violation in the interference of decays with and without mixing has been observed for the neutral $K$ system (${\cal I}m\ \varepsilon_K\neq0$). It is expected to be an effect of ${\cal O}(1)$ in various $B$ decays. For such cases, the contribution from CP violation in mixing is clearly negligible. For decays that are dominated by a single CP violating phase (for example, $B\rightarrow\psi K_S$ and $K_L\rightarrow\pi^0\nu\bar\nu$), so that the contribution from CP violation in decay is also negligible, $a_{f_{\rm CP}}$ is cleanly interpreted in terms of purely electroweak parameters. Explicitly, ${\cal I}m\lambda_{f_{{\rm CP}}}$ gives the difference between the phase of the $B-\bar B$ mixing amplitude ($2\phi_B$) and twice the phase of the relevant decay amplitude ($2\phi_f$) (see eq. (\ref{lamfCP})): \begin{equation}\label{intCKM} {\cal I}m\lambda_{f_{{\rm CP}}}=-\eta_{f_{{\rm CP}}}\sin[2(\phi_B+\phi_f)]. \end{equation} \subsection{Indirect vs. Direct CP Violation} The terms indirect CP violation and direct CP violation are commonly used in the literature. While various authors use these terms with different meanings, the most useful definition is the following: \begin{itemize} \item[(i)] {\bf Indirect CP violation} refers to CP violation in meson decays where the CP violating phases can all be chosen to appear in $\Delta F=2$ (mixing) amplitudes. \item[(ii)] {\bf Direct CP violation} refers to CP violation in meson decays where some CP violating phases necessarily appear in $\Delta F=1$ (decay) amplitudes. \end{itemize} Examining eqs. (\ref{inmixin}) and (\ref{solveqp}), we learn that CP violation in mixing is a manifestation of indirect CP violation. Examining eqs. (\ref{indecay}) and (\ref{defAf}), we learn that CP violation in decay is a manifestation of direct CP violation. Examining eqs. (\ref{ininter}) and (\ref{deflam}), we learn that the situation concerning CP violation in the interference of decays with and without mixing is more subtle. For any single measurement of ${\cal I}m\lambda_f\neq0$, the relevant CP violating phase can be chosen by convention to reside in the $\Delta F=2$ amplitude ($\phi_f=0$, $\phi_B\neq0$ in the notation of eq. (\ref{lamfCP})), and then we would call it indirect CP violation. Consider, however, the CP asymmetries for two different final CP eigenstates (for the same decaying meson), $f_a$ and $f_b$. Then, a non-zero difference between ${\cal I}m\lambda_{f_a}$ and ${\cal I}m\lambda_{f_b}$ requires that there exists CP violation in $\Delta F=1$ processes ($\phi_{f_a}- \phi_{f_b}\neq0$), namely direct CP violation. Experimentally, both direct and indirect CP violation have been established. Below we will see that $\varepsilon_K$ signifies indirect CP violation while $\varepsilon^\prime_K$ signifies direct CP violation. Theoretically, most models of CP violation (including the Standard Model) have predicted that both types of CP violation exist. There is, however, one class of models, that is {\it superweak models} \cite{WolSW,WinSW,SWSW,WWSW}, that predict only indirect CP violation. The measurement of $\varepsilon^\prime_K\neq0$ has excluded this class of models. \subsection{The $\varepsilon_K$ and $\varepsilon^\prime_K$ Parameters} Historically, a different language from the one used by us has been employed to describe CP violation in $K\rightarrow\pi\pi$ and $K\rightarrow\pi\ell\nu$ decays. In this section we `translate' the language of $\varepsilon_K$ and $\varepsilon_K^\prime$ to our notations. Doing so will make it easy to understand which type of CP violation is related to each quantity. The two CP violating quantities measured in neutral $K$ decays are \begin{equation}\label{defetaij} \eta_{00}={\vev{\pi^0\pi^0|{\cal H}|K_L}\over\vev{\pi^0\pi^0|{\cal H}|K_S}},\ \ \ \eta_{+-}={\vev{\pi^+\pi^-|{\cal H}|K_L}\over\vev{\pi^+\pi^-|{\cal H}|K_S}}. \end{equation} Define for $(ij)=(00)$ or $(+-)$ \begin{equation}\label{epsamp} A_{ij}=\vev{\pi^i\pi^j|{\cal H}|K^0},\ \ \ \bar A_{ij}=\vev{\pi^i\pi^j|{\cal H}|\bar K^0}, \end{equation} \begin{equation}\label{epslam} \lambda_{ij}=\left({q\over p}\right)_K{\bar A_{ij}\over A_{ij}}. \end{equation} Then \begin{equation}\label{etapqA} \eta_{00}={1-\lambda_{00}\over1+\lambda_{00}},\ \ \ \eta_{+-}={1-\lambda_{+-}\over1+\lambda_{+-}}. \end{equation} The $\eta_{00}$ and $\eta_{+-}$ parameters get contributions from CP violation in mixing ($|(q/p)|_K\neq1$) and from the interference of decays with and without mixing (${\cal I}m\lambda_{ij}\neq0$) at ${\cal O}(10^{-3})$ and from CP violation in decay ($|\bar A_{ij}/A_{ij}|\neq1$) at ${\cal O}(10^{-6})$. There are two isospin channels in $K\rightarrow\pi\pi$ leading to final $(2\pi)_{I=0}$ and $(2\pi)_{I=2}$ states: \begin{equation}\label{twoisZ} \langle\pi^0\pi^0|=\sqrt{1\over3}\langle(\pi\pi)_{I=0}|- \sqrt{2\over3}\langle(\pi\pi)_{I=2}|, \end{equation} \begin{equation}\label{twoisT} \langle\pi^+\pi^-|=\sqrt{2\over3}\langle(\pi\pi)_{I=0}|+ \sqrt{1\over3}\langle(\pi\pi)_{I=2}|. \end{equation} The fact that there are two strong phases allows for CP violation in decay. The possible effects are, however, small (on top of the smallness of the relevant CP violating phases) because the final $I=0$ state is dominant (this is the $\Delta I=1/2$ rule). Defining \begin{equation}\label{defAI} A_I=\vev{(\pi\pi)_I|{\cal H}|K^0},\ \ \ \bar A_I=\vev{(\pi\pi)_I|{\cal H}|\bar K^0}, \end{equation} we have, experimentally, \begin{equation}\label{AtwoAzer} |A_2/A_0|\approx1/20. \end{equation} Instead of $\eta_{00}$ and $\eta_{+-}$ we may define two combinations, $\varepsilon_K$ and $\varepsilon^\prime_K$, in such a way that the possible effects of CP violation in decay (mixing) are isolated into $\varepsilon^\prime_K$ ($\varepsilon_K$). The experimental definition of the $\varepsilon_K$ parameter is \begin{equation}\label{defepsex} \varepsilon_K\equiv{1\over3}(\eta_{00}+2\eta_{+-}). \end{equation} To zeroth order in $A_2/A_0$, we have $\eta_{00}=\eta_{+-}=\varepsilon_K$. However, the specific combination (\ref{defepsex}) is chosen in such a way that the following relation holds to {\it first} order in $A_2/A_0$: \begin{equation}\label{defepsth} \varepsilon_K={1-\lambda_0\over1+\lambda_0}, \end{equation} where \begin{equation}\label{deflamz} \lambda_0=\left({q\over p}\right)_K\left({\bar A_0\over A_0}\right). \end{equation} Since, by definition, only one strong channel contributes to $\lambda_0$, there is indeed no CP violation in decay in (\ref{defepsth}). It is simple to show that ${\cal R}e\ \varepsilon_K\neq0$ is a manifestation of CP violation in mixing while ${\cal I}m\ \varepsilon_K\neq0$ is a manifestation of CP violation in the interference between decays with and without mixing. Since experimentally $\arg\varepsilon_K\approx\pi/4$, the two contributions are comparable. It is also clear that $\varepsilon_K\neq0$ is a manifestation of indirect CP violation: it could be described entirely in terms of a CP violating phase in the $M_{12}$ amplitude. The experimental definition of the $\varepsilon^\prime_K$ parameter is \begin{equation}\label{defepspex} \varepsilon^\prime_K\equiv{1\over3}(\eta_{+-}-\eta_{00}). \end{equation} The theoretical expression is \begin{equation}\label{defepspth} \varepsilon^\prime_K\approx{1\over6}(\lambda_{00}-\lambda_{+-}). \end{equation} Obviously, any type of CP violation which is independent of the final state does not contribute to $\varepsilon^\prime_K$. Consequently, there is no contribution from CP violation in mixing to (\ref{defepspth}). It is simple to show that ${\cal R}e\ \varepsilon^\prime_K\neq0$ is a manifestation of CP violation in decay while ${\cal I}m\ \varepsilon^\prime_K\neq0$ is a manifestation of CP violation in the interference between decays with and without mixing. Following our explanations in the previous section, we learn that $\varepsilon^\prime_K\neq0$ is a manifestation of direct CP violation: it requires $\phi_2-\phi_0\neq0$ (where $\phi_I$ is the CP violating phase in the $A_I$ amplitude defined in (\ref{defAI})). \section{CP VIOLATION IN THE STANDARD MODEL} \subsection{Introduction} Within the Standard Model, CP violation can only arise from the Yukawa interactions: \begin{equation}\label{Hint} -{\cal L}_Y=Y^d_{ij}{\overline {Q^I_{Li}}}\phi d^I_{Rj} +Y^u_{ij}{\overline {Q^I_{Li}}}\tilde\phi u^I_{Rj} +Y^\ell_{ij}{\overline {L^I_{Li}}}\phi\ell^I_{Rj}. \end{equation} The various fermion representations of $SU(3)_{\rm C}\times SU(2)_{\rm L} \times U(1)_{\rm Y}$ are denoted here by \begin{equation}\label{MSSMrep} Q_i^I(3,2)_{+1/6},\ \ \bar u_i^I(\bar3,1)_{-2/3},\ \ \bar d_i^I(\bar3,1)_{+1/3},\ \ L_i^I(1,2)_{-1/2},\ \ \bar\ell_i^I(1,1)_{+1}, \end{equation} and the Higgs representation is $\phi(1,2)_{+1/2}$ ($\tilde\phi=i\sigma_2\phi^*$). There are 27 complex parameters in the three Yukawa matrices, but not all of them are physical. If the Yukawa couplings are switched off, the Standard Model has (in addition to a discrete CP symmetry) a global $U(3)^5$ symmetry. But the subgroup of $U(1)_B\times U(1)_e\times U(1)_\mu\times U(1)_\tau$ remains a symmetry of the Standard Model even in the presence of non-zero Yukawa couplings. Since a $3\times3$ unitary matrix has three real and six imaginary parameters, we conclude that $15$ real and $26$ imaginary parameters in the Yukawa matrices are not physical, leaving twelve real and one imaginary physical parameters. It is easy to identify the physical parameters in the mass basis. Nine of the real parameters are the charged fermion masses. All other parameters are related to the CKM matrix $V_{\rm CKM}$, that is the quark mixing matrix that parametrizes the charged gauge boson interactions \cite{Cabi,KoMa}: \begin{equation}\label{Wmas} -{\cal L}_{W^\pm}={g\over\sqrt2}{\overline {u_{Li}}}\gamma^\mu (V_{\rm CKM})_{ij}d_{Lj} W_\mu^++{\rm h.c.}. \end{equation} The unitary $V_{\rm CKM}$ can be parametrized with three real mixing angles and a single phase. The single irremovable phase in the CKM matrix is the only source CP violation within the Standard Model. In the Wolfenstein parametrization of $V_{\rm CKM}$, the four mixing parameters are $(\lambda,A,\rho,\eta)$ with $\lambda=|V_{us}|=0.22$ playing the role of an expansion parameter and $\eta$ representing the CP violating phase \cite{WOLpar}: \begin{equation}\label{WCKM} V=\pmatrix{1-{\lambda^2\over2}&\lambda&A\lambda^3(\rho-i\eta)\cr -\lambda&1-{\lambda^2\over2}&A\lambda^2\cr A\lambda^3(1-\rho-i\eta)&-A\lambda^2&1\cr}+{\cal O}(\lambda^4). \end{equation} The fact that there is a single CP violating parameter in the SM can be seen also in another useful way. The unitarity of the CKM matrix leads to various relations among the matrix elements, {\it e.g.} \begin{equation}\label{Unitds} V_{ud}V_{us}^*+V_{cd}V_{cs}^*+V_{td}V_{ts}^*=0, \end{equation} \begin{equation}\label{Unitsb} V_{us}V_{ub}^*+V_{cs}V_{cb}^*+V_{ts}V_{tb}^*=0, \end{equation} \begin{equation}\label{Unitdb} V_{ud}V_{ub}^*+V_{cd}V_{cb}^*+V_{td}V_{tb}^*=0. \end{equation} Each of the three relations (\ref{Unitds})$-$(\ref{Unitdb}) requires the sum of three complex quantities to vanish and so can be geometrically represented in the complex plane as a triangle. These are ``the unitarity triangles", though the term ``unitarity triangle" is usually reserved for the relation (\ref{Unitdb}) only. It is a surprising feature of the CKM matrix that all unitarity triangles are equal in area. For any choice of $i,j,k,l=1,2,3$, one can define a quantity $J$ according to \cite{Jarl} \begin{equation}\label{defJ} {\cal I}m[V_{ij}V_{kl}V_{il}^*V_{kj}^*]=J\sum_{m,n=1}^3\epsilon_{ikm}\epsilon_{jln}. \end{equation} Then, the area of each unitarity triangle equals $|J|/2$ while the sign of $J$ gives the direction of the complex vectors around the triangles. CP is violated in the Standard Model only if $J\neq0$. The quantity $J$ can then be taken as the CP violating parameter of the SM. The area of the triangles is then related to the size of the Standard Model CP violation. The relation between Jarlskog's measure of CP violation $J$ and the Wolfenstein parameters $(\lambda,A,\eta)$ is given by \begin{equation}\label{JAre} J\simeq \lambda^6 A^2\eta. \end{equation} The rescaled unitarity triangle is derived from (\ref{Unitdb}) by (a) choosing a phase convention such that $(V_{cd}V_{cb}^*)$ is real, and (b) dividing the lengths of all sides by $|V_{cd}V_{cb}^*|$. Step (a) aligns one side of the triangle with the real axis, and step (b) makes the length of this side 1. The form of the triangle is unchanged. Two vertices of the rescaled unitarity triangle are thus fixed at (0,0) and (1,0). The coordinates of the remaining vertex correspond to the Wolfenstein parameters $(\rho,\eta)$. Depicting the rescaled unitarity triangle in the $(\rho,\eta)$ plane, the lengths of the two complex sides are \begin{equation}\label{RbRt} R_u\equiv\sqrt{\rho^2+\eta^2}={1\over\lambda} \left|{V_{ub}\over V_{cb}}\right|,\ \ \ R_t\equiv\sqrt{(1-\rho)^2+\eta^2}={1\over\lambda} \left|{V_{td}\over V_{cb}}\right|. \end{equation} The three angles of the unitarity triangle are denoted by $\alpha,\beta$ and $\gamma$ \cite{DDGN}: \begin{equation}\label{abcangles} \alpha\equiv\arg\left[-{V_{td}V_{tb}^*\over V_{ud}V_{ub}^*}\right],\ \ \ \beta\equiv\arg\left[-{V_{cd}V_{cb}^*\over V_{td}V_{tb}^*}\right],\ \ \ \gamma\equiv\arg\left[-{V_{ud}V_{ub}^*\over V_{cd}V_{cb}^*}\right]. \end{equation} They are physical quantities and, we will soon see, can be independently measured by CP asymmetries in $B$ decays. To make predictions for future measurements of CP violating observables, we need to find the allowed ranges for the CKM phases. There are three ways to determine the CKM parameters (see {\it e.g.} \cite{HarNir}): \begin{itemize} \item[(i)] {\bf Direct measurements} are related to SM tree level processes. At present, we have direct measurements of $|V_{ud}|$, $|V_{us}|$, $|V_{ub}|$, $|V_{cd}|$, $|V_{cs}|$, $|V_{cb}|$ and $|V_{tb}|$. \item[(ii)] {\bf CKM Unitarity} ($V_{\rm CKM}^\dagger V_{\rm CKM}={\bf 1}$) relates the various matrix elements. At present, these relations are useful to constrain $|V_{td}|$, $|V_{ts}|$, $|V_{tb}|$ and $|V_{cs}|$. \item[(iii)] {\bf Indirect measurements} are related to SM loop processes. At present, we constrain in this way $|V_{tb}V_{td}|$ (from $\Delta m_B$ and $\Delta m_{B_s}$) and $\delta_{\rm KM}$ (from $\varepsilon_K$). \end{itemize} When all available data is taken into account, we find \cite{NirCern,PlSc,Plas,GNPS,BaBar}: \begin{equation}\label{recons} -0.15\leq\rho\leq+0.35,\ \ \ +0.20\leq\eta\leq+0.45, \end{equation} \begin{equation}\label{abcons} 0.4\leq\sin2\beta\leq0.8,\ \ \ -0.9\leq\sin2\alpha\leq1.0,\ \ \ 0.23\leq\sin^2\gamma\leq1.0. \end{equation} Of course, there are correlations between the various parameters. The full information in the $(\rho,\eta)$ and in the $(\sin2\alpha, \sin2\beta)$ plane is given in fig. 1. \begin{figure} \centerline{$(a)$} \centerline{ \psfig{file=ut98.ps,width=370pt,bbllx=0pt,bblly=410pt,bburx=612pt,bbury=740pt }} \centerline{$(b)$} \centerline{ \psfig{file=ss98.ps,width=320pt,bbllx=0pt,bblly=160pt,bburx=612pt,bbury=653pt }} \caption{The present allowed range $(a)$ in the $\rho-\eta$ plane and $(b)$ in the $\sin2\alpha-\sin2\beta$ plane using constraints from $|V_{cb}|$, $|V_{ub}/V_{cb}|$, $\Delta m_{B_d}$, $\varepsilon_K$ and $\Delta m_{B_s}$. For the methods and the data used in this analysis, see ref. [45]. \label{presentRE}} \end{figure} Since the Standard Model contains only a single independent CP-violating phase, all possible CP-violating effects in this theory are very closely related. Consequently, the pattern of CP-violations in $B$ decays is strongly constrained. The goal of $B$ factories is to test whether this pattern occurs in Nature. If no new physics is discovered, we will have a much improved determionation of the CKM parameters. The impact of measurements of CP asymmetries in $B$ decays and of the rate of $K_L\rightarrow\pi\nu\nu$ is demonstrated in fig. 2. \begin{figure} \centerline{$(a)$} \centerline{ \psfig{file=ut180fb.ps,width=370pt,bbllx=0pt,bblly=410pt,bburx=612pt,bbury=700pt }} \centerline{$(b)$} \centerline{ \psfig{file=kpi.ps,width=370pt,bbllx=0pt,bblly=260pt,bburx=612pt,bbury=700pt }} \caption{(a) The effect of the CP asymmetries in $B$ decays on the constraints in the $\rho-\eta$ plane. We use hypothetical ranges appropriate for 180 fb$^{-1}$ integrated luminosity in the $B$-factories. (b) The effect of the $K\rightarrow\pi\nu\bar\nu$ measurements on the constraints in the $\rho-\eta$ plane. We use the hypothetical ranges BR($K^+\rightarrow\pi^+\nu\bar\nu)=(1.0\pm0.1)\times10^{-10}$, BR($K_L\rightarrow\pi^0\nu\bar\nu)=(3.0\pm0.3)\times10^{-11}$. For the methods and the data used in this analysis, see ref. [45]. \label{kpnnRE}} \end{figure} \subsection{CP Violation in Semileptonic Decays of Neutral $B$ Mesons} In the $B_d$ system we expect model independently that $\Gamma_{12}\ll M_{12}$. Moreover, within the SM and assuming that the box diagram (with a cut) is appropriate to estimate $\Gamma_{12}$, we can actually calculate the two quantities from quark diagrams \cite{BKUS}. The calculation gives \begin{equation}\label{SMGtoM} {\Gamma_{12}\over M_{12}}=-{3\pi\over2}{1\over f_2(m_t^2/m_W^2)} {m_b^2\over m_t^2}\left(1+{8\over3}{m_c^2\over m_b^2} {V_{cb}V_{cd}^*\over V_{tb}V_{td}^*}\right). \end{equation} This confirms our order of magnitude estimate, $|\Gamma_{12}/M_{12}| \lsim10^{-2}$. The deviation of $|q/p|$ from unity is proportional to ${\cal I}m(\Gamma_{12}/M_{12})$ which is even further suppressed by another order of magnitude: \begin{equation}\label{AbsqpSM} 1-\left|{q\over p}\right|={1\over2}{\cal I}m{\Gamma_{12}\over M_{12}}= {4\pi\over f_2(m_t^2/m_W^2)}{m_c^2\over m_t^2}{J\over|V_{tb}V_{td}^*|^2} \sim10^{-3}. \end{equation} Note that the suppression comes from the $(m_c^2/m_t^2)$ factor. The last term is the ratio of the area of the unitarity triangle to the length of one of its sides squared, so it is ${\cal O}(1)$. In contrast, for the $B_s$ system, where (\ref{AbsqpSM}) holds except that $V_{td}$ is replaced by $V_{ts}$, there is an additional suppression from $J/|V_{tb}V_{ts}^*|^2\sim10^{-2}$ (see the corresponding unitarity triangle). The above estimate of CP violation in mixing suffers from large uncertainties (of order 30\% \cite{Alek}\ or even higher \cite{WolGot}) related to the use of a quark diagram to describe $\Gamma_{12}$ . We remind the reader that (\ref{AbsqpSM}) gives the leading contribution to the CP asymmetry in semileptonic decays: \begin{equation}\label{SMSLB} a_{\rm SL}\approx2(1-|q/p|). \end{equation} The smallness of the predicted asymmetry will make its measurement a rather challenging task. \subsection{CP Violation in Hadronic Decays of Neutral $B$ Mesons} In the previous subsection we estimated the effect of CP violation in mixing to be of ${\cal O}(10^{-3})$ within the Standard Model, and $\leq{\cal O}(|\Gamma_{12}/M_{12}|)\sim10^{-2}$ model independently (for a recent discussion, see \cite{RaSu}). In semileptonic decays, CP violation in mixing is the leading effect and therefore it can be measured through $a_{\rm SL}$. In purely hadronic $B$ decays, however, CP violation in decay and in the interference of decays with and without mixing is $\geq {\cal O}(10^{-2})$. We can therefore safely neglect CP violation in mixing in the following discussion and use \begin{equation}\label{qpSM} {q\over p}=-{M_{12}^*\over|M_{12}|}= {V_{tb}^*V_{td}\over V_{tb}V_{td}^*}\omega_B. \end{equation} (From here on we omit the convention-dependent quark phases $\omega_q$ defined in eq. (\ref{CPofq}). Our final expressions for physical quantities are of course unaffected by such omission.) A crucial question is then whether CP violation in decay is comparable to the CP violation in the interference of decays with and without mixing or negligible. In the first case, we can use the corresponding charged $B$ decays to observe effects of CP violation in decay. In the latter case, CP asymmetries in neutral $B$ decays are subject to clean theoretical interpretation: we will either have precise measurements of CKM parameters or be provided with unambiguous evidence for new physics. The question of the relative size of CP violation in decay can only be answered on a channel by channel basis, which is what we do in this section. Most channels have contributions from both tree- and three types of penguin-diagrams, the latter classified according to the identity of the quark in the loop, as diagrams with different intermediate quarks may have both different strong phases and different weak phases \cite{BSS}. On the other hand, the subdivision of tree processes into spectator, exchange and annihilation diagrams is unimportant in this respect since they all carry the same weak phase. While quark diagrams can be easily classified in this way, the description of $B$ decays is not so neatly divided into tree and penguin contributions once long distance physics effects are taken into account. Rescattering processes can change the quark content of the final state and confuse the identification of a contribution. There is no physical distinction between rescattered tree diagrams and long-distance contributions to the cuts of a penguin diagram. While these issues complicate estimates of various rates, they can always be avoided in describing the weak phase structure of $B$-decay amplitudes. The decay amplitudes for $b\rightarrow q \bar q q^\prime$ can always be written as a sum of three terms with definite CKM coefficients: \begin{equation}\label{defCKM} A(q \bar q q^\prime)= V_{tb}V^*_{tq^\prime}P^t_{q^\prime} + V_{cb}V_{cq^\prime}^*(T_{c\bar c q^\prime}\delta_{qc} + P^c_{q^\prime})+ V_{ub}V_{uq^\prime}^*(T_{u \bar u q^\prime}\delta_{qu} + P^u_{q^\prime}). \end{equation} Here $P $ and $T$ denote contributions from tree and penguin diagrams, excluding the CKM factors. As they stand, the $P$ terms are not well defined because of the divergences of the penguin diagrams. Only differences of penguin diagrams are finite and well defined. However already we see that diagrams that can be mixed by rescattering effects always appear with the same CKM coefficients and hence that a separation of these terms is not needed when discussing weak phase structure. Now it is useful to use eqs. (\ref{Unitsb}) and (\ref{Unitdb}) to eliminate one of the three terms, by writing its CKM coefficient as minus the sum of the other two. In the case of $q \bar q s$ decays it is convenient to remove the $V_{tb} V^*_{ts}$ term. Then \begin{equation}\label{ccstype} A(c\bar c s)= V_{cb}V_{cs}^*(T_{c\bar cs}+P^c_s-P^t_s) +V_{ub}V^*_{us}(P^u_s-P^t_s), \end{equation} \begin{equation}\label{uustype} A(u\bar u s)= V_{cb}V_{cs}^*(P^c_s-P^t_s) +V_{ub}V^*_{us}(T_{u\bar us}+P^u_s-P^t_s), \end{equation} \begin{equation}\label{ssstype} A(s\bar s s)= V_{cb}V_{cs}^*(P^c_s-P^t_s)+V_{ub}V^*_{us}(P^u_s- P^t_s). \end{equation} In these expressions only differences of penguin contributions occur, which makes the cancellation of the ultraviolet divergences of these diagrams explicit. Furthermore, the second term has a CKM coefficient that is much smaller, by ${\cal O}(\lambda^2)$, than the first. Hence this grouping is useful in classifying the expected CP violation in decay. In the case of $q \bar q d $ decays the three CKM coefficients are of similar magnitude. The convention is then to retain the $V_{tb}V^*_{td}$ term because, in the Standard Model, the phase difference between this weak phase and half the mixing weak phase is zero. Thus only one unknown weak phase enters the calculation of the interference between decays with and without mixing. We can choose to eliminate which of the other terms does not have a tree contribution. In the cases $q=s$ or $d$, since neither has a tree contribution either term can be removed. Thus we write \begin{equation}\label{ccdtype} A(c\bar c d)=V_{tb}V_{td}^*(P^t_d-P^u_d) + V_{cb}V_{cd}^*(T_{c\bar c d}+P^c_d-P^u_d), \end{equation} \begin{equation}\label{uudtype} A(u\bar u d)=V_{tb}V_{td}^*(P^t_d-P^c_d) +V_{ub}V_{ud}^*(T_{u\bar u d}+P^u_d - P^c_d), \end{equation} \begin{equation}\label{ssdtype} A(s\bar s d)=V_{tb}V_{td}^*(P^t_d-P^u_d)+V_{cb}V_{cd}^*(P^c_d-P^u_d). \end{equation} Again only differences of penguin amplitudes occur. Furthermore the difference of penguin terms that occurs in the second term would vanish if the charm and up quark masses were equal, and thus is GIM suppressed \cite{GIM}. However, even in modes with no tree contribution, $(s\bar s d)$, the interference of the two terms can still give significant CP violation. The penguin processes all involve the emission of a neutral boson, either a gluon (strong penguins) or a photon or $Z$ boson (electroweak penguins). Excluding the CKM coefficients, the ratio of the contribution from the difference between a top and light quark strong penguin diagram to the contribution from a tree diagram is of order \begin{equation}\label{pengtree} r_{PT} = { P^t-P^{light}\over T_{q \bar q q^\prime }}\approx{\alpha_s\over12\pi}\ln{m_t^2\over m_b^2}. \end{equation} This is a factor of ${\cal O}(0.03)$. However this estimate does not include the effect of hadronic matrix elements, which are the probability factor to produce a particular final state particle content from a particular quark content. Since this probability differs for different kinematics, color flow and spin structures, it can be different for tree and penguin contributions and may partially compensate the coupling constant suppression of the penguin term. Recent CLEO results on $BR(B\rightarrow K\pi)$ and $BR(B\rightarrow\pi\pi)$ \cite{CLEOpen}\ suggest that the matrix element of penguin operators is indeed enhanced compared to that of tree operators. The enhancement could be by a factor of a few, leading to \begin{equation}\label{pentre} r_{PT}\sim\lambda^2-\lambda. \end{equation} (Note that $r_{PT}$ does not depend on the CKM parameters. We use powers of the Wolfenstein parameter $\lambda$ to quantify our estimate for $r_{PT}$ is order to simplify the comparison between the size of CP violation in decay and CP violation in the interference between decays with and without mixing.) Electroweak penguin difference terms are even more suppressed since they have an $\alpha_{\rm EM}$ or $\alpha_W$ instead of the $\alpha_s$ factor in (\ref{pengtree}), but certain $Z$-contributions are enhanced by the large top quark mass and so can be non-negligible. We thus classify $B$ decays into four classes. Classes (i) and (ii) are expected to have relatively small CP violation in decay and hence are particularly interesting for extracting CKM parameters from interference of decays with and without mixing. In the remaining two classes, CP violation in decay could be significant and the neutral decay asymmetries cannot be cleanly interpreted in terms of CKM phases. \begin{itemize} \item[(i)] Decays dominated by a single term: $b\rightarrow c\bar cs$ and $b\rightarrow s\bar ss$. The Standard Model cleanly predicts very small CP violation in decay: ${\cal O}(\lambda^4-\lambda^3)$ for $b\rightarrow c\bar cs$ and ${\cal O}(\lambda^2)$ for $b\rightarrow s\bar ss$. Any observation of large CP asymmetries in charged $B$ decays for these channels would be a clue to physics beyond the Standard Model. The corresponding neutral modes have cleanly predicted relationships between CKM parameters and the measured asymmetry from interference between decays with and without mixing. The modes $B\rightarrow\psi K$ and $B\rightarrow\phi K$ are examples of this class. \item[(ii)] Decays with a small second term: $b\rightarrow c\bar cd$ and $b\rightarrow u \bar ud$. The expectation that penguin-only contributions are supressed compared to tree contributions suggests that these modes will have small effects of CP violation in decay, of ${\cal O}(\lambda^2-\lambda)$, and an approximate prediction for the relationship between measured asymmetries in neutral decays and CKM phases can be made. Examples here are $B\rightarrow\ DD$ and $B\rightarrow\pi\pi$. \item[(iii)] Decays with a suppressed tree contribution: $b\rightarrow u \bar us$. The tree amplitude is suppressed by small mixing angles, $V_{ub}V_{us}$. The no-tree term may be comparable or even dominate and give large interference effects. An example is $B\rightarrow\rho K$. \item[(iv)] Decays with no tree contribution: $b\rightarrow s\bar s d$. Here the interference comes from penguin contributions with different charge 2/3 quarks in the loop and gives CP violation in decay that could be as large as 10\% \cite{Flei,BuFl}. An example is $B\rightarrow KK$. \end{itemize} Note that if the penguin enhancement is significant, then some of the decay modes listed in class (ii) might actually fit better in class (iii). For example, it is possible that $b\rightarrow u\bar ud$ decays have comparable contributions from tree and penguin amplitudes. On the other hand, this would also mean that some modes listed in class (iii) could be dominated by a single penguin term. For such cases an approximate relationship between measured asymmetries in neutral decays and CKM phases can be made. \subsection{CP Violation in the Interference Between $B$ Decays With and Without Mixing} Let us first discuss an example of class (i), $B\rightarrow\psi K_S$. A new ingredient in the analysis is the effect of $K-\bar K$ mixing. For decays with a single $K_S$ in the final state, $K-\bar K$ mixing is essential because $B^0\rightarrow K^0$ and $\bar B^0\rightarrow\bar K^0$, and interference is possible only due to $K-\bar K$ mixing. This adds a factor of \begin{equation}\label{qpK} \left({p\over q}\right)_K={V_{cs}V_{cd}^*\over V_{cs}^*V_{cd}}\omega_K^* \end{equation} into $(\bar A/A)$. The quark subprocess in $\bar B^0\rightarrow\psi\bar K^0$ is $b\rightarrow c\bar cs$ which is dominated by the $W$-mediated tree diagram: \begin{equation}\label{ApsiK} {\bar A_{\psi K_S}\over A_{\psi K_S}}= \eta_{\psi K_S} \left({V_{cb}V_{cs}^*\over V_{cb}^*V_{cs}}\right) \left({V_{cs}V_{cd}^*\over V_{cs}^*V_{cd}}\right)\omega_B^*. \end{equation} The CP-eigenvalue of the state is $\eta_{\psi K_S} = -1$. Combining (\ref{qpSM}) and (\ref{ApsiK}), we find \begin{equation}\label{lampsiK} \lambda(B\rightarrow\psi K_S)=-\left({V_{tb}^*V_{td}\over V_{tb}V_{td}^*}\right)\left({V_{cb}V_{cs}^*\over V_{cb}^*V_{cs}}\right) \left({V_{cd}^*V_{cs}\over V_{cd}V_{cs}^*}\right) \ \Longrightarrow\ {\cal I}m\lambda_{\psi K_S}=\sin(2\beta). \end{equation} The second term in (\ref{ccstype}) is of order $\lambda^2r_{PT}$ for this decay and thus eq. (\ref{lampsiK}) is clean of hadronic uncertainties to ${\cal O}(10^{-3})$. Consequently, this measurement can give the theoretically cleanest determination of a CKM parameter, even cleaner than the determination of $|V_{us}|$ from $K\rightarrow\pi\ell\nu$. (If BR($K_L\rightarrow\pi\nu\bar\nu$) is measured, it will give a comparably clean determination of $\eta$.) A second example of a theoretically clean mode in class (i) is $B\rightarrow\phi K_S$. The quark subprocess involves FCNC and cannot proceed via a tree level SM diagram. The leading contribution comes from penguin diagrams. The two terms in eq. (\ref{ssstype}) are now both differences of penguins but the second term is CKM suppressed and thus of ${\cal O}(\lambda^2)$ compared to the first. Thus CP violation in the decay is at most a few percent, and can be neglected in the analysis of asymmetries in this channel. The analysis is similar to the $\psi K_S$ case, and the asymmetry is proportional to $\sin(2\beta)$. The same quark subprocesses give theoretically clean CP asymmetries also in $B_s$ decays. These asymmetries are, however, very small since the relative phase between the mixing amplitude and the decay amplitudes ($\beta_s$ defined below) is very small. The best known example of class (ii) is $B\rightarrow\pi\pi$. The quark subprocess is $b\rightarrow u\bar ud$ which is dominated by the $W$-mediated tree diagram. Neglecting for the moment the second, pure penguin, term in eq. (\ref{uudtype}) we find \begin{equation}\label{Apipi} {\bar A_{\pi\pi}\over A_{\pi\pi}}=\eta_{\pi\pi}{V_{ub}V_{ud}^*\over V_{ub}^*V_{ud}}\omega_B^*. \end{equation} The CP eigenvalue for two pions is $+1$. Combining (\ref{qpSM}) and (\ref{Apipi}), we get \begin{equation}\label{lampipi} \lambda(B\rightarrow\pi^+\pi^-)=\left({V_{tb}^*V_{td}\over V_{tb}V_{td}^*}\right) \left({V_{ud}^*V_{ub}\over V_{ud}V_{ub}^*}\right) \ \Longrightarrow\ {\cal I}m\lambda_{\pi\pi}=\sin(2\alpha). \end{equation} The pure penguin term in eq. (\ref{uudtype}) has a weak phase, $\arg(V_{td}^*V_{tb})$, different from the term with the tree contribution, so it modifies both ${\cal I}m\lambda_{\pi\pi}$ and (if there are non-trivial strong phases) $|\lambda_{\pi\pi}|$. The recent CLEO results mentioned above suggest that the penguin contribution to $B\rightarrow\pi\pi$ channel is significant, probably $10\%$ or more. This then introduces CP violation in decay, unless the strong phases cancel (or are zero, as suggested by factorization arguments). The resulting hadronic uncertainty can be eliminated using isospin analysis \cite{GrLo}. This requires a measurement of the rates for the isospin-related channels $B^+ \rightarrow \pi^+ \pi^0$ and $B^0 \rightarrow \pi^0\pi^0$ as well as the corresponding CP-conjugate processes. The rate for $\pi^0\pi^0$ is expected to be small and the measurement is difficult, but even an upper bound on this rate can be used to limit the magnitude of hadronic uncertainties \cite{GrQu}. Related but slightly more complicated channels with the same underlying quark structure are $B\rightarrow\rho^0\pi^0$ and $B\rightarrow a_1^0\pi^0$. Again an analysis involving the isospin-related channels can be used to help eliminate hadronic uncertainties from CP violations in the decays \cite{LNQS,SnQu}. Channels such as $\rho\rho$ and $a_1\rho$ could in principle also be studied, using angular analysis to determine the mixture of CP-even and CP-odd contributions. The analysis of $B\rightarrow D^+D^-$ proceeds along very similar lines. The quark subprocess here is $b\rightarrow c\bar cd$, and so the tree contribution gives \begin{equation}\label{lamDD} \lambda(B\rightarrow D^+D^-)= \eta_{D^+D^-}\left({V_{tb}^*V_{td}\over V_{tb}V_{td}^*}\right) \left({V_{cd}^*V_{cb}\over V_{cd}V_{cb}^*}\right) \ \Longrightarrow\ {\cal I}m\lambda_{DD}=-\sin(2\beta), \end{equation} where we used $\eta_{D^+D^-}=+1$. Again, there are hadronic uncertainties due to the pure penguin term in (\ref{ccdtype}), but they are estimated to be small. In all cases the above discussions have neglected the distinction between strong penguins and electroweak penguins. The CKM phase structure of both types of penguins is the same. The only place where this distinction becomes important is when an isospin argument is used to remove hadronic uncertainties due to penguin contributions. These arguments are based on the fact that gluons have isospin zero, and hence strong penguin processes have definite $\Delta I$. Photons and $Z$-bosons on the other hand contribute to more than one $\Delta I$ transition and hence cannot be separated from tree terms by isospin analysis. In most cases electroweak penguins are small, typically no more than ten percent of the corresponding strong penguins and so their effects can safely be neglected. However in cases (iii) and (iv), where tree contributions are small or absent, their effects may need to be considered. (A full review of the role of electroweak penguins in $B$ decays has been given in ref. \cite{ewpenreview}.) \subsection{Unitarity Triangles} One can obtain an intuitive understanding of the Standard Model CP violation in the interference between decays with and without mixing by examining the unitarity triangles. It is instructive to draw the three triangles, (\ref{Unitds}), (\ref{Unitsb}) and (\ref{Unitdb}), knowing the experimental values (within errors) for the various $|V_{ij}|$. In the first triangle (\ref{Unitds}), one side is of ${\cal O}(\lambda^5)$ and therefore much shorter than the other, ${\cal O}(\lambda)$, sides. In the second triangle (\ref{Unitsb}), one side is of ${\cal O}(\lambda^4)$ and therefore shorter than the other, ${\cal O}(\lambda^2)$, sides. In the third triangle (\ref{Unitdb}), all sides have lengths of ${\cal O}(\lambda^3)$. The first two triangles then almost collapse to a line while the third one is open. Let us examine the CP asymmetries in the leading decays into final CP eigenstates. For the $B$ mesons, the size of these asymmetries ({\it e.g.} ${\cal I}m\lambda_{\psi K_S}$) depends on $\beta$ because it gives the difference between half the phase of the $B-\bar B$ mixing amplitude and the phase of the decay amplitudes. The form of the third unitarity triangle, (\ref{Unitdb}), implies that $\beta={\cal O}(1)$, which explains why these asymmetries are expected to be large. It is useful to define the analog phases for the $B_s$ meson, $\beta_s$, and the $K$ meson, $\beta_K$: \begin{equation}\label{bbangles} \beta_s\equiv\arg\left[-{V_{ts}V_{tb}^*\over V_{cs}V_{cb}^*}\right],\ \ \ \beta_K\equiv\arg\left[-{V_{cs}V_{cd}^*\over V_{us}V_{ud}^*}\right]. \end{equation} The angles $\beta_s$ and $\beta_K$ can be seen to be the small angles of the second and first unitarity triangles, (\ref{Unitsb}) and (\ref{Unitds}), respectively. This gives an intuitive understanding of why CP violation is small in the leading $K$ decays (that is $\varepsilon_K$ measured in $K\rightarrow\pi\pi$ decays) and is expected to be small in the leading $B_s$ decays ({\it e.g.} $B_s\rightarrow\psi\phi$). Decays related to the short sides of these triangles are rare but could exhibit significant CP violation. Actually, the large angles in the (\ref{Unitds}) triangle are approximately $\beta$ and $\pi-\beta$, which explains why CP violation in $K\rightarrow\pi\nu\bar\nu$ is related to $\beta$ and expected to be large. The large angles in the (\ref{Unitsb}) triangle are approximately $\gamma$ and $\pi-\gamma$. This explains why the CP asymmetry in $B_s\rightarrow\rho K_S$ is related to $\gamma$ and expected to be large. (Note, however, that this mode gets comparable contributions from penguin and tree diagrams and does not give a clean CKM measurement \cite{BuFl}.) \section{CP VIOLATION BEYOND THE STANDARD MODEL} The Standard Model picture of CP violation is rather unique and highly predictive. In particular, we would like to point out the following features: \begin{itemize} \item[(i)] CP is broken explicitly. \item[(ii)] All CP violation arises from a single phase, that is $\delta_{\rm KM}$. \item[(iii)] The measured value of $\varepsilon_K$ requires that $\delta_{\rm KM}$ is of order one. (In other words, CP is not an approximate symmetry of the Standard Model.) \item[(iv)] The values of all other CP violating observables can be predicted. In particular, CP violation in $B\rightarrow\psi K_S$ (and similarly various other CP asymmetries in $B$ decays), and in $K\rightarrow\pi\nu\bar\nu$ are expected to be of order one. \end{itemize} The commonly repeated statement that CP violation is one of the least tested aspects of the Standard Model is well demonstrated by the fact that none of the above features necessarily holds in the presence of New Physics. In particular, there are viable models of new physics ({\it e.g.} certain supersymmetric models) with the following features: \begin{itemize} \item[(i)] CP is broken spontaneously. \item[(ii)] There are many CP violating phases (even in the low energy effective theory). \item[(iii)] CP is an approximate symmetry, with all CP violating phases small (usually $10^{-3}\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\;\phi_{\rm CP}\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\; 10^{-2}$). \item[(iv)] Values of CP violating observables can be predicted and could be very different from the Standard Model predictions (except, of course, $\varepsilon_K$). In particular, ${\cal I}m\lambda_{\psi K_S}$ and $a_{\pi\nu\bar\nu}$ could both be $\ll1$. \end{itemize} We emphasize that various extensions of the SM modify its predictions of CP violation in various ways. The models of approximate CP described above are only one example of how the predictions can be dramatically violated. To understand how the Standard Model predictions could be modified by New Physics, we will focus on CP violation in the interference between decays with and without mixing. As explained above, it is this type of CP violation which, due to its theoretical cleanliness, may give unambiguous evidence for New Physics most easily. \subsection{CP Violation as a Probe of Flavor Beyond the Standard Model} Let us consider five specific CP violating observables. \begin{itemize} \item[(i)] ${\cal I}m\lambda_{\psi K_S}$, the CP asymmetry in $B\rightarrow\psi K_S$. This measurement will cleanly determine the relative phase between the $B-\bar B$ mixing amplitude and the $b\rightarrow c\bar cs$ decay amplitude ($\sin2\beta$ in the Standard Model). The $b\rightarrow c\bar cs$ decay has Standard Model tree contributions and therefore is very unlikely to be significantly affected by new physics. On the other hand, the mixing amplitude can be easily modified by new physics. We parametrize such a modification by a phase $\theta_d$: \begin{equation}\label{apksNP} {\cal I}m\lambda_{\psi K_S}=\sin[2(\beta+\theta_d)]. \end{equation} \item[(ii)] ${\cal I}m\lambda_{\phi K_S}$, the CP asymmetry in $B\rightarrow\phi K_S$. This measurement will cleanly determine the relative phase between the $B-\bar B$ mixing amplitude and the $b\rightarrow s\bar ss$ decay amplitude. The $b\rightarrow s\bar ss$ decay has only Standard Model penguin contributions and therefore is sensitive to new physics. We parametrize the modification of the decay amplitude by a phase $\theta_A$ \cite{GrWo}: \begin{equation}\label{aphksNP} {\cal I}m\lambda_{\phi K_S}=\sin[2(\beta+\theta_d+\theta_A)]. \end{equation} \item[(iii)] $a_{\pi\nu\bar\nu}$, the CP violating ratio of $K\rightarrow \pi\nu\bar\nu$ decays: \begin{equation}\label{defapnn} a_{\pi\nu\bar\nu}={\Gamma(K_L\rightarrow\pi^0\nu\bar\nu)\over \Gamma(K^+\rightarrow\pi^+\nu\bar\nu)}. \end{equation} This measurement will cleanly determine the relative phase between the $K-\bar K$ mixing amplitude and the $s\rightarrow d\nu\bar\nu$ decay amplitude. The experimentally measured small value of $\varepsilon_K$ requires that the phase of the $K-\bar K$ mixing amplitude is not modified from the Standard Model prediction. (More precisely, it requires that the phase in the mixing amplitude is very close to the phase in the $s\rightarrow d\bar uu$ decay amplitude.) On the other hand, the decay, which in the Standard Model is a loop process with small mixing angles, can be easily modified by new physics. \item[(iv)] ${\cal I}m(\lambda_{K^-\pi^+})$, the CP violating quantity in $D\rightarrow K^-\pi^+$ decay. The time-dependent ratios between the doubly-Cabibbo-suppressed and the Cabibbo-allowed $D\rightarrow K\pi$ decay rates are given by \begin{eqnarray}\label{deflD} {\Gamma[D^0(t)\rightarrow K^+\pi^-]\over \Gamma[D^0(t)\rightarrow K^-\pi^+]}\simeq&\ |\lambda|^2+ {(\Delta m_D)^2\over4}t^2+{\cal I}m(\lambda^{-1}_{K^+\pi^-})t,\\ {\Gamma[\bar D^0(t)\rightarrow K^-\pi^+]\over \Gamma[\bar D^0(t)\rightarrow K^+\pi^-]}\simeq&\ |\lambda|^2+ {(\Delta m_D)^2\over4}t^2+{\cal I}m(\lambda_{K^-\pi^+})t, \end{eqnarray} where we used the fact that CP violation in mixing is expected to be small and, consequently, \begin{equation}\label{lKpi} |\lambda^{-1}_{K^+\pi^-}|=|\lambda_{K^-\pi^+}|\equiv\lambda. \end{equation} The ratio \begin{equation}\label{defaD} a_{D\rightarrow K\pi}={{\cal I}m(\lambda_{K^-\pi^+})\over|\lambda_{K^-\pi^+}|} \end{equation} depends on the relative phase between the $D-\bar D$ mixing amplitude and the $c\rightarrow d\bar su$ decay amplitude. Within the Standard Model, this decay channel is tree level. It is unlikely that it is affected by new physics. On the other hand, the mixing amplitude can be easily modified by new physics \cite{BKS}. \item[(v)] $d_N$, the electric dipole moment of the neutron. We did not discuss this quantity so far because, unlike CP violation in meson decays, flavor changing couplings are not necessary for $d_N$. In other words, the CP violation that induces $d_N$ is {\it flavor diagonal}. It does in general get contributions from flavor changing physics, but it could be induced by sectors that are flavor blind. Within the Standard Model (and ignoring the strong CP angle $\theta_{\rm QCD}$), the contribution from $\delta_{\rm KM}$ arises at the three loop level and is at least six orders of magnitude below the experimental bound \cite{PDG}\ $d_N^{\exp}$, \begin{equation}\label{dnexp} d_N^{\rm exp}=1.1\times10^{-25}\ e\ {\rm cm}. \end{equation} \end{itemize} The various CP violating observables discussed above are sensitive then to new physics in the mixing amplitudes for the $B-\bar B$ and $D-\bar D$ systems, in the decay amplitudes for $b\rightarrow s\bar ss$ and $s\rightarrow d\nu\bar\nu$ channels and to flavor diagonal CP violation. If information about all these processes becomes available and deviations from the Standard Model predictions are found, we can ask rather detailed questions about the nature of the new physics that is responsible to these deviations: \begin{itemize} \item[(i)] Is the new physics related to the down sector? the up sector? both? \item[(ii)] Is the new physics related to $\Delta B=1$ processes? $\Delta B=2$? both? \item[(iii)] Is the new physics related to the third generation? to all generations? \item[(iv)] Are the new sources of CP violation flavor changing? flavor diagonal? both? \end{itemize} It is no wonder then that with such rich information, flavor and CP violation provide an excellent probe of new physics. \section{Supersymmetry} A generic supersymmetric extension of the Standard Model contains a host of new flavor and CP violating parameters. (For reviews on supersymmetry see refs. \cite{nilles,haberkane,barbieri,HaberTASI,RamoTASI,BaggTASI,SMart,LBH}. The following chapter is based on \cite{GNR}.) The requirement of consistency with experimental data provides strong constraints on many of these parameters. For this reason, the physics of flavor and CP violation has had a profound impact on supersymmetric model building. A discussion of CP violation in this context can hardly avoid addressing the flavor problem itself. Indeed, many of the supersymmetric models that we analyze below were originally aimed at solving flavor problems. As concerns CP violation, one can distinguish two classes of experimental constraints. First, bounds on nuclear and atomic electric dipole moments determine what is usually called the {\it supersymmetric CP problem}. Second, the physics of neutral mesons and, most importantly, the small experimental value of $\varepsilon_K$ pose the {\it supersymmetric $\varepsilon_K$ problem}. In the next two subsections we describe the two problems. Then we describe various supersymmetric flavor problems and the ways in which they address the supersymmetric CP problem. Before turning to a detailed discussion, we define two scales that play an important role in supersymmetry: $\Lambda_S$, where the soft supersymmetry breaking terms are generated, and $\Lambda_F$, where flavor dynamics takes place. When $\Lambda_F\gg\Lambda_S$, it is possible that there are no genuinely new sources of flavor and CP violation. This leads to models with exact universality, which we discuss in section IV.C. When $\Lambda_F\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\;\Lambda_S$, we do not expect, in general, that flavor and CP violation are limited to the Yukawa matrices. One way to suppress CP violation would be to assume that CP is an approximate symmetry of the full theory (namely, CP violating phases are all small). We discuss this scenario in section IV.D. Another option is to assume that, similarly to the Standard Model, CP violating phases are large, but their effects are screened, possibly by the same physics that explains the various flavor puzzles. Such models, with Abelian or non-Abelian horizontal symmetries, are described in section IV.E. It is also possible that CP violating effects are suppressed because squarks are heavy. This scenario is also discussed in section IV.E. Some concluding comments are given in section IV.F. \subsection{The Supersymmetric CP Problem} One aspect of supersymmetric CP violation involves effects that are flavor preserving. Then, for simplicity, we describe this aspect in a supersymmetric model without additional flavor mixings, {\it i.e.} the minimal supersymmetric standard model (MSSM) with universal sfermion masses and with the trilinear SUSY-breaking scalar couplings proportional to the corresponding Yukawa couplings. (The generalization to the case of non-universal soft terms is straightforward.) In such a constrained framework, there are four new phases beyond the two phases of the Standard Model ($\delta_{\rm KM}$ and $\theta_{\rm QCD}$). One arises in the bilinear $\mu$-term of the superpotential, \begin{equation}\label{muterm} W=\mu H_uH_d, \end{equation} while the other three arise in the soft supersymmetry breaking parameters $m_{\tilde g}$ (the gaugino mass), $A$ (the trilinear scalar coupling) and $m_{12}^2$ (the bilinear scalar coupling): \begin{equation}\label{sSUSYb} {\cal L}=-{1\over2}m_{\tilde g}\tgl{\tilde g}-A(Y^u QH_u\bar u -Y^d QH_d\bar d-Y^e LH_d\bar\ell)-m_{12}^2 H_uH_d+{\rm h.c.}, \end{equation} where ${\tilde g}$ are the gauginos and $Y$ are Yukawa matrices. Only two combinations of the four phases are physical \cite{DGH,DiTh}. In the absence of (\ref{muterm}) and (\ref{sSUSYb}), there are two additional global $U(1)$ symmetries in the MSSM, an $R$ symmetry and a Peccei-Quinn symmetry. This means that one could treat the various dimensionful parameters in (\ref{muterm}) and (\ref{sSUSYb}) as spurions which break the symmetries, thus deriving selection rules. The appropriate charge assignments are: \begin{equation}\label{spur} \matrix{&m_{\tilde g}&A&m_{12}^2&\mu&H_u&H_d&Q\bar u&Q\bar d&L\bar\ell\cr U(1)_{\rm PQ}&0&0&-2&-2&1&1&-1&-1&-1\cr U(1)_{\rm R}&-2&-2&-2&0&1&1&1&1&1\cr} \end{equation} Physical observables can only depend on combinations of the dimensionful parameters that are neutral under both $U(1)$'s. There are three such independent combinations: $m_{\tilde g}\mu(m_{12}^2)^*$, $A\mu(m_{12}^2)^*$ and $A^* m_{\tilde g}$. However, only two of their phases are independent, say \begin{equation}\label{phiAB} \phi_A=\arg(A^* m_{\tilde g}),\ \ \ \phi_B=\arg(m_{\tilde g}\mu(m_{12}^2)^*). \end{equation} In the more general case of non-universal soft terms there is one independent phase $\phi_{A_{i}}$ for each quark and lepton flavor. Moreover, complex off-diagonal entries in the sfermion mass matrices may represent additional sources of CP violation. The most significant effect of $\phi_A$ and $\phi_B$ is their contribution to electric dipole moments (EDMs). For example, the contribution from one-loop gluino diagrams to the down quark EDM is given by \cite{BuWy,PoWi}: \begin{equation}\label{ddsusy} d_d=M_d{e\alpha_3\over 18\pi\tilde m^4}\left( |A m_{\tilde g}|\sin\phi_A+\tan\beta|\mu m_{\tilde g}|\sin\phi_B\right), \end{equation} where we have taken $m^2_Q\sim m^2_D\sim m^2_{\tilde g}\sim\tilde m^2$, for left- and right-handed squark and gluino masses. We define, as usual, $\tan\beta = \vev{H_u}/\vev{H_d}$. Similar one-loop diagrams give rise to chromoelectric dipole moments. The electric and chromoelectric dipole moments of the light quarks $(u,d,s)$ are the main source of $d_N$ (the EDM of the neutron), giving \cite{FPT}\ \begin{equation}\label{dipole} d_N\sim 2\, \left({100\, {\rm GeV}\over \tilde m}\right )^2 \sin \phi_{A,B}\times10^{-23}\ e\, {\rm cm} \end{equation} where, as above, $\tilde m$ represents the overall SUSY scale. The present experimental bound, $d_N<1.1\times 10^{-25}e\, {\rm cm}$ \cite{smith,altarev}, is then violated for ${\cal O}(1)$ phases, unless the masses of superpartners are above ${\cal O}(1\ TeV)$. Alternatively for light SUSY masses, the new phases should be $<{\cal O}(10^{-2})$. Notice however that one may consider the actual bound weaker than this, due to the theoretical uncertainty in the estimate of the hadronic matrix elements that lead to eq. (\ref{dipole}) \cite{florellis}. With this caveat, whether the phases are small or squarks are heavy, a fine-tuning of order $10^{-2}$ seems to be required, in general, to avoid too large a $d_N$. This is {\it the Supersymmetric CP Problem} \cite{BuWy,PoWi,EFN,Barr}. In addition to $d_N$, the SUSY CP phases contribute to atomic and nuclear EDMs (see a detailed discussion in ref. \cite{FPT}). The former are also sensitive to phases in the leptonic sector. The latter give additional constraints on the quark sector phases. For instance, the bound on the nuclear EDM of ${}^{199} {\rm Hg}$ is comparable to the one given by $d_N$. In practice, these additional bounds on SUSY CP phases are not stronger than those from $d_N$ (at least in the quark sector). However, since there are significant theoretical uncertainties in the calculation of nuclear EDMs, it is important to measure as many as possible of them to obtain more reliable bounds. \subsection{The Supersymmetric $\varepsilon_K$ Problem} The contribution to the CP violating $\varepsilon_K$ parameter in the neutral $K$ system is dominated by diagrams involving $Q$ and $\bar d$ squarks in the same loop \cite{DNW,GaMa,HKT,GMS,GGMS}. The corresponding effective four-fermi operator involves fermions of both chiralities, so that its matrix elements are enhanced by ${\cal O}(m_K/m_s)^2$ compared to the chirality conserving operators. For $m_{\tilde g}\simeq m_Q \simeq m_D= \tilde m$ (our results depend only weakly on this assumption) and focusing on the contribution from the first two squark families, one gets (we use the results in ref. \cite{GGMS}) \begin{equation}\label{epsKSusy} \varepsilon_K={5\ \alpha_3^2 \over 162\sqrt2}{f_K^2m_K\over\tilde m^2\Delta m_K}\left [\left({m_K\over m_s+m_d}\right)^2+{3\over 25}\right] {\cal I}m\left\{{(\delta m_Q^2)_{12}\over m_Q^2} {(\delta m_D^2)_{12}\over m_D^2}\right\}, \end{equation} where $(\delta m_{Q,D}^2)_{12}$ are the off diagonal entries in the squark mass matrices in a basis where the down quark mass matrix and the gluino couplings are diagonal. These flavor violating quantities are often written as $(\delta m_{Q,D}^2)_{12}=V_{11}^{Q,D}\delta m_{Q,D}^2 V_{21}^{Q,D*}$, where $\delta m_{Q,D}^2$ is the mass splitting among the squarks and $V^{Q,D}$ are the gluino coupling mixing matrices in the mass eigenbasis of quarks and squarks. Note that CP would be violated even if there were two families only \cite{NirSusy}. (There are also contributions involving the third family squarks via the (13) and (23) mixings. In some cases the third family contribution actually dominates.) Using the experimental value of $\varepsilon_K$, we get \begin{equation}\label{epsKScon} {(\Delta m_K\varepsilon_K)^{\rm SUSY}\over(\Delta m_K\varepsilon_K)^{\rm EXP}}\sim10^7 \left ({300 \ {\rm GeV}\over\tilde m}\right)^2 \left({m^2_{Q_2}-m^2_{Q_1}\over m_Q^2}\right) \left({m^2_{D_2}-m^2_{D_1}\over m_D^2}\right) |K_{12}^{dL}K_{12}^{dR}|\sin\phi, \end{equation} where $\phi$ is the CP violating phase. In a generic supersymmetric framework, we expect $\tilde m={\cal O}(m_Z)$, $\delta m_{Q,D}^2/m_{Q,D}^2={\cal O}(1)$, $K_{ij}^{Q,D}={\cal O}(1)$ and $\sin\phi={\cal O}(1)$. Then the constraint (\ref{epsKScon}) is generically violated by about seven orders of magnitude. (Four-fermi operators with same chirality fermions give a smaller effect. The resulting $\varepsilon_K$-bounds are therefore weaker by about one order of magnitude.) Eq. (\ref{epsKScon}) also shows what are the possible ways to solve the supersymmetric $\varepsilon_K$ problem: \begin{itemize} \item[(i)] Heavy squarks: $\tilde m\gg300\ GeV$; \item[(ii)] Universality: $\delta m_{Q,D}^2\ll m_{Q,D}^2$; \item[(iii)] Alignment: $|K_{12}^d|\ll1$; \item[(iv)] Approximate CP: $\sin\phi\ll1$. \end{itemize} The $d_N$ problem (see eq. (\ref{dipole})) is solved by either heavy squarks or approximate CP. \subsection{Exact Universality} Both supersymmetric CP problems are solved if, at the scale $\Lambda_S$, the soft supersymmetry breaking terms are universal and the genuine SUSY CP phases $\phi_{A,B}$ vanish. Then the Yukawa matrices represent the only source of flavor and CP violation which is relevant in low energy physics. This situation can naturally arise when supersymmetry breaking is mediated by gauge interactions at a scale $\Lambda_S\ll\Lambda_F$ \cite{DiNe,DNeS,DNNS,DNiS}. In the simplest scenarios, the $A$-terms and the gaugino masses are generated by the same SUSY and $U(1)_R$ breaking source (see eq. (\ref{spur}). Thus, up to very small effects due to the {\it standard} Yukawa matrices, $\arg(A)=\arg(m_{\tilde g})$ so that $\phi_A$ vanishes. In specific models also $\phi_B$ vanishes in a similar way \cite{DNeS,DNiS}. It is also possible that similar boundary conditions occur when supersymmetry breaking is communicated to the observable sector up at the Planck scale \cite{CAN,BFS,HLW,EKN,LaRo}. The situation in this case seems to be less under control from the theoretical point of view. Dilaton dominance in SUSY breaking, though, seems a very interesting direction to explore \cite{KaLo,BML}. The most important implication of this type of boundary conditions for soft terms, which we refer to as {\it exact universality} \cite{DiGe,Saka}, is the existence of the SUSY analogue of the GIM mechanism which operates in the SM. The CP violating phase of the CKM matrix can feed into the soft terms via Renormalization Group (RG) evolution only with a strong suppression from light quark masses \cite{DGH}. With regard to the supersymmetric CP problem, gluino diagrams contribute to quark EDMs as in eq. (\ref{ddsusy}), but with a highly suppressed effective phase, {\it e.g.} \begin{equation}\label{gim} \phi_{A_d}\sim (t_S/16 \pi^2)^4 Y_t^4 Y_c^2 Y_b^2 J. \end{equation} Here $t_S=\log (\Lambda_S/M_W)$ arises from the RG evolution from $\Lambda_S$ to the electroweak scale, the $Y_i$'s are quark Yukawa couplings (in the mass basis), and $J\simeq 2\times 10^{-5}$ is defined in eq. (\ref{defJ}). A similar contribution comes from chargino diagrams. The resulting EDM is $d_N\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\; 10^{-31}\ e\ {\rm cm}$. This maximum can be reached only for very large $\tan\beta\sim60$ while, for small $\tan\beta \sim 1$, $d_N$ is about 5 orders of magnitude smaller. This range of values for $d_N$ is much below the present ($\sim10^{-25}\ e$ cm) and foreseen ($\sim 10^{-28}\ e$ cm) experimental sensitivities \cite{RoStru,BeVi,IMSS,ACW}. With regard to the supersymmetric $\varepsilon_K$ problem, the contribution to $\varepsilon_K$ is proportional to ${\cal I}m(V_{td}V_{ts}^*)^2 Y_t^4 (t_S/16 \pi^2)^2$, giving the same GIM suppression as in the SM. This contribution turns out to be small \cite{DGH}: \begin{equation}\label{unieps} |\varepsilon_K^{\rm SUSY}|\sim 6 \times 10^{-6} \left [{J\,{\cal R}e (V_{td}V_{ts}^*)\over10^{-8}}\right]\left[{300{\rm GeV}\over\tilde m}\right]^2 \left [ {\ln (\Lambda_S/m_W)\over 5}\right ]^2. \end{equation} The value $t_S=5$ is typical to gauge mediated supersymmetry breaking, but (\ref{unieps}) remains negligible for any scale $\Lambda_S\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\; M_{\rm Pl}$ (namely $t_S\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\; 35$). The supersymmetric contribution to $D-\bar D$ mixing is similarly small and we expect no observable effects. For the $B_d$ and $B_s$ systems, the largest SUSY contribution to the mixing comes from box diagrams with intermediate charged Higgs and the up quarks. It can be up to ${\cal O}(0.2)$ of the SM amplitude for $\Lambda_S=M_{\rm Pl}$ and $\tan\beta = {\cal O}(1)$ \cite{BBMR,GNO,Nihe,BaKo}, and much smaller for large $\tan\beta$. The contribution is smaller in models of gauge mediated SUSY breaking where the mass of the charged Higgs boson is typically $\lower.8ex\hbox{$\sim$}\kern-.75em\raise.45ex\hbox{$>$}\; 300\ GeV$ \cite{DNNS}\ and $t_S\sim5$. The SUSY contributions to $B_s-\bar B_s$ and $B_d -\bar B_d$ mixing are, to a good approximation, proportional to $(V_{tb}V_{ts}^*)^2$ and $(V_{tb}V_{td}^*)^2$, respectively, like in the SM. Then, regardless of the size of these contributions, the relation $\Delta m_{B_d}/\Delta m_{B_s} \sim |V_{td}/V_{ts}|^2$ and the CP asymmetries in neutral $B$ decays into final CP eigenstates are the same as in the SM. \subsection{Approximate CP Symmetry} Both supersymmetric CP problems are solved if CP is an approximate symmetry, broken by a small parameter of order $10^{-3}$. This is one of the possible solutions to CP problems in the class of supersymmetric models with $\Lambda_F\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\;\Lambda_S$, where the soft masses are generically not universal, so that we do not expect flavor and CP violation to be limited to the Yukawa matrices. (Of course, some mechanism has also to suppress the real part of the $\Delta S=2$ amplitude by a sufficient amount.) Most models where soft terms arise at the Planck scale ($\Lambda_S\sim M_{\rm Pl}$) belong to this class. If CP is an approximate symmetry, we expect also the SM phase $\delta_{\rm KM}$ to be $\ll 1$. Then the standard box diagrams cannot account for $\varepsilon_K$ which should arise from another source. In supersymmetry with non-universal soft terms, the source could be diagrams involving virtual superpartners, mainly squark-gluino box diagrams. Let us call $(M_{12}^K)^{\rm SUSY}$ the supersymmetric contribution to the $K-\bar K$ mixing amplitude. Then the requirements ${\cal R}e (M_{12}^K)^{\rm SUSY}\lower.8ex\hbox{$\sim$}\kern-.8em\raise.45ex\hbox{$<$}\;\Delta m_K$ and ${\cal I}m(M_{12}^K)^{\rm SUSY}\sim\varepsilon_K\Delta m_K$ imply that the generic CP phases are $\geq{\cal O}(\varepsilon_K)\sim 10^{-3}$. Of course, $d_N$ constrains the relevant CP violating phases to be $\lsim10^{-2}$. If all phases are of the same order, then $d_N$ must be just below or barely compatible with the present experimental bound. A signal should definitely be found if the accuracy is increased by two orders of magnitude. The main phenomenological implication of these scenarios is that CP asymmetries in $B$ meson decays are small, perhaps ${\cal O}(\varepsilon_K)$, rather than ${\cal O}(1)$ as expected in the SM. Also the ratio $a_{\pi\nu\bar\nu}$ (see (\ref{defapnn})) is very small, in contrast to the Standard Model where it is expected to be of $O(\sin^2\beta)$. Explicit models of approximate CP were presented in refs. \cite{abefre,BaBa,Eyal}. The fact that the Standard Model and the models of approximate CP are both viable at present is related to the fact that the mechanism of CP violation has not really been tested experimentally. The only measured CP violating observale, that is $\varepsilon_K$, is small. Its smallness could be related to the `accidental' smallness of CP violation for the first two quark generations, as is the case in the Standard Model, or to CP being an approximate symmetry, as is the case in the models discussed here. Future measurements, particularly of processes where the third generation plays a dominant role (such as $a_{\psi K_S}$ or $a_{\pi\nu\bar\nu}$), will easily distinguish between the two scenarios. While the Standard Model predicts large CP violating effects for these processes, approximate CP would suppress them too. The distinction between the Standard Model and Supersymmetry could also be made -- though less easily -- in measurements of CP violation in neutral $D$ decays and of the electric dipole moments of the neutron. Here, the GIM mechanism of the Standard Model is so efficient that CP violating effects are unobservable in both cases. In contrast, the flavor breaking in supersymmetry might be much stronger, and then the approximate CP somewhat suppresses the effects but to a level which is perhaps still observable. \subsection{Approximate Horizontal Symmetries} Another option is to assume that, similarly to the Standard Model, CP violating phases are large, but their effects are screened, possibly by the same physics that explains the various flavor puzzles. This usually requires Abelian or non-Abelian horizontal symmetries. Two ingredients play a major role here: selection rules that come from the symmetry and holomorphy of Yukawa and $A$-terms that comes from the supersymmetry. With Abelian symmetries, the screening mechanism is provided by {\it alignment} \cite{NiSe,LNSb,RaNi}, whereby the mixing matrices for gaugino couplings have very small mixing angles, particularly for the first two down squark generations. With non-Abelian symmetries, the screening mechanism is {\it approximate universality}, where squarks of the two families fit into an irreducible doublet and are, therefore, approximately degenerate \cite{DKL,PoSe,HaMu,PoTo,BDH,CHM,Zurab,Raby,Gali}. In all of these models, it is difficult to avoid $d_N\gsim10^{-28}$ e cm. As far as the third generation is concerned, the signatures of Abelian and non-Abelian models are similar. In particular, they allow observable deviations from the SM predictions for CP asymmetries in $B$ decays. In some cases, non-Abelian models give relations between CKM parameters and consequently predict strong constraints on these CP asymmetries. For the two light generations, only alignment allows interesting effects. In particular, it predicts large CP violating effects in $D-\bar D$ mixing \cite{NiSe,LNSb}. Thus, it allows for $a_{D\rightarrow K\pi}={\cal O}(1)$ and, in particular, for ${\cal I}m(\lambda^{-1}_{K^+\pi^-})\neq {\cal I}m(\lambda_{K^-\pi^+})$ which will signify new CP violating phases (see eqs. (\ref{deflD})). Finally, it is possible that CP violating effects are suppressed because squarks are heavy. If the masses of the first and second generations squarks $m_i$ are larger than the other soft masses, $m_i^2\sim 100\, \tilde m^2$ then the Supersymmetric CP problem is solved and the $\varepsilon_K$ problem is relaxed (but not eliminated) \cite{PoTo,DKL}. This does not necessarily lead to naturalness problems, since these two generations are almost decoupled from the Higgs sector. Notice though that, with the possible exception of $m_{\tilde b_R}^2$, third family squark masses cannot naturally be much above $m_Z^2$. If the relevant phases are of $O(1)$, the main contribution to $d_N$ comes from the third family via the two-loop induced three-gluon operator \cite{Weintg}, and it is roughly at the present experimental bound when $m_{\tilde t_{L,R}}\sim 100\ GeV$. Models with the first two squark generations heavy have their own signatures of CP violation in neutral meson mixing \cite{CKLN}. The mixing angles relevant to $D-\bar D$ mixing are similar, in general, to those of models of alignment (if alignment is invoked to explain $\Delta m_K$ with $m^2_{Q,D}\lsim20\ TeV$). However, since the $\tilde u$ and $\tilde c$ squarks are heavy, the contribution to $D-\bar D$ mixing is one to two orders of magnitude below the experimental bound. This may lead to the interesting situation that $D-\bar D$ mixing will first be observed through its CP violating part \cite{WolfD}. In the neutral $B$ system, ${\cal O}(1)$ shifts from the Standard Model predictions of CP asymmetries in the decays to final CP eigenstates are possible. This can occur even when the squarks masses of the third family are $\sim1\ TeV$ \cite{CKNS}, since now mixing angles can naturally be larger than in the case of horizontal symmetries (alignment or approximate universality). \subsection{Some Concluding Comments} The conclusion from our discussion of supersymmetric flavor models is that measurements of CP violation will provide us with an excellent probe of the flavor and CP structure of supersymmetry. This is clearly demonstrated in Table I. \begin{table} \[ \begin{array}{|c|c|c|c|c|c|} \hline {\rm Model} & d_N/d_N^{\rm exp} & \theta_d & \theta_A & a_{D^0\rightarrow K^-\pi^+} & a_{K\rightarrow\pi\nu\bar\nu} \\ \hline {\rm Standard\ Model} & \lsim10^{-6} & 0 & 0 & 0 & {\cal O}(1) \\ {\rm Exact Universality} & \lsim10^{-6} & 0 & 0 & 0 & ={\rm SM} \\ {\rm Approximate\ CP} & \sim10^{-1} & -\beta & 0 & {\cal O}(10^{-3}) & {\cal O}(10^{-5}) \\ {\rm Alignment} & \gsim10^{-3} & {\cal O}(0.2) & {\cal O}(1) & {\cal O}(1) & \approx{\rm SM} \\ {\rm Approx.\ Universality} & \gsim10^{-2} & {\cal O}(0.2) & {\cal O}(1) & 0 & \approx{\rm SM} \\ {\rm Heavy\ Squarks} & \sim10^{-1} & {\cal O}(1) & {\cal O}(1) & {\cal O}(10^{-2}) & \approx{\rm SM} \\ \hline \end{array} \] \vskip 12pt \caption {CP violating observables in various classes of Supersymmetric flavor models.} \end{table} The unique features of CP violation are well demonstrated by examining the CP asymmetry in $B\rightarrow\psi K_S$, ${\cal I}m\lambda_{\psi K_S}$, and CP violation in $K\rightarrow\pi \nu\bar\nu$, ${\cal I}m\lambda_{\pi\nu\bar\nu}$. Model independently, ${\cal I}m\lambda_{\psi K_S}$ measures the relative phase between the $B-\bar B$ mixing amplitude and the $b\rightarrow c\bar cd$ decay amplitude (more precisely, the $b\rightarrow c\bar cs$ decay amplitude times the $K-\bar K$ mixing amplitude), while ${\cal I}m\lambda_{\pi\nu\bar\nu}$ measures the relative phase between the $K-\bar K$ mixing amplitude and the $s\rightarrow d\nu\bar\nu$ decay amplitude. We would like to emphasize the following three points: \begin{itemize} \item[(i)] {\it The two measurements are theoretically clean to better than} ${\cal O}(10^{-2})$. Thus they can provide the most accurate determination of CKM parameters. \item[(ii)] {\it As concerns CP violation, the Standard Model is a uniquely predictive model}. In particular, it predicts that the seemingly unrelated ${\cal I}m\lambda_{\psi K_S}$ and ${\cal I}m\lambda_{\pi\nu\bar\nu}$ measure the same parameter, that is the angle $\beta$ of the unitarity triangle. \item[(iii)] {\it In the presence of New Physics, there is in general no reason for a relation between ${\cal I}m\lambda_{\psi K_S}$ and ${\cal I}m\lambda_{\pi\nu\bar\nu}$}. Therefore, a measurement of both will provide a sensitive probe of New Physics. \end{itemize} \bigskip \noindent {\bf Acknowledgements:} \smallskip \noindent My understanding of CP violation has benefitted from numerous discussions with Helen Quinn. Large parts of this review are based on previous reviews written in collaboration with her. Parts of this review are based on contributions to the BaBar Physics Book that were written in collaboration with Gerald Eigen, Ben Grinstein, Stephane Plaszczynski and Marie-Helene Schune. Other parts of this review are based on a review written in collaboration with Yuval Grossman and Riccardo Rattazzi. Special thanks go to Stephane Plaszczynski and Marie-Helene Schune for performing the CKM fit and preparing the plots for this review and to Joao Silva for explaining to me various subtleties concerning phase conventions and for useful comments on the manuscript. This manuscript is based on sets of lectures given in the school on flavor and gauge hierarchies (Cargese, 1998) and in the school on B and CP within and beyond the Standard Model (KIAS, Seoul, 1999). I thank Pierre Binetruy, and Seongyoul Choi and Kiwoon Choi for their hospitality in the respective schools. Y.N. is supported in part by the United States $-$ Israel Binational Science Foundation (BSF) and by the Minerva Foundation (Munich).
\section{Introduction} If it is desired to transmit a signal across the Galaxy so that another, unknown, recipient may detect it there are two basic types of transmission patterns, an omnidirectional signal that may be detected anywhere, or a beamed signal that can only be detected by those in the beam. Similarly, in the time domain, a signal may either be transmitted continuously or, with the same energy expenditure, a more powerful signal may be transmitted for a shorter period of time. Providing that the recipient knows where and when a signal is coming from, a beamed brief, and hence stronger, signal would be easier to detect. However, for such a transmission scheme to be feasible, the problem is for a transmitter and a recipient, one or both unknown to the other, to find a strategy that will enable the transmitter and receiver to transmit and observe at the right time and location. A strategy to achieve transmitter/receiver synchronization that has been considered by a number of authors is to utilize natural astronomical events, see, for example, Pace \& Walker (1975), Tang (1976), McLaughlin (1977), Makovetskii (1978), Pace (1979), Gruber \& Pfleiderer (1982), Tang (1981), Siebrand (1982), and Lemarchand (1994). In the simplest scheme omnidirectional signals would be transmitted at the occurrence of some particular event such as a nova outburst, maximum flux of a long period variable, specific binary phase or supernova occurrence. A signal would then be detected at the Earth delayed by a time corresponding to the difference between the event/Earth distance and the event/transmitter + transmitter/Earth distances. The time delay is thus given by: \begin{equation} \Delta T = (R_s - D + (R_s^2 + D^2 - 2R_sD cos \theta )^{1/2})/c \end{equation} where $R_s$ is the distance to the synchronizing astrophysical event, $D$ is the distance to the transmitter, and $\theta$ is the angular separation as viewed from the Earth. A further refinement is to transmit in a direction exactly or approximately away from the event which both reduces the time difference and gives a preferred direction in space. The use of one such locally dramatic event in particular, SN 1987A, is considered by Lemarchand (1994). The increased probability of detecting a signal if synchronizers are used is considered by McLaughlin (1977). To date there has been no definite detection of a signal in the Search for Extra-Terrestrial Intelligence (SETI). However, there have been a few detections of non-repeating signals that have generated some interest such as the ``wow" signal found at Ohio State Radio Observatory (Dixon 1985, Gray 1994) and the strongest events from the META survey which appear to preferentially lie in the Galactic plane (Horowitz \& Sagan 1993). While these may simply be noise or arise from natural astrophysical phenomena they could conceivably be genuine extra-terrestrial artificial signals that are transient either because transmission is intermittent or caused by interstellar scintillation (Cordes, Lazio, \& Sagan 1997). In this paper the use of one particular type of natural synchronizing signal is considered - the phenomenon of gamma-ray bursts (GRBs). These appear to posses a number of important advantages over other possible astrophysical events and their use in SETI is advocated for, in particular, targeted observations of relatively nearby stars. A brief summary of the phenomenology of gamma-ray bursts and their observations is given in Section 2, followed in Section 3 by a brief comparison of the use of GRBs to some other possible synchronizers, and in Section 4 two possible strategies for utilizing gamma-ray bursts in extraterrestrial communication are discussed. \section{Gamma-ray Bursts} GRBs are found to occur istrotropically over the sky (e.g. Briggs 1993) and the BATSE detector on board the Compton Gamma-Ray Observatory (CGRO) detects approximately one burst per day. GRBs are typically rather rapid events with log(T90), where T90 is the time within which 90\% of the flux from a burst is contained, showing a bimodal distribution with peaks at about 0.3 and 20 seconds (Fishman et al. 1994). Afterglows in the X-ray, optical, and radio bands have also recently been detected (see e.g. the review by M\'esz\'aros\ 1998) and the optical and X-ray afterglows have been seen to decay away with power-law indices of between roughly 1 to 2. The GRB intensity distribution shows that brighter bursts follow a peak flux distribution with a power-law index of ${-3/2}$ as expected for a uniform spatial distribution (e.g. Fenimore et al. 1993). However, fainter bursts show a flattening curve for which a simple interpretation is that we are observing the ``edge'' of the GRB distribution. This interpretation is complicated though by the unknown range of luminosities that GRBs may display. Although GRBs were once considered to arise from phenomena arising on Galactic neutron stars, the detection within the last year of so of a number of afterglows associated with host galaxies has enabled their distances to be firmly established as ``cosmological'', i.e. at very large distances. For the first three cases where a redshift was measured, values of Z = 0.835, 3.418 and 0.966 were obtained for GRBs 970508, 971214 and 980703 respectively (Metzger et al. 1997, Kulkarni et al. 1998, Djorgovski et al. 1998). At these large distances the implied luminosities of GRBs are extremely large. For example, the implied isotropic gamma-ray luminosity of GRB 971214 is \sqig3$\times$10$^{53}$ ergs s$^{-1}$ (Kulkarni et al. 1998). Even if the probable significant beaming is taken account of, the energy involved is at least comparable to, and likely exceeds, that associated with supernovae. The astrophysics of gamma-ray bursts is currently very poorly understood and current explanations to account for GRBs include merging neutron stars (e.g. M\'esz\'aros\ 1998) and hypernovae (Paczy\'nski 1998). However, the mechanism that causes GRBs is {\em not} important for their use in synchronizing communication across large distances. Although the BATSE detector is more sensitive than earlier generations of instruments it has only limited precision in locating bursts and and even bright bursts are not positioned to much better than a few degrees (Pendleton et al. 1999). Progress in precisely locating GRBs has recently come from the instruments onboard the SAX satellite. In addition to a GRB detector, SAX carries X-ray instrumentation consisting of a Wide-Field Camera (WFC) and a suite of Narrow Field instruments. For those bursts that occur within the field of view of the WFC, a position accurate to 3 to 8 arc minutes can be obtained, follow-up observations of X-ray afterglows with the narrow field instruments several hours later can then yield positions to about one arcminute (e.g. In't Zand et al. 1998). The All Sky Monitor on-board the Rossi X-ray Timing Explorer has also provided some locations for GRBs to a few arc minutes (Smith et al. 1998). It is this provision of arc minute accuracy positions that has lead to the discovery of the optical afterglows which can then make it possible to obtain positions accurate to better than an arc second. In the near future, the HETE-II mission (Ricker 1997) should provide positions accurate to 10 arc seconds to arc minutes for about 30 bursts per year. A further possible future GRB mission is Swift (Gehrels et al. 1999). Swift is designed to produce positions accurate to better than an arc second (if optical emission is also detected) within better than 90 seconds and would view an area of 2 steradians. The sensitivity of the Swift GRB detector is such that it is expected that it will detect at least 300 bursts per year. If funded it is intended that Swift would be launched in 2003. Systems of GRB detectors in spacecraft spread across the solar system in interplanetary networks (see e.g. Hurley et al. 1994) can also provide locations by comparing the time of arrival of a burst at the various spacecraft. Generally this type of network has not so far given very rapid determinations of source location, this can be caused by, for example, infrequent contacts with the spacecraft from ground stations or the difficulties of merging data obtained with a variety instruments in possibly differing formats. \section{Comparison with Some Other Astrophysical Synchronizers} Probably the most important property of a synchronizer is that it should, in some sense, be obvious and so likely to be used by both the transmitter and receiver. Arguably the key property in making a phenomenon obvious is its luminosity. In addition, it is desirable for the phenomenon to occur at a sufficiently rapid rate to facilitate the use of a large number of these events. Further, it should be possible to determine their time of occurrence precisely to make it possible to use a brief artificially transmitted signal. After GRBs, supernovae are the next most energetic phenomena known in the universe. However, it is considerably more difficult to detect all supernovae down to a specified flux limit than is the case for GRBs. The positions of supernovae are typically derived from optical observations and are thus known to high precision. However, the optical light curves of supernovae are relatively slowly varying and so are significantly less sharp time markers. Supernovae could perhaps be of better use if larger numbers could have their exact times of onset measured via, for example, a neutrino pulse (cf. SN 1987A Bionata et al. 1987, Hirata et al. 1987). Note that it has been suggested that a small subset of GRBs may be caused by the unusual Type Ic supernova class based on a possible association between SN1998bw, which had a redshift of only 0.008, and GRB980425 (e.g. Kippen et al. 1998). Novae and other variable stars have also been proposed as possible synchronizers. Novae are significantly less energetic than supernovae and, in the case of Galactic novae, it is also typically necessary to accurately know the distance to the nova as well as the transmitter to be able to calculate time delays. Variable stars form a broad diverse ``class'' of objects. Their much lower luminosities and varied nature makes them less obvious as the synchronizers that would be universally used. Until recently, the relatively poor precision with which gamma-ray burst locations were determined was arguably a problem with their use as synchronizers as this yields a large uncertainty in the time delay (eqn. 1). However, this limitation is disappearing with new generations of instruments augmented by the discovery of, in particular, optical afterglows. As the most energetic of all currently known natural phenomena this alone draws attention to GRBs. They occur at rapid rates, one a day even with BATSE level technology, are easily detectable, have short durations making them ``sharp'' time markers, and also occur at sufficiently large distances that it is not necessary to actually know these distances to calculate time delays. Hence, no other class of proposed phenomenon appears to posses any obvious properties which makes it a better candidate than GRBs for use as a synchronizer. \section{Use of GRBs for SETI} It is now considered how GRBs may be used as synchronizers in SETI and two basic transmission strategies are investigated: (i) Targeted Signaling. In this case the transmitter sends a signal promptly after the detection of a GRB in one or more directions ``close'' (i.e. within an angle $\theta$) to the opposite direction of the GRB at target(s) previously decided to be of interest. The beam width is likely to be made as small as possible. (ii) All-sky signaling. The transmitter sends a signal promptly after the detection of a GRB in a direction exactly opposite from the GRB with a beam half-width $\theta$. For a GRB located at a very large distance from both the transmitter and receiver, as appropriate for a GRB, the time delay between the detection of a GRB and the prompt transmitted signal is given by the simplified expression:\\ $\Delta T = (1 - cos\theta) D/c$ For small angles $\Delta T = D \theta^2/c$, for $\theta$ in radians. Note that $D$ and $\theta$ formally refer to the distance and angular separation when the burst arrives at the transmitter rather than the receiver. \subsection{Targeted Signaling} It is not possible to know in advance the angle within which the transmitter will decide to broadcast to particular locations downstream from the GRB. The smaller the angle utilized the longer it will take to transmit to all targets. However, for large angles the time delays will be larger and harder to accurately calculate for the receiver and more targets may have to be transmitted to for any particular GRB. A non-exhaustive list of factors which could be used by the transmitter to decide which places to target, in likely order of the distance at which they could be detected by the transmitter, is: \begin{itemize} \item{Detection of a solar type star} \item{Detection of a planetary system with Jupiter mass planet(s)} \item{Detection of terrestrial mass planet(s)} \item{Detection of life through the presence of, for example, an Oxygen atmosphere} \item{Detection of intelligent life through, for example, radio transmissions} \end{itemize} The down-stream angle(s) used by a transmitter will presumably depend on the number of targets that are thought to be potential hosts for receivers for the signal. For a large number of possible targets small angles would perhaps be likely to be used. Conversely, at the small number extreme, the Earth would be the only target transmitted to even if the Earth was exactly upstream from the GRB event. In order for the receiver to be able to calculate the time delay from a potential transmitter it is necessary to know the distance to that transmitter and the angular distance from the GRB. The current best set of stellar distances comes from HIPPARCOS parallax measurements (Perryman et al. 1997) which gave values accurate to $\sim$ 1 milli-arcsecond. In the the near future NASA's Space Interferometry Mission (SIM, Unwin et al. 1998) and ESA's GAIA (Gilmore et al. 1998) may yield large numbers of parallaxes with precisions better than $\sim$ 10 micro-arcseconds. For an illustration of the resulting time delays, and the errors involved in calculating these, the three classes of objects in the Project Phoenix targeted survey are considered (Henry et al. 1995) and presented in Table 1. Time delays are given for the three maximum distances corresponding to the ``Nearest 100'' (D $<$ 7.2 pc), ``Best \& Brightest'' (D $<$ 20 pc), and ``G Dwarf'' (D $<$ 50pc) targets as well as for distances of 100 pc and 1000 pc. Errors in calculating time delays due to uncertainties in distance measurements from HIPPARCOS and a future GAIA/SIM class mission are listed. Time delay errors due to GRB position determinations accurate to 1\arcsec\ and 10\arcsec\ are also given as appropriate to future GRB missions. For the closest class it appears feasible to use even 10\arcsec\ accuracy GRB locations and HIPPARCOS distances for moderately large offset angles. For the largest distances considered in this table of 1000 pc, at angles of only 1$^{\circ}$\ even parallax measurements accurate to 10 micro-arcsecond yield uncertainties in the arrival of a signal of almost 2 days. For all three of these Project Phoenix target classes GAIA/SIM class parallax measurements combined with good GRB measurements give errors on time delays that are modest ($<$ 1 day) for offset angles up to 5$^{\circ}$. \subsection{All-sky Signaling} If the intention is to eventually broadcast to the entire sky with no preference for particular locations there are two competing constraints:\\ (i) The use of a narrow beam width will produce both a larger gain and also reduce the maximum possible time delay at the receiver between the detection of a GRB and the transmitted signal. \\ (ii) The narrower the beam the longer it will take to cover the entire sky. One consideration may lead to a ``natural'' beam width for transmissions. If the transmitter wishes to broadcast out to a certain distance $D$, then a beam width can be chosen such that the maximum time delay at $D$ would be equal to the average interval between GRBs for the intensity level that they are using to select GRBs. i.e. $ \Delta T = 1 / R $ where $R$ is the GRB/transmission rate. This yields $ \theta\ = (c/DR)^{1/2} $ Note that basing a beam width on this consideration means that using additional lower intensity GRBs would not alter the time taken to illuminate the entire sky as the time delay and area of sky illuminated both depend on $\theta^{2}$. However, using additional bursts does result in a narrower beam and hence larger gains and smaller maximum time delays. The assumed transmission distance, $D$ might vary depending on the transmission direction, for example transmissions perpendicular or parallel to the Galactic plane. In addition, the beam width could potentially also be altered depending on the luminosity of a gamma-ray burst so that the maximum time delay at a certain distance would be the mean recurrence time for bursts of that luminosity or greater. The time taken to illuminate an area equal to the entire sky, although with overlap, is given by $ T = 4 \pi / (R B) $ where B is the area of the beam in steradians. Hence the ``natural'' beam width gives $T = 4D/c$. Thus, for any significant distance, a considerable time is required to illuminate the entire sky. Alternatively, if this ``natural'' beam width is not used then the number of GRB locations that needs to be monitored is $\sim R \Delta T$ = $ R D \theta^2/c $. \section{Conclusion} Gamma-ray bursts posses a number of properties that make them very good candidates for synchronizers that could aid in the search for brief beamed extraterrestrial signals. GRB positions are now starting to be obtained with sufficient precision to make them useful for targeted searches. This is augmented by precise measurements of stellar parallaxes from satellite borne instruments, and it is anticipated that even better parallax measurements will be obtained in the relatively near future. GRB locations and times are rapidly available from the GCN (GRB Coordinates Network, Barthelmy et al. 1998) and this is likely to continue with future missions which also aids in their use. For all sky surveys, the case for using GRBs as target positions to observe may be a lot weaker. However, there appears to be little to lose by pointing a detector as soon as possible at GRB locations and monitoring that location for some time. Additionally, even the positions currently provided by BATSE may be used for this purpose by a detector that has a field of view of a few degrees. \newpage
\section{Introduction} It is known that Neutron stars or Pulsars have strong magnetic fields of $\sim 10^8$ Tesla in their vicinity, while certain White Dwarfs have magnetic fields $\sim 10^2$ Tesla. If we were to use conventional arguments that when a sun type star with a magnetic field $\sim 10^{-4}$ Tesla contracts, there is conservation of magnetic flux, then we are lead to magnetic fields for Pulsars and White Dwarfs which are a few orders of magnitude less than the required values\cite{r1}.\\ We will now argue, that in the light of recent results that below the Fermi temperature, the degenerate electron gas obeys a semionic statistics, that is a statistics in between the Fermi-Dirac and Bose-Einstein, it is possible to deduce the correct magnetic fields for Neutron stars and White Dwarfs. Moreover this will also enable us to deduce the correct magnetic field of a planet like the earth. \section{Anomalous Behaviour Below the Fermi Temperature} In recent years, it has been realized that under specific conditions, for example low dimensionality or sub Fermi temperatures, Fermions exhibit an anomalous character - they obey statistics inbetween the Fermi-Dirac and Bose-Einstein statistics\cite{r2,r3}.\\ Let us specifically consider the case of sub Fermi temperatures (cf.ref.\cite{r3}). To notice the anomalous behaviour in a simple way we observe that in this case as is known\cite{r4} the assembly fills up each and every single particle energy level below the Fermi energy, with the Fermionic occupation number $1$. The density of states in momentum space is given by $d^3p$, exactly as in the case of Bosons. Whence we obtain the well known result \begin{equation} \epsilon_F = \frac{\hbar^2}{2m} \left(\frac{6\pi^2}{v}\right)^{2/3}\label{e1} \end{equation} where $\epsilon_F$ is the Fermi energy. The result for Phonons which obey Bose-Einstein statistics is identical to equation (\ref{e1}) (cf.ref.\cite{r3}).\\ The anomalous behaviour can also be seen as follows: We have for the energy density $e,$ in this case \begin{equation} e \propto \int^{p_F}_o \frac{p^2}{2m} d^3p \propto T^{2.5}_F\label{e2} \end{equation} where $p_F$ is the Fermi momentum and $T_F$ is the Fermi temperature. On the other hand, it is known that \cite{r5,r6} in $n$ dimensions we have, \begin{equation} e \propto T^{n+1}_F\label{e3} \end{equation} (For the case $n = 3,$ (\ref{e3}) is identical to the Stefan-Boltzmann law). Comparison of (\ref{e3}) and (\ref{e2}) shows that the assembly behaves with the fractal dimensionality $1.5$.\\ Let us now consider an assembly of $N$ electrons. As is known, if $N_+$ is the average number of particles with spin up, the magnetisation per unit volume is given by \begin{equation} M = \frac{\mu (2 N_+ - N)}{V}\label{e4} \end{equation} where $\mu$ is the electron magnetic moment. At low temperatures, in the usual theory, $N_+ \approx \frac{N}{2}$, so that the magnetisation given in (\ref{e4}) is very small. On the other hand, for Bose-Einstein statistics we would have, $N_+ \approx N$. With the above semionic statistics we have, \begin{equation} N_+ = \beta N, \frac{1}{2} < \beta < 1,\label{e5} \end{equation} If $N$ is very large, this makes an enormous difference in (\ref{e4}). Let us use (\ref{e4}) and (\ref{e5}) for the case of Neutron stars. \section{Magnetism of Neutron Stars and White Dwarfs} In this case, as is well known, we have an assembly of degenerate electrons at temperatures $\sim 10^7K$, (cf.for example \cite{r4}). So the considerations of Section 2 apply. In the case of a Neutron star we know that the number density of the degenerate electrons, $n \sim 10^{31}$ per c.c.\cite{r7,r8}. So using (\ref{e4}) and (\ref{e5}) and remembering that $\mu \approx 10^{-20}G,$ the magnetic field near the Pulsar is $\sim 10^{11}G \leq 10^8$ Tesla, as required.\\ As mentioned earlier some White Dwarfs also have magnetic fields. If the White Dwarf has an interior of the dimensions of a Neutron star, with a similar magnetic field, then remembering that the radius of a White Dwarf is about $10^3$ times that of a Neutron star, its magnetic field would be $10^{-6}$ times that of the neutron star, which is known to be the case.\\ \section{Discussion} It is quite remarkable that the above mechanism can also explain the magnetism of the earth. As is known the earth has a solid core of radius of about 1200 kilometers and temperature about 6000 K\cite{r9}. This core is made up almost entirely of Iron $(90 \%)$ and Nickel $(10 \%)$. It can easily be calculated that the number of particles $N \sim 10^{48}$, and that the Fermi temperature $\sim 10^5$. In this case we can easily verify using (\ref{e4}) and (\ref{e5}) that the magnetic field near the earth's surface $\sim 1G$, which is indeed the case.\\ It may be mentioned that the anomalous Bosonic behaviour given in (\ref{e5}) would imply a sensitivity to external magnetic influences which could lead to effects like magnetic flips or reversals.\\ To see this, we observe that the number of electrons, with spin aligned along a magnetic field $B$ which is introduced, where, $$B < < \epsilon_F/2\mu,$$ is given by (cf.ref.\cite{r4}), using Fermi-Dirac statistics, $$N_+ \approx \frac{N}{2} (1+\frac{3\mu B}{2\epsilon_F})$$ That is, $\beta$ in (\ref{e5}) is given by $$\beta \approx \frac{1}{2},$$ and the introduction of the field $B$, does not lead to a significant magnetic field in (\ref{e4}). But, if as in Section 2, $\beta \ne \frac{1}{2}$, but rather, $\beta > \frac{1}{2}$, then in view of the fact that $N$ is very large, the contribution from (\ref{e4}) could be significant.\\ Indeed in the case of the earth magnetic reversals do take place from time to time and are as yet not satisfactorily explained\cite{r10}.
\section{Introduction} It is generally believed that the picture of space-time as a manifold ${\cal M}$ should break down at very short distances of the order of the Planck length. One possible approach to the description of physical phenomena at small distances is based on noncommutative geometry of space-time. There have been investigations in the context of Connes' approach \cite{Connes} to gravity and the Standard Model of electroweak and strong interactions \cite{ConnesL,Chams} and in the framework of the string theory \cite{strings-nc}. Another approach starting from study of a relation between measurements at very small distances and black hole formations has been developed in the pioneering works \cite{DoplicherFR}. One more possibility is based on Quantum Group theory (see, {\em e.g.,\ } \cite{ChaichianDbk}). The essence of the noncommutative geometry consists in reformulating first the geometry in terms of commutative algebras and modules of smooth functions, and then generalizing them to their noncommutative analogs. If the notions of the noncommutative geometry are used directly for the description of the space-time, the notion of points as elementary geometrical entity is lost and one may expect that an ultraviolet cutoff appears. As is well known from the standard quantum mechanics, a quantization of any compact space, in particular a sphere, leads to finite-dimensional representations of the corresponding operators, so that in this case any calculation is reduced to manipulations with finite-dimensional matrices and thus there is simply no place for UV-divergences (see \cite{Ber,Hoppe,GKP3} and refs. therein). Things are not so easy in the case of noncompact manifolds. The quantization leads to infinite-dimensional representations and we have no guarantee that noncommutativity of the space-time coordinates removes UV-divergences. In our preceding paper \cite{CDP98} we have shown that ultraviolet behaviour of a field theory on a noncommutative space-time is sensitive to the topology of the space-time, namely to its compactness. We considered theories on a two-dimensional plane with Heisenberg-like commutation relations among coordinates (see also \cite{DoplicherFR,Filk}) and on a noncommutative cylinder. While the former retains the divergent tadpoles (as an ordinary QFT), the latter proves to be UV-finite. We argued that the underlying reason for such a UV-behaviour of the models is related to the properties of the complete coordinate-momentum quantum mechanical algebra and to the fact that the momenta degrees of freedom are associated to the fully {\it noncompact} Heisenberg-Weyl group manifold in the first case and to the cylinder in the second case (the cylinder has one {\it compact} dimension). Using these qualitative arguments, we supposed that the quantum field theory constructed on the $q$-deformed plane \cite{VaksmanK,SchuppWZ,BonecciCGST} with $E_q(2)$-symmetry also has UV-divergences. We have proved that indeed there are no kinematical reasons for this model to be UV-finite: the Green function of the free theory on the $q$-plane is singular. Moreover, we have shown that the interaction with an external field does produce divergent tadpole. However, in the paper \cite{CDP98} we used decomposition of the fields on the $q$-plane in the so-called distorted plane waves ($q$-deformed exponential functions). This makes difficult matching the $q$-deformed field theory with the corresponding firstly quantized quantum mechanics of particles on the q-deformed plane and due to the absence of the additivity property for the $q$-exponentials, makes an explicit calculation of nontrivial ({\em e.g.,\ } ${\varphi}^4$-) vertices impossible. Thus the results of \cite{CDP98} have left open the possibility that the complete interacting theory on the $q$-plane is UV-finite because of (dynamical) properties of the corresponding ${\varphi}^4$-vertices. In this paper we use another decomposition of the fields, namely, the decomposition in partial waves, similar to the recently proposed ``spherical field theory'' \cite{Lee} on commutative spaces. This decomposition together with the Haar ($E_q(2)$-invariant) measure and $q$-deformed integral used for the definition of the field theoretical action, allows to present the field theory on the $q$-deformed plane as a lattice theory of infinite number of interacting one-dimensional fields (partial waves). The resulting field theoretical degrees of freedom are in transparent correspondence with the spectrum of operators in the firstly quantized version of the model. The calculation of the tadpole with the account of the ${\varphi}^4$-vertex shows that UV-properties of the theory on the $q$-deformed plane are even worse than those on the ordinary commutative plane. This fact confirms the conclusion of the paper \cite{CDP98} that the very transition to the noncommutative space-times does not guarantee UV-finiteness. The example of the plane with the most simple and natural Heisenberg-like commutation relations among coordinates was used in \cite{CDP98} also for study of symmetry transformations of noncommutative space-times with Lie algebra commutation relations for coordinates. The noncommutative coordinates prove to be tensor operators, and we considered concrete examples of the corresponding transformations of localized states (analog of space-time point transformations). In this paper, we extend this consideration to the much more involved case of quantum group coaction on noncommutative space-times. More precisely, we derive the rules of transformations of particle states induced by the coaction of a quantum group. The paper is organized as follows. In section 2 we consider firstly quantized theory of particles on the $q$-deformed plane $P^{(2)}_q$ with $E_q(2)$-symmetry. We derive representations of the algebras of coordinates and momenta on the $q$-plane and find spectra of the relevant operators. In section 3 the field theory (second quantization) on $P^{(2)}_q$ is introduced and presented in the form of infinite number of interacting partial waves defined on a one-dimensional lattice, the partial wave at the sites of the lattice (interpreted as creation and annihilation operators) being in one-to-one correspondence with spectra of the quantum mechanical operators found in section 2. Calculation of a tadpole diagram shows that the model has even more severe UV-divergences than the standard two-dimensional scalar ${\varphi}^4$-theory. In section 4 we are interested in transformation properties of a system on $P^{(2)}_q$ under the coaction of the quantum group $E_q(2)$. The point is that now the coordinates ${\bar{z}}, z$ are noncommuting operators and $E_q(2)$ provides only existence of coaction, {{\em i.e.,}\ } homomorphism of the algebra of functions on $P^{(2)}_q$ into the direct product $E_q(2)\otimes P^{(2)}_q$ of algebras of functions on the quantum group and plane. Then the question is: how does this coaction influence states of a quantum system on $P^{(2)}_q$? In other words, if a system is in some state $\p$ (say, with a definite value of one of the coordinate operators, $z$ or ${\bar{z}}$ or some their combination) we are interested in determination of the state after the $E_q(2)$-group coaction. In subsection 4.1 we clarify a general formulation of this problem and then (in subsection 4.2) give an explicit answer for the $E_q(2)$ group. Section 5 is devoted to the summary of the results. \section{Quantum mechanics on the noncommutative plane with quantum $E_q(2)$ group symmetry} In this paper we consider Quantum Mechanics induced by a quantum group structure. Recall that in the case of ordinary Lie group $G$, the group structure defines a unique symplectic structure on the cotangent bundle $T^*_G$ to the group manifold $G$ (see, {\em e.g.,\ } \cite{LibermannM}) and, hence, the corresponding canonical quantization (via substitution of Poisson brackets by the corresponding commutators). A similar construction with necessary generalizations, can be carried out for Lie-Poisson groups, which after the quantization procedure become quantum groups (see, {\em e.g.,\ } review in \cite{ChaichianDbk} and refs. therein). In fact, the quantization of a system on a Lie group cotangent bundle $T^*_G$ corresponds to choice of the group manifold as a configuration space ({{\em i.e.,}\ } group parameters as space coordinates) and left- (or right-)invariant vector fields on $G$ (elements of the corresponding Lie algebra) as quantum mechanical momenta. Instead of using a whole group $G$, one can start from some of its coset (homogeneous) space $G/H$, where $H$ is a subgroup $H\subset G$. In this approach the basic problem of Quantum Mechanics, {{\em i.e.,}\ } determination of possible representations of canonical operators, is reduced to mathematically well-developed problem of construction of regular representation (or quasi-regular, if one deals with a homogeneous space) and its decomposition into irreducible parts (see, {\em e.g.,\ } \cite{BarutR}). In some particular cases this general construction becomes rather simple and quite familiar from elementary course on Quantum Mechanics. For example, let us consider a two-dimensional Euclidean group $E(2)=U(1)\times\hspace{-1.1 em}\supset T_2$ containing rotations and translations of a two-dimensional plane. Its homogeneous space $P^{(2)}=E(2)/U(1)$ is the Euclidean plane with the metric $\eta_{ij}={\rm{\,diag\,}}\{+1,+1\}$ which is invariant with respect to $E(2)$-transformations. This configuration space is parameterized by two coordinates $x_1, x_2$, while left-invariant fields tangent to this homogeneous space are nothing but usual derivatives which up to the factor $-\i\hbar$ correspond to the standard momentum operators. As is well-known, any representation of the algebra of coordinates and left-invariant vector fields on $P^{(2)}$ is unitary equivalent to this representation by the coordinate functions and derivatives in the Hilbert space ${\cal H}=\cL^2(\R^2)$ of square integrable functions. States $\p(x)\in{\cal H}$ are transformed according to representations of $E(2)$ in the Hilbert space ${\cal H}$. Since the coordinate operators are commuting, their eigenvalues are transformed under an action of $E(2)$-group as their classical counterparts: in the convenient complex notation $$ z=x_1+\i x_2\ ,\qquad \bar z=x_1-\i x_2\ , $$ the $E(2)$ transformations read as \begin{eqnarray} z\,\rightarrow\, z'&=&vz+t\,\nonumber\\[-1mm] &&\label{1.3}\\[-1mm] \bar z\,\rightarrow\, \bar z'&=&\bar v\bar z+\bar t\,\nonumber \end{eqnarray} where $v,\ \bar v$ subjected to the constraint $\bar vv=1$, define two-dimensional rotation group $U(1)$ and $t=t_1+\i t_2\ , \bar t=t_1-\i t_2$ parameterize translations. For the simple case of the quantum mechanical systems on the Euclidean plane the underlying mathematics related to cotangent bundle structures, regular representations {\em etc.\ } seems to be redundant. But for generalizations to more complicated homogeneous spaces, in particular, with non-zero curvature and nontrivial topology, the group theoretical methods become quite actual and powerful. In this work we are going to study another generalization: instead of starting from ordinary $E(2)$, we shall use its quantum version $E_q(2)$ \cite{VaksmanK,SchuppWZ,BonecciCGST}. Though in the case of quantum groups and corresponding quantum homogeneous spaces (definition of the latter see, {\em e.g.,\ } in \cite{BonecciCGST}) group parameters (coordinates) become noncommutative, the general scheme of quantization still can be applied. The role of momentum operators is now attributed the $q$-deformed left- (or right-) invariant generalizations of vector fields (see, {\em e.g.,\ } \cite{Zumino}). Thus the Planck constant $\hbar$ enters, as usual, the commutation relations (CR) for momenta and coordinates, while the group deformation parameter $q$ governs nontrivial coordinate-coordinate and momentum-momentum CR. Therefore, first of all we have to construct possible representations of this {\it combined} $q$-deformed algebra of noncommuting coordinates and momenta. For the particular case which we consider in this paper ($q$-deformed quantum Euclidean plane $P^{(2)}_q$) this is not a very complicated problem and we shall consider it in this section. We start from the quantum group $E_q(2)$ generated by elements ${\bar{v}},\ v,\ {\bar{t}},\ t$ with the defining relations \cite{VaksmanK} \begin{equation}\begin{array}{lllll} {\bar{v}} v = v{\bar{v}} = 1\ ,& \qquad &t{\bar{t}} = q^2 {\bar{t}} t\ ,& \\[3mm] vt= q^2tv\ ,& \qquad &{\bar{v}} t = q^{-2}t{\bar{v}} &\qquad& q\in{\Ibb R}\ .\end{array} \label{2.1} \end{equation} Other commutation relations follow from the involution: $v^{\dagger} ={\bar{v}},\ t^{\dagger}={\bar{t}}$. The comultiplication has the form \begin{equation} \begin{array}{cc} \Delta v = v \otimes v\ , & \Delta{\bar{v}} = {\bar{v}} \otimes {\bar{v}}\ , \\[2mm] \Delta t = v \otimes t + t \otimes {\leavevmode{\rm 1\mkern -5.4mu I}}\ , & \Delta{\bar{t}} = {\bar{v}} \otimes {\bar{t}} + {\bar{t}} \otimes {{\leavevmode{\rm 1\mkern -5.4mu I}}}\ . \end{array} \label{2.2} \end{equation} The explicit form of other basic maps for $E_q(2)$ (antipode, counity) will not be used in what follows. The unitary element $v$ can be parameterized with the help of the symmetric element $\theta} \newcommand{\T}{\Theta$: \begin{equation}\begin{array}{lcl} v={\rm{e}}^{\i\theta} \newcommand{\T}{\Theta}\ ,&\qquad&\theta} \newcommand{\T}{\Theta^{\dagger}=\theta} \newcommand{\T}{\Theta\ ,\\[2mm] \ &\ &\Delta\theta} \newcommand{\T}{\Theta=\theta} \newcommand{\T}{\Theta\otimes{\leavevmode{\rm 1\mkern -5.4mu I}}+{\leavevmode{\rm 1\mkern -5.4mu I}} \otimes \theta} \newcommand{\T}{\Theta\ .\end{array} \label{2.4} \end{equation} The corresponding quantum universal enveloping algebra (QUEA) ${\cal U}_qe(2)$ is generated by the elements $J,\ \bar T,\ T$ which are dual to the generators $\theta} \newcommand{\T}{\Theta,\ {\bar{t}},\ t$ of the algebra $E_q(2)$ and, as a result of the duality, satisfy the following commutation relations \begin{equation}\begin{array}{lll} [J,T]=\i T\ ,&\qquad& [J,\bar T]=-\i \bar T\ ,\\[2mm] T\bar T=q^2\bar T T&&\end{array} \label{2.6} \end{equation} (comultiplication and the other basic maps are also defined by the duality). The left action of elements from QUEA ${\cal U}_q L$ of an arbitrary Lie algebra $L$ on elements of the corresponding quantum group $G_q$ is defined by the expressions \begin{equation} \ell (X)f=({\rm{id}\,}\otimes X)\cdot\Delta f\equiv\sum_if^i_{(1)} \langle\langle X,f^i_{(2)}\rangle\rangle\ , \label{2.8} \end{equation} or \begin{equation} \lambda} \newcommand{\LD}{\Lambda (X)f=(S(X)\otimes {\rm{id}\,})\cdot\Delta f\equiv\sum_i\langle\langle S(X),f^i_{(1)}\rangle\rangle f^i_{(2)}\ , \label{2.8a} \end{equation} where $X\in{\cal U}_q L,\ f,f^i_{(1,2)}\in G_q$, $\langle\langle\cdot,\cdot\rangle\rangle$ denotes the duality contraction, $S(X)$ is antipode and where the comultiplication in $G_q$ is presented in the form $\Delta f= \sum_if^i_{(1)}\otimes f^i_{(2)}$. An explicit calculation of this left action in the case of $E_q(2)$ shows that the operators $\bar T,T\in{\cal U}_qe(2)$ act on elements of $E_q(2)$ generated by ${\bar{t}},\ t$ exactly in the same way as the $q$-deformed derivatives $\bar{\partial}_q,\ \partial_q$. In fact, the elements ${\bar{t}},\ t$ generate the $q$-deformed analog $P^{(2)}_q= E_q(2)/U_q(1)$ of the homogeneous space $P^{(2)}$, {{\em i.e.,}\ } generate the algebra of functions on quantum Euclidean plane \cite{BonecciCGST}. We shall denote elements of the algebra $P^{(2)}_q$ by ${\bar{z}},\ z$ to distinguish them from elements ${\bar{t}}, \ t$ of the algebra $E_q(2)$. The elements ${\bar{z}},\ z$ and $\bar{\partial}_q,\ \partial_q$ defines the $q$-deformed algebra of functions on $P^{(2)}_q$ together with the $q$-deformed left-invariant vector fields (derivatives). Its defining relations read as \begin{eqnarray} &&z{\bar{z}}=q^2{\bar{z}} z\ ,\qquad \partial_q{\bar{\partial}}_q=q^2{\bar{\partial}}_q\partial_q\nonumber\\[3mm] &&\partial_q z=1+q^{-2}z\partial_q\ ,\qquad{\bar{\partial}}_q {\bar{z}}=1+q^{2}{\bar{z}}{\bar{\partial}}_q\ ,\label{2.9}\\[3mm] &&{\bar{\partial}}_q z=q^{2}z{\bar{\partial}}_q\ ,\qquad\partial_q {\bar{z}}=q^{-2}{\bar{z}}\partial_q\ ,\nonumber \end{eqnarray} (the commutation relation for the $q$-derivatives is just the rewritten commutation relation for $\bar T,\ T$ \r{2.6} and those for the $q$-derivatives and coordinates are derived from \r{2.8}). If we put $q=1$ and define $p=-\i\hbar\partial,\ \bar p=-\i\hbar{\bar{\partial}}$, the relations \r{2.9} become the usual canonical commutation relations for a particle in two-dimensional space. The requirement of consistency with antipode dictates the following conjugation rule for the $q$-derivatives \cite{SchuppWZ} \begin{equation} \partial_q^{\dagger}=-q^2{\bar{\partial}}_q\ ,\qquad{\bar{\partial}}_q^{\dagger}=-q^{-2}\partial_q\ . \label{2.10} \end{equation} We consider the relations \r{2.9} as a $q$-deformation of the canonical commutation relations and is going to construct their representation in a Hilbert space. To this aim let us introduce the operators $N$ and $\bar{N}$ defined by the relations \begin{equation} [N;q^{-2}]=z\partial_q\ ,\qquad [\bar{N};q^{2}]={\bar{z}}{\bar{\partial}}_q\ , \label{2.11} \end{equation} \begin{equation} [X;q^{\a}]\equiv\frac{q^{\a X}-1}{q^{\a}-1}\ . \label{q-number} \end{equation} These operators have simple commutation relations \begin{eqnarray} &&q^{\a N}z=q^\a zq^{\a N}\ ,\qquad q^{\a \bar{N}}{\bar{z}}=q^\a {\bar{z}} q^{\a \bar{N}} \ ,\nonumber\\[-1mm] && \label{2.15} \\[-1mm] &&q^{\a N}\partial_q=q^{-\a}\partial_q q^{\a N}\ ,\qquad q^{\a\bar{N}}{\bar{\partial}}_q=q^{-\a}{\bar{\partial}}_q q^{\a\bar{N}}\ .\nonumber \end{eqnarray} Using \r{2.10} and $z^{\dagger}={\bar{z}}$, we find \begin{equation} N^{\dagger}=-\bar{N}-1\ ,\qquad \bar{N}^{\dagger}=-N-1\ . \label{2.16} \end{equation} The operators $q^{2\bar{N}},\ q^{2N}$ allow to construct commuting pairs of conjugate operators: \begin{eqnarray} &&\bar Z= q^{N-\bar{N}}{\bar{z}}\ ,\qquad Z= zq^{N-\bar{N}}\ ,\nonumber\\[-1mm] &&\label{2.20}\\[-1mm] &&\bar P= qq^{-(N-\bar{N})}{\bar{\partial}}_q\ ,\qquad P= -q^{-1}\partial_qq^{-(N-\bar{N})}\ ,\nonumber \end{eqnarray} with the commutation relations \begin{eqnarray} &&P\bar Z=\bar Z P\ ,\qquad \bar Z Z=Z\bar Z\ ,\qquad ZP=1+q^{2}PZ \ ,\nonumber\\[-1mm] &&\label{2.22}\\[-1mm] &&\bar P Z=Z\bar P\ ,\qquad \bar P P=P\bar P\ ,\qquad \bar P\bar Z=1+q^{2}\bar Z\bar P\ .\nonumber \end{eqnarray} If we were given only the algebra of the operators ${\bar{z}},\ z,\ {\bar{\partial}}_q,\ \partial_q$, we would reasonably name the commuting operators $\bar Z,\ Z$ by coordinates and $\bar P,\ P$ by the corresponding lattice momenta and then deal with two independent (commuting with each other) one-dimensional algebras on the $q$-lattice. However, fields in NC-QFT depend on noncommutative ($q$-commuting) coordinates ${\bar{z}},\ z$ which are more suitable to trace a result of coaction by $E_q(2)$. We have found convenient to use the hermitian and unitary combination of the coordinate operators: \begin{equation} r^2\equiv z{\bar{z}}\ \ \ \mbox{(hermitian)},\qquad u\equiv\sqrt{{\bar{z}} z^{-1}} \ \ \ \mbox{(unitary)}, \label{2.37a} \end{equation} together with $q^{(\bar{N}-N)}$ (hermitian operator) and $q^{2(\bar{N}+N+1)}$ (unitary operator) as a basic set of the phase space operators. The commutation relations for this set of operators read as \begin{equation}\begin{array}{lll} [q^{2(\bar{N}-N)},r^2]=0\ ,&\qquad &[q^{(\bar{N}+N+1)},u]=0\ ,\\[3mm] r^2u=q^2u r^2\ , & \qquad & [q^{2(\bar{N}-N)},q^{(\bar{N}+N+1)}]=0\ ,\\[3mm] q^{2(\bar{N}-N)}u=q^2u q^{2(\bar{N}-N)}\ , & \qquad & q^{(\bar{N}+N+1)}r^2=q^2r^2q^{(\bar{N}+N+1)}\ ,\end{array} \label{2.38} \end{equation} Now we are ready to construct a representation of this algebra in the space $\ell^2$ ({{\em i.e.,}\ } infinite dimensional matrix representation): \begin{eqnarray} r^2\mid n,m\rangle_{r_0,l_0} & = & r^2_0q^{2n}\mid n,m\rangle_{r_0,l_0} \nonumber\\[3mm] q^{2(\bar{N}-N)}\mid n,m\rangle_{r_0,l_0} & = & l_0q^{2m}\mid n,m\rangle_{r_0,l_0}\ . \label{2.39} \end{eqnarray} \begin{eqnarray} u\mid n,m\rangle_{r_0,l_0} & = & \mid n+1,m+1\rangle_{r_0,l_0}\ , \nonumber\\[3mm] q^{(\bar{N}+N+1)}\mid n,m\rangle_{r_0,l_0}& = &\mid n+1,m\rangle_{r_0,l_0}\ . \label{2.40} \end{eqnarray} The constants $r_0$ and $l_0$ mark different representations and from the eigenvalues of $r^2$ and $q^{2(\bar{N}-N)}$ it follows that in the ranges $[r_0,q^4 r_0)$ and $[l_0,q^4l_0)$ the representations are nonequivalent. The matrices $r^2,q^{2(\bar{N}-N)}$ are hermitian and $u,\ q^{(\bar{N}+N+1)}$ are unitary with respect to the scalar product defined by $$ {}_{r_0,l_0}\langle n,m\mid n',m'\rangle_{r_0,l_0}=\delta_{nn'}\delta_{mm'}\ . $$ Thus we have obtained that states of a particle on the quantum plane are characterized by discrete values of its radius-vector and discrete values of the operator $q^{2(\bar{N}-N)}$ which is obviously related to deformation of the angular momentum operator. Indeed, from \r{2.11} we conclude that the operator \begin{equation} J_q\equiv[\bar{N}-N;q^2]=\frac{q^{2(\bar{N}-N)}-1}{q^{2}-1}\ , \label{2.41} \end{equation} (which differs from $q^{2(\bar{N}-N)}$ by multiplication and shifting by the constants) in the continuum limit $q\ra1$ becomes the ordinary angular momentum operator. Therefore it is natural to consider $J_q$ as an appropriate deformation of the latter. Of course, discreteness of values of an angular momentum operator is not peculiar feature of $q$-deformed systems but general property of all quantum systems. Analogously, the natural $q$-deformation of the dilatation operator reads as \begin{equation} D_q\equiv[(\bar{N}+N+1);q^2]=\frac{q^{2(\bar{N}+N+1)}-1}{q^{2}-1}\ , \label{2.42} \end{equation} Another possibility for the construction of representations of the algebra \r{2.38} which will be convenient for us in the next section is to construct the representation in the basis of the unitary operators \begin{equation} u\equiv{\bar{z}} z^{-1}\ ,\qquad\mbox{and}\qquad q^{2(\bar{N}+N+1)}\ , \label{2.43} \end{equation} which commute with each other and, hence, have common eigenvalues. This basis, certainly, less suitable for construction of a matrix representation of the kind presented above since the two other (hermitian) operators do not shift an eigenvector of the operators \r{2.43} exactly into another eigenvector. However, this basis proves to be more suitable for the study of transformations of the states under the coaction of the quantum group $E_q(2)$ which we shall carry out in the section 4. \section{Quantum field theory on $P_q^{(2)}$ as a one-dimensional lattice theory for an infinite set of interacting fields} In this section we shall introduce the scalar ${\varphi}^4$-field theory on the noncommutative plane $P_q^{(2)}$ and present it in the form of infinite number of interacting partial waves defined on a one-dimensional lattice. \subsection{Preliminaries on ``the spherical field theory''} The starting idea of the spherical field theory \cite{Lee} in a usual commutative space-time is the representation of a $d$-dimensional Euclidean field theory as a theory for an infinite set of one-dimensional interacting fields. In what follows we shall confine ourselves with the simplest case of two-dimensional scalar theory. The initial action is quite standard: \begin{equation} S=\int\,d^2x\,\left[\big(\partial_i\bar{\varphi}\big)\big(\partial_i{\varphi}\big) +\mu^2\bar{\varphi}\vf+\frac{\lambda} \newcommand{\LD}{\Lambda}{2}\big(\bar{\varphi}\vf\big)^2 -j\bar{\varphi}-\bar j{\varphi}\right]\ . \label{pw1} \end{equation} Decomposing ${\varphi}(x)$ and $j(x)$ into partial waves \begin{eqnarray} {\varphi}(x)&=&{\varphi}(r,\a)=\frac{1}{\sqrt{2\pi}}\sum_{N=-\infty}^\infty{\varphi}_N(r) {\rm{e}}^{\i N\a}\ , \label{pw2}\\[3mm] j(x)&=&j(r,\a)=\frac{1}{\sqrt{2\pi}}\sum_{N=-\infty}^\infty j_N(r) {\rm{e}}^{-\i N\a}\ , \label{pw2a} \end{eqnarray} one can rewrite \r{pw1} as \begin{eqnarray} S&=&\sum_{N=-\infty}^\infty \int_0^\infty\,dr\left[r \pad{\bar{\varphi}_N}{r}\pad{{\varphi}_N}{r}+\frac{\mu^2 r^2+N^2}{r}\bar{\varphi}_N{\varphi}_N -rj_N\bar{\varphi}_N-r\bar j_N{\varphi}_N\right]\nonumber\\[3mm] &+&\frac{\lambda} \newcommand{\LD}{\Lambda}{2}\sum_{N,M,K,L=-\infty}^\infty\int_0^\infty\,dr\,r\, \Big(\bar{\varphi}_N{\varphi}_M\bar{\varphi}_K{\varphi}_L\,\d_{N-M+K-L,0}\Big)\ . \label{pw3} \end{eqnarray} Let $\widetilde{G}(k)$ denote the usual Green function in the momentum representation \begin{equation} \widetilde{G}(k)=\int\,d^2x\,{\rm{e}}^{\i \vec{k}\vec{x}}\<0|\bar{\varphi}(x){\varphi}(0)|0\>\ .\label{pw4} \end{equation} Then the propagator for the $N$-th partial wave proves to be \begin{equation} \<0|\bar{\varphi}_N(r_1){\varphi}_N(r_2)|0\>=\int\,dk\,k\,J_{|N|}(kr_1) J_{|N|}(kr_2)\widetilde{G}(k)\ . \label{pw5} \end{equation} Here $J_N(kr)$ is the Bessel function of the first kind and $k\equiv\sqrt{k_1^2+k_2^2}$. For the scalar field theory the propagator has the form \begin{equation} \widetilde{G}(k)=\frac{1}{k^2+\mu ^2}\ , \label{pw6} \end{equation} so that \r{pw5} gives \begin{eqnarray} G_N(r_1,r_2)&\equiv&\<0|\bar{\varphi}_N(r_1){\varphi}_N(r_2)|0\>\nonumber\\[3mm] &=&\int\,dk\,k\,J_{|N|}(kr_1)J_{|N|}(kr_2)\frac{1}{k^2+\mu ^2}\nonumber\\[3mm] &=&\theta(r_1-r_2)K_{\left|N\right|}(\mu r_1)I_{\left|N\right| }(\mu r_2) +\theta (r_2-r_1)K_{\left|N\right| }(\mu r_2)I_{\left| N\right| }(\mu r_1)\ , \label{pw7} \end{eqnarray} where $\theta} \newcommand{\T}{\Theta(r)$ is the step-function; $I_N,\ K_N$ are the modified Bessel functions of the first and second kind respectively. The principal aim of the ``spherical field theory'' (SQFT) is the development of a nonperturbative approach to calculations in the standard QFT. We are interested in UV-behaviour of perturbation expansion in the quantum field theory on the noncommutative plane (NC-QFT) which we are going to present in the form similar to the SQFT. Thus to make a comparison, let us first find out how the UV-divergences of the ordinary (two-dimensional, scalar) field theory reveal themselves in SQFT. To this aim, we consider the tadpole diagram depicted in figure~\ref{F:tadp.SQFT}. \begin{figure}\centering \unitlength=1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt} \begin{picture}(35.00,24.00) \put(5.00,10.00){\line(1,0){30.00}} \put(20.00,10.00){\circle*{1.00}} \put(20.00,17.00){\circle{14.00}} \put(10.00,7.00){\makebox(0,0)[cc]{$G_K$}} \put(30.00,7.00){\makebox(0,0)[cc]{$G_K$}} \put(20.00,7.00){\makebox(0,0)[cc]{$r$}} \put(20.00,21.00){\makebox(0,0)[cc]{$G_N$}} \put(9.00,10.00){\vector(1,0){2.00}} \put(27.00,10.00){\vector(1,0){2.00}} \put(19.00,24.00){\vector(1,0){1.00}} \end{picture} \caption{\rm The tadpole diagram in the spherical field theory \label{F:tadp.SQFT}} \end{figure} This diagram is proportional to the factor \begin{eqnarray} &&\sum_{N=-\infty}^\infty G_N(r,r)=\sum_{N=-\infty}^\infty K_{\left| N\right| }(\mu r)I_{\left| N\right| }(\mu r)\nonumber\\[3mm] &&\ \ =K_0(\mu r)I_0(\mu r) +2\sum_{N=1}^\infty K_{\left| N\right| }(\mu r)I_{\left| N\right| }(\mu r)\ . \label{pw8} \end{eqnarray} It is seen that the Green function $G_N(r_1,r_2)$ for a fixed partial wave is not singular at coinciding arguments: $G_N(r,r)$ has well-defined values for any $r$ and $N=0,\pm1,\pm2,...\,.$ However, the tadpole diagram is still divergent: the divergence appears in the summation over the angular momentum numbers $N$. Indeed, let us for simplicity consider small values of $\mu r\ll 1$, so that we can use the asymptotic expressions: \begin{eqnarray*} \left.I_N(\mu r)\right|_{\mu r\ll 1}&\approx&\frac{1}{N!} \left(\frac{\mu r}{2}\right)^N\ ,\qquad N=0,\pm1,\pm2,...\ ,\nonumber\\[3mm] \left.K_N(\mu r)\right|_{\mu r\ll 1}&\approx&\frac{(N-1)!}{2} \left(\frac{2}{\mu r}\right)^N\ ,\qquad |N|\geq1\ ,\nonumber\\[3mm] \left.K_0(\mu r)\right|_{\mu r\ll 1}&\approx&\ln\frac{2}{\gamma} \newcommand{\G}{\Gamma\mu r} \ ,\qquad \mbox{($\gamma} \newcommand{\G}{\Gamma$ is the Euler constant)}\ . \end{eqnarray*} Thus for $\mu r\ll 1$ the tadpole is proportional to $$ \sum_{N=-\infty}^\infty G_N(r,r)=\ln\frac{2}{\gamma} \newcommand{\G}{\Gamma\mu r}+\sum_{N=1}^\infty \frac{1}{N}\longrightarrow\infty\ , $$ so that it is (logarithmically) divergent as it should be. Of course, the same is true for any values of $\mu r$, though an explicit demonstration of this fact becomes more involved. In order to circumvent such a calculation with the special (Bessel) functions, we can present the action \r{pw3} in a modified form. First, if we are interested only in UV-properties of the model, we can drop out the mass term. However, the massless theory in two dimension has the infra-red divergences which are also logarithmic for the tadpole diagram and may distort the true picture of the UV-behaviour of the model. Thus, in the massless case we need some IR-regularization and to achive it, let us introduce into the (massless) action \r{pw3} the additional term of the form \begin{equation} S^{(IR)}=\int_0^\infty\,dr\, \frac{\sigma} \renewcommand{\S}{\Sigma^2}{r}\bar{\varphi}_N(r){\varphi}_N(r)\ . \label{exp1} \end{equation} Now the free action reads as \begin{equation} S^{(0)}=\sum_{N=-\infty}^\infty \int_0^\infty\,dr\left[r \pad{\bar{\varphi}_N}{r}\pad{{\varphi}_N}{r}+\frac{N^2+\sigma} \renewcommand{\S}{\Sigma^2}{r}\bar{\varphi}_N{\varphi}_N \right]\ , \label{exp2} \end{equation} and after an introduction of the new coordinate $y$ defined by the relation \begin{equation} y=\ln(\mu r)\ ,\qquad -\infty < y < \infty\ , \label{exp3} \end{equation} it acquires the simple form of the standard one-dimensional scalar action: \begin{equation} S_0=\sum_{N=-\infty}^\infty \int_{-\infty}^\infty\,dy\left[ \pad{\bar{\varphi}_N}{y}\pad{{\varphi}_N}{y}+(N^2+\sigma} \renewcommand{\S}{\Sigma^2)\bar{\varphi}_N{\varphi}_N\right]\ .\label{exp4} \end{equation} The free Green function for this action can be easily found by the use of the Fourier\ transform and proves to be the following \begin{equation} G_N(y_1-y_2)=\frac{1}{2M}{\rm{e}}^{-M|y_1-y_2|}\ ,\qquad M\equiv\sqrt{N^2+\sigma} \renewcommand{\S}{\Sigma^2}\ , \label{exp4a} \end{equation} or, in terms of the initial radial coordinates, \begin{equation} G_N(r_1,r_2)=\frac{1}{2M}\left[\theta} \newcommand{\T}{\Theta(r_1-r_2)\left(\frac{r_2}{r_1}\right)^M +\theta} \newcommand{\T}{\Theta(r_2-r_1)\left(\frac{r_1}{r_2}\right)^M\right]\ . \label{exp5} \end{equation} This explicit expression shows immeditely that the tadpole diagram in figure~\ref{F:tadp.SQFT} is proportional to the sum \begin{equation} \sum_{N=-\infty}^\infty G_N(r,r)= \sum_{N=-\infty}^\infty\frac{1}{\sqrt{N^2+\sigma} \renewcommand{\S}{\Sigma^2}}\ , \label{exp6} \end{equation} and, hence, logarithmically divergent. The IR-regularization parameter $\sigma} \renewcommand{\S}{\Sigma$ is inessetial for large $N$ and does not influence on the UV-behaviour of the model (as it should be). The form \r{exp4} of the action and the UV-behaviour analysis following it will be useful for us in the noncommutative case as well. We are going to show that the ${\varphi}^4$-theory on the $q$-plane can be rewritten as a theory of the partial waves on a one-dimensional (nonequidistant) lattice with the behaviour in $N$ (angular momentum number) being even worse, so that UV-divergences of the NC-QFT are even more severe than those of the usual scalar theory. \subsection{Quantum field theory on the $q$-plane and its partial wave decomposition} Let us consider the generalization of the two-dimensional scalar field theory, induced by the noncommutativity \r{2.9} of the space coordinates on the plane $P^{(2)}_q$. The field action for ${\varphi}(z,{\bar{z}})$ can be defined with the help of the $q$-deformed Haar (invariant) measure \cite{Koelink}. For a function $F_N(z,{\bar{z}})=z^Nf(z{\bar{z}})$ on $P^{(2)}_q$ one defines the linear functional ($q$-integral): \begin{eqnarray} H_{r_0}[F_N]&\equiv&\int^q\,d^2z\,z^Nf(z{\bar{z}})\nonumber\\[3mm] &\equiv&\d_{N,0}r^2_0(q^2-1)\sum_{k=-\infty}^\infty q^{2k}f(q^{2k}r^2_0) \label{pw9} \end{eqnarray} ($r_0$ labels the nonequivalent representations \r{2.39} of the $q$-deformed coordinate-momentum algebra). In formula \r{pw9} and in what follows we assume, for definiteness, that $q^2>1$ (the quite similar construction can be carried out for $q<1$, {\em cf.\ } \r{corr19}). In order that the sum on the right hand side of \r{pw9} be meaningful, the function $f(z{\bar{z}})$ must satisfy an appropriate conditions at infinity \cite{Koelink}. We shall assume that the set of the fields on the $q$-plane which we consider below does satisfy this condition. Notice that if $N<0$ in \r{pw9}, the integrand can be rewritten as ${\bar{z}}^{|N|}f'(z{\bar{z}})$ ($f'(z{\bar{z}})$ is some modification of the function $f(z{\bar{z}})$). Using the $q$-integral \r{pw9}, we can define the action on the $q$-plane as the straightforward generalization of the usual ${\varphi}^4$-action \r{pw1}: \begin{eqnarray} S_q&=&\int^q\,d^2z\,\bigg[-\bar{\varphi}(z,{\bar{z}})\partial_q\bar{\partial}_q{\varphi}(z,{\bar{z}}) +\mu^2\bar{\varphi}(z,{\bar{z}}){\varphi}(z,{\bar{z}})\nonumber\\[3mm] &&+\frac{\lambda} \newcommand{\LD}{\Lambda}{2}\bigg(\bar{\varphi}(z,{\bar{z}}){\varphi}(z,{\bar{z}})\bigg)^2 -j(z,{\bar{z}})\bar{\varphi}(z,{\bar{z}})-\bar j(z,{\bar{z}}){\varphi}(z,{\bar{z}})\bigg]\ . \label{pw10} \end{eqnarray} Now our aim is to rewrite the action \r{pw10} in the form similar to \r{pw3} where the integral over radial variables is substituted by the sum \r{pw9} while the sum over the angular momentum numbers remains in the $q$-deformed case too. To achieve this, we decompose a field on $P^{(2)}_q$ into terms with definite eigenvalues of the $q$-deformed angular momentum operator $J_q$ ({\em cf.\ } \r{2.41}): \begin{equation} {\varphi}(z,{\bar{z}})=\sum_{N=-\infty}^\infty z^Nr^{-N}{\varphi}_N(r) \ ,\qquad r^2\equiv z{\bar{z}}\ . \label{pw11} \end{equation} Again, as in the case of the expression \r{pw9}, it is worth to notice that the terms with $N<0$ in the sum \r{pw11} can be rewritten in the form with positive powers of ${\bar{z}}$: \begin{equation} z^Nr^{-N}=q^{-|N|(|N|-1)}{\bar{z}}^{|N|}r^{-|N|}= q^{-N(N+1)}{\bar{z}}^{-N}r^{N}\ ,\qquad (N<0)\ . \label{pw12} \end{equation} Here we have used the relation: \begin{equation} z^N{\bar{z}}^N=q^{N(N-1)}(z{\bar{z}})^N\equiv q^{N(N-1)}r^{2N}\ . \label{pw13} \end{equation} The next step is the substitution of the decomposition \r{pw11} into the action \r{pw10} and then use of the definition \r{pw9} to convert the action into a lattice one. In order to do this, one needs the following commutation relations which are derivable from \r{2.9}: \begin{eqnarray} &&\sqrt{z}\sqrt{{\bar{z}}}=\sqrt{q}\sqrt{{\bar{z}}}\sqrt{z}\ ,\qquad z^Nr^{-N}=q^{-N(N-1/4)}{\bar{z}}^{-N/2}z^{N/2}\ ,\nonumber\\[3mm] &&zr=qrz\ ,\qquad {\bar{z}} r=q^{-1}r{\bar{z}}\ ,\nonumber\\[3mm] &&\partial_q\sqrt{{\bar{z}}}=q^{-1}\sqrt{{\bar{z}}}\partial_q\ ,\qquad \partial_q\sqrt{z}=\frac{1}{q^{-1}+1}\frac{1}{\sqrt{z}}+q^{-1}\sqrt{z}\partial_q\ ,\nonumber\\[3mm] &&\bar{\partial}_q\sqrt{z}=q\sqrt{z}\bar{\partial}_q\ ,\qquad \bar{\partial}_q\sqrt{{\bar{z}}}=\frac{1}{q+1}\frac{1}{\sqrt{{\bar{z}}}}+q\sqrt{{\bar{z}}}\bar{\partial}_q \ , \label{pw14}\\[3mm] &&\partial_q z^{N/2}=\frac{[N;q^{-1}]}{q^{-1}+1}z^{N/2-1}+q^{-N}z^{N/2}\partial_q\ ,\nonumber\\[3mm] &&\bar{\partial}_q{\bar{z}}^{-N/2}=-\frac{[N;q^{-1}]}{q(q+1)}{\bar{z}}^{-N/2-1}+q^{-N} {\bar{z}}^{-N/2}\bar{\partial}_q\ ,\nonumber\\[3mm] &&\partial_q\bar{\partial}_q f(r)=\frac{1}{2+q+q^{-1}}\left[\frac{1}{r}D^{(r)}_{q^{-1}} rD^{(r)}_{q}\right]f(r)\ . \nonumber \end{eqnarray} The Jackson derivatives in the last line are defied as follows: \begin{eqnarray} D^{(r)}_{q^{-1}}f(r)&=&\frac{f(q^{-1}r)-f(r)}{r(q^{-1}-1)}\ ,\nonumber\\[-1mm] && \label{corr0}\\[-1mm] D^{(r)}_{q}f(r)&=&\frac{f(qr)-f(r)}{r(q-1)} \nonumber \end{eqnarray} (these definitions imply that we are working with the representation where the operator $r$ is diagonal). Use of these relations together with \r{pw9} and \r{pw11} allows to present the action \r{pw10} in the form of the Jackson integral over the radial variable $r^2$: \begin{eqnarray} S_q&=&\sum_{N=-\infty}^\infty\int^J\,dr^2\, \Bigg\{q^{-N(N+1)}\bar{\varphi}_{N}(r^2)\bigg[-\frac{q^3}{(q+1)^2}\Delta_q\nonumber\\[3mm] &&+q^{2(N+1)}\frac{[N;q^{-1}]^2}{(q+1)^2r^2} -q^{N+1}\frac{[N;q^{-1}]}{q+1)}\Big(q^2 D_{q^2}^{(r^2)}- D_{q^{-2}}^{(r^2)} \Big)\bigg]{\varphi}_{N}(r^2)\nonumber\\[3mm] &&+\mu^2 q^{-N(N+1)}\bar{\varphi}_{N}(r^2){\varphi}_{N}(r^2)\nonumber\\[3mm] &&+\frac{\lambda} \newcommand{\LD}{\Lambda}{2}\sum_{M,K=-\infty}^\infty q^{2k}q^{-N^2-K^2-M^2-MK+NK+NM-M-K}\nonumber\\[3mm] &&\times\bar{\varphi}_{N}(r^2){\varphi}_{N}(q^{2(N-M)}r^2) \bar{\varphi}_{N}(q^{2(N-M)}r^2){\varphi}_{N}(r^2)\nonumber\\[3mm] &&+ q^{-N(N+1)}\bar{j}_{N}(r^2){\varphi}_{N}(r^2)+q^{-N(N+1)} \bar{\varphi}_{N}(r^2)j_{N}(r^2)\Bigg\}\ . \label{pw15m} \end{eqnarray} Here the Jackson integral for and arbitrary function $f(r^2)$ is defined in the standard way (see, {\em e.g.,\ } \cite{ChaichianDbk}): \begin{equation} \int^J\,dr^2\,f(r^2)\,\stackrel{\rm def}{\equiv}\, r_0^2|q^2-1|\sum_{k=-\infty}^{\infty}q^{2k} f(q^{2k}r^2_0)\ , \label{corr1} \end{equation} and the Jackson derivatives are defined by \r{corr0} and by the following similar relations: \begin{eqnarray} D^{(r^2)}_{q^{-2}}f(r^2)&=&\frac{f(q^{-2}r^{2})-f(r^{2})}{r^{2}(q^{-2}-1)} \ ,\nonumber\\[-1mm] && \label{corr2}\\[-1mm] D^{(r^2)}_{q^{2}}f(r^{2})&=&\frac{f(q^{2}r^{2})-f(r^{2})}{r^{2}(q^{2}-1)} \ .\nonumber \end{eqnarray} The radial part $\Delta_q$ of the $q$-deformed Laplacian reads as \begin{equation} \Delta_q=\frac{1}{r}D^{(r)}_{q^{-1}}rD^{(r^2)}_{q}\ . \label{corr3} \end{equation} Notice that the expression \r{pw15m} for the action $S_q$ in term of the Jackson integral is equally correct both for the case $q>1$ and for $q<1$. For definiteness, we continue to discuss the case $q>1$. The consideration for the case $q<1$ is essentially the same and we shall present for it only the result ({\em cf.\ } \r{corr19}). Since both the Jackson integral and the Jakson derivatives turn into their nondeformed (continuous) counterparts in the limit $q\ra1$, it is readily seen that the action \r{pw15m} becomes in this limit the usual action \r{pw3} for the two-dimensional scalar theory in the polar coordinates. Now we proceed to study the UV-behaviour of the field theory on the $q$-deformed plane. Therefore, we again, similarly to the nondeformed case in the preceding subsection ({\em cf.\ } \r{exp1}), substitue the mass term with the IR-regularizing term \begin{equation} S_q^{(IR)}=\sum_{N=-\infty}^\infty\int^J\,dr^2\, \frac{q^{-N(N+1)}\sigma} \renewcommand{\S}{\Sigma^2}{r^2}\bar{\varphi}_{N}(r^2){\varphi}_{N}(r^2)\ . \label{corr5} \end{equation} The exponential dependence of the fields in \r{pw15m},\r{corr1} on the space (discrete) variable $k$ inspires to make the substitution similar to that for the nondeformed model ({\em cf.\ } \r{exp3}) and to denote: \begin{eqnarray*} &&\bar{\varphi}_{Nk}\equiv\bar{\varphi}_N(q^{2k}r_0^2)\ ,\qquad {\varphi}_{Nk}\equiv{\varphi}_N(q^{2k}r_0^2)\ ,\nonumber\\[3mm] &&\bar j_{Nk}\equiv\bar j_N(q^{2k}r_0^2)\ ,\qquad j_{Nk}\equiv j_N(q^{2k}r_0^2)\ . \end{eqnarray*} In this notation the action \r{pw15m} in which the mass term is subsituted by \r{corr5}, acquires the form of an action for infinite number of scalar fields on a one-dimensional lattice: \begin{eqnarray} S_q&=&S_q^{(0)}+S_q^{(int)}+S_q^{(e.s.)}\ , \label{corr6}\\[3mm] S_q^{(0)}&=&\sum_{N=-\infty}^\infty\,A_N\,\sum_{k=-\infty}^\infty \left[\frac{({\varphi}_{Nk+1}-{\varphi}_{Nk})^2}{a} +aM_N^2\bar{\varphi}_{Nk}{\varphi}_{Nk}\right]\ , \label{corr7}\\[3mm] S_q^{(int)}&=&\frac{\lambda} \newcommand{\LD}{\Lambda r_0^2}{2}(q^2-1) \sum_{M,N,K=-\infty}^\infty\sum_{k=-\infty}^\infty q^{-N^2-k^2-M^2-MK+NK+NM-M-K}q^{2k}\nonumber\\[3mm] &&\phantom{+\frac{\lambda} \newcommand{\LD}{\Lambda r_0^2}{2}(q^2-1) \sum_{M,N,K=-\infty}^\infty}\times\,\bar{\varphi}_{Mk}{\varphi}_{N(k+M-N)} \bar{\varphi}_{K(k+M-N)}{\varphi}_{(M-N+K)k}\ , \label{corr8}\\[3mm] S_q^{(e.s.)}&=&r_0^2(q^2-1)\sum_{N,k=-\infty}^\infty q^{2k}q^{-N(N+1)}(\bar j_{Nk}{\varphi}_{Nk}+\bar{\varphi}_{Nk}j_{Nk})\ ,\nonumber \end{eqnarray} where \begin{eqnarray} &&A_N=q^{-N^2+4}\ , \label{corr9}\\[3mm] &&M_N^2=\frac{A_N}{R_N}\ , \label{corr10}\\[3mm] &&R_N=q^{-N(N+1)+4}\left(\frac{[N;q]^2}{(q+1)^2}+ \frac{\sigma} \renewcommand{\S}{\Sigma^2}{q^4}\right)\ , \label{corr11}\\[3mm] &&a=q^2-1\ . \label{corr12} \end{eqnarray} It is obvious that $A_N>0,\ R_N>0$ for all $N=0,\pm1,\pm2,...$. This justifies the definition \r{corr10} and shows that the quadratic part of the Euclidean action \r{pw10} is positively defined. The latter fact, in turn, provides that the generating functional for Green functions in the model with the action \r{pw10} (or, in the lattice form, \r{corr6}-\r{corr8}) given by the infinite dimensional integral (discrete lattice analog of the path integral): \begin{equation} {\bm{\cZ}}[j]=\frac{\int\,\prod_{N,k}\,d\bar{\varphi}_{Nk}\,d{\varphi}_{Nk}\,\exx{-S}} {\int\,\prod_{N,k}\,d\bar{\varphi}_{Nk}\,d{\varphi}_{Nk}\, \exx{-\left.S\right|_{j=\bar j=0}}} \ , \label{pw18} \end{equation} can be calculated by the perturbation expansion. A few remarks are in order: \ben \item The fields $\bar{\varphi}_N,\ {\varphi}_N$ at points $q^{2k}r_0,\ k=0,\pm1,\pm2,...$, {{\em i.e.,}\ } the quantities $\bar{\varphi}_{Nk},\ {\varphi}_{Nk}$, can be considered as the creation and annihilation operators of particles on the quantum plane $P^{(2)}_q$ in the states \r{2.39}. Thus the field model with the action \r{corr6} is the secondary quantized theory of the particles on the quantum $q$-plane $P_q^{(2)}$. \item The quadratic part \r{corr7} of the action \r{corr6} has the standard form of the lattice scalar theory ({\em cf.\ } {\em e.g.,\ } \cite{Creutz}), so that we can use the standard method of the Fourier\ transform in order to diagonalize it. \item As a result of the nonequidistance of the $q$-lattice, the mass term in \r{pw15m} (the second line) looks as if the model interacts with the external field, {{\em i.e.,}\ } the mass term contains additional factors $q^{2k}$ under the sign of the sum. This makes diagonalization of the complete quadratic part of the action \r{pw15m} rather involved problem. \item It is rather striking result that the action on the $q$-plane is not only lattice-like but also nonlocal, as is seen from the interaction term of the action in the form \r{corr8}. \end{enumerate} The quadratic part $S^{(0)}_q$ of the action can be diagonalized by performing the Fourier transform \begin{equation} {\varphi}_{Nk} = \int^{\pi /a}_{-\pi /a}\,\frac{dp}{2\pi}\,{\rm{e}}^{\i apk}\, \widetilde{{\varphi}}_N (p)\ . \end{equation} Then \begin{equation} S_q^{(0)}=\sum_{N=-\infty}^\infty\,A_N\,\int^{\pi /a}_{-\pi /a}\, \frac{dp}{2\pi}\,\widetilde{\bar{\varphi}}_N (p) \left[\frac{2}{a^2}\Big(1-\cos(ap)\Big) +M^2_N \right]\widetilde{{\varphi}}_N (p)\ , \end{equation} and the free Green function $G_N(k-m)$ has the usual for the lattice field theories form (see, {\em e.g.,\ } \cite{Creutz}): \begin{equation} G_N(k,m)=\int_{-\pi/a}^{\pi/a}\,\frac{dp}{2\pi}\, \frac{{\rm{e}}^{\i p(k-m)}}{A_N\left[M_N^2+\frac{2}{a^2}\left(1- \cos(ap)\right)\right]}\ . \label{corr14} \end{equation} Together with the ${\varphi}^4$-vertex of the action \r{corr8}, this free propagator defines the Feynman rules for the model under consideration which are depicted in figure~\ref{F:feyn.rules}. \begin{figure}\centering \unitlength=1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt} \begin{picture}(54.00,33.00)(50.00,00.00) \put(10.00,10.00){\rule{6.00\unitlength}{2.00\unitlength}} \put(10.00,11.00){\circle*{2.00}} \put(16.00,11.00){\circle*{2.00}} \put(9.00,11.00){\line(-1,-1){4.00}} \put(17.00,11.00){\line(1,1){4.00}} \put(17.00,11.00){\line(1,-1){4.00}} \put(9.00,11.00){\line(-1,1){4.00}} \put(7.00,9.00){\vector(1,1){1.00}} \put(19.00,9.00){\vector(1,-1){1.00}} \put(19.00,13.00){\vector(-1,-1){1.00}} \put(7.00,13.00){\vector(-1,1){1.00}} \put(5.00,5.00){\makebox(0,0)[cc]{${\scriptstyle M-N+K}$}} \put(5.00,17.00){\makebox(0,0)[cc]{${\scriptstyle M}$}} \put(21.00,5.00){\makebox(0,0)[cc]{${\scriptstyle K}$}} \put(21.00,17.00){\makebox(0,0)[cc]{${\scriptstyle N}$}} \put(7.00,11.00){\makebox(0,0)[cc]{${\scriptscriptstyle k}$}} \put(19.00,11.00){\makebox(0,0)[lc]{${\scriptscriptstyle (k+M-N)}$}} \put(5.00,30.00){\vector(1,0){8.00}} \put(13.00,30.00){\line(1,0){8.00}} \put(13.00,33.00){\makebox(0,0)[cc]{${\scriptstyle N}$}} \put(64.00,30.00){\makebox(0,0)[lc]{$G_N(k,l)$}} \put(6.00,28.00){\makebox(0,0)[cc]{${\scriptstyle k}$}} \put(20.00,28.00){\makebox(0,0)[cc]{${\scriptstyle l}$}} \put(64.00,11.00){\makebox(0,0)[lc] {$\frac12\lambda(q^2-1)q^{2k}q^{-N^2-K^2-M^2-MK+NK+MN-M-K}$}} \put(46.00,30.00){\makebox(0,0)[cc]{$\Leftrightarrow$}} \put(46.00,11.00){\makebox(0,0)[cc]{$\Leftrightarrow$}} \end{picture} \caption{\rm Feynman rules (free propagator and the nonlocal vertex) for the scalar theory on the noncommutative plane $P^{(2)}_q$\label{F:feyn.rules}} \end{figure} We shall not carry out detailed perturbative calculations: because of the nonequidistance of the $q$-lattice such calculations (especially in the case of nonzero mass) prove to be rather cumbersome. Notice, however, that these peculiarities of the $q$-lattice seems to be not a difficulty for computer simulations. In this paper we shall demonstrate only that the UV-divergences retain in the scalar field theory on the $q$-deformed plane. Let us consider the tadpole diagrams presented in figure~\ref{F:tadp}. \begin{figure}\centering \unitlength=1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt} \begin{picture}(59.00,34.00) \put(10.00,12.00){\rule{6.00\unitlength}{2.00\unitlength}} \put(10.00,13.00){\circle*{2.00}} \put(16.00,13.00){\circle*{2.00}} \put(9.00,13.00){\line(-1,-1){4.00}} \put(17.00,13.00){\line(1,-1){4.00}} \put(7.00,11.00){\vector(1,1){1.00}} \put(19.00,11.00){\vector(1,-1){1.00}} \put(5.00,7.00){\makebox(0,0)[cc]{${\scriptstyle N}$}} \put(21.00,7.00){\makebox(0,0)[cc]{${\scriptstyle N}$}} \put(7.00,13.00){\makebox(0,0)[cc]{${\scriptscriptstyle l}$}} \put(19.00,13.00){\makebox(0,0)[cc]{${\scriptscriptstyle l}$}} \put(13.00,16.50){\oval(12.00,7.00)[]} \put(12.00,20.00){\vector(1,0){2.00}} \put(13.00,22.00){\makebox(0,0)[cc]{${\scriptstyle M}$}} \put(13.00,3.00){\makebox(0,0)[cc]{a)}} \put(48.00,15.00){\rule{2.00\unitlength}{6.00\unitlength}} \put(49.00,15.00){\circle*{2.00}} \put(49.00,21.00){\circle*{2.00}} \put(49.00,27.00){\circle{10.00}} \put(49.00,32.00){\vector(1,0){1.00}} \put(49.00,34.00){\makebox(0,0)[cc]{${\scriptstyle M}$}} \put(49.00,24.00){\makebox(0,0)[cc]{${\scriptscriptstyle l}$}} \put(49.00,14.00){\line(-2,-1){10.00}} \put(49.00,14.00){\line(2,-1){10.00}} \put(43.00,11.00){\vector(2,1){2.00}} \put(53.00,12.00){\vector(2,-1){2.00}} \put(51.00,15.00){\makebox(0,0)[lc]{${\scriptscriptstyle l+M-N}$}} \put(39.00,7.00){\makebox(0,0)[cc]{${\scriptstyle N}$}} \put(59.00,7.00){\makebox(0,0)[cc]{${\scriptstyle N}$}} \put(49.00,3.00){\makebox(0,0)[cc]{b)}} \end{picture} \caption{\rm Two types of the tadpole diagrams in the model under consideration\label{F:tadp}} \end{figure} We confine our consideration to the diagram~\ref{F:tadp}.a because its analysis is a bit easier than that for the diagram~\ref{F:tadp}.b. Using the generating functional \r{pw18} and the Feynman rules we find the following expression for the tadpole~\ref{F:tadp}.a: \begin{equation} \frac12\lambda} \newcommand{\LD}{\Lambda(q^2-1)q^{-M(M+1)}\sum_{l=-\infty}^\infty q^{2l}G_M(k,l)G_M(l,m)\sum_{N=-\infty}^\infty q^{-N(N+1)}G_N(0)\ .\label{pw20} \end{equation} The partial wave propagator $G_M(0)$ is finite at the coincident arguments (similar to the case of the ordinary ${\varphi}^4$-model on a commutative plane,{\em cf.\ } \r{pw7},\r{pw8} and \r{exp4a},\r{exp5}) and is given by the relation \r{corr14}. The simple integration yields \begin{equation} G_M(0)=\frac{1}{\sqrt{R_N(4A_N+a^2R_N)}}\ . \label{corr15} \end{equation} The divergence again appears in the summation over the angular momentum numbers: as is seen from \r{pw20}, the tadpole contribution is proportional to the factor \begin{eqnarray} \lefteqn{\sum_{N=-\infty}^\infty q^{-N(N+1)}G_N(0)}\nonumber\\[3mm] &&= \sum_{N=-\infty}^\infty\frac{q^2-1}{q^4\sqrt{[(q^N-1)^2+(q^2-1)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4] [(q^N+1)^2+(q^2-1)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4]}}\ . \label{corr16} \end{eqnarray} The terms in this series have the following asymptotics: \begin{eqnarray} \frac{q^2-1}{q^4\sqrt{[(q^N-1)^2+(q^2-1)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4] [(q^N+1)^2+(q^2-1)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4]}}&\limar{N}{\infty}& (q^2-1)q^{-(N+4)}\ , \nonumber\\[-1mm] && \label{corr17} \\[-1mm] \frac{q^2-1}{q^4\sqrt{[(q^N-1)^2+(q^2-1)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4] [(q^N+1)^2+(q^2-1)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4]}}&\limar{N}{-\infty}& \frac{q^2-1}{q^4+(q^2-1)^2\sigma} \renewcommand{\S}{\Sigma^2}\ . \nonumber \end{eqnarray} The series has the {\it linear} divergence at the lower limit (the IR-regularization parameter again, as in the nondeformed case, is inessential in the $N\,\rightarrow\,\pm\infty$ limit). Notice that in the nondeformed limit $q\ra1$ the series \r{corr16} turns into the following one \begin{equation} \sum_{N=-\infty}^\infty\frac{1}{\sqrt{N^2+\sigma} \renewcommand{\S}{\Sigma^2}}\ , \label{corr18} \end{equation} and coincides with the nondeformed result ({\em cf.\ } \r{exp6}) (logarithmic divergence). Quite similar calculation in the case $q<1$ shows that the tadpole diagram~\ref{F:tadp}.a is proportional to the series \begin{equation} \sum_{N=-\infty}^\infty\frac{1-q^2}{q^4\sqrt{[(1-q^N)^2+(1-q^2)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4] [(q^N+1)^2+(1-q^2)^2\sigma} \renewcommand{\S}{\Sigma^2/q^4]}}\ , \label{corr19} \end{equation} which has the linear divergence at the upper limit. Thus the perturbation theory for the ${\varphi}^4$-model on the noncommutative plane $P^{(2)}_q$ with the action constructed with the help of the $E_q(2)$--quantum group invariant measure contains the UV-divergences and, hence, can not be considered as a regularization of the usual scalar field theory on the commutative plane. \section{Transformation of states on the noncommutative plane $P^{(2)}_q$ induced by the coaction of the quantum Euclidean group $E_q(2)$} The central problem of this section is to determine transformation properties of a system on $P^{(2)}_q$ under a coaction of the quantum group $E_q(2)$. In the subsection 4.1 we shall clarify a general formulation of this problem and in subsection 4.2 give an explicit answer for $E_q(2)$ group. In particular, we shall show that this coaction leads to nonlocal transformations of states. \subsection{Transformation of states in noncommutative geometry induced by a quantum group coaction} Let a quantum group $G_q$ coact on a noncommutative space ${\cal X}_q$, i.e. there exists the homomorphic map \begin{equation} \delta\,: \ Fun_q({\cal X})\longrightarrow Fun_q(G){\otimes} Fun_q({\cal X})\ , \label{1} \end{equation} (the algebra $Fun_q({\cal X})$ of functions on ${\cal X}_q$ is the configuration space subalgebra of the algebra of all operators of the given quantum system). It is natural to say that the system is invariant with respect to the quantum group transformations if all the properties of the system are independent on the coaction map $\delta$. In other words, the algebra $Fun_q({\cal X})$ can be realized as the subalgebra of multiple tensor product $Fun_q(G){\otimes} Fun_q(G){\otimes}...{\otimes} Fun_q({\cal X})$ and no measurements can distinguish the description based on the algebras with different numbers of the factors $Fun_q(G)$. At first sight, this definition of symmetry transformations may look unusual but, in fact, it is a direct generalization of commutative transformations. Indeed, usual action of a group $G$ of transformations of a manifold ${\cal M}$ on a function $f\in Fun({\cal M})$ is defined by the equality \begin{equation} T_gf(x):=f(g^{-1}x)\ ,\qquad g\in G,\ x\in{\cal M}\ . \label{1a} \end{equation} The right hand side of this definition can be considered as the function defined on $G\times{\cal M}$. In other words, the transformations $T$ defines the map $$ T\,:\ Fun({\cal M})\longrightarrow Fun(G){\otimes} Fun({\cal M})\ . $$ More customary map $\phi\,:G{\otimes}{\cal M}\,\rightarrow\,{\cal M}$ is defined for points of the manifolds, which play the role of the dual set of states for the commutative algebra of observables (functions) on usual manifolds. Returning to the transformations with noncommutative parameters, let us define the map \cite{Demichev} which is dual to the transformations \r{1} of observables (operators), {{\em i.e.,}\ } \begin{equation} {\cal S}\,: {\cal H}_{G_q}{\otimes}{\cal H}_{{\cal X}_q}\longrightarrow{\cal H}_{{\cal X}_q}\ , \label{2} \end{equation} where ${\cal H}_{G_q}$ and ${\cal H}_{{\cal X}_q}$ are the Hilbert spaces of {\it all} representations of the algebras $Fun_q(G)$ and $Fun_q({\cal X})$. The intertwining operator ${\cal S}$ is implicitly defined by the equation \begin{equation} \langle\langle\delta A|\Psi{\otimes}\psi\rangle\rangle =\langle\langle A|{\cal S} (\Psi,\psi )\rangle\rangle \ , \label{2a} \end{equation} where $\Psi$ is arbitrary vector from ${\cal H}_G$, $\psi$ is arbitrary vector from ${\cal H}_{\cal X}$ and ${\cal S} (\Psi,\psi )\in{\cal H}_{\cal X}$ and the duality relation $\langle\langle A|\psi\rangle\rangle\,:{\cal O}{\otimes}{\cal H}_{\cal O}\,\rightarrow\,{\ibb C}$ between an operator $A$ from some algebra ${\cal O}$ and a vector $\psi$ from the Hilbert space ${\cal H}_{\cal O}$ of the representations of this algebra, is defined by the ordinary mean value of $A$ in the state $\psi\,:\ \langle\langle A|\psi\rangle\rangle=\langle\psi|A|\psi\rangle$. In fact, the usual definition \r{1a} of the action of (classical, commutative) transformation groups in the space of functions on some homogeneous manifold ${\cal M}$ also has the general form \r{2a}. Indeed, in this case the duality relation between the algebra $Fun({\cal M})$ and states, i.e. points of ${\cal M}$, is defined as follows $$ \langle\langle f|x\rangle\rangle=f(x)\ ,\qquad f\in Fun({\cal M}),\ x\in{\cal M}\ . $$ The same is true for the group manifold: $$ \langle\langle T|g\rangle\rangle=T_g\ ,\qquad T\in Fun(G),\ g\in G\ . $$ Thus \r{1a} can be represented in the form $$ \langle\langle\delta f|g{\otimes} x\rangle\rangle = \langle\langle T{\otimes} f|g{\otimes} x\rangle\rangle =T_gf(x) =\langle\langle f|{\cal S} (g,x)\rangle\rangle=\langle\langle f|g^{-1}x\rangle\rangle=f(g^{-1}x)\ , $$ where the third equality follows from \r{2a} and in this special case ${\cal S} (g,x)=g^{-1}x$. {}From \r{2} it follows that the matrix elements of the operator ${\cal S}$ in a chosen bases of ${\cal H}_{G_q}{\otimes}{\cal H}_{{\cal X}_q}$ and ${\cal H}_{{\cal X}_q}$ play the role of generalized Clebsch-Gordan coefficients (GCGC). If the multiple index (set of quantum numbers) $\{m\}$ defines basis vectors $\psi_{\{m\}}$ of ${\cal H}_{{\cal X}_q}$, and the set $\{K\}$ defines basis $\Psi_{\{K\}}$ of ${\cal H}_{G_q}$, one can write \begin{equation} \psi'={\cal S} (\Psi_{\{K\}},\psi_{\{m\}})=\sum_{\{l\}} C^{\{K\}}_{\{m\}\{l\}} \psi_{\{l\}} \ , \label{3.5} \end{equation} where $C^{\{K\}}_{\{m\}\{l\}}$ are the set of GCGC. In this formula the vector $\psi_{\{m\}}$ is a transformed state on the quantum plane, and the vector $\Psi_{\{K\}}$ (analog of a point on a grou manifold in the case of ordinary Lie groups) defines ``parameters'' of the trasformation of $\psi_{\{m\}}$. One can apply analogous consideration to the very quantum group $G_q$ which coacts on itself \begin{equation} M'^i_{\ j}=\Delta M^i_{\ j}=M^i_{\ k}{\otimes} M^k_{\ j}\ , \label{3.7} \end{equation} This leads to the corresponding transformation of vectors in ${\cal H}_{G_q}$ \begin{equation} \Psi'={\cal S} (\Psi_{\{K\}},\Psi_{\{N\}})\equiv {\cal S}_{\Psi_{\{K\}}}(\Psi_{\{N\}}) = \sum_{\{L\}} C^{\{K\}}_{\{N\}\{L\}} \Psi_{\{L\}} \ , \label{3.8} \end{equation} Two subsequent coactions of the form \r{3.7} induce composition of the transformations \r{3.8} and general properties of algebra representations provide its associativity (or, equivalently, this follows from the coassociativity of Hopf algebras). This means that the transformations \r{3.8} form the {\it semigroup}. The trivial representation $\Psi_{\{0\}}\in{\cal H}_{G_q}$ correspond to the identity transformation. However, there is no inverse transformation for arbitrary ${\cal S}_{\Psi_{\{K\}}}$. This means that the transformations \r{3.8},\r{3.5} do not form a group. The map ${\cal S}$ satisfies the obvious consistency condition which can be expressed as a requirement of commutativity of the diagram in figure~\ref{cons.rel} (in other words, an equivalence of the different ways through the diagrams along the arrows). \begin{figure} \centering \unitlength=1.00mm \begin{picture}(68.00,38.00) \put(10.00,32.00){\makebox(0,0)[cc]{${\cal H}_{G_q}{\otimes}{\cal H}_{{\cal X}_q}$}} \put(65.00,32.00){\makebox(0,0)[cc]{${\cal H}_{G_q}{\otimes}{\cal H}_{{\cal X}_q}$}} \put(37.00,36.00){\makebox(0,0)[cc]{$\d A$}} \put(10.00,28.00){\thicklines\vector(0,-1){19.00}} \put(10.00,4.00){\makebox(0,0)[cc]{${\cal H}_{{\cal X}_q}$}} \put(65.00,28.00){\thicklines\vector(0,-1){19.00}} \put(65.00,4.00){\makebox(0,0)[cc]{${\cal H}_{{\cal X}_q}$}} \put(4.00,19.00){\makebox(0,0)[cc]{${\cal S}$}} \put(37.00,8.00){\makebox(0,0)[cc]{$A$}} \put(68.00,19.00){\makebox(0,0)[cc]{${\cal S}$}} \put(20.00,32.00){\thicklines\vector(1,0){35.00}} \put(18.00,4.00){\thicklines\vector(1,0){40.00}} \end{picture}\caption{\protect\small The diagrammatic representation of the consistency relation for the Generalized Clebsch-Gordan map ${\cal S}$.} \label{cons.rel} \end{figure} For the generalized Clebsch-Gordan coefficients this consistency relation reads as \begin{equation} \sum_{L,n}C^L_{nr}\< \Psi_L|M^i_{\ j}|\Psi_K\>\< \psi_n|x^j|\psi_m\> = \sum_{l}C^K_{ml}\< \psi_r|x^i|\psi_l\>\ . \label{3.10} \end{equation} (for the convenience and shortness we use Dirac bracket notation and drop curly brackets indicating that $K,m,...$ are multi-indices). The equations \r{3.10} must be completed by the normalization conditions which follow from the normalization of vectors $|K\>$ and $|m\>$. Notice that the equations \r{3.10} are an analog of the recursion equations used for determination of ordinary Clebsch-Gordan coefficients of $SU(2)$ group. However, in the case of quantum groups the problem of explicit solution of these equation proves to be much more difficult \cite{Demichev}. \subsection{Representations of the algebra of functions on $E_q(2)$ and transformations of states on $P_q^{(2)}$} In this subsection we shall be interested in construction of the cotransformations of the coordinate subalgebra on a quantum plane and therefore, for shortness, drop the quantum numbers related to the angular momentum $J_q$. According to the discussion in the preceding subsection, a transformation of a state on a quantum plane depends on a vector from representation space for the algebra of functions on the group $E_q(2)$. Thus we need explicit construction of representations of the algebra $E_q(2)$ ({\em cf.\ } \r{2.1}). These representations have been presented in \cite{VaksmanK}: in slightly different (more "physical") form they read as \begin{eqnarray} && t={\rm{e}}^{\sqrt{\eta} a}\ ,\qquad {\bar{t}}={\rm{e}}^{\sqrt{\eta} a^{\dagger}}\ ,\qquad q^2={\rm{e}}^{\eta} \ ,\nonumber\\[-1mm] && \label{4.4}\\[-1mm] && v={\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}{\rm{e}}^{\sqrt{\eta}(a-a^{\dagger})}\ ,\qquad {\phi}} \newcommand{\F}{{\Phi}\in[0,2\pi]\ ,\nonumber \end{eqnarray} where $a,\ a^{\dagger}$ are usual creation and annihilation operators \begin{equation} [a,a^{\dagger}]=1\ , \label{4.5} \end{equation} and hence can be represented in any well known way ({\em e.g.,\ } coordinate, Bargmann-Fock, coherent state, infinite matrix representations). The factor ${\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}$ in \r{4.4} is an eigenvalue of the central element $I$ of the algebra $E_q(2)$ \begin{equation} I=q^{-1}v{\bar{t}} t^{-1}\ , \label{4.6} \end{equation} so that different values of ${\phi}} \newcommand{\F}{{\Phi}\in[0,2\pi]$ separate different (though identical) irreducible representations of the $E_q(2)$-algebra. Notice that $I$ is unitary operator $I^{\dagger} I=1$, that is why its eigenvalues are parameterized by ${\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}$. For an explicit construction we may use the basis of coherent states \begin{eqnarray} &&a\ket{\z}=\z \ket{\z}\ ,\qquad\z\in{\ibb C}\ , \label{4.8}\\[3mm] &&\<\z'|\z\>={\rm{e}}^{\bar \z'\z}\ ,\nonumber \end{eqnarray} so that \begin{eqnarray} t\ket{\z}&=&{\rm{e}}^{\sqrt{\eta}\z}\ket{\z}\ ,\nonumber\\[-1mm] &&\label{4.9}\\[-1mm] {\bar{t}}\ket{\z}&=&\ket{\z+\sqrt{\eta}}\ ,\nonumber\\[3mm] v\ket{\z}&=&{\rm{e}}^{-\eta/2+\i{\phi}} \newcommand{\F}{{\Phi}}{\rm{e}}^{\sqrt{\eta} \z}\ket{\z-\sqrt{\eta}}\ ,\qquad {\bar{v}}=v^{-1}\ .\label{4.10} \end{eqnarray} The algebra of functions on $P^{(2)}_q$ generated by ${\bar{z}},\ z$ has not a central element and its unique representation has the form \r{4.9} where ${\bar{t}},\ t$ are substituted by ${\bar{z}},\ z$. According to the general discussion in the preceding subsection, the coaction of $E_q(2)$ on $P^{(2)}_q$ \begin{eqnarray} {\bar{z}}\,\rightarrow\,{\bar{z}}'&=&\d{\bar{z}}\equiv {\bar{v}}{\otimes}{\bar{z}}+{\bar{t}}{\otimes}{\leavevmode{\rm 1\mkern -5.4mu I}}\ ,\nonumber\\[-1mm] &&\label{4.11}\\[-1mm] z\,\rightarrow\, z'&=&\d z\equiv v{\otimes} z+ t{\otimes}{\leavevmode{\rm 1\mkern -5.4mu I}}\ ,\nonumber \end{eqnarray} induces transformations of states from the representation space ${\cal H}_{P^{(2)}_q}$ of $P^{(2)}_q$ depending on a state from the representation space ${\cal H}_{E_q(2)}$ of the quantum group $E_q(2)$: \begin{eqnarray} \ket{\xi}\,\rightarrow\,\ket{\xi'}&=&{\cal S}\bigg(\ket{{\phi}} \newcommand{\F}{{\Phi},\z},\ket{\xi}\bigg)\ ,\label{4.12}\\[3mm] &&\ket{{\phi}} \newcommand{\F}{{\Phi},\z}\in{\cal H}_{E_q(2)}\ ,\qquad \ket{\xi},\ket{\xi'}\in{\cal H}_{P^{(2)}_q}\ . \nonumber \end{eqnarray} Explicitly this map can be written as follows \begin{equation} \ket{\xi'}=\int\,d\z\,{\rm{e}}^{-\bar\xi\xi}\ket{\xi}\<\xi|{\phi}} \newcommand{\F}{{\Phi},\z;\xi'\>\ ,\label{4.14} \end{equation} where the generalized Clebsch-Gordan coefficients for the coaction of $E_q(2)$ on $P^{(2)}_q$ are denoted as \begin{equation} \<\xi|{\phi}} \newcommand{\F}{{\Phi},\z;\xi'\>\equiv\<\xi|{\cal S}|{\phi}} \newcommand{\F}{{\Phi},\z;\xi'\>\ . \label{4.14a} \end{equation} Thus to define transformations properties of the states on $P^{(2)}_q$ we should calculate the GCGC \r{4.14a}. The consistency relation given by the commutativity of the diagram in figure~\ref{cons.e2} leads to the set of equations \begin{eqnarray} &&\<\xi|{\phi}} \newcommand{\F}{{\Phi},\z;\xi'\>={\rm{e}}^{-\sqrt{\eta}\z}\<\xi+\sqrt{\eta}|{\phi}} \newcommand{\F}{{\Phi},\z;\xi'\> -{\rm{e}}^{-\eta/2+\i{\phi}} \newcommand{\F}{{\Phi}}{\rm{e}}^{\sqrt{\eta}\xi'}\<\xi|{\phi}} \newcommand{\F}{{\Phi},\z-\sqrt{\eta};\xi'\>\nonumber\\[3mm] &&\<\xi|{\phi}} \newcommand{\F}{{\Phi},\z;\xi'\>={\rm{e}}^{-\sqrt{\eta}{\bar{\xi}}}\<\xi|{\phi}} \newcommand{\F}{{\Phi},\z+\sqrt{\eta};\xi'\> -{\rm{e}}^{-\eta/2-\i{\phi}} \newcommand{\F}{{\Phi}}{\rm{e}}^{-\sqrt{\eta}(\z+{\bar{\xi}}}\<\xi|{\phi}} \newcommand{\F}{{\Phi},\z+\sqrt{\eta};\xi'+\sqrt{\eta}\> \ ,\label{4.15}\\[3mm] &&\<\xi+\sqrt{\eta}|{\phi}} \newcommand{\F}{{\Phi},\z;\xi'\>={\rm{e}}^{\eta/2} {\rm{e}}^{\sqrt{\eta}(\z-{\bar{\xi}}+\xi')}\<\xi|{\phi}} \newcommand{\F}{{\Phi},\z-\sqrt{\eta};\xi'-\sqrt{\eta}\>\ ,\nonumber \end{eqnarray} which must be accompanied by normalization conditions. As we mentioned in the preceding subsection, it is not easy to solve this equations straigthforwardly. To circumevent the problem it is helpful to consider the basis where the primitive elements ${\bar{v}},v\in E_q(2)$ are diagonal. The point is that a cotransformation for primitive elements has the form of a cotransformation for an ordinary Lie algebra, thus the consistency conditions for them also have a most simple form. \begin{figure} \centering \unitlength=1.00mm \begin{picture}(68.00,38.00) \put(10.00,32.00){\makebox(0,0)[cc]{${\cal H}_{E_q(2)}{\otimes}{\cal H}_{E_q(2)}$}} \put(65.00,32.00){\makebox(0,0)[cc]{${\cal H}_{E_q(2)}{\otimes}{\cal H}_{E_q(2)}$}} \put(37.00,36.00){\makebox(0,0)[cc]{$\d v,\d{\bar{t}},\d t$}} \put(10.00,28.00){\thicklines\vector(0,-1){19.00}} \put(10.00,4.00){\makebox(0,0)[cc]{${\cal H}_{E_q(2)}$}} \put(65.00,28.00){\thicklines\vector(0,-1){19.00}} \put(65.00,4.00){\makebox(0,0)[cc]{${\cal H}_{E_q(2)}$}} \put(4.00,19.00){\makebox(0,0)[cc]{${\cal S}$}} \put(37.00,8.00){\makebox(0,0)[cc]{$v,{\bar{t}},t$}} \put(68.00,19.00){\makebox(0,0)[cc]{${\cal S}$}} \put(23.00,32.00){\thicklines\vector(1,0){29.00}} \put(18.00,4.00){\thicklines\vector(1,0){40.00}} \end{picture}\caption{\protect\small The diagrammatic representation of the consistency relation for the Generalized Clebsch-Gordan map ${\cal S}$ in case of coaction of the group $E_q(2)$ on itself.} \label{cons.e2} \end{figure} Proceeding in this way, let us construct representation in the basis of the primitive element $v$. This is easy to do taking into account that in the parameterization \r{2.4} the commutation relations for the $E_q(2)$ takes the form \begin{eqnarray} &&[\theta} \newcommand{\T}{\Theta,\rho^2]=-\i2\eta\rho^2\ , \label{4.17}\\[3mm] && [\theta} \newcommand{\T}{\Theta,\nu^2]= 0 , \label{4.18} \end{eqnarray} where we have introduced, in analogy with the algebra $P^{(2)}_q$ ({\em cf.\ } \r{2.37a}) the operators: \begin{equation} \rho^2\equiv {\bar{t}} t\ ,\qquad \nu^2\equiv {\bar{t}} t^{-1}\ . \label{4.18a} \end{equation} The only nontrivial commutation relations \r{4.17} is equivalent to that for $igl(1,{\Ibb R})$ Lie algebra (Lie algebra of translations and dilatations on a line). Representations of this algebra are well known (see, {\em e.g.,\ } \cite{Vilenkin}) and this allows to write immediately the required representation with $v={\rm{e}}^{\i\theta} \newcommand{\T}{\Theta}$ being diagonal: \begin{eqnarray} vf_{\phi}} \newcommand{\F}{{\Phi}(x)&=&{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}{\rm{e}}^{-2\i\eta x}f_{\phi}} \newcommand{\F}{{\Phi}(x)\ ,\nonumber\\[3mm] \rho^2f_{\phi}} \newcommand{\F}{{\Phi}(x)&=&{\rm{e}}^{-\eta}{\rm{e}}^{-\i\partial}f_{\phi}} \newcommand{\F}{{\Phi}(x)\ , \label{4.24}\\[3mm] \nu^2f_{\phi}} \newcommand{\F}{{\Phi}(x)&=&{\rm{e}}^{\eta/2}{\rm{e}}^{2\i\eta x}f_{\phi}} \newcommand{\F}{{\Phi}(x)\ . \end{eqnarray} From the form of the operators it is clear that the variable $x$ takes values on a circle: $x\in [0,2\pi/\eta]$. Thus the functions $f_{\phi}} \newcommand{\F}{{\Phi}(x)$ are defined on the circle and form the Hilbert space with the scalar product \begin{equation} \<f_{1{\phi}} \newcommand{\F}{{\Phi}}(x),f_{2{\phi}} \newcommand{\F}{{\Phi}}(x)\>=\int_0^{2\pi/\eta}\,dx\, \bar f_{1{\phi}} \newcommand{\F}{{\Phi}}(x)f_{2{\phi}} \newcommand{\F}{{\Phi}}(x)\ . \label{4.21} \end{equation} The consistency relations (commutativity of the diagram in figure~\ref{cons.e2}) for the generators $t,\ \bar t$ result again in still rather complicated recursion relations: \begin{eqnarray*} &&\Delta t=v{\otimes} t+t{\otimes}{\leavevmode{\rm 1\mkern -5.4mu I}}\qquad\Rightarrow\nonumber\\[3mm] &&\left[{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{-2\i\eta(x_1+x_2)}{\rm{e}}^{\i\partial_2/2}+ {\rm{e}}^{-\i\eta x_1}{\rm{e}}^{\i\partial_1/2}-{\rm{e}}^{-\i\eta x}{\rm{e}}^{-\i\partial/2}\right] \<{\phi}} \newcommand{\F}{{\Phi},x|{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2\>=0\ ,\nonumber\\[3mm] &&\Delta {\bar{t}}={\bar{v}}{\otimes}{\bar{t}}+{\bar{t}}{\otimes}{\leavevmode{\rm 1\mkern -5.4mu I}}\qquad\Rightarrow\nonumber\\[3mm] &&\left[{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{2\i\eta(x_1+x_2)}{\rm{e}}^{\i\partial_2/2}+ {\rm{e}}^{\i\eta x_1}{\rm{e}}^{\i\partial_1/2}-{\rm{e}}^{\i\eta x}{\rm{e}}^{-\i\partial/2}\right] \<{\phi}} \newcommand{\F}{{\Phi},x|{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2\>=0\ . \end{eqnarray*} But the relation for the primitive element ${\bar{v}}$ has quite simple form \begin{eqnarray} &&\Delta {\bar{v}}={\bar{v}}{\otimes} {\bar{v}}\qquad\Rightarrow\nonumber\\[3mm] &&{\rm{e}}^{-\i({\phi}} \newcommand{\F}{{\Phi}_1-2\eta x_1+{\phi}} \newcommand{\F}{{\Phi}_2-2\eta x_2)}\<{\phi}} \newcommand{\F}{{\Phi},x|{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2\> ={\rm{e}}^{-\i({\phi}} \newcommand{\F}{{\Phi}-2\eta x)}\<{\phi}} \newcommand{\F}{{\Phi},x|{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2\>\ . \label{4.25} \end{eqnarray} Here $\<{\phi}} \newcommand{\F}{{\Phi},x|{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2\>$ denotes GCGC for the algebra $E_q(2)$ in the realization \r{4.24} and we used the basis $|{\phi}} \newcommand{\F}{{\Phi},x\>$ of eigenfunctions of the operator ${\rm{e}}^{2\i\eta x}$ which are, of course, $\d$-functions on the circle: \begin{equation} |{\phi}} \newcommand{\F}{{\Phi},x_0\>=\d^{(S)}(x-x_0)=\frac{1}{2\pi\eta} \sum_{k=-\infty}^\infty {\rm{e}}^{2\i\eta k(x-x_0)} \ . \label{4.26} \end{equation} The relation \r{4.25} shows that the eigenvalues $x_1,x_2$ of vectors in the representations ${\phi}} \newcommand{\F}{{\Phi}_1$ and ${\phi}} \newcommand{\F}{{\Phi}_2$ under the tensor product sign and the eigenvalue $x$ in an irreducible part ${\phi}} \newcommand{\F}{{\Phi}$ of the resulting representation are connected by the relation \begin{equation} -{\phi}} \newcommand{\F}{{\Phi}_1+2\eta x_1-{\phi}} \newcommand{\F}{{\Phi}_2+2\eta x_2 =-{\phi}} \newcommand{\F}{{\Phi}+2\eta x+2\pi n\ ,\qquad n=0,1,2,...\label{4.27} \end{equation} This is the analog of additivity of the magnetic quantum number in the case of $su(2)$ Lie algebra: $m=m_1+m_2$ ($m_1,m_2$ are $J^{(1)}_3,J^{(2)}_3$-eigenvalues of two spins to be summed up and $m$ is an eigenvalue of $J_3=J^{(1)}_3+J^{(2)}_3$). Let us consider the action of the central operator $I$ on the direct product of two representations \begin{equation} \Delta I\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\ . \label{c1} \end{equation} Acting in addition by the intertwining operator ${\cal S}$ and denoting $$ \ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\equiv\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\ , $$ we have \begin{eqnarray} {\cal S}\Delta I\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2}&=&I{\cal S}\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\nonumber\\[3mm] &=&I\int\,d{\phi}} \newcommand{\F}{{\Phi}\,dy\,\ket{{\phi}} \newcommand{\F}{{\Phi},y}\bra{{\phi}} \newcommand{\F}{{\Phi},y}{\cal S}\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\nonumber\\[3mm] &=&I\int\,d{\phi}} \newcommand{\F}{{\Phi}\,\ket{{\phi}} \newcommand{\F}{{\Phi},y}C({\phi}} \newcommand{\F}{{\Phi};{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2)\nonumber\\[3mm] &=&\int\,d{\phi}} \newcommand{\F}{{\Phi}\,{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}\ket{{\phi}} \newcommand{\F}{{\Phi},y}C({\phi}} \newcommand{\F}{{\Phi};{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2)\ , \label{c2} \end{eqnarray} where according to \r{4.27} we have defined \begin{equation} \bra{{\phi}} \newcommand{\F}{{\Phi},y}{\cal S}\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\equiv \d^{(S)}\Big({\phi}} \newcommand{\F}{{\Phi}-{\phi}} \newcommand{\F}{{\Phi}_1-{\phi}} \newcommand{\F}{{\Phi}_2+2\eta(x_1+x_2-y)\Big) C({\phi}} \newcommand{\F}{{\Phi};{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2)\ . \label{c3} \end{equation} On the other hand, consider the concrete realization of the operators $I$ and $\Delta I$ in $\cL^2(S)$ and in the tensor product $\cL^2(S){\otimes}\cL^2(S)$, respectively. To this aim we need an explicit form of $\Delta\nu^2$ and $\Delta\rho^2$. Calculation in the representation \r{4.24} gives the result: \begin{eqnarray} \Delta \rho^2&=&{\leavevmode{\rm 1\mkern -5.4mu I}}{\otimes}\rho^2+\rho^2{\otimes}{\leavevmode{\rm 1\mkern -5.4mu I}}+{\bar{v}} t{\otimes}{\bar{t}}+{\bar{t}} v{\otimes} t\nonumber\\[3mm] &=&{\rm{e}}^{-\eta}{\rm{e}}^{-\i\partial}\bigg[{\rm{e}}^{-\i\widetilde\partial} +{\rm{e}}^{\i\widetilde\partial}+{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{-\eta/2}{\rm{e}}^{\i\eta x} +{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{\eta/2}{\rm{e}}^{-\i\eta x}\bigg]\ , \label{4.33}\\[3mm] \Delta \nu^2&=&\bigg({\bar{v}} t^{-1}{\otimes}{\bar{t}}+\nu^2{\otimes}{\leavevmode{\rm 1\mkern -5.4mu I}}\bigg) \bigg({\leavevmode{\rm 1\mkern -5.4mu I}}{\otimes}{\leavevmode{\rm 1\mkern -5.4mu I}}+vt^{-1}{\otimes} t\bigg)^{-1}\nonumber\\[3mm] &=&{\rm{e}}^{\eta/2}{\rm{e}}^{\i\eta (x+\widetilde x)} \frac{1+{\rm{e}}^{-\eta/2}{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{\i\eta x}{\rm{e}}^{\i\widetilde\partial}} {1+{\rm{e}}^{-\eta/2}{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{-\i\eta x}{\rm{e}}^{\i\widetilde\partial}}\ . \label{4.34} \end{eqnarray} Here we used the change of variables \begin{equation} x=x_1+x_2\ ,\qquad \widetilde x=x_1-x_2\ , \label{4.31} \end{equation} inspired by the equality \r{4.27}. Now we can easily calculate the comultiplication for the central operator: \begin{eqnarray} \Delta I&=&q^{-1}\Delta v\Delta\nu^2\nonumber\\[3mm] &=&{\rm{e}}^{\i({\phi}} \newcommand{\F}{{\Phi}_1+{\phi}} \newcommand{\F}{{\Phi}_2)}{\rm{e}}^{-\i\eta x}{\rm{e}}^{\i\eta \widetilde x} \frac{1+{\rm{e}}^{-\eta/2}{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{\i\eta x}{\rm{e}}^{\i\widetilde\partial}} {1+{\rm{e}}^{-\eta/2}{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{-\i\eta x}{\rm{e}}^{\i\widetilde\partial}}\ . \label{4.35} \end{eqnarray} In spite of their rather cumbersome form it is easy to check that the operators $\Delta \rho^2$ and $\Delta \nu^2$ indeed satisfy the commutation relations of the $E_q(2)$ algebra. In fact, this immediately follows from their general form which can be written as follows \begin{eqnarray*} \Delta \rho^2&=&{\rm{\,const\,}}\cdot{\rm{e}}^{-\i\partial}F(x,\widetilde\partial)\ ,\nonumber\\[3mm] \Delta \nu^2&=&{\rm{\,const\,}}\cdot \Delta I {\rm{e}}^{2\i\eta x}\ \end{eqnarray*} (here $F(x,\widetilde\partial)$ is a function of only $x$ and $\widetilde\partial$ explicit form of which is given in \r{4.33}). It is clear that if we start from some eigenvector of the operator $\nu^2$ \begin{equation} \nu^2\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}={\rm{e}}^{\eta/2}{\rm{e}}^{2\i\eta x_2}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\ \label{4.36} \end{equation} ({\em cf.\ } \r{4.24}) then, after the comultiplication, we have \begin{equation} \Delta\nu^2\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}={\rm{e}}^{\eta/2}{\rm{e}}^{-\i({\phi}} \newcommand{\F}{{\Phi}_1+{\phi}} \newcommand{\F}{{\Phi}_2)} {\rm{e}}^{2\i\eta(x_1+x_2)}\Delta I\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\ .\label{4.37} \end{equation} Since in any irreducible component the central element is proportional to the unity operator \begin{equation} \Delta IP_{\phi}} \newcommand{\F}{{\Phi}\bigg(\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\bigg)= {\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}P_{\phi}} \newcommand{\F}{{\Phi}\bigg(\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\bigg) \label{4.37a} \end{equation} ($P_{\phi}} \newcommand{\F}{{\Phi}$ is a projector onto the irreducible component of the representation space corresponding to the central element eigenvalue ${\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}$), the relation \r{4.37} shows that for the ${\phi}} \newcommand{\F}{{\Phi}$-component the eigenvalue of $\Delta\nu^2$ is \begin{equation} {\rm{e}}^{\eta/2}{\rm{e}}^{\i({\phi}} \newcommand{\F}{{\Phi}-{\phi}} \newcommand{\F}{{\Phi}_1-{\phi}} \newcommand{\F}{{\Phi}_2+2\eta x)}\ ,\qquad x=x_1+x_2\ . \label{4.37b} \end{equation} This expression shows how the initial eigenvalue ${\rm{e}}^{\eta/2}{\rm{e}}^{2\i\eta x_2}$ of the operator $\nu^2$ is transformed under the coaction of $E_q^{(2)}$ with the state $\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}$ defining the ``parameters'' of the transformations ({\em cf.\ } \r{3.5}, \r{3.8}). The thing which we have to do now is to find the decomposition of an arbitrary function $f(x_1,x_2)=f(x,\widetilde x)\in{\cal H}_{E_q(2)}{\otimes}{\cal H}_{E_q(2)}$ into eigenvectors of $\Delta I$, {{\em i.e.,}\ } into irreducible components. To this aim we must solve the eigenvalue equation \begin{equation} \Delta Ig_{\phi}} \newcommand{\F}{{\Phi}(x,\widetilde x)= {\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}}g_{\phi}} \newcommand{\F}{{\Phi}(x,\widetilde x)\ . \label{4.40} \end{equation} The solution written in terms of the Fourier\ transform \begin{equation} g_{\phi}} \newcommand{\F}{{\Phi}(x,\widetilde x)=\sum_{k=-\infty}^\infty \widetilde g_{\phi}} \newcommand{\F}{{\Phi}(x,k){\rm{e}}^{\i k\widetilde x}\ ,\label{3.40a} \end{equation} has the form \begin{equation} \widetilde g_{\phi}} \newcommand{\F}{{\Phi}(x,k)={\rm{e}}^{\i({\phi}} \newcommand{\F}{{\Phi}-{\phi}} \newcommand{\F}{{\Phi}_1-{\phi}} \newcommand{\F}{{\Phi}_2+\eta x)k}{\rm{e}}^{\i d(x,k)}\ , \label{4.46} \end{equation} $d(x,k)$ being defined by the simple recursion relation \begin{equation} d(x,k+1)=d(x,k)-\lambda} \newcommand{\LD}{\Lambda(x,k+1)\ ,\qquad d(x,0)=0\ , \label{4.45} \end{equation} where \begin{equation} {\rm{e}}^{\i\lambda} \newcommand{\LD}{\Lambda(x,k)}\equiv \frac{1+{\rm{e}}^{-\eta/2}{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{\i\eta x}{\rm{e}}^{\eta k}} {1+{\rm{e}}^{-\eta/2}{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{-\i\eta x}{\rm{e}}^{\eta k}} \ . \label{4.42} \end{equation} The solution of this recursion is obvious: \begin{eqnarray} &&d(x,k)=-\sum_{n=1}^k\lambda} \newcommand{\LD}{\Lambda(x,n)\ ,\qquad k>0\ ,\nonumber\\[3mm] &&d(x,k)=\sum_{n=0}^{k+1}\lambda} \newcommand{\LD}{\Lambda(x,n)\ ,\qquad k<0\ ,\nonumber\\[3mm] &&{\rm{e}}^{\i d(x,k)}=\prod_{n=1}^k \frac{1+{\rm{e}}^{-\eta/2}{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{-\i\eta x}{\rm{e}}^{\eta n}} {1+{\rm{e}}^{-\eta/2}{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{\i\eta x}{\rm{e}}^{\eta n}} \ , \qquad k>0\ , \label{cc1}\\[3mm] &&{\rm{e}}^{\i d(x,k)}=\prod_{n=1}^k \frac{1+{\rm{e}}^{-\eta/2}{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{\i\eta x}{\rm{e}}^{\eta n}} {1+{\rm{e}}^{-\eta/2}{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}_1}{\rm{e}}^{-\i\eta x}{\rm{e}}^{\eta n}} \ , \qquad k<0\ , \label{cc2} \end{eqnarray} The solution \r{4.46}, of course, is not square integrable function but a distribution (like momentum eigenvectors in ordinary Quantum Mechanics). Any function $f(x,k)$ can be presented as a Fourier\ integral of the functions $g_{\phi}} \newcommand{\F}{{\Phi}(x,k)$ \begin{eqnarray} f(x,k)&=&\frac{1}{2\pi}\int\,d{\phi}} \newcommand{\F}{{\Phi}\,c({\phi}} \newcommand{\F}{{\Phi})g_{\phi}} \newcommand{\F}{{\Phi}(x,k)\nonumber\\[3mm] &=&\frac{1}{2\pi}{\rm{e}}^{-\i({\phi}} \newcommand{\F}{{\Phi}_1+{\phi}} \newcommand{\F}{{\Phi}_1-\eta x-c(x,k))} \int\,d{\phi}} \newcommand{\F}{{\Phi}\,c({\phi}} \newcommand{\F}{{\Phi})g_{\phi}} \newcommand{\F}{{\Phi}(x,k){\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi} k}\ , \label{4.49}\\[3mm] c({\phi}} \newcommand{\F}{{\Phi})&=&\sum_k{\rm{e}}^{\i({\phi}} \newcommand{\F}{{\Phi}_1+{\phi}} \newcommand{\F}{{\Phi}_1-\eta x-c(x,k))}{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi} k}f(x,k)\ . \label{4.50} \end{eqnarray} If we start from eigenvectors of the operator $\nu^2$ (in each component of the tensor product $\cL^2(S){\otimes}\cL^2(S)$) \begin{eqnarray*} f(x,\widetilde{x})&\equiv&\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}= \d^{(S)}(x_1-X_1)\d^{(S)}(x_2-X_2)\nonumber\\[3mm] &=&\frac{1}{(2\pi\eta)^2} \sum_{m,n=-\infty}^\infty{\rm{e}}^{\i\eta[m(x-X)+n(\widetilde{x}-\widetilde{X})]}\ , \qquad X=X_1+X_2\,,\ \ \widetilde{X}=X_1-X_2\ \end{eqnarray*} ($X_1,X_2$ are the representation variables, while $x_1,x_2$ labels eigenvalues of the operator $\nu^2$: $\nu^2\d^{(S)}(x_i-X_i)= \exx{\eta/2+2\i\eta x_i} \d^{(S)}(x_i-X_i)$), its decomposition over the eigenfunctions $g_{\phi}} \newcommand{\F}{{\Phi}$ has the form \r{4.49} with the coefficients: \begin{equation} c({\phi}} \newcommand{\F}{{\Phi},x,{\phi}} \newcommand{\F}{{\Phi}_1,{\phi}} \newcommand{\F}{{\Phi}_2)=\d^{(S)}(x-X)\sum_{k=-\infty}^\infty {\rm{e}}^{\i({\phi}} \newcommand{\F}{{\Phi}_1+{\phi}} \newcommand{\F}{{\Phi}_2-{\phi}} \newcommand{\F}{{\Phi}-\eta x)k+\i d(x,k)}{\rm{e}}^{-2\i\eta\widetilde{X}k}\ . \label{c4} \end{equation} Thus \begin{equation} \Delta I \ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}=\frac{1}{2\pi}\int\,d{\phi}} \newcommand{\F}{{\Phi}\,{\rm{e}}^{\i{\phi}} \newcommand{\F}{{\Phi}} c({\phi}} \newcommand{\F}{{\Phi},x,{\phi}} \newcommand{\F}{{\Phi}_1,{\phi}} \newcommand{\F}{{\Phi}_2)g_{\phi}} \newcommand{\F}{{\Phi}(x,\widetilde{x})\ . \label{c5} \end{equation} Comparison of \r{c5} and \r{c2} allows to read off the expression for GCGC of the quantum group $E_q(2)$: \begin{eqnarray} \bra{{\phi}} \newcommand{\F}{{\Phi},y}{\cal S}\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2}&\equiv&\d^{(S)} \Big({\phi}} \newcommand{\F}{{\Phi}-{\phi}} \newcommand{\F}{{\Phi}_1-{\phi}} \newcommand{\F}{{\Phi}_2+2\eta(x_1+x_2-y)\Big)C({\phi}} \newcommand{\F}{{\Phi};{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2)\ ,\nonumber\\[3mm] C({\phi}} \newcommand{\F}{{\Phi};{\phi}} \newcommand{\F}{{\Phi}_1,x_1;{\phi}} \newcommand{\F}{{\Phi}_2,x_2)&=&c({\phi}} \newcommand{\F}{{\Phi},x,{\phi}} \newcommand{\F}{{\Phi}_1,{\phi}} \newcommand{\F}{{\Phi}_2)\ , \label{c6} \end{eqnarray} and $c({\phi}} \newcommand{\F}{{\Phi},x,{\phi}} \newcommand{\F}{{\Phi}_1,{\phi}} \newcommand{\F}{{\Phi}_2)$ is given by \r{c4} and \r{cc1},\r{cc2}. The corresponding coaction on vectors on $P^{(2)}_q$ have quite the same form with the only restriction ${\phi}} \newcommand{\F}{{\Phi}_2=0$. Now we are ready to answer the question about transformations of the coordinate operators eigenvalues. The operator ${u^2}={\bar{z}} z^{-1}\in P_q^{(2)}$ ({\em cf.\ } \r{2.37a}) has the eigenvectors similar to those of $\nu^2$: \begin{equation} {u^2}\ket{x_2}={\rm{e}}^{\eta/2}{\rm{e}}^{2\i\eta x_2}\ket{x_2}\ , \label{4.50a} \end{equation} After the coaction $\d$ by $E_q(2)$-quantum group, we obtain the operator $\d{u^2}$ acting on vectors $\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{x_2}\in{\cal H}_{E_q(2)} {\otimes}{\cal H}_{P^{(2)}_q}$ and with a structure which is quite similar to $\Delta\nu^2$ ({\em cf.\ } \r{4.37}) \begin{equation} \big(\d{u^2}\big)\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{x_2}={\rm{e}}^{\eta/2}{\rm{e}}^{-\i{\phi}} \newcommand{\F}{{\Phi}_1} {\rm{e}}^{2\i\eta(x_1+x_2)}\Delta I\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{{\phi}} \newcommand{\F}{{\Phi}_2,x_2}\ .\label{4.51} \end{equation} This implies that the resulting vector can be decomposed into irreducible parts and, simultaneously, into vectors with definite values of the coordinate ${u^2}$ as follows: \begin{equation} \big(\d{u^2}\big)\,\ket{{\phi}} \newcommand{\F}{{\Phi}_1,x_1}\ket{x_2}=\frac{1}{2\pi}{\rm{e}}^{\eta/2} \int\,d{\phi}} \newcommand{\F}{{\Phi}\,c({\phi}} \newcommand{\F}{{\Phi},x,{\phi}} \newcommand{\F}{{\Phi}_1,{\phi}} \newcommand{\F}{{\Phi}_2) {\rm{e}}^{\i({\phi}} \newcommand{\F}{{\Phi}-{\phi}} \newcommand{\F}{{\Phi}_1+2\eta x)}g_{\phi}} \newcommand{\F}{{\Phi}(x,\widetilde x)\ . \label{4.55} \end{equation} The expression \r{4.55} presents the form of transformation of position eigenvectors on a quantum plane $P^{(2)}_q$ and shows that the coaction of $E_q(2)$ induces nonlocal transformations of the states. \section{Conclusion} We have shown that transition to a noncommutative $q$-deformed plane does not lead to an ultraviolet regularization of the scalar ${\varphi}^4$-quantum field theory. We start from the firstly quantized theory of quantum particles on the noncommutative plane. Then we have defined quantum fields depending on noncommutative coordinates and the field theoretical action using the quantum analog of the Haar ($E_q(2)$-invariant) measure on the noncommutative plane. With the help of the partial wave decomposition we have shown that this quantum field theory can be considered as a second quantization of the particle theory on the noncommutative plane and that it has (contrary to the common belief) even more severe ultraviolet divergences than its counterpart on the usual commutative plane. We have discussed symmetry transformations on the noncommutative spaces and the induced transformations of the states. In the case of Lie algebra-like spaces the coordinates form a tensor operator ${\widehat x}_i\,\rightarrow\, {\widehat x}'_i=M_{ij}{\widehat x}_j+b_i={\widehat U}_g{\widehat x}_i{\widehat U}_g^{-1}$ and states of the field system are transformed by the operator ${\widehat U}_g$. We considered the example of such transformations for the case of noncommutative Euclidean and Minkowski planes in the preceding paper \cite{CDP98}. In the $q$-deformed case, we have shown that the quantum group coaction on a coordinate algebra induces nonlocal transformations of states in the coordinate space. These transformations are defined by the generalized Clebsch-Gordan coefficients, describing decomposition of tensor products of representations of algebras of functions on quantum spaces and representations of the corresponding quantum group. In other words, the coaction puts in correspondence to a pair of states on a group algebra $G_q$ and on a quantum space ${\cal X}_q$ some new state on the quantum space. We have considered such transformations for the case of the $q$-deformed plane with the $E_q(2)$-symmetry. \vspace{5mm} {\bf Acknowledgements} The financial support of the Academy of Finland under the Projects No. 44129 is greatly acknowledged. A.D.'s work was partially supported also by RFBR-98-02-16769 grant and P.P.'s work by VEGA project 1/4305/97.
\section*{\center SEARCH FOR THE NEUTRINO MAGNETIC MOMENT IN THE NON-EQUILIBRIUM REACTOR ANTINEUTRINO ENERGY SPECTRUM} \Large {\hspace{1.5 cm} V.I. Kopeikin, L.A. Mikaelyan, V.V. Sinev} \\ \normalsize {\it \center RRC "Kurchatov Institute", Kurchatov Sq., 1, Moscow-123182, Russia.\\ Resume of seminar talks given at Kurchatov Institute, March 1999 y.} \hspace{3 cm} \vspace{1em}\\ \righthyphenmin=3 We study the time evolution of the typical nuclear reactor antineutrino energy spectrum during reactor ON period and the decay of the residual antineutrino spectrum after reactor is stopped. We find that relevant varia-tions of the soft recoil electron spectra produced via weak and magnetic ${\widetilde {\nu}}_{e},e$ scattering process can play a significant role in the current and planned searches for the neutrino magnetic moment at reactors. \section{Introduction} \large Efforts are currently being done to observe the neutrino magnetic moment below the limit ${\mu}_{\nu}<2\cdot10^{-10}{\mu}_{B}$ that was found in previous ${\widetilde {\nu}}_{e},e$ scattering experiments at SAVANNAH RIVER, KRASNOYARSK and ROVNO reactors [1]. The KURCHATOV-PNPI collaboration is plan-ning for KRASNOYARSK new studies of low kinetic energy recoil electrons in ${\widetilde {\nu}}_{e},e$ experiment with a Si semiconductor multi detector. The MUNU collaboration experiment at BUGEY with a gas TPC chamber is in the final state of preparation [2]. The dominant contribution to the soft recoil electron produced in the ${\widetilde {\nu}}_{e},e$ scattering comes from the low energy part of the reactor ${\widetilde {\nu}}_{e}$ energy spectrum. This part of the spectrum is strongly time dependent: it never comes to saturation during the reactor operating run and does not vanish after the reactor is shut down, the time when the background is usually measured. Here we consider the time evolution of the typical reactor ${\widetilde {\nu}}_{e}$ energy spectrum and discuss relevant variations of the ${\widetilde {\nu}}_{e},e$ scattering recoil electron spectra. \section{TIME VARIATION OF THE REACTOR \\ANTINEUTRINO SPECTRUM} 1. Three components contribute the reactor ${\widetilde {\nu}}_{e}$ energy spectrum \\ ${\rho}(E)/fiss\cdot MeV$: \begin{equation} {\rho}(E)=^{F}{\rho}(E) + ^{U}{\rho}(E) + {\Delta}{\rho}(E). \end{equation} Here, the term $^{F}{\rho}(E)$ represents the radiation of the $^{235}U$, $^{239}Pu$, $^{238}U$ and $^{241}Pu$ fission fragments. The second term stems from the chain of the ${\beta}$-decays which follow neutron radiative capture in $^{238}U$: \begin{equation} ^{238}U(n,{\gamma})^{239}U \frac{\beta}{23.5 min}\to ^{239}Np \frac{\beta}{2.36 days}\to ^{239}Pu \end{equation} The last term in Eq.(1) accounts for the antineutrinos (and neutrinos) induced by the neutron interactions with other materials in the reactor core. As discussed in Ref.[3] this term adds no more than 1\% to the total reactor ${\widetilde {\nu}}_{e}$ flux and is disregarded here. Till recently the term $^{F}{\rho}$ has traditionally been identified with the reactor ${\widetilde {\nu}}_{e}$ spectrum. The contribution of the chain (2) antineutrinos is however quite sizable for all reactors where neutrino experiments are running or planned. In the ROVNO, BUGEY and CHOOZ PWR-type reactors about 1.2 ${\widetilde {\nu}}_{e}$ per fission come from this source. \vspace{0.2 cm} 2. For each of the four isotopes $^{235}U$, $^{239}Pu$, $^{238}U$ and $^{241}Pu$ the evolution of the neutrino spectra ${\rho}(E,t)$ have been calculated vs time t since the beginning of the fission process. The subsequent decay of the spectra during reactor OFF period have been followed. The base used for these calculations involves data on 571 fission fragments, data on nuclear isomers and delayed neutron emission. \vspace{0.2 cm} 3. Calculations show that in PWR reactors about 2/3 of all antineu-trinos belong typically to the energy range below E = 1,5 MeV. This part of the ${\widetilde {\nu}}_{e},e$ spectrum ${\rho}(E,330)$ and it's component due to fission fragments ${\rho}(E,330)$ at the end of the reactor 330 day ON period are presented in Fig.1. The evolution of the ${\widetilde {\nu}}_{e},e$ spectrum during PWR reactor ON period and its decay after the reactor is shut down is illustrated in Fig.2a,b. \section{RECOIL ELECTRON ENERGY SPECTRA} 1. The recoil-electron spectra $S^{W}$(T) and $S^{M}$(T) in $cm^{2}/MeV\cdot fiss.$ units, (T is the recoil-electron kinetic energy) for weak (W) and magnetic (M) scattering of reactor antineutrinos are found by convolution the ${\widetilde {\nu}}_{e},e$ spectra ${\rho}(E)$ with the differential cross sections for monoenergy antineu-trino: \begin{equation} \frac{d{\sigma}^{W}}{dT}=g^{2}_{F}\frac{m}{2{\pi}}\cdot [4x^{4}+(1+2x^{2})(1-\frac{T}{E})^{2}-2x^{2}(1+2x^{2})\frac{mT}{E^{2}}] \end{equation} \begin{equation} \frac{d{\sigma}^{M}}{dT}={\pi}r^{2}_{0}\frac{{\mu}^{2}_{\nu}}{{\mu}^{2}_{B}}(\frac{1}{T}-\frac{1}{E}), \end{equation} where m is the electron mass, $g^{2}_{F}\frac{m}{2{\pi}}=4.31\cdot 10^{-45} cm^{2}/MeV$, $x=sin^{2}{\theta}_{W}=0.232$ is the Weinberg parameter, ${\pi}r^{2}_{0}=2.495\cdot 10^{-25} cm^{2}$. \vspace{0.2 cm} 2. Calculated recoil-electron spectra $S^{W}$(T,330) and $S^{M}$(T,330) at the end of the reactor ON period are shown in Fig.3. In searches for the neutrino magnetic moment, weak ${\widetilde {\nu}}_{e},e$ scattering plays the role of the reactor-correlated background. We note, that in order to keep this background at sufficiently low level one should try to study recoil electrons at not too high energies. As an example, the recoil electron energies $T > 100$ keV seem to be "high" to search for ${\mu}_{\nu} = 2\cdot 10^{-11}{\mu}_{B}$ while the range $T < 700$ keV is tolerably low for ${\mu}_{\nu} = 5\cdot 10^{-11}{\mu}_{B}$. Calculated time variations of the recoil-electron spectra during PWR reactor ON and OFF periods for weak and magnetic scattering are presented in Fig.4a,b. \section{DISCUSSION AND CONCLUSIONS} In practice the situation is not as simple as presented above. There occur deviations from the standard operating schedule: reactor can be stopped for a few days, or it can operate at a reduced level of power etc. For each particular experiment a comprehensive analysis of the reactor operation data should be carried out including all details. The main result of this study is that effects due to neutrino relaxa-tion play not negligible role in sensitive searches for the neutrino magne-tic moment at reactors. \section*{Acknowledgments} This work is supported by RFBR, projects 97-02-16031, 96-15-96640.
\section{Introduction} 4U\,1700-37 is a massive eclipsing X-ray binary consisting of a compact accreting object embedded in the wind of HD153919, a supergiant O star (Jones et al. 1973). Orbiting every 3.4 days, the compact object has an X-ray spectrum reminiscent of an accreting highly magnetized neutron star system (Haberl et al. 1989), although no X-ray pulsations have ever been confirmed. The X-ray flux is significantly less than expected on the basis of standard wind accretion theory. The most recent estimates of the mass of HD~153919 and its companion are from Heap \& Corcoran (1992), and Rubin et al. (1996). Heap \& Corcoran (1992) propose a mass for HD~153919 of ${\rm 50 \pm 2 M_{\odot}}$, with a companion mass of ${\rm 1.8 \pm 0.4 M_{\odot}}$. Rubin et al. (1996) reanalyzed the system parameters using Monte Carlo methods, finding masses of 30$^{+11}_{-7}$ M$_{\odot}$ and 2.6 $^{+2.3}_{-1.4}$ M$_{\odot}$ for HD~153919 and its companion, respectively. Both estimates are, at their lower ranges, consistent with the compact object being a neutron star, but a black hole cannot be excluded. Since pulsations have never been detected, further evidence for the object's nature must be sought from its X-ray spectrum. In general, X-ray pulsar spectra require an exponential cutoff to the continuum between $\sim$10--20~keV, unlike black hole spectra, which require either a single power-law or an ultra-soft component plus hard power-law tail (e.g., Tanaka \& Lewin 1995). Additionally, about half of the sample of 23 X-ray pulsars observed by {\it Ginga} show cyclotron features in their spectra, usually at energies between 10--60~keV (Makishima \& Mihara 1992; Mihara 1995), indicating the presence of strong magnetic fields. \section{Previous X-ray spectroscopy of 4U\,1700-37} The X-ray spectrum of 4U\,1700-37 has been studied using many different satellites, including HEAO-1 and {\it Einstein} (White et al. 1983), and EXOSAT (Haberl et al. 1989). The spectral shape is highly reminiscent of other X-ray pulsars (White et al. 1983), being well described by a power-law with photon index, $\alpha$, of $\sim$0.15, modified by a high-energy cutoff above 6--7~keV. Haberl \& Day (1992) confirmed the above results with {\it Ginga} observations, also finding that $\alpha$ varies between $-$0.5 and 1.0. Haberl et al. (1994) show that the 0.1--2.4~keV ROSAT Position Sensitive Proportional Counter (PSPC) spectrum is consistent with a thermal bremsstrahlung model with a temperature, kT, fixed at 0.74~keV, as originally found by {\it Ginga} (Haberl \& Day 1992), plus a hydrogen column density of $3.6 \times 10^{22}$~atom~cm$^{-2}$. Haberl \& Day (1992) also found that the best-fit values of $\alpha$ varied depending on the photoelectric column density. Since this is physically unrealistic, the authors show that the spectral variations could also be modeled using a two-component scattering model, consisting of two power-laws with the same slopes but variable absorption, one of which represents a less absorbed scattered component. The spectrum above 20 keV was examined using BATSE data by Rubin et al. (1996, and references therein) who confirmed earlier findings that it can be represented by a thermal bremsstrahlung model with kT $\sim$25~keV, out to 120~keV. There have been no reported changes in the shape of the high energy spectrum due to orbital phase or source intensity. While the above observations are individually useful, the launch of BeppoSAX, with its very wide spectral bandpass, offers the first opportunity to study the entire X-ray spectrum of 4U\,1700-37 simultaneously. BeppoSAX is proving to be a highly capable mission for discovering and studying cyclotron lines (e.g., Dal Fiume et al. 1998). For the first time, therefore, we present a single analysis of the entire spectrum, obtained at one epoch, spanning 0.5--200~keV. \section{Observations} The X-ray astronomy satellite BeppoSAX (Boella et al. 1997a) contains four coaligned Narrow Field Instruments, or NFI. Results from the Low-Energy Concentrator Spectrometer (LECS; 0.1--10~keV; Parmar et al. 1997), Medium-Energy Concentrator Spectrometer (MECS; 1.3--10~keV; Boella et al. 1997b), High Pressure Gas Scintillation Proportional Counter (HPGSPC; 5--120~keV; Manzo et al. 1997) and the Phoswich Detection System (PDS; 15--300~keV; Frontera et al. 1997) are presented here. The MECS consisted of three identical grazing incidence telescopes with imaging gas scintillation proportional counters in their focal planes. The LECS uses an identical concentrator system as the MECS, but utilizes an ultra-thin entrance window and a driftless configuration to extend the low-energy response to 0.1~keV. The non-imaging HPGSPC consists of a single unit with a collimator that is alternatively rocked on- and off-source. The non-imaging PDS consists of four independent units arranged in pairs each having a separate collimator. Each collimator can be alternatively rocked on- and off-source. 4U\,1700-37 was observed by BeppoSAX between 1997 April 1 11:21 and 22:15~UTC. This interval corresponds to orbital phases 0.44--0.58, where mid-eclipse of the X-ray source occurs at phase 0.0, using the ephemeris of Rubin et al. (1996). Good data were selected from intervals when the elevation angle above the Earth's limb was $>$$4^{\circ}$ and when the instrument configurations were nominal, using the SAXDAS 1.3.0 data analysis package. The standard collimator dwell time of 96~s for each on- and off-source position was used, together with rocking angles of 180\arcmin\ and 210\arcmin\ for the HPGSPC and PDS, respectively. The exposures in the LECS, MECS, HPGSPC, and PDS instruments are 12.2~ks, 23.7~ks, 11.2~ks, and 10.7~ks, respectively. LECS and MECS data were extracted centered on the position of 4U\,1700-37 using radii of 8\arcmin\ and 4\arcmin, respectively. Background subtraction in the imaging instruments was performed using standard files, but is not critical for such a bright source. Background subtraction in the non-imaging instruments was carried out using data from the offset intervals. The background subtracted count rates in the LECS, MECS, HPGSPC and PDS were 5.3, 18.5, 40.5, and 27.0~s$^{-1}$, respectively. \subsection{Spectral fits} The 4U\,1700-37 spectrum was investigated by simultaneously fitting data from all the NFI. The LECS and MECS spectra were rebinned to oversample the full width half maximum of the energy resolution by a factor 3 and to have additionally a minimum of 20 counts per bin to allow use of the $\chi^2$ statistic. Data was selected in the energy ranges 0.5--5.0~keV (LECS) and 1.8--10.0~keV (MECS) where the instrument responses are well determined. The HPGSPC and PDS data were rebinned using standard procedures in the energy ranges 7--40~keV and 15--200~keV, respectively. The photoelectric absorption cross sections of Morrison \& McCammon (1983) and the solar abundances of Anders \& Grevesse (1989) are used throughout. The spectrum was first fit with an absorbed power-law model, including an iron line, but with no cutoff or soft component. The iron line energy, ${\rm E_{Fe}}$, was fixed at 6.5 keV, but the width, ${\rm \sigma_{Fe}}$, and normalization (and all other parameters) were allowed to vary. Unsurprisingly, given the simplicity of the model, only a very poor fit with a reduced $\chi^{2}$ of 47.0 for 280 degrees of freedom (dof) was obtained. Matters are significantly improved if a high energy cutoff and a low energy thermal bremsstrahlung component are included, resulting in a reduced $\chi^{2}$ of 2.00 for 276 dof. The best-fit thermal bremsstrahlung component (see Haberl et al. 1994) has a kT of $0.2 \pm 0.1$~keV. Both continuum components suffer low-energy absorption, ${\rm N_H}$, of $(5.1 \pm 0.2) \times 10^{20}$~atom~cm$^{-2}$. The line width was consistent with a narrow line and so was fixed at a value (0.1~keV) much smaller than the instrumental resolution. This gives an equivalent width, EW, of $120 \pm 20$~eV for the iron line. The functional form of the cutoff above energy ${\rm E_{cut}}$ is ${\rm \exp[(E_{cut} - E)/E_{fold}]} $, where ${\rm E_{cut}}$ and ${\rm E_{fold}}$ are the cutoff and folding energies, respectively. The count rate spectrum for this model is shown in the left panels of Fig.~1, together with the contributions to $\chi$$^{2}$. A significant contribution to the existing poor fit quality is residual structure in the PDS spectrum at $\sim$37~keV (see the left hand panels of Fig.~1). Note that the same structure is not evident in the HPGSPC spectrum. Since this feature is reminiscent of a cyclotron resonance line, we added a single cyclotron component (the {\sc cyclabs} model in {\sc xspec}) to the model and re-fitted the data. The fit quality improves sharply, with a reduced $\chi$$^{2}$ of 1.42 for 273 dof. This improved fit is shown in the right hand panels of Fig.~1. The best-fit values of ${\rm E_{cut}}$ and ${\rm E_{fold}}$ are $5.9 \pm 0.2$~keV and $23.9 \pm 0.5$~keV, respectively. These values are consistent with those of $6.6 \pm 0.7$~keV and $21.1 ^{+4.3} _{-3.3}$~keV found by Haberl et al. (1989) using EXOSAT data. Haberl \& Day (1992) determined a slightly higher cutoff energy using {\it Ginga}, measuring ${\rm E_{cut}}$ = 7.6~keV, and ${\rm E_{fold}}$ = 19.5~keV, with no uncertainties quoted. Nonetheless, given that these observations were made with different missions at different epochs, the broad agreement is gratifying and shows that the cutoff is a necessary component in the spectrum. \begin{figure*} \begin{center} \hbox{ \epsfig{file=1700_spec13.ps, angle=-90, width=8.5cm, clip=} \hspace{0.5cm} \epsfig{file=1700_spec23.ps, angle=-90, width=8.5cm, clip=}} \caption{The NFI 0.5--200 keV spectrum of 4U\,1700-37. The left panels show the fit to an absorbed power-law and low-energy thermal bremsstrahlung model together with an iron emission line at 6.5~keV and an exponential cutoff beginning near 6~keV. In the right panels a broad cyclotron absorption feature at $\sim$37~keV is included in the model. The contributions to $\chi^{2}$ are plotted in the lower panels} \label{fig:spectrum} \end{center} \end{figure*} \begin{table} \caption[]{Best-Fit BeppoSAX spectral parameters. Norm${\rm _{PL}}$ and Norm${\rm _{brem}}$ are the power-law and thermal bremsstrahlung normalizations in units of Photon keV$^{-1}$~cm$^{-2}$~s$^{-1}$. A${\rm _{FE}}$ is the total number of iron line photons in units of cm$^{-2}$~s$^{-1}$. All uncertainties are quoted at 90\% confidence. The values without uncertainties were held fixed in the fits} \begin{flushleft} \begin{tabular}{ll} \hline\noalign{\smallskip} Parameter & Value \\ \noalign{\smallskip\hrule\smallskip} ${\rm N_{H}}$ (10$^{22}$~atom~cm$^{-2}$) & 5.1 $\pm$ 0.2 \\ $\alpha$ & $1.07 ^{+0.02} _{-0.03}$ \\ Norm${\rm _{PL}}$ & $0.17 \pm 0.08$ \\ ${\rm E_{FE}}$ (keV) & 6.5 \\ ${\rm \sigma_{FE}}$ (keV) & 0.1 \\ A${\rm _{FE}}$ &$(2.7 \pm 0.2) \times 10^{-3}$ \\ EW${\rm _{FE}}$ (keV) & $0.120 \pm 0.020$ \\ kT (keV) & 0.2 $\pm$ 0.1 \\ Norm${\rm _{brem}}$ & 180$^{+480} _{-120}$\\ ${\rm E_{cut}}$ (keV) & 5.9 $\pm$ 0.2 \\ ${\rm E_{fold}}$ (keV) & 23.9 $\pm$ 0.5 \\ ${\rm E_{cyc}}$ (keV) & 36.6 $\pm$ 1.0 \\ ${\rm \sigma_{cyc}}$ (keV) & 11$^{+5}_{-3}$ \\ \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \label{tab:bestfit} \end{table} The best-fit energy, ${\rm E_{cyc}}$, and width, ${\rm \sigma_{cyc}}$, of the feature are $36.6 \pm 1.0$~keV and $11 ^{+5}_{-3}$~keV, respectively. This width is a factor $\sim$2.5 broader then expected from the correlation between these parameters observed by Dal Fiume et al. (1998) for five other massive X-ray binaries. This suggests that the feature may arise through incorrect modeling of the continuum. This is supported by the lack of detection in the HPGSPC. Variations in PDS performance, which could produce a similar spectral feature if the background and source spectra had slightly different gains, are however excluded. The feature is still present if the continuum is modeled with a broken power-law together with a high-energy cutoff, as suggested by Dal Fiume et al. (1998). Following Mihara (1995) and Dal Fiume et al. (1998), we also generated a Crab ratio spectrum, by dividing the PDS data by those obtained from a PDS observation of the Crab Nebula, which is time-invariant and has a smoothly varying E$^{-2.1}$ continuum in this energy range. This technique is useful in minimizing instrumental and model dependent effects. While the Crab ratio spectrum indicates some structure at $\sim$37~keV, it does not show the pronounced deficit which is the signature of the cyclotron features discussed in Dal Fiume et al. (1998). We conclude that while the suggestion of a cyclotron line is intriguing, and results in a significantly lower value of $\chi^{2}$, we cannot exclude the possibility that the feature is an artefact of the fitting process. Further improvements in fit quality require a better description of the spectral region covered by the LECS, since the main remaining contribution to $\chi^{2}$ is a feature at $\sim$1~keV (see Fig.~1). This feature has not been seen in broad-band spectra of 4U\,1700-37 before, since the {\it Ginga} observations did not cover the energy region below 1.5~keV, but it may also be evident in the residuals of the ROSAT PSPC spectrum presented in Haberl et al. (1994). Since Rubin et al. (1996) show that the BATSE spectrum above 20~keV is consistent with a thermal bremsstrahlung model, we examined the PDS data separately to check for consistency with this earlier study. The result of Rubin et al. (1996) is confirmed, with the PDS spectrum being described by a thermal bremsstrahlung with kT = $25.9 \pm 0.3$~keV, in agreement with the temperature of 25~keV derived from the BATSE data, for a reduced $\chi^{2}$ of 2.16 for 46~dof. The addition of a broad cyclotron line at $\sim$40~keV improves the fit significantly giving a reduced $\chi^{2}$ of 1.03 for 43~dof. \subsection{The nature of the compact object in 4U\,1700-37} The broadband BeppoSAX spectrum presented here is very similar to the X-ray spectrum of an accreting pulsar. The underlying shape is an absorbed power-law sharply modified $\mathrel{\hbox{\rlap{\lower.55ex \hbox {$\sim$}$6.0~keV by a cutoff. Although the photon index of $\sim$1.0 is rather higher than the average reported by previous authors, it lies within the range of observed variation exhibited by 4U\,1700-37, and is at the upper range of the values for various pulsars summarized in White et al. (1983). In what respects does the 4U\,1700-37 spectrum differ from, or resemble that, of a black hole? Four diagnostics of black hole candidates (none of them individually conclusive) are listed in Tanaka \& Lewin (1995). These are: (1) ultrasoft spectra when luminous, (2) high-energy power-law tails, (3) bimodal spectral states, and (4) millisecond variability and flickering in the hard state. 4U\,1700-37 possesses none of these attributes. In particular, steep, cutoff spectra (${\rm E_{cut}} \mathrel{\hbox{\rlap{\lower.55ex \hbox {$\sim$}$20~keV) are a feature of neutron stars rather than black holes (White et al. 1988), while black holes (at least in outburst) generally possess hard tails, emitting significant flux $\mathrel{\hbox{\rlap{\lower.55ex \hbox {$\sim$}$100~keV (see e.g., Ballet et al. 1994). 4U\,1700-37 resembles neither the ultrasoft ``high-state'', nor the hard power-law ``low-state'' spectrum of the canonical black hole candidate Cyg X-1. \section{Discussion} We have demonstrated that the broadband X-ray spectrum of 4U\,1700-37 is qualitatively similar to that of other accreting neutron star X-ray pulsars. A feature at $\sim$37~keV may be modeled as a broad cyclotron absorption line, but its presence is uncertain due to uncertainties in continuum modeling and instrument performance. Such features have only been reported in about half of the known X-ray pulsars -- and in some cases only at marginal significance. The lack of pulsations remains puzzling, and may suggest (as has been argued by previous authors) that the magnetic field is intrinsically weak or aligned with the rotation axis. The neutron star's progenitor may have been very massive, suggesting that there may be a range of masses over which progenitors form both black holes and neutron stars, or that black holes are formed only through a limited range of progenitor masses, above which the remnants are again neutron stars. \begin{acknowledgements} The BeppoSAX satellite is a joint Italian-Dutch programme. We thank Fabio Favata, Matteo Guainazzi, and Tim Oosterbroek for helpful discussions. We also thank the referee, Frank Haberl, for suggestions which helped improve the paper. \end{acknowledgements}
\section{Introduction} As the interaction energy for a heavy-ion reaction increases above the Coulomb barrier, the reaction grazing angle moves to more forward angles so that, by a value of about 20\% above the barrier, very small quasielastic and single-nucleon transfer reaction yields are expected at larger scattering angles. Instead, the incident flux that might be expected to scatter to larger angles, corresponding to smaller impact parameters, is trapped by the formation of a compound nucleus. For lighter systems, this compound system subsequently decays by light particle and $\gamma$-ray emission, with a very small heavy-fragment $(A > 4)$ emission component. The experimental observation of significant large-angle, elastic-scattering cross sections at energies well above the Coulomb barrier has therefore been viewed with considerable interest. Initially seen in systems of comparable target and projectile masses \cite{ob75}, anomalous large-angle elastic cross sections have been found in a number of systems. An early explanation for these yields \cite{ob75} was in terms of a possible elastic-transfer mechanism where, simply, the target and projectile were viewed to exchange identities. Alternatively, optical model calculations have also been found to predict enhanced back-angle yields when weak absorption is assumed for the grazing partial waves. Complicating the picture is the observation that not all systems are found to demonstrate the elastic scattering enhancements. Also, some systems which exhibit the phenomena involve the scattering of particles of very different masses. As a further complication, many of the systems demonstrating a large-angle elastic scattering enhancement are also found to show resonant behavior in scattering and transfer reaction excitation functions. An overview of the anomalous large-angle scattering (ALAS) phenomenon is given in the review paper of Braun-Munzinger and Barrette \cite{bmb82}. Although several mechanisms could be identified as possible reasons for enhanced large angle yields in elastic and quasielastic channels, it was still a surprise when Shapira {\it et al.} \cite{sfgsd79} discovered similar large-angle yield enhancements for strongly energy-damped inelastic channels of the $^{12}$C+ $^{28}$Si reaction. The energy spectra for these channels were found to be peaked at values that one would expect if the $^{12}$C and $^{28}$Si fragments are emitted at rest from a sticking configuration of the compound system. Further, angular distributions of constant $d\sigma/d\theta_{c.m.}$ suggest a long-lived intermediate complex. These features led to the suggestion that an orbiting, dinucleus configuration is formed that decays back to the entrance channel. Weak absorption of the grazing partial waves, as previously suggested in the elastic case, is necessary to avoid having the orbiting configuration spreading into the compound nucleus states. However, whereas for the elastic scattering enhancements it was only necessary to invoke weak absorption for the grazing partial waves, the orbiting phenomenon suggests weak absorption for even lower partial waves with values between the critical angular momentum for fusion and the reaction grazing angular momentum. After the discovery of orbiting in the $^{12}$C+$^{28}$Si system, similar enhancements of large-angle, binary-reaction yields were subsequently observed in somewhat heavier systems. While studying the strong resonance behaviors found in excitation functions of the $^{28}$Si+$^{28}$Si \cite{bdp81,bbg81,bs83} and $^{24}$Mg+$^{24}$Mg \cite{zkbsh83} elastic and inelastic channels, a significant non-resonant background yield was discovered in the energy spectra of these channels at higher excitation energies. This yield was found to extend to larger angles. Other evidence for damped, binary yields was found in a study by Grotowski {\it et al.} \cite{gmp84} of symmetric mass fragments from the $^{12}$C+$^{40}$Ca, $^9$Be+$^{40}$Ca and $^6$Li+$^{40}$Ca reactions. Using coincidence techniques, they established the existence of significant decay strength to the mass-symmetric channels for these systems. It is unlikely that this behavior could arise from a direct reaction mechanism because of the large difference in mass asymmetry between the entrance and exit channels. Angular distribution and energy spectra measurements were shown to be consistent with the fission decay of the respective compound systems. Although these measurements were done at relatively forward angles, the observed angular dependence would suggest significant large-angle cross sections. Concurrent with the experimental measurements of energy-damped, binary reaction yields in lighter systems came the development of finite-nuclear-range corrections to macroscopic-energy calculations for these systems. By taking better account of the role of the nuclear surface in determining nuclear binding energies, these calculations made it possible for the first time to obtain reasonable estimates of fission barrier heights in lighter systems \cite{sierk86}. Whereas the standard rotating liquid drop model indicates high fission barriers, thereby precluding the prediction of strong fission competition in these systems, the barrier energies found with the new calculations suggest that significant fission competition might occur. The possibility of fission competition in light systems raises the question of how one can distinguish between the orbiting and fusion-fission mechanisms. This is difficult to answer, although a clearer understanding of the respective processes has been emerging through a number of different studies, as will be discussed in this report. Some of the conceptual differences between the two mechanisms are indicated schematically in Fig. 1. Both processes are seen to involve near-grazing impact parameters with, however, the fusion-fission mechanism viewed as proceeding through the formation of a fully equilibrated compound nucleus. The fission decay of the compound nucleus is determined by the phase space available at a ``transition" configuration and, for light systems, can lead to significant population in many of the energetically allowed mass channels. For the orbiting mechanism, the system becomes trapped in a more deformed configuration than that of the compound nucleus and is inhibited from spreading into the compound nucleus states. This allows for significant decay back to the binary channels. Although orbiting can be considered as a deep-inelastic scattering mechanism, the rapid mass equilibration that is thought to occur in light systems can result in the population of channels with significantly different mass asymmetry than the entrance channel. Still, it is expected that the orbiting mechanism will retain a greater memory of the entrance channel than the fusion-fission process. The strong resonance behavior observed in excitation-function measurements of large-angle elastic and inelastic scattering yields in several light systems suggests that some structural aspect of the compound system may be strongly influencing these yields. This behavior is most pronounced in the $^{24}$Mg+$^{24}$Mg system \cite{zkbsh83} where measurements \cite{qzkp90} have indicated a resonance spin close to, or exceeding, the grazing angular momentum for the reaction. A possible explanation is that a highly deformed, metastable configuration for the compound system is formed. The relationship between the resonance behavior and the fission process, the latter of which is observed to occur at some level in all of the reactions studied at energies above the Coulomb barrier, is still an open question. A review of the resonance aspects of heavy-ion breakup reactions, including a discussion of earlier electrofission measurements, is found in the review article of Fulton and Rae \cite{fr90}. This paper will review the status of our experimental and theoretical understanding of energy-damped, binary-decay processes in light systems. Although the relationship between these processes and that responsible for heavy-ion resonance behavior will be explored, the general topic of heavy-ion resonances in light systems will not be discussed and the reader is referred to several review papers that already cover this general topic \cite{fr90,eb85,gps95}. In Sec. \ref{sec:exp} we will explore the general experimental signatures of the binary yields observed in lighter systems. In Sec. 3 the experimental techniques used to study these behaviors will be presented and some of the experimental difficulties highlighted. In Sec. 4 the various models that have been developed to explain the observed yields are discussed. In Sec. 5 we will develop the experimental systematics by presenting a system-by-system discussion of current data, comparing these data to the model calculations. In Sec. 6 we will present some of the open problems and suggest possible measurements that might address these problems. Finally, in Sec. 7 we will summarize the present review. \section{Experimental Behaviors of Binary Decay Yields} \label{sec:exp} \setcounter{section}{2} \setcounter{equation}{0} One way to characterize nuclear reactions is in terms of the number of nucleons exchanged between the incident particles. Alternatively, one can consider the degree to which the kinetic energy of the incident channel is transferred to internal excitation of the outgoing fragments. Clearly, there can be a significant correlation between the number of nucleons transferred and the energy damping that occurs in the reaction. In our discussion of the general reaction properties, we use a somewhat arbitrary division that is motivated by the experimental observables of mass, angle, and energy excitation of the fragments. We start with a discussion of the large-angle elastic scattering yields. We then consider the situation where one observes some energy damping, and possibly mass flow, but still do not have angular distributions of constant $d\sigma/d\theta_{c.m.}$ that would signify a very long lived complex. Finally, we consider the situation where all degrees of freedom of the system appear to have reached full equilibration. \subsection{Large-angle elastic and quasi-elastic scattering} After the first observations in heavy-ion reactions involving near-identical particle systems \cite{ob75} of large-angle elastic-scattering yields in excess of the optical model expectations, these backward-angle enhancements were found to be a relatively common feature of collisions between p-shell and sd-shell nuclei \cite{bmb82}. Figure 2 shows an angular distribution for the $^{16}$O+$^{28}$Si reaction at $E_{lab}=55$ MeV. The dashed curve is the prediction of an optical model calculation using the strong absorption E18 potential of ref. \cite{cdgz76}. The large angle yield is far in excess of the model calculation and demonstrates a highly oscillatory behavior. As shown in the insert, the angular distribution suggests a single orbital angular momentum of $\ell=26\hbar$ may dominate the large-angle yields. An excitation function of the backward angle yields for this system shows strong resonance-like enhancements. This behavior might result from the formation of nuclear molecular configurations \cite{eb85,gps95,tmc82}, to an orbiting phenomenon \cite{s82} related to the dynamics of the interaction potential, or to compound-elastic scattering, that is, the fission decay of the compound nucleus back to the elastic channel. Two or more of these mechanisms may coexist. The explanation for the resonance-like structures that are often found associated with backward angle elastic and quasi-elastic yields in light systems might be found in the interplay between interaction barriers and energy dissipation mechanisms. The observed large-angle cross sections suggest the formation of a long-lived nuclear molecule or dinucleus system that fragments at a large deformation. This is consistent with the sticking model of deep inelastic collision processes for which two colliding nuclei adhere to each other and undergo finite rotation before separating, with the possible transformation of orbital angular momentum into spin of the outgoing fragments and with a possible loss of kinetic energy of relative motion. The resonances might correspond to particularly simple configurations of the dinucleus complex. In a 1982 review of the ``anomalous" large-angle scattering phenomenon (ALAS), however, Braun-Munzinger and Barrette \cite{bmb82} have shown the difficulty of attributing a specific reaction mechanism to the resonance behavior. The current focus is to establish a global systematics for the occurrence of ALAS behavior based on the occurrence of weak absorption of the near-grazing partial waves \cite{baa94}. \subsection{ Deep-inelastic scattering yields} Deep-inelastic reactions between heavy nuclei $(A_{target} + A_{projectile}>50)$ typically involve substantial rearrangement of the nuclear matter of the incident particles, with strong damping of the entrance-channel's kinetic energy and transfer of orbital angular momenta to the spins of the outgoing fragments \cite{v78,sh84}. The composite system may reach a configuration close to the compound-nucleus saddle point before subsequent elongation and fission. Wilczynski \cite{w73} has developed a qualitative interpretation of deep-inelastic reactions in terms of a nuclear orbiting process with dissipation of energy resulting from frictional forces. A schematic illustration of the progression from quasielastic scattering to nuclear orbiting in heavy-ion collisions is given in Fig. 3 (adopted from ref. \cite{w73}). The two dimensional plot shows a contour of maximum cross section (``Wilczynski diagram") that contains the components of the three possible classical trajectories, as drawn in the figure. Trajectories 1 and 2 correspond to near-side and far-side scattering from the target nucleus. In trajectory 3 the dinucleus system survives long enough to complete close to, or more than, one full rotation. It is associated with an ``orbiting" process and leads to emission spectra and emission probabilities that are independent of angle \cite{sh84}. That is, there is complete energy damping resulting in the final kinetic energy of the fragments being close to the relative potential energy at the point where the fragments are closest together. Also, the fragments' angular distributions of $d\sigma/d\Omega$ follow a $1/\sin \theta_{c.m.}$ angular dependence (corresponding to angular distributions of constant $d\sigma/d\theta_{c.m.}$) suggesting the occurrence of a long-lived dinucleus object in a sticking configuration. This early classical picture of the orbiting phenomenon has been developed further using the classical models proposed, for example, by Bondorf {\it et al.} \cite{bss74} and Gross and Kalinowski \cite{gk78} for heavy-ion damped reactions. The significant cross sections ($>10$ mb) observed for large-angle inelastic scattering yields of reactions such as $^{20}$Ne+$^{12}$C \cite{sfgsd79}, $^{27}$Al+ $^{16}$O \cite{sfg80}, $^{28}$Si+ $^{12}$C \cite{s82}, and $^{24}$Mg+$^{12}$C \cite{gdbh90} were initially interpreted in terms of the formation of a dinucleus, orbiting configuration. These large yields were found to be very surprising since any system in this mass range and making such close contact, as suggested by an orbiting behavior, was expected to fuse into a compound nucleus. The picture that emerges is one where the interacting ions are trapped in a pocket of the ion-ion potential that results from the combined effects of the Coulomb, centrifugal, and nuclear interactions. The orbiting complex may act as a ``doorway" to fusion, but weak absorption may also allow decay back to the binary channels, possibly after significant mass flow and rotation of the nuclear configuration. The description of an equilibrium orbiting model will be developed in Sec. \ref{sec:eqorb}. In subsequent investigations it has been shown that at least some of the yield attributed to orbiting may correspond instead to the process of fusion followed by fission. Still, experimental evidence for a distinct orbiting process in light systems will be presented with the discussion of the $^{28}$Si+$^{12}$C and $^{28}$Si+$^{14}$N reactions. A non-compound origin is surmised for the reaction yields in these systems based on their entrance-channel dependence \cite{rk85}. \subsection{Fully energy-damped binary yields} \subsubsection{Angular Distributions} Enhanced large-angle yields in heavy-ion reactions reflect a relatively long reaction time allowing significant rotation of the intermediate complex formed in the reaction. For the mechanism of fusion followed by fission, the evolving intermediate complex is taken to pass through a fully equilibrated compound nucleus. In the orbiting picture, a long-lived dinucleus configuration is envisioned. In both cases the binary yields are assumed to correspond to peripheral reactions involving incident angular momenta close to the critical angular momentum for fusion $\ell_{cr}$. In considering the fission of heavier mass systems than those discussed here, angular distribution data have yielded significant insight on the details of the breakup mechanism \cite{sk97}. The ``reference" distributions for these studies are those predicted by the standard transition-state model of fission \cite{vh73}. Using heavy-ion reactions to populate the systems, it is commonly found that the orbital angular momenta {\bf $L$} are large compared to the projectile and target spins, resulting in the projection of the total angular momentum {\bf $J$} on the space fixed z-axis, the beam direction, to be small. Assuming a Gaussian distribution of {\bf $K$}, the projection of {\bf $J$} on the symmetry axis of the fissioning system, and in the limit of zero projectile and target spins, the expected fission angular distribution becomes \cite{vh73} \begin{equation} \begin{array}{rll} W(\theta)\propto & \Sigma^\infty_{J=0}(2J+1)T_J & \\ & \times {\displaystyle \Sigma^J_{K=-J} {(2J+1)|d^J_{M=0,K}(\theta)|^2 \exp (-K^2 /2 K^2_0)\over \sum\limits^J_{K=-J}\exp (-K^2/2K^2_0)} } ,\\ \end{array} \label{eqn:ang} \end{equation} where the $T_J$'s are the transmission coefficients giving the combined probability for formation and fission decay of a compound nucleus with $J=L$ and $d^J_{MK}$ is the reduced rotation matrix. The variance of the $K$ distribution can be expressed as \[K^2_0=I_{eff}T/\hbar^2 ,\] where \[{1\over I_{eff}}={1\over I_\parallel}-{1\over I_\perp}\] and $I_\perp$ and $I_\parallel$ are the moments of inertia about axes perpendicular and parallel to the symmetry axis, respectively. The ``transition'' state is typically taken as the saddle point, defined as the configuration where the potential energy of the compound system as a function of deformation and mass asymmetry is at its maximum. The temperature at the saddle point $T$ is found by $T=\sqrt{E_x/a}$, where $E_x$ is the energy of internal excitation at the saddle point and $a$ is the level density parameter. Deviations between the observed angular distributions and those predicted by the standard transition-state model using the above expression for $W(\theta)$ have been used as evidence for faster processes than that of complete fusion followed by fission. Deep-inelastic scattering mechanisms, where the characters of the incident nuclei are maintained, can also lead to large angle yields if a dinucleus configuration is reached that survives for a significant fraction of the rotation period. If a large number of partial waves contribute to the reaction, the resulting angular distributions become relatively featureless. For heavy-ion reactions, where the angular momenta associated with processes leading to large angle yields tends to be high, the resulting angular distributions approach the classical limit associated with systems emitting fragments perpendicular to the rotation spin vector, with ${d\sigma / d\theta_{c.m.}} = C$ or, equivalently, ${d\sigma / d\Omega}={C' / \sin \theta_{c.m.}}$, where $C$ and $C'$ are angle-independent normalization constants. For the systems considered in this report, the differences between the angular distributions predicted by the standard transition state model and the ${d\sigma / d\Omega}={C' / \sin \theta_{c.m.}}$ distribution that characterizes a classical, long-lived orbiting configuration are only evident at very forward and backward angles. This is illustrated for the $^{35}$Cl+$^{12}$C reaction at $E_{lab}= 180$ MeV \cite{bdhd89,bdhf92}. Angular distributions of fission-like fragments are shown in Fig. 4 for the inclusive yields to different mass channels. One finds distributions of constant ${d\sigma / d\theta_{c.m.}}$ for each of the measured isotopes. In Fig. 5 the angular distribution data in ${d\sigma / d\Omega}$ for the Ne channel are compared to both a $1/\sin \theta_{c.m.}$ angular dependence and the angular dependence predicted by the transition state model (using Eqn. \ref{eqn:ang}). For this latter calculation the transmission coefficients for the fission yields $T_J$ were based on the transition-state calculation to be discussed later in this report and the level density parameter was taken as $a = A_{CN}/8$. The two models show overlapping behavior in the region of the experimental data and significant differences between the calculations are only observed at very forward and backward angles. Unfortunately, the cross sections at angles near to 0$^\circ$ and 180$^\circ$ are also very difficult to measure. Large elastic scattering yields can obscure the forward angle data and low particle energies can make the large angles yields difficult to detect. In general it has not been possible in lighter systems to deduce information about the time scales of the more strongly damped processes based on angular distribution data. A more rapid rise in the energy-inclusive cross sections at forward angles than given by a 1/sin$\theta_{c.m.}$ distribution indicates a shorter lifetime of the composite system, with breakup occurring within the first revolution of the system. Such lifetimes are incompatible with the formation of an equilibrated compound nucleus, but may still reflect significant energy damping within a deep-inelastic mechanism. By measuring the forward peaked distributions, it is possible to estimate the lifetime of the intermediate nuclear complex using a diffractive, Regge-pole model. The distributions are fitted by \begin{equation} {d\sigma\over d\Omega}={C\over \sin \theta_{c.m.}} \left\{ {{\rm e}^{- \theta_{c.m.}/\omega t} +{\rm e}^{-(2\pi - \theta_{c.m.})/\omega t}}\right\} . \label{eqn:life} \end{equation} This expression describes the decay of a rotating dinucleus with angular velocity $\omega ={\hbar \ell / \mu R^2}$ where $\mu$ represents the reduced mass of the system, $\ell$ its angular momentum (which should fall somewhere between the grazing $\ell_g$ and critical $\ell_{cr}$ angular momentum), and $R$ represents the distance between the two centers of the dinucleus. Small values of the ``life angle" $\alpha(\equiv \omega\tau$) lead to forward peaked angular distributions associated with fast processes, whereas large values of $\alpha$, associated with longer times as compared to the dinucleus rotation period $\tau$, are consequently associated with longer lived configurations and lead to more isotropic angular distributions. In the limiting case of a very long-lived configuration, the distributions approach a $d\sigma / d\Omega \propto 1 / \sin \theta_{c.m.}$ dependence. One way to characterize the sensitivity of the angular distributions to the ``life angle" $\alpha$ is by considering the anisotropy function $R(\theta)$, where \[R(\theta)={d\sigma\over d\Omega}(\theta,\alpha)/ {d\sigma\over d\Omega}(90^o,\alpha) . \] The functional dependence of $R(\theta)$ with $\alpha$ is shown in Fig. 6 for $\theta = 10^o$ and $170^o$. The corresponding anisotropy for an angular distribution with ${d\sigma / d\Omega}\propto {1 / \sin \theta_{c.m.}}$ is shown by the dashed line. The plot suggests that an anisotropy measurement might be sensitive to the time scale of fast processes $ (\alpha < 90^o)$, but is less sensitive to the time scale of strongly damped processes with $\alpha > 120^o$. Measurements of excitation energy integrated yields of specific mass channels show, in some systems, angular distributions which reflect the contributions of both short- and long-lived processes. As an example, the observed distributions for inclusive, energy-summed yields for the $^{16}$O+$^{11}$B reaction at $E_{lab}=64$ MeV are shown in Fig. 7 \cite{aacf94}. It should be noted that for this particular reaction, the isotope pairs C-N, B-O, and F-Be correspond to specific exit channels. From this figure one finds a contribution of forward peaked processes (N and F fragments) with $\alpha<70^o$, corresponding to direct mass transfer, with slower processes (C, B, and Be fragments) presenting large ``life angles", with $\alpha \gg 180^o$. \subsubsection{Energy Spectra} For fully energy-damped reactions, the energy spectra are peaked at large negative $Q$-values, with the cross section maxima typically occurring at a value of total kinetic energy in the exit channel that corresponds to the relative potential energy of the two nascent fragments in a near-touching configuration. This observation is consistent with the idea that the composite system passes through a stationary point in the potential energy surface where the relative kinetic energy is at a minimum. The stationary point could be the saddle point of a compound nucleus undergoing fission or, alternatively, the dinucleus configuration of an orbiting binary complex. The total kinetic energy of the fragments (TKE) in the exit channel is then expected to equal the sum of the nuclear and Coulomb potential energies and the rotational energy of the rotating complex at the stationary point, with \[TKE = V_{nucl}(d) + V_{Coul}(d) + V_{rot}(d) ,\] and $V_{nucl}$, $V_{coul}$, and $V_{rot}$ are the nuclear, Coulomb, and relative rotational energy between the two nascent breakup fragments at a separation $d$ corresponding to the stationary point. In general, the sticking limit is assumed in calculating the rotational energy contribution. The total kinetic energy of the outgoing fragments is therefore sensitive to the deformation of the system at the time of scission \cite{s82,e76,bmcg76,cbmc77,nnegg77,a80,sgmd80,rdg85}. In general, the most probable TKE values are well described by assuming stationary shapes consistent with the predicted saddle-point configurations of the nuclear systems \cite{s91}. The exceptions to this rule tend to occur for nuclear systems where anomalous large-angle elastic scattering (ALAS) cross sections are observed. It should be noted that even though large kinetic energy damping is one of the experimental signatures used to distinguish a damped reaction process from direct reactions in which only a small part of the initially available kinetic energy is dissipated, it is not necessarily the case that a given fusion-fission or orbiting reaction event will have a large energy loss. The distribution of the $TKE$ values about the peak value can be large and, in some cases, may lead to the population of the ground states of the outgoing fragments, as would be the case for compound elastic scattering. The dependence of the average energy release in fission on atomic number and mass of the compound nucleus has been widely studied for heavy systems. A linear relationship between the most probable TKE release value with $Z^2/A^{1/3}$ of the fissioning nucleus has been established by the systematic work of Viola {\it et al.} \cite{vkw85}, with this systematics more recently extended to the lighter systems in ref.\cite{tt92}. The compiled $TKE$ values \cite{bmetc95} of the symmetric fission fragments produced in light heavy-ion systems are shown in Fig. 8. The original Viola systematics \cite{vkw85} (dashed line) are capable of describing the whole data set except in the case of low $Z$ fissioning nuclei. In these very light systems, the diffuse nature of the nuclear surface and the associated perturbations of the necking degree of freedom, as calculated using the finite-range, liquid-drop model \cite{ajs85}, results in a change in the predicted slope, leading to vanishing $TKE$ values as $Z$ approaches zero. This effect is observed experimentally and is reproduced by the solid curve in Fig. 8, which is calculated using \cite{tt92}: \[TKE = Z^2/(aA^{1/3}+bA^{-1/3}+cA^{-1})\] where the values of the fitting parameters are a=9.39 MeV$^{-1}$, b=-58.6 MeV$^{-1}$ and c=226 MeV$^{-1}$, respectively. \section{Experimental Arrangements} \setcounter{section}{3} \setcounter{equation}{0} Several considerations influence the experimental techniques used to study the processes considered in this report. In a given measurement most of the processes that compete with those being studied correspond to more peripheral collisions and result in yields at angles near to or forward of the grazing angle. This requires that angular distributions be extended to larger angles where the more peripheral mechanisms have little yield, although with the possible experimental complication of the energy of the reaction fragments falling below the identification thresholds of the detectors being used. To provide a complete description of the reaction process, it becomes necessary to determine, for each fragment, the mass(A), atomic number (Z), energy ($E$) (or velocity ($v$)), and the emission angle ($\theta$) of the particle. In cases where the fragments are emitted in binary decays, the conservation laws for mass, energy, and momentum can be used to reduce the number of quantities needed to be measured to fully specify the reaction. If one or both of the fragments is at an energy above its particle emission threshold, however, there is likely to be a secondary light particle emitted before the fragment is detected. The effects of such secondary light particle emission need to be considered. In the simplest arrangements for detecting single particles, element identification (Z) and energy measurement ($E$) is generally accomplished by determining differential energy loss $\Delta E$ and the total energy $E$ in a telescope consisting of a gas $\Delta E$ counter backed by a solid-state detector. The $\Delta E$ signal is proportional to the square of the charge. By stopping the particle in the solid state detector, the residual energy is also determined and thus the total energy can be deduced. Alternatively, a Bragg-curve detector can supply the information on $Z$ and $E$ \cite{ku94}. In either case, the desired cross sections tend to be low (typically, $d\sigma/d\Omega < $5 mb/sr), requiring large solid-angle detectors and, possibly, necessitating the use of position-sensitive detectors for angle determination. For measurements done at pulsed beam facilities it is possible to determine the energy and mass of a particle through time-of-flight and total energy measurement. In this case a single, bare Si(surface-barrier) detector can be used to measure the yields. The detection of low energy particles may still be difficult, however, as a consequence of multiple scattering effects of the particle leaving the detector and the dependence of the timing signal on the energy and charge of the particle incident on the detector. In general, depending on the preferred detection technique, experiments have either measured the charge or mass of the outgoing fragments, but not both simultaneously. As an alternative to the singles measurement of the reaction products, for a binary decay it is also possible to determine the mass and energy of the outgoing fragments by measuring the scattering angles of both fragments and the relative timing of the two fragments. This technique has proven to be quite valuable since it allows for the use of inexpensive, large-area, position-sensitive avalanche detectors (PPACs). Figure 9 illustrates the basic variables of the measurement. The two outgoing particles are detected at laboratory scattering angles of $\theta_1$ and $\theta_2$, respectively, in detectors located at distances of $d_1$ and $d_2$ from the target. The relative time of arrival at the detectors, $\Delta t=t_2-t_1$, can be measured directly using a time-to-amplitude circuit, or can be deduced from measurements of the individual flight times. In the non-relativistic limit, the linear momenta of the two particles in the laboratory system, $\vec p_1$ and $\vec p_2$, can be determined from the total laboratory momentum $\vec p_o$ and the scattering angles, with \[ p_1=p_o{\sin (\theta_2)\over \sin (\theta_1+\theta_2)}, \; {\rm and} \] \[ p_2=p_o{\sin (\theta_1)\over \sin (\theta_1+\theta_2)} .\] The masses of the two scattered particles can be related to the above quantities and the projectile and target masses, $A_{proj}$ and $A_{targ}$, with \[A_2={(d_1 {\displaystyle A_{proj}+A_{targ}\over {\displaystyle p_1}}+\Delta t )\over {\displaystyle ({d_2\over p_2}+ {d_1\over p_1}) }},\; {\rm and}\] \[A_1=A_{proj} + A_{targ} - A_2.\] It is then possible to determine the energies of the two particles and the corresponding reaction $Q$-value using the derived masses, with \[E_1 = E_{lab}{A_{proj}\over A_1} [{\sin (\theta_2)\over \sin (\theta_1+\theta_1)}]^2\; ,\] \[E_2=E_{lab}{A_{proj}\over A_2} [{\sin (\theta_1)\over \sin (\theta_1+\theta_1)}]^2\; ,\; {\rm and}\] \[Q=E_1+E_2-E_{lab}\; ,\] where $E_{lab}$ is the laboratory beam energy. If single mass resolution is achieved (the typical situation for the lighter systems being considered), then the reaction $Q$-value can be subsequently determined by using only the position information and the known masses. In this case, the excellent position resolution that can be achieved with the PPAC's can result in $Q$-value resolutions of $<200$ keV, with the limiting resolution determined largely by multiple-scattering effects as the particles leave the target. The coincidence detection of both reaction fragments can also be useful in cases where one or both of the fragments is formed with sufficient excitation energy to result in a secondary emission of a light particle (typically an $\alpha$ particle). On average, the emitted light particles result in a distribution of fragment velocities about their original, pre-evaporation values. In this case, the deduced mass distribution reflects the pre-evaporation values. If, in addition, the fragment mass is measured through an alternative method, one can determine the extent to which the secondary evaporation modifies the observed mass distribution from the original, preevaporation distribution \cite{setc89}. An essential element of most model calculations of the binary decay yields is an estimate of the spin distribution of the compound nucleus, as usually determined through measurement of the total fusion cross section. In general, the experimental determination of the fusion cross section is obtained by measuring the evaporation residues yields and, in the case of a fusion-fission analysis, adding in the fission yields. When pulsed beams are available, the evaporation-residue yields can be measured using bare, Si(surface barrier) detectors. Otherwise, it is possible to employ time-of-flight telescopes using micro-channel-plate detectors (MCP), or to use a set of $E$-$\Delta$E telescopes. Figure 10 shows the experimental arrangement used in a study of the binary yields from the $^{24}$Mg+ $^{24}$Mg reaction \cite{hsfp94} that has many of the elements discussed above. The Bragg-curve detector \cite{ku94} determined the charge and energy of outgoing fragments. A large multigrid avalanche counter (MGAC 2) \cite{wplf87} determined the angles of the particles entering the Bragg-curve detector. A second avalanche counter (MGAC 1) was used to detect the coincident fragments and determine their angles. Finally, a number of Si(surface barrier) detectors were used for normalization purposes and to obtain a measure of the evaporation-residue cross section. The use of angular distribution data to separate different reaction components requires that these distributions be measured over as large an angular range as possible. For a given fragment, energy thresholds set by the detectors may severely restrict the largest angle that can be reasonably measured. In the case of a binary decay, however, it may be possible to use a measurement of the corresponding recoiling fragment to extend the angular distribution. Figure 11 demonstrates this simple concept for the $^{18}$O+ $^{10}$B reaction going to the $^{15}$N+ $^{13}$C exit channel. The ``beamlike" particle ($^{15}$N) going to forward center-of-mass angles is detected at forward laboratory angles, as indicated in the velocity diagram. Although ``reverse kinematics" also results in forward scattering angles when the $^{15}$N particle undergoes large center-of-mass deflection, the laboratory energy of the particle is now small, complicating particle identification. At this point, however, the ``targetlike" $^{13}$C particle is scattered to forward angles and can be easily detected. Combining the measurements for the $^{15}$N and $^{13}$C results in a more complete angular distribution for this channel. Although this procedure is only valid in cases where secondary light-fragment emission can be ignored, it has been found to be very useful in the analysis of lighter systems \cite{aacf94,atst93}. In studies where the interplay between the nuclear structure and reaction dynamics is investigated, one important technique has been to identify the $\gamma$-decay cascade from the excited final fragments. In this case, the low reaction cross sections mandate experimental arrangements with both large particle and $\gamma$-ray detection efficiencies. The identified particles are used for channel selection and Doppler correction of the $\gamma$-ray spectra. Such measurements have been done in studies of the $^{32}$S+ $^{24}$Mg \cite{shp94} and $^{36}$Ar+ $^{12}$C \cite{fsd96} reactions. \section{Model calculations} \setcounter{section}{4} \setcounter{equation}{0} \subsection{ Statistical Models} It took only a short time after the experimental observation of neutron-induced fission for a quantitative description of the process to be developed in terms of the transition-state model \cite{bw39}. In this model, where fission competes with $\gamma$-ray and light-particle emission in the de-excitation of the compound nucleus, the probability for fission is determined by the most restrictive phase space (level density) encountered by the fissioning system between the equilibrated compound-nucleus configuration and the exit channel. This ``transition state" is usually taken as the configuration where the macroscopic potential energy reaches its maximum value, the saddle point, although discrepancies observed in heavier systems with the angular dependence described by Eqn. 2.1 have also led to the consideration of the more deformed scission point as the transition state \cite{pdb84,rhs84}. In lighter systems, where the saddle- and scission-point configurations are expected to be very close and little damping is expected as the system proceeds between the two, there is less reason to expect significant differences between calculations done at the saddle and scission points than for heavier systems where the shapes of the saddle- and scission-point configurations are quite different and significant damping can occur in moving between the two. Indeed, calculations based on saddle-point \cite{s91} and scission-point \cite{mbnm97} transition-state configurations are found to give equivalent results in lighter systems. In this section the transition-state formalism will be reviewed and the saddle-point \cite{s91} and scission-point \cite{mbnm97} calculations compared. The development of saddle-point calculations in lighter systems was significantly delayed from the comparable development for heavy-system fission because of the difficulty in accounting for the finite range and diffuse nuclear surface effects that strongly influence the macroscopic energies of these systems. In light nuclei, the saddle-point shapes correspond to two deformed spheroids separated by a well-developed neck region, with the surfaces of the two spheroids coming within close proximity of one another. This results in strong surface effects. The development of the finite-range model \cite{sierk86}, however, has made it possible to extend the saddle-point calculations to very light systems. Within this model, saddle-point energies are found by determining the stationary points of the compound nucleus potential-energy surface as a function of spin and constrained mass asymmetry. The shape parameterization, discussed by Nix in Ref. \cite{jrn68}, consists of three connected quadratic surfaces of revolution. The calculations explicitly account for the diffuse nature of the nuclear surface and the finite range of the nuclear interaction. Since the development of the finite-range model, one remaining difficulty in applying its results to the fission of lighter systems has been the lengthy process of performing individual calculations for all of the necessary spin and mass-asymmetry dependent saddle-points that might be encountered as the system breaks apart. Whereas for heavier systems the mass-asymmetry dependence of the macroscopic energy favors symmetric breakup, in lighter systems an asymmetric breakup is favored. This problem has been somewhat ameliorated with the development of a simple, double-spheroid parameterization of these energies \cite{s91} . In this double-spheroid picture, the saddle-point energy is given by \[ V_{saddle}(J_{CN}, \eta) = V_C + V_n + V_r + V_o ,\] where $J_{CN}$ is the spin of the compound nucleus, $\eta (= 1-2A_L/A_{CN})$ is the mass asymmetry, $V_C$ is the Coulomb energy between two deformed spheroids, $V_n$ is the nuclear interaction energy given by the two-body, finite-range-model potential of Krappe, Nix, and Sierk \cite{kns79}, and $V_r$ is the total rotational energy of the double-spheroid configuration, using a diffuse-surface corrected moment of inertia. $A_L$ is the atomic mass of the lighter of the two fission fragments. The final term, $V_o$, accounts for the influence of the saddle-point neck as well as other, shape-independent aspects of the macroscopic energy. The spheroid geometry is adjusted to reproduce the full finite-range-model calculations of Sierk \cite{sierk86}. A comparison of saddle-point energies found using the double-spheroid model and energies obtained from full macroscopic energy calculations is shown in Fig. 12 for several systems. Figure 13 (from Ref. \cite{s91}) compares some typical saddle-points shapes obtained with the full calculations (solid curves) and the corresponding double-spheroid shapes (dashed curves). The double-spheroid parameterization is found to closely reproduce the saddle-point energies found with the full calculations. Moreover, the shapes obtained with the double-spheroid model are similar (excluding the neck) to the corresponding finite-range-model configurations. Assuming that the saddle- and scission-point configurations are similar, this suggests that the relative energy of the two spheroids calculated within this model should be related to the total kinetic energy $E^{tot}_K$ of the fragments in the exit channel. A schematic diagram of the energy balance for a fusion-fission reaction is shown in Fig. 14, taken from Ref. \cite{shp94}. Starting from the entrance channel with incident center-of-mass energy $E_{c.m.}$, the Coulomb and centrifugal energy dominated entrance barrier must first be overcome to form a compound nucleus with excitation energy $E^*_{CN}$ and spin $J$. The effective excitation of the compound nucleus $E^*_{eff}$, which enters in calculating the density of compound nucleus states, is expected to be somewhat less than $E^*_{CN}$ (by an amount $\Delta_{eff}$) since, at these high energies, it is assumed that the virtual ground states corresponds to the macroscopic-energy ground state \cite{fp77}. The most restricted region of phase space, where the density of states becomes a minimum, occurs at the saddle point. In Fig. 14, $\varepsilon$ is taken as the kinetic energy associated with the radial motion at the saddle point. This term further reduces the density of saddle-point states. A saddle-point shell correction $\Delta V_{shell}$ is also taken to influence the fission decay probabilities. The excitation energy available at the saddle point $u_J$, which determines the corresponding level density $\rho_f$, is then given by \[u_j=E^*_{CN}-V_{saddle}(J,\eta)- \Delta V_{shell}-\Delta_{eff}-\varepsilon \] The probability for the compound nucleus of spin $J$ to break up to a fission channel of mass asymmetry $\eta$ is proportional to the level density $\rho_f$ above the corresponding spin $J$ saddle point. Some authors have taken the scission point as the transition state, in which case the corresponding energy of the two fragments at scission (not shown) is used to determine the transition-state level density. It will be shown that the very small energy difference $\delta$ believed to exist between the saddle and scission points leads to very similar fission predictions of the saddle- and scission-point models. Assuming $\delta \approx 0$, the total kinetic energy in the exit channel, and, correspondingly, the reaction $Q$-value, can then be related to the relative energy of the fragments at the saddle/scission configuration. In this case, taking $\varepsilon=0$ for the most probable decay probability (corresponding to the highest level density), \[E^{tot}_K=V_C+V_n+{\hbar^2\over 2\Im_{rel}}\ell (\ell +1)\] with \[\ell={\Im_{rel}\over \Im_{tot}} J_{CN}.\] The relative moment of inertia of the two spheroids is given by $\Im_{rel}=\mu r^2$ and, taking the moments of inertia of the individual deformed spheroids as $\Im_1$ and $\Im_2$, respectively, the total moment of inertia for the two spheroid configuration is then given by \[\Im_{tot}=\Im_1+\Im_2+\Im_{rel}.\] Calculation of fission cross sections in the statistical models is based on the Hauser-Feshbach formalism. For a compound nucleus of spin $J$ that is populated with a partial fusion cross section of $\sigma_J$, the partial fission cross section is given in terms of the ratio of the fission decay width $\Gamma^{fis}_J$ to the total decay width for this spin $\Gamma^{tot}_J$, with \[\sigma^{fis}_J={\Gamma^{fis}_J\over \Gamma_J^{tot}} \sigma_J\; .\] The fusion partial cross section for formation of a compound nucleus of spin $J$ from projectile and target nuclei of spins $J_p$ and $J_t$, respectively, at center-of-mass energy $E_{c.m.}$ is given by \[\sigma_J=\pi\mathchar'26\mkern-10mu\lambda ^2{2J+1\over (2J_p+1)(2J_t+1)}\sum\limits^{J_p+J_t}_{S=|J_p-J_t|} \sum\limits^{J+S}_{\ell=|J-S|} T_\ell (E_{c.m.})\; ,\] with \[\sigma^{tot}_{fus}=\sum\limits^\infty_{J=0}\sigma_J.\] A simple and commonly used method of representing the fusion transmission coefficient is to take \[T_\ell(E_{c.m.}) = {1\over 1+exp \{[\ell -\ell_{cr}(E_{c.m.})]/\Delta\}} ,\] where $\ell_{cr}$ is the critical angular momentum for fusion and $\Delta$ is the diffuseness of the fusion $\ell$-distribution. The critical angular momentum for fusion $\ell_{cr}$ can either be obtained from fusion model calculations or by adjusting its value to achieve consistency with measured evaporation-residue cross sections. Most of the lighter system calculations have been done using $\Delta=1\hbar$. For the calculation of the total decay width $\Gamma_{tot}$, it is assumed that the deexcitation of the compound nucleus is through the emission of neutrons, protons, alpha particles, $\gamma$ rays and/or fission fragments. Then \[ \Gamma_{tot}=\Gamma_n + \Gamma_p +\Gamma_\alpha+ \Gamma_\gamma + \Gamma_{fis} .\] Both the saddle-point calculations of Sanders \cite{s91} and the recent scission-point calculations of Matsuse {\it et al.} \cite{mbnm97} use the code CASCADE \cite{fp77} for calculating the partial widths for the three light particles and $\gamma$ rays. Here, the partial width $\Gamma_x$ for particle $x (x = n, p,$ or $\alpha$) of spin $s_x$ to be emitted from the compound nucleus of excitation energy $E^*_{CN}$ and spin $J_{CN}$ to form an evaporation-residue nucleus ER of excitation energy $E^*_{ER}$ and spin $J_{ER}$ is given by \[ \Gamma_x=\int{\rho_{ER}(E^*_{ER} -\Delta_{eff},J_{ER})\over 2\pi \rho_{CN}(E^*_{CN}-\Delta_{eff}, J_{CN})} \sum\limits^{J_{ER}+S_x}_{S=|J_{ER}-S_x|} \sum\limits^{J_{CN}+S}_{\ell=|J_{CN}-S|} T^x_\ell (\varepsilon_x)d\varepsilon_x .\] The integral is over all kinetic energies of the emitted light particle $\varepsilon_x$, and $\rho_{CN}$ and $\rho_{ER}$ are the level densities of the compound nucleus and resulting evaporation residue, respectively. The transmission coefficients $T_\ell^x(\varepsilon_x)$ are obtained from optical-model calculations using average parameters. For the higher excitation energies involved in the fission process, it is expected that shell effects should have little influence on the level densities and hence an effective macroscopic energy ground state is used to calculate these densities \cite{fp77}. The parameter $\Delta_{eff}$ determines the zero point of the effective excitation energy (see Fig. 14), with \[\Delta_{eff}(MeV)=E_B(Z,A)-E^{macro}_B(Z,A).\] Here $E_B$ is the measured binding energy of the nucleus and $E^{macro}_B$ is the corresponding macroscopic energy \cite{s91,fp77}. The partial width for $\gamma$ decay assumes decay through the giant dipole resonance. \subsubsection{Saddle-point model} In lighter systems, the mass asymmetry dependence of the fission barrier favors the decay into mass asymmetric exit channels. This can be seen in a plot of the calculated saddle-point energies $V_{saddle}$ as a function of the fragment mass for the $^{56}$Ni compound system in Fig. 15. In this figure, $A_{fragment}=28$ corresponds to symmetric breakup into two $^{28}$Si fragments. Typical saddle-point shapes are shown for spin $\ell =0\hbar$ at two different mass asymmetries and for the symmetric barrier at $\ell = 36\hbar$. At low spins the very steep rise in the saddle-point barrier energy in going to more symmetric configurations leads to a strong favoring of light-fragment evaporation over heavy-fragment fission of the compound nucleus. It is only at higher spin values that the more symmetric breakup channels become able to compete in the compound-system decay. In developing a model for the fission decay of light systems, it therefore becomes important to consider the decay widths to specific exit channels, corresponding to different mass asymmetries at the saddle point, with the total width given by the sum over the widths to the individual channels: \[\Gamma_f = \sum\limits_{A_L}\sum\limits_{Z_L} \Gamma_f(Z_L, A_L).\] Here the partial widths are denoted by the charge $Z_L$ and mass $A_L$ of the lighter fragment. The corresponding mass asymmetry $\eta$ is given by $\eta=1-2(A_L/A_{CN})$. In the saddle-point model \cite{s91} the fission widths $\Gamma(Z_L,A_L)$ are obtained by considering the level density above the mass-asymmetry dependent saddle point, with \[\Gamma_f(Z_L,A_L) = \int\limits^{20 MeV}_{\epsilon=0} {\rho_f(u_J)\over 2\pi\rho_{CN}(E^*_{CN}-\Delta_{eff}, J_{CN})} T^f_{J_{CN}}(\varepsilon)d\varepsilon ,\] and where the transmission coefficients have a sharp cutoff form, with \[ \begin{array}{rll} T^f_{J_{CN}}&(\varepsilon)=& \\ &\left\{ \begin{array}{ll} 1\; {\rm for}\; \varepsilon \leq E^*_{CN}-V_{saddle}(J_{CN},\eta)-\Delta V_{shell}(J_{CN},Z_L, A_L)-\Delta_{eff} \\ 0\; {\rm for}\; \varepsilon > E^*_{CN}-V_{saddle}(J_{CN},\eta)-\Delta V_{shell}(J_{CN},Z_L, A_L)-\Delta_{eff} & \\ \end{array} \right . \end{array} .\] The integration is over the energy of radial motion $\varepsilon$ and is insensitive to the upper limit assuming that this limit is sufficiently large. $V_{saddle}(J_{CN},\eta)$ is the spin- and mass-asymmetry dependent saddle-point energy with respect to the macroscopic-energy ground state of the compound nucleus. The importance of the compound nucleus spin in determining the competition between light particle evaporation and fission to heavier fragments is illustrated in Fig. 16 where the partial cross section distribution for fusion (solid line) and fission (shaded region), as defined by binary decay to a channel where the lighter fragment has mass $A_L\geq 6$, are shown for the $^{32}$S+ $^{24}$Mg reaction at $E_{lab}$=121 MeV. In this system, fission is found to only compete at incident partial waves close to the critical angular momentum for fusion. Although shell effects are not expected to influence the level density of the equilibrated compound nucleus, at the much ``colder" saddle-point configuration these effects can be important. This is evidenced by a strong isotopic dependence for the fission cross sections that is inconsistent with a smooth dependence of the potential-energy surface on the mass-asymmetry parameter $\eta$. As a first approximation of the shell corrections at the saddle point, a term $\Delta V_{shell}(Z_L,A_L)$ has been added to the barrier energies based on the sum of the Wigner energy corrections \cite{mn81} for the two nascent fragments: \[\Delta V_{shell}(Z_L,A_L)= W(Z,A_L) + W(Z_{CN}-Z_L, A_{CN} -A_L)\] with \[ W(Z,A)=(36 {\rm MeV}) \left[ |{A- 2Z\over A}| + \left\{ {1/A, Z\; {\rm and}\; N\; {\rm odd\; and\; equal}\atop 0,\; {\rm otherwise}} \right] \right\} \] This correction to the potential energy surface has the effect of enhancing the fission cross section to channels where both fragments have N=Z. To calculate the level densities of the compound nucleus and saddle-point configurations, a Fermi-gas formula \cite{bm81} is used, with \[ \rho(u,J)={2J+1\over 12}\sqrt{a_x} \left[{\hbar^2\over 2\Im}\right]^{3/2} u^{-2}\exp (2\sqrt{a_x u})\] and \[ u= \left\{ \begin{array}{ll} E^*_{ER} - {\hbar^2\over 2\Im}J (J+1)-\Delta, & {\rm ER} \\ E^*_{CN}-V(J_{CN},\eta)- \Delta V_{shell}(Z_L, A_L)-\Delta_{eff} - \varepsilon, & {\rm saddle\; point} \end{array} \right . \] For the evaporation residues (ER), the level-density parameter $a_x =a_n$ and the spin $J$ and energy offset $\Delta$ correspond to the evaporation residue. The saddle-point densities are calculated with $a_x=a_f$ and $J=J_{CN}$. Most of the calculations done for light systems have used $a_n=A_{ER}/(8 {\rm MeV})$ and $a_f=A_{CN}/(8 {\rm MeV})$. In light nuclear systems, the fission breakup of the compound nucleus is often to final fragments where the density of states is low. This can result in considerable structure in the excitation spectra of the fission fragments. To explore the population of states in the final fragments while retaining the saddle-point as the ``transition" state, a procedure has been developed \cite{shp94} to calculate the population of specific mutual excitations assuming a stochastic process. Within the transition-state method, population of a given saddle-point level already corresponds to a commitment to fission into a particular mass partition. The partial cross section for the population of the compound nucleus with spin $J$ that subsequently undergoes fission to mass asymmetry $\eta$ is taken as $\sigma_{FF}(J,\eta)$ and is calculated using the transition-state model based on the saddle-point phase space. The cross section for populating a specific mutual excitation ($\beta_1,\beta_2)$ is then given by \[ \sigma(\beta_1,\beta_2)= \sum\limits_J\sigma_{FF}(J,\eta) { \sum\limits_{\ell_{out}} [\beta_1\times \beta_2]_{J,\ell_{out}} P(\eta,J, \varepsilon) \over \sum\limits_{\lambda_1,\lambda_2,\ell_{out}} [\lambda_1\times \lambda_2]_{J,\ell_{out}} P(\eta, J, \varepsilon)} ,\] where $[\lambda_1\times \lambda_2]_{J,\ell_{out}}$ represents the sum of the possible spin couplings between the two fragments in states $\lambda_1$ and $\lambda_2$ with orbital angular momentum $\ell_{out}$ and coupling to compound-nucleus spin J, and $P(\eta,J,\varepsilon )$ is the probability of the compound nucleus of spin $J$ to fission with mass asymmetry $\eta$ and radial kinetic energy $\varepsilon$. This probability depends implicitly on $\ell_{out}$ through $\varepsilon$. The radial kinetic energy $\varepsilon$ can be expressed in terms of the characteristic energies of the reaction with \[ \varepsilon = E_{c.m.} + Q_o-V_{rel}(\ell_{out}, \eta)+\delta - E_x.\] The significance of each of these energies is shown schematically in Fig. 14. Here, $E_{c.m.}$ is the center-of-mass energy in the entrance channel, $Q_o$ is the ground-state $Q$ value, $V_{rel}(\ell_{out}, \eta)$ is the relative energy of the two spheroids that comprise the saddle-point shape, $\delta$ is the energy loss that occurs in moving from the saddle to scission configurations, and $E_x$ is the mutual excitation of the final fragments. In light systems $\delta$ is expected to be small. Figure 17 from ref. \cite{shp94} compares the results of this calculation to the observed excitation spectra for the $^{24}$Mg($^{32}$S,$^{28}$Si) $^{28}$Si reaction at $E_{c.m.}=51.0$ and 54.5 MeV. The bold-line histograms show the experimental results. The dotted line histograms show the predicted spectra using all known levels in $^{28}$Si up to the 14.339 MeV excitation. The thin-line histograms are the predicted spectra for only the particle-bound levels. Since the experimental results were obtained using a kinematic coincidence technique that discriminates against excitations where one or both of the populated states subsequently emits a light particle, the thin-line histograms are expected to more faithfully represent the experimental situation. The observed structure is very well reproduced by the calculations. The structure that is observed at higher excitation energies can be attributed to groupings of mutual excitations with high channel spins. The essential validity of the predicted population pattern was confirmed by a measurement of the $\gamma$-rays in coincidence with the fission fragments for this system \cite{shp94}, as will be discussed in Sec. \ref{sec:exp44to56}. One of the interesting features of the energy spectrum calculation is that it suggests that higher excitation energies correspond to smaller compound nucleus spins. The calculation also suggests significant alignment of the spin and orbital angular momenta at higher excitations, with the possibility of anti-aligned configurations being prevalent at lower excitation energies. This is shown in Fig. 18 for the calculation shown in Fig. 17. The average values of the compound nucleus spin, orbital angular momentum, and channel spin is shown for each mutual excitation of the $^{28}$Si+$^{28}$Si fission channel. \subsubsection {Scission-point model} In the Extended Hauser-Feshbach, scission-point model, the partial fission decay width is determined by the product of the level densities in the final nuclei, with \begin{eqnarray*} \Gamma_f(Z_L,A_L) & = & \sum\limits_{(I_L,I_H)I}\; \sum\limits_{(L,I)J}\int\int\int {\rho_{I_L}(E^*_L)\rho_{I_H}(E^*_H)\over 2\pi\rho_{CN}(E^*_{CN},J_{CN})} \\ & & \times \delta (E^*_L + E_H^* + \varepsilon + Q - E^*_{CN})T_L (\varepsilon)d E^*_L d E^*_H d\varepsilon \end{eqnarray*} and where $E^*_L$ and $E_H^*$ are the excitation energies in the lighter and heavier fragments, respectively, $I_L$ and $I_H$ are the spins of the fragments, $\varepsilon$ is the relative energy of the fragments in the exit channel, and $Q$ is the $Q$-value for the binary breakup. The delta function assures energy conservation. For low-lying excitations of the fragments, the integrals are replaced by summations over the discrete energy levels. In the development of the Extended Hauser-Feshbach model, various assumptions have been made for the level densities and transmission coefficients needed to evaluate the above expression. In what is perhaps the most systematic study of this model as it applies to light systems, the level density expression is taken as given above for the saddle-point model and with level density parameters given by $a=A/(8 {\rm MeV})$ \cite{mbnm97}. The transmission coefficients for this study are evaluated by using the simplified formula \[ T_L(E)={1\over 1+\exp \left\{ \left[ V(L)-E\right] /\Delta_s \right\} },\] where the diffuseness parameter $\Delta_s$ is typically taken to equal 0.5 MeV. It has further been found possible to obtain good agreement with experiment using a simple parametric expression of the barrier height $V(L)$ at the scission point, with \[ V(L) = V_{Coul}+{\hbar^2 \over 2\mu_f R^2_S}\ell (\ell +1) ,\] where $\mu_f$ is the reduced mass of the decaying complex fragments. The scission point $R_S$ is estimated by using the radii $R_L=r_s A^{1/3}_L$ and $R_H=r_s A^{1/3}_H$ of two spherical fragments of mass number $A_L$ and $A_H$ separated by a distance d, with \[R_S=R_L+R_H +d.\] The ``neck" parameter d for the reactions covered by this report is found to vary from 2.5 to 3.5 fm using $r_s$ = 1.2 fm. The Coulomb energy is calculated using the expression \[ V_{Coul}=Z_LZ_He^2/R_S ,\] where $Z_L$ and $Z_H$ are the atomic numbers of the lighter and heavier fragments, respectively. In general, it has been found possible to reproduce the experimental results by varying the neck parameter $d$ in a systematic manner, leading to predicted fission cross sections comparable to those obtained with the saddle-point model. With this adjustment, the moments of inertia for the two calculations are also similar, supporting the assertion that, in lighter nuclear systems, the scission- and saddle -point configurations are similar. Mass distribution calculations based on the saddle-point \cite{s91} and scission-point models \cite{vkw85,bdhf93}, summing the fission and evaporation-residue components for a given mass channel, are compared to experimental cross sections obtained for the $^{35}$Cl+$^{12}$C reaction at $E_{lab} = 200$ MeV and for the $^{23}$Na+$^{24}$Mg reaction at $E_{lab}$ = 89.1 MeV \cite{bdhf92,bdhf93} in Fig. 19. These two reactions reach the common $^{47}$V compound systems at approximately the same excitation energy of $E^*_{CN}$= 64 MeV. The observed and calculated values for the average total kinetic energies for the two reactions are shown in Fig. 20, although here the extended Hauser-Feshbach results for the $^{23}$Na+ $^{24}$Mg reaction are not available. Both calculations are found to give comparably good agreement with the experimental results. In both cases, the overall fission+evaporation-residue cross sections observed experimentally were used to determine the $\ell_{cr}$ values for the two systems. Otherwise, in the saddle-point calculations, all of the remaining parameters are set by the general systematics of fission in light systems. In the case of the extended Hauser-Feshbach calculations, the neck separation parameter $d$ was adjusted to optimize the fit to the fission data. However, as previously indicated, this parameter is found to vary uniformly as a function of the mass of the compound system. The saddle-point calculations have been extended to even lighter systems, most notably by the S\~ao Paulo group exploring nuclei in the mass A=28 region \cite{aacf94}, using the double-spheroid model to extrapolate saddle-point energies to these systems. Again, good agreement is achieved between the calculations and experimental results. \subsection{Equilibrium orbiting model} \label{sec:eqorb} The binary reaction yields for a number of light systems, including those for the $^{24}$Mg+ $^{12}$C \cite{gdbh90}, $^{28}$Si+ $^{12}$C \cite{s82,ssf84}, $^{24}$Mg+ $^{16}$O \cite{kbg92}, and $^{28}$Si+ $^{14}$N \cite{sah87,ssa88} reactions, have been interpreted in terms of the formation and decay of long-lived, rotating, dinucleus complexes. A quantitative model for this behavior has been developed by Shivakumar, Ayik and collaborators \cite{ssa88,ass88} based on nucleon transport theory. However, because of the long lifetime of the orbiting complex, as indicated by the angular distribution data, the equilibrium limit of the more general transport model is used in comparisons with data. In this limit, the model is somewhat similar to the saddle-point model of fission, although the ``equilibrium orbiting" model provides a more unified description of the fusion and deep-inelastic orbiting processes. At energies near the Coulomb barrier, the collision of two heavy ions is governed by a barrier potential comprised of nuclear, Coulomb and centrifugal terms. The equilibrium orbiting model is based on the observation that at low energies, and for all partial-waves up to some maximum value $\ell_{pocket}$, the interaction potential exhibits a pocket as a function of the distance coordinates. This pocket allows for trapping of the incident particles and the subsequent fusion or orbiting behavior of the composite system. For even higher angular momenta with $\ell_{pocket}\leq \ell \leq \ell_{max}$, trapping still occurs as frictional forces can reduce the relative energy and angular momentum to values of $\ell\leq \ell_{pocket}$. The value of the maximum angular momentum $\ell_{max}$ is set by when the centrifugal energy reaches a value that prevents the incident particles from closing to within a critical distance where the nuclear surfaces overlap. The trapped, dinucleus complex can either evolve with complete amalgamation into a fully equilibrated compound nucleus or, alternatively, escape into a binary exit channel by way of orbiting trajectories. Orbiting can therefore be described in terms of the formation of a long-lived dinucleus complex which acts as a ``doorway" state to fusion with a strong memory of the entrance channel. The observed orbiting yields correspond to the fragmentation of the dinucleus complex during the early stages of the interaction, while long-lived complexes relax towards the mono-nuclear shape of the compound nucleus. A full description of the exchange of mass and charge between the two interacting ions requires solving coupled transport equations describing the mass flow. The equilibrium model \cite{ass88} is an approximation to the full transport theory that assumes the probability of fragmentation into a channel $\chi\equiv (N,Z)$ and angular momentum $\ell$ can be expressed as the product of two terms. The first term is a distribution function that is based on the density of states in $\chi$ available to the incident channel, normalized with respect to the density of states for all channels, including the fusion channel. The second term is a transitional probability that is assumed to be slowly varying with energy and is thus taken as a constant. The defining potential energy surface for the dinucleus complex is written as \[ \begin{array}{rll} U_J(N,Z;R) = &V_n(N,Z;R)+V_C(N,Z;R)& \\ &+ {\displaystyle {\hbar^2\over 2{\Im}_{tot}(N,Z;R)} }J(J+1)+Q(N,Z),& \end{array}\] where $V_n$ is the nuclear interaction (taken as the empirical proximity potential of Bass\cite{rb77}), $V_C$ is the Coulomb interaction, $\Im_{tot}(N,Z;R)$ is the total system moment of inertia in the sticking limit, $J$ is the total angular momentum, and $Q(N,Z)$ is the ground state $Q$-value of channel $(N,Z)$ with respect to the entrance channel. For long-lived systems the decay distribution function $P_\ell(N,Z)$ is obtained from the ratio of the density of states evaluated at the saddle point (conditional saddle) of the potential, for a given $\ell$-value and exit channel, to the sum over all open channels of the state densities at the respective saddle points and the corresponding state densities evaluated at the configurations corresponding to the minimum of the $\ell$-dependent potential-energy for each open channel. The total binary fragmentation probability is given by summing over the open channels, with \[P_\ell=\sum\limits_{N,Z}P_\ell (N,Z)\] and the corresponding fusion probability is then (1-$P_\ell$). It is therefore possible to evaluate both the binary fragment and fusion probabilities within a consistent formalism. The total production of a fragment (N,Z) is given by the sum of all $\ell$-partial wave contributions up to the maximum angular momentum $\ell_{max}$ for which the system can become trapped in a pocket of the interaction potential: \[\sigma_\ell (N,Z)={\pi\over k^2}\sum\limits^{\ell_{max}}_{\ell = 0}(2\ell +1)P_\ell (N,Z)\] A corresponding expression gives the fusion cross section: \[\sigma_{fus}={\pi\over k^2}\sum\limits^{\ell_{max}}_{\ell = 0}(2\ell +1)(1-P_\ell)\] The final total kinetic energy is determined by the sum of the nuclear, Coulomb, and rotational energies at the conditional saddle point, with: \[E^{tot}_{K,\ell}(N,Z)=U_o(N,Z;R_S)+ {\hbar^2\ell (\ell +1)f^2\over 2 \Im_{rel}(N,Z;R_S)}-Q(N,Z) ,\] where $f = \Im_{rel}(N,Z;R_S) / \Im_{tot}(N,Z;R_S)$ and $\Im_{rel}$ and $\Im_{tot}$ are the relative and total moments of inertia at the saddle point configuration with radius $R_s$, respectively. The nuclear and Coulomb contributions to the potential energy $U_o(N,Z;R_s)$ are evaluated at $\ell=0$. The average kinetic energy in the exit-channel, taking into account the channel cross section, is found using: \[<TKE(N,Z)>={\pi\over k^2}\sum\limits^{\ell_{max}}_{\ell=0} (2\ell +1)E^{tot}_{K,\ell} {P_\ell(N,Z)\over P(N,Z)} ,\] with \[ P(N,Z)={\pi\over k^2}\sum\limits^{\ell_{max}}_{\ell=0} (2\ell+1)P_\ell(N,Z) .\] The equilibrium orbiting model has been used to successfully explain both the observed cross sections and TKE values of the fully damped fragments for several lighter nuclear systems. It has been found that the adjustment of a single strength parameter in the Bass parameterization \cite{rb77} of the nuclear potential leads to good agreement for both orbiting and total fusion yields. This is shown in Figs. 21 and 22 \cite{sah87} for the $^{28}$Si+$^{12}$C reaction \cite{s82,ssf84}. In Fig. 21 the measured total kinetic energies for the carbon, nitrogen, and oxygen channels are compared to the experimental results. Figure 22 compares the measured and calculated cross sections for the same three channels as well as for the total evaporation-residue cross section. In both of these figures the orbiting calculations are indicated by the dotted curves. For comparison, the results of the transition-state, saddle-point calculation discussed earlier are also shown in Figs. 21 and 22 by the solid curves. The parameters used for the transition-state calculations are the same as those used to successfully describe the experimental cross sections in other light systems. For the $^{28}$Si+ $^{12}$C reaction, however, the transition-state calculations tend to underestimate the observed binary yields. This is seen in the figures where the total fusion cross section has been adjusted to achieve the best general agreement of the total kinetic energies and binary reaction cross sections with experiment. This leads, however, to fission-model predictions of evaporation residue cross sections which are high compared to the experimental results and predicted carbon cross sections that are too small. In other reactions of very light systems, such as for the $^{28}$Si+ $^{14}$N reaction, the equilibrium orbiting and transition-state models result in comparable agreement with the data. For heavier systems ($A_{CN}\geq 47$), the simple form of the equilibrium orbiting model is found to be less successful \cite{bdhf92,setc89}. This may reflect, however, some of the simplifying assumptions of the calculations, such as the calculation of moments of inertia based on touching, spherical fragments and the use of ground-state $Q$-values for the driving potential energy surface. To achieve a satisfactory agreement for the $^{32}$S+ $^{24}$Mg reaction within the equilibrium orbiting model \cite{setc89}, for example, it has been found necessary to introduce deformation effects by assuming a moment of inertia more similar to that of the nuclear saddle point. In this case the equilibrium orbiting and scission-point models employ very similar phase-space arguments. The shortcomings of the equilibrium model for orbiting does not imply that the presence of an orbiting mechanism, as distinct from fission, can be ruled out in some systems. The alternative statistical models also fail to explain several important features of the orbiting reaction data, such as the strong entrance channel dependence in the $^{40}$Ca system \cite{kbg92} and spin distributions and alignments measurements for the $^{28}$Si+ $^{12}$C reaction \cite{rsh91}. \subsection{NOC} The relationships among the different reaction mechanisms such as fusion-fission, orbiting, and heavy-ion resonances are still not fully understood. It has been suggested \cite{rrb84}, for example, that the heavy-ion resonance phenomenon observed in some light nuclear systems may reflect a strong shell correction to the fission potential energy surface and, as such, be related to the statistical fission mechanism. In the absence of a fully consistent picture relating the different reaction mechanisms, considerable effort has gone into developing the systematics of these processes. In general, the fully energy-damped yields are observed with a cross section comparable to the predictions of the transition-state calculations. The occurrence of a distinctly different orbiting or resonance component to the reaction cross section seems to be correlated with the number of open reaction channels (NOC). In lighter systems, for example, the even-even C+C, C+O, and O+O reactions all show strong resonance behavior and also have small NOC values. Underlying this correlation may be a relationship between the NOC for a given reaction and the ``surface transparency" of the reaction \cite{ha81}. The NOC calculations have been systematically extended to heavier systems \cite{baa94} and a strong correlation found between the anomalous large-angle resonant structure and small NOC values. For example, systems populating the $^{40}$Ca compound nucleus, for which there is significant evidence of an orbiting process \cite{kbg92}, have low values for the NOC corresponding to the grazing partial waves. Alternatively, the energy-damped yields for several systems populating the $^{47}$V system \cite{bdhd89,bdhf92,bdhf93}, where the NOC values are large, are consistent with a fusion followed by fission picture. The NOC for a given system is obtained by a triple summation over all possible two-body mass partitions in the exit channels, over all possible angular momentum couplings and, finally, over the allowed excitations of the two fragments: \[N^J(E_{c.m.})=\sum\limits_{A_1+A_2=A_{cn}}\; \sum\limits_{J = L+I_1+I_2}\; \sum\limits_{E_r=E^*_{CN}-E_1-E_2-Q_{12}}T_L(E_r) .\] In this expression, $E_{c.m.}$ and $E_{CN}^*$ are the center-of-mass energies of the incident particle and the excitation energy of the compound system, respectively. $A_1, A_2, A_{CN}$ are the mass numbers of the outgoing fragments and the compound system. $I_1, I_2$ and $L$ are the intrinsic spins of the fragments and the orbital angular momentum of their relative motion. $Q_{12}$ is the reaction $Q$-value of the decay into the fragments. $E_1, E_2$ and $E_r$ are the intrinsic excitation energies of the fragments and the energy available to their relative motion, respectively. $T_L(E_r)$ is the transmission coefficient of the outgoing channel as a function of angular momentum and relative energy. The transmission coefficients have been calculated using the semi-classical, parabolic barrier penetration approximation: \[T_L(E_r)=1/\left[1+\exp\left[{2\pi(E_L-E_r)\over \hbar\omega_L}\right]\right]\] where $E_L=V_L(R_B)$. The rotational frequency, \[\hbar\omega_L=\hbar\sqrt{ \left[ ({d^2V_L(R)\over dR^2})|_{R=R_B}\right] / \mu } ,\] is related to the curvature of the outer barrier. In this expression, $\mu$ is the reduced mass and $V_L(R)$ is the sum of the Coulomb, centrifugal and nuclear potentials. In more recent calculations a macroscopic proximity form has been used for the nuclear interaction \cite{baa94}, rather than the Saxon-Woods form employed in earlier calculations. The sum over the energy sharing between the two fragments employs discrete energy levels of the fragments at lower energies, where these are known, and an angular momentum dependent level density expression at higher energies \cite{baa94}. The sensitivity of the NOC calculations to the choice of level density expression and to the transmission coefficients is discussed in Ref. \cite{baa94}. Although the quantitative results are sensitive to these choices, the qualitative behavior is not, thus preserving the predictive value of the calculations. The expression for $N^J$ is similar to the denominator appearing in the Hauser-Feshbach formalism \cite{hf52} for the compound nucleus. However, the NOC calculations include direct reaction channels in addition to the evaporation channels. This allows the phase space calculation to be extended to values of the incident angular momentum greater than the critical angular momentum for fusion $L_{gr}$. In order to compare different systems, it is useful to normalize $N^J$ by the corresponding incident flux. The quantity N/F is defined, with \[N/F=N^J(E_{c.m.})/F^J(E_{c.m.})\; .\] Here, $F^J(E_{c.m.})$ is the incident flux for the total angular momentum J, as given by \[F^J(E_{c.m.})={\pi\over k^2}\sum\limits_{J=L+I_1+I_2}g_J T_L(E_{c.m.}) ,\] where $E_{c.m.}=\hbar^2k^2/2\mu,$ $g_J=(2J+1)/\{(2I_1+1)(2I_2+1)\},$ and $I_1$ and $I_2$ are the intrinsic spins of the incident particles. Examples of the NOC calculations for a number of lighter systems is given in Fig. 23. The N/F values for the different systems are shown as a function of $L_{gr}$. Each of the curves shows a characteristic minimum at an $L_{gr}$ value that corresponds to an energy well above the corresponding Coulomb barrier. The initial drop in N/F as a function of $L_{gr}$ is due to the increasing difficulty as $L_{gr}$ increases of dissipating the angular momentum brought into the compound system solely through the evaporation of light particles. The subsequent rise in N/F occurs when an increasing number of direct channels (such as single and mutual inelastic excitation, nucleon and $\alpha$-transfer and, finally, deep-inelastic orbiting and fusion-fission processes) become accessible. These channels are activated at somewhat higher energies because of their reaction $Q$-values. A strong correlation has been demonstrated between systems where N/F is small and the occurrence of quasi-molecular resonances in light- and medium-light heavy-ion reactions \cite{baa94}. For example, it is observed from Fig. 23 that the $^{12}$C+ $^{12}$C and $^{12}$C+ $^{16}$O systems, where prominent resonant behavior \cite{baa94} has been observed, both show very low minimum values of N/F $(N/F\approx 10^{-1}$). On the other hand, for systems such as B+O, where the NOC values are much larger $(N/F>10^4)$, there is no compelling evidence for a strongly energy-damped reaction component other than what can be accounted for through the fusion-fission mechanism \cite{aacf94}. Typical NOC calculations for a number of heavier systems with $36\leq A_{CN}\leq 48$ are shown in Fig. 24. These systems can be classified into two groups, with the $^{24}$Mg+$^{24}$Mg reaction having an intermediate behavior. The $^{24}$Mg+ $^{12}$C, $^{24}$Mg+$^{16}$O and $^{28}$Si+ $^{12}$C systems, which are composed of $\alpha$-particle-like nuclei, belong to the first group and have NOC minima corresponding to $N/F\approx 10$. For these systems, strongly oscillatory angular distributions are observed in the backward-angle elastic scattering cross sections \cite{baa94} and there is evidence of a non-statistical orbiting mechanism \cite{s82,gdbh90,rk85}. The second group of reactions, with large values of N/F, show little evidence of an excess yield beyond that expected from the statistical fission mechanism. The $^{24}$Mg+$^{24}$Mg system appears to be much more surface-transparent at large $L_{gr}$ in comparison to $^{23}$Na+ $^{24}$Mg, for instance, but never achieves the low N/F values of the reactions populating the $^{40}$Ca compound system. One distinguishing feature of the reaction is that the ``quasi-molecular resonance window", i.e., the region of relatively low N/F values, corresponds to quite high values of $L_{gr}$. Very narrow, high-spin resonances have been observed in elastic and inelastic scattering measurements for this system \cite{zkbsh83,qzkp90} with strong correlation among the various inelastic channels. These strong, heavy-ion resonance features seem to coexist with ``normal", statistical fusion-fission yields from the $^{48}$Cr compound nucleus, as discussed in Ref. \cite{hsfp94}. The resonances, however, correspond to compound nucleus spin values \cite{qzkp90} that approach the N/F minimum, but are large compared to dominant spins expected to lead to fusion-fission yields. This circumstance of having a small number of open channels for the grazing partial waves and, in addition, having the angular momentum characterizing these partial waves being greater than the dominant angular momenta for the statistical fission competition may contribute to the strength of the observed resonance structure. \subsection{Dynamical model of intermediate mass fragment emission} Fission in light systems can be considered as a dynamical process consisting of the gradual shape change of the compound system with the formation of a neck which subsequently narrows and breaks. It is therefore tempting to include dynamical effects into the statistical model calculations to have a better understanding of the fission process and to provide some information on the time scales which are involved during the reaction. The statistical model developed by Dhara et al. \cite{dbbk93} is based on the transition-state picture and involves solving the classical equations of motion to follow the change of shape of the composite system. The fission dynamics are therefore explicitly considered while evaluating the relevant physical observables. In the absence of any precise knowledge of the energy sharing between the intrinsic excitation and collective degrees of freedom, it is assumed that a random fraction of the initial compound-nucleus excitation energy goes into collective degrees of freedom to generate the fission dynamics. A standard proximity potential is used, with non-conservative frictional forces introduced by a viscosity term whose magnitude has a temporal dependence. The time evolution of this viscosity term has a strong impact on the dynamics of the fission process. The fission probability is calculated by Monte Carlo simulation for a large number of trajectories. The spin transferred to the binary fragments are computed in the sticking limit. The model still uses phase space arguments to determine the fusion-fission cross sections, however, in a manner that is quite similar to the transition-state model approach \cite{s91,lgm75}. The two models differ primarily in their parametric expressions describing the nuclear shapes and the saddle-point energies. Calculations using the dynamical model have been shown to reproduce reasonably well the experimental results obtained for the $^{31}$P compound system \cite{bbb91,bbm95,3b96}. \subsection{Generalized liquid drop model} \label{sec:ternary} Within the macroscopic energy models, the fission saddle-point is found by studying the potential energy surface of the compound system as a function of a specific shape parameterization. One such parameterization has been developed by Royer and Remaud \cite{rr84} involving a class of shapes that evolves from a spherical mononucleus to two touching spherical fragments with a deeply creviced neck. A nuclear proximity potential is used to account for the surface interaction of the neck configuration. An interesting aspect of this parameterization is that it is easily generalized to a three fragment configuration, allowing the study of a possible ternary fission valley \cite{gr95}. The possibility of a ternary fission path has been explored through model calculations for the $^{48}$Cr compound system. In this study it is found that a prolate aligned $^{16}$O+ $^{16}$O+ $^{16}$O molecule configuration results in a minimum of the potential energy as a function of deformation for high spins. Such behavior could result in an enhanced cross section for ternary fission to the three $^{16}$O final channel. A shell stabilized, three $^{16}$O chain configuration of $^{48}$Cr has also been suggested by unconstrained $\alpha$-cluster model calculations by Rae and Merchant \cite{rm92}. An experimental search for the three $^{16}$O breakup channel of $^{48}$Cr, as populated through the $^{24}$Mg+ $^{24}$Mg reaction, has failed to find evidence for this behavior, however \cite{mmetc96}. \section{System-by-system Discussion} The compound nuclei considered in this report range in mass from $19\leq A_{CN}\leq 80$. The low mass limit is set by the experimental difficulty of unfolding the fully energy damped yields from the more peripheral reaction yields in even lighter systems. The high end of the mass range corresponds to the transition region where, for the partial waves that lead to fission in heavy-ion reactions, the potential energy surface becomes relatively flat as a function of mass asymmetry. For systems with mass lower than $A_{CN} \approx$ 80, the potential energy surface favors fission of the compound system into fragments of unequal mass. For heavier systems, the symmetric mass fission of the compound nucleus is favored in the absence of strong shell effects. Tables 1 and 2 summarize the experiments that have been performed in lighter systems where fully energy-damped yields have been studied. In the following discussion, a somewhat arbitrary division is made between systems with $A_{CN} < 32$, $32 \leq A_{CN}\leq 44$, $44<A_{CN}\leq 56$, and $56<A_{CN}\leq 80$. The systems with $A_{CN} < 32$ are the most difficult ones in which to experimentally establish the fission cross sections. They also create the greatest challenge to the fission-model calculations because of the difficulty of calculating fission barriers for these very light systems and the increasing influence of shell corrections on these barriers. The mass range $32 \leq A_{CN} \leq 44$ encompasses the systems for which there is the greatest evidence of a dinucleus orbiting mechanism that is distinctly different from the expectations of the fission model calculations. In the mass region $44<A_{CN}\leq 56$, there is good overall agreement of the expectations of fission transition state model with the observed fully energy damped yields, although an additional heavy-ion resonance behavior is observed for several systems in this mass range. A shift to symmetric mass fission starts to become evident towards the higher end of the mass region with $56<A_{CN}\leq 80$. \begin{table} \caption{Systems studied with $A_{CN}\leq 40.$} \begin{tabular}{|lccccccc|} \hline $A_{CN}$& Reaction$^a$& $E_{lab}$& Detected& Tech-& C or& $NOC$ & Ref. \\ && (MeV) &Fragments&nique&S$^b$&(min.)&\\ \hline $^{19}$F&$^9$Be+$^{10}$B&10-40&$5\leq Z\leq 9$& $\Delta$E-E&S& $1.4\times 10^{-3}$ &\cite{faa93}\\ $^{20}$F&$^{9}$Be+$^{11}$B&10-40&$5\leq Z\leq 9$& $\Delta$E-E&S& $1.6\times 10^{-2}$&\cite{faa93}\\ $^{20}$Ne&$^{10}$B+$^{10}$B&15-50&$5\leq Z \leq 9$& $\Delta$E-E&S& $7\times 10^{-3}$&\cite{caaf91}\\ $^{21}$Ne&$^{10}$B+$^{11}$B&15-50&$5\leq Z\leq 9$& $\Delta$E-E&S& $2\times 10^{-3}$&\cite{caaf91}\\ $^{22}$Ne&$^{11}$B+$^{11}$B&15-50&$5\leq Z\leq 9$& $\Delta$E-E&S &$2\times 10^{-3}$&\cite{caaf91}\\ $^{26}$Al&$^{16}$O+$^{10}$B&22-64& $5\leq Z\leq 9$& $\Delta$E-E&S&$2\times 10^1$&\cite{aacf94}\\ $^{27}$Al&$^{16}$O+$^{11}$B&22-64&$5\leq Z\leq 9$& $\Delta$E-E&S &$5\times 10^0$&\cite{aacf94}\\ &$^{17}$O+$^{10}$B&22-64&$5\leq Z\leq 9$& $\Delta$E-E&S&$1\times10^2$&\cite{aacf94}\\ $^{28}$Al&$^{17}O+^{11}B$&22-64&$5\leq Z\leq 9$& $\Delta$E-E&S&$9\times 10^{1}$&\cite{aacf94}\\ &$^{18}$O+$^{10}$B&22-63&$5\leq Z\leq 9$&$\Delta$E-E &S&$5\times 10^2$&\cite{aacf94}\\ &$^{19}$F+$^9$Be&56&$5\leq Z\leq 9$&$\Delta$E-E& S,C&$3\times 10^{-3}$&\cite{aacf94}\\ $^{29}$Al&$^{18}$O+$^{11}$B&22-64&$5\leq Z\leq 9$& $\Delta$E-E&S,C&$3\times 10^1$&\cite{aacf94}\\ $^{31}$P&$^{12}$C+$^{19}$F&96&$3\leq Z\leq 11$& $\Delta$E-E&S&$1\times 10^2$&\cite{3b96}\\ &$^7Li+^{nat}Mg$&47&$4\leq Z\leq 9$&$\Delta$E-E&S &$7\times 10^3$&\cite{bbm95}\\ &$a+^{27}$Al&60&$4\leq Z\leq 8$ &$\Delta$E-E& S& $<$100&\cite{bbb91}\\ $^{32}$S&$^{20}Ne+^{12}C$&50-80&Z=6&$\Delta$E-E &S&$7\times 10^0$&\cite{sfgsd79}\\ $^{36}$Ar&$^{20}$Ne+$^{16}$O&70-160&$6\leq Z\leq 9$ &$\Delta$E-E&S&$1\times 10^1$&\cite{sdc83}\\ &$^{24}$Mg+$^{12}$C&90-126&$5\leq Z\leq 9$& $\Delta$E-E&S&$1\times 10^1$&\cite{gdbh90}\\ $^{40}$Ca& $^{24}$Mg+$^{16}$O&75-115&$6\leq Z\leq 8$ &$\Delta$E-E&S&$9\times 10^0$&\cite{rk85,kbg92}\\ &$^{28}$Si+$^{12}$C&99-180&$5\leq Z\leq 8$& $\Delta$E-E&S&$2\times 10^1$&\cite{s82,ssf84}\\ &$^{20}$Ne+$^{20}$Ne&70-160&$8\leq Z\leq 10$& $\Delta$E-E&C&$7\times 10^1$ &\cite{sdc83}\\ \hline \end{tabular} $^a$ Projectile + Target \\ $^b$ C--Coincidence, S--Singles \end{table} \normalsize \begin{table} \caption{Systems studied with $A_{CN}>40$.} \begin{tabular}{|lccccccc|} \hline $A_{CN}$& Reaction$^a$& $E_{lab}$& Detected& Tech-& C or& $NOC$ & Ref. \\ && (MeV) &Fragments&nique&S$^b$&(min.)&\\ \hline $^{42}$Sc&$^{28}$Si+$^{14}$N &100-170&$3\leq Z\leq 10$&$\Delta$E-E& S&$8\times 10^3$&\cite{sss86}\\ $^{43}$Sc&$^{27}$Al+$^{16}$O&120-160 &$6\leq Z\leq 9$&$\Delta$E-E&S &$7\times 10^2$&\cite{sfg80}\\ $^{44}$Ti&$^{24}$Mg+$^{20}$Ne&71-120& $16\leq A\leq 32$&kin$^\dagger$ &C&$3\times 10^2$&\cite{bzb95}\\ &$^{28}$Si+$^{16}$O&61,79& $6\leq Z\leq 8$&$\Delta$E-E&S& $1\times 10^1$&\cite{oliv96}\\ &$^{32}$S+$^{12}$C&80-124& $6\leq Z\leq 8$&$\Delta$E-E&S&$6\times 10^2$&\cite{ocd79}\\ $^{46}$V&$^6$Li+$^{40}$Ca&153&$9\leq Z\leq 11$& $\Delta$E-E&S&$>$106&\cite{gmp84}\\ $^{47}$V&$^{23}$Na+$^{24}$Mg&89.1&$5\leq Z\leq 10$& $\Delta$E-E&S&$4\times 10^4$&\cite{rsc91}\\ &$^{31}$P+$^{16}$O&135.6&$6\leq Z\leq 8$& $\Delta$E-E&S&$1\times 10^4$&\cite{bdhf93}\\ &$^{35}$Cl+$^{12}$C&180-280&$5\leq Z\leq 12$& $\Delta$E-E,&S,C &$7\times 10^4$&\cite{bdhf92,bdhd89},\\ & & & &kin & & &\cite{alot96}\\ $^{48}$Cr&$^{24}$Mg+$^{24}$Mg&88.8&$6<A<24$& kin&C&$5\times 10^2$&\cite{hsfp94}\\ &$^{28}$Si+$^{20}$Ne&87-128&$16\leq A\leq 32$ &kin$^\dagger$&C&$6\times 10^2$&\cite{bzb95}\\ &$^{20}$Ne+$^{28}$Si&78.0&$20\leq A\leq 24$& kin&C&$6\times 10^2$&\cite{fsd96,cnrs94}\\ &$^{36}$Ar+$^{12}$C&187.7&$6\leq A\leq 24$&kin& C&$3\times 10^3$&\cite{fsd96}\\ $^{49}$Cr&$^9$Be+$^{40}$Ca&141&$9\leq Z\leq 11$& $\Delta$E-E&S&$>106$&\cite{gmp84}\\ $^{49}$V&$^{37}$Cl+ $^{12}$C&150,180&$5\leq Z\leq 11$ &$\Delta$E-E&S,C& $>106$&\cite{yno89}\\ $^{52}$Fe&$^{12}$C+$^{40}$Ca&74-186& $9\leq Z\leq 11$&$\Delta$E-E&S&$8\times 10^4$& \cite{gmp84}\\ $^{56}$Ni&$^{28}$Si+$^{28}$Si&85-150&A=28& kin&C&$2\times 10^1$&\cite{betts81}\\ &$^{32}$S+$^{24}$Mg&121-142&$12\leq A\leq 28$& tof,&S,C&$1\times 10^4$&\cite{setc89,shp94},\\ & & & &kin,$\gamma$ & & &\cite{sbjk90}\\ &$^{16}$O+ $^{40}$Ca&69-87&$20\leq A\leq 28$& tof&S&$4\times 10^5$&\cite{sba86}\\ $^{59}$Cu&$^{35}$Cl+$^{24}$Mg&275&$5\leq Z\leq 12$& $\Delta$E-E&S,C&$>10^6$&\cite{nbm96}\\ $^{60}$Ni&$^{16}$O+$^{44}$Ca&69-87& $20\leq A\leq 30$&tof&S&$>10^6$&\cite{sba86}\\ $^{64}$Zn&$^{16}$O+$^{48}$Ti&118&$5\leq Z\leq 15$& $\Delta$E-E&S,C&$>10^6$&\cite{yno89}\\ &$^{37}$Cl+$^{27}$Al&162-200&$5\leq Z\leq 15$& $\Delta$E-E&S,C&$>10^6$&\cite{yno89}\\ $^{78}$Sr&$^{28}$Si+$^{50}$Cr&150& $12\leq A\leq 58$&$\Delta$E-E,&S&$>10^6$& \cite{esps89}\\ & & & & tof & & &\\ $^{80}$Zr&$^{40}$Ca+$^{40}$Ca&197,231&$7\leq A\leq 62$&tof&S&$>10^6$&\cite{esps89}\\ \hline \end{tabular} $^a$ Projectile + Target \\ $^b$ C--Coincidence, S--Singles \\ $^\dagger$ Only low-lying excitation studied. \end{table} \subsection{$A_{CN} < 32$} In describing heavy-ion reaction behavior in terms of statistical models, the dominant component of the deformation-dependent binding energy of the compound system is usually obtained through macroscopic energy calculations, that is, extensions of the rotating liquid drop model. Shell corrections to the calculated energies are usually handled in only a very approximate fashion. These corrections can be difficult to obtain for the relevant shapes of the compound system and, moreover, the statistical phase space relevant for the fission process is generally required at an excitation energy of the compound system where shell corrections may already be strongly attenuated. The lower mass limit for the validity of macroscopic description of light nuclear systems and statistical behavior of very light heavy-ion reactions has been recently pursued by investigating reactions involving light s-d shell nuclei. One of the interesting features that can be probed by these studies is the effect of the channel spin on light ion reaction mechanisms. Among the systems studied, the choice of $^{10,11}$B nuclei was based on the fact that very high channel spins $\{ ^{10}$B($3^+$) and $^{11}$B($3/2^+$)$\}$ are involved when compared to the values of the grazing angular momenta. This has two significant consequences: 1.) the phase space in the exit channel is enlarged and 2.) the composite system angular momentum is increased. There are, however, experimental difficulties encountered in the study of lighter systems. All of the model calculations require some estimate of the distributions of incident orbital angular momenta contributing to the fully energy damped binary yields. This generally requires a measurement of the fusion cross sections for the reactions---in the fission models, the damped binary yields define the high end of the fusion partial wave distribution whereas, for the orbiting models, the higher fusion partial waves are seen as competing with the orbiting mechanism. Identification of fusion reaction yields involving very light heavy ion reactions is complicated, however, by the difficulty of identifying the corresponding evaporation residues. Frequently, a given element can be produced by nucleon or massive transfer, sequential decay of a compound nucleus or by a binary decay of the composite system. The production of $^8$Be residues, which subsequently decay by breaking apart into two $\alpha$ particles, may not be negligible. To unfold the different processes it is necessary to look for differences in their kinematics. Here, velocity spectra can be very useful. An example of the unfolding procedure is shown in Fig. 25 for the $^9$Be+ $^{11}$B reaction at $E_{lab}$=37 MeV and $\theta_{lab}=8^o$ \cite{faa93}. At lower recoil velocities the spectrum is consistent with a statistical model calculation using the code LILITA \cite{cbd84}, as shown by the solid curve. The additional component, shown by the dashed curve, can be attributed to a more peripheral reaction mechanism. Strongly energy damped yields have been investigated in the $^{10,11}$B+$^{10,11}$B reactions simultaneously with fusion-evaporation residue yields \cite{caaf91}. For the $^{10,11}B+ ^{11}B$ channels, the observed energy spectra and excitation functions of the evaporation residues are consistent with the general fusion systematics in this mass range. However, a significant inhibition of the fusion cross section is observed in the $^{10}$B+ $^{10}$B entrance channel (see Fig. 26). At least part of the missing flux may be directed to strongly energy damped decay yields in the Z=5 channel, which are found to be strongly enhanced. These yields demonstrate binary characteristics according to their velocity distributions (which are peaked at higher velocities than expected for evaporation residues) and isotropic angular distributions \cite{caaf91}. A similar enhancement in the $^{11}$B channel is not observed in either the $^{10,11}$B+ $^{11}$B or the $^9$Be+ $^{10,11}$B reactions up to $E_{lab}/A$= 4 MeV. The mechanism by which the very high entrance-channel spin of the $^{10}$B+ $^{10}$B system may play a role in the anomalous behavior observed for this system is not clear. Although higher compound-nucleus spins would be expected to favor binary fission over light particle emission, the behavior is also similar to that observed in somewhat heavier systems (with zero channel spin) and attributed to a dinucleus ``orbiting" mechanism. The experimental situation can be somewhat simpler for reactions involving mass asymmetric entrance channels. Here the use of inverse kinematics can lead to clear identification of the target-like products corresponding to processes that are fully energy damped and emitted at large center-of-mass angles. This approach has been used in an investigation of the $^{16,17,18}$O+ $^{10,11}$B, $^{19}$F + $^9$Be reactions leading to the $^{26-28}$Al compound nuclei \cite{aacf94}. Contributions from binary reaction process were clearly identified in the Li, Be, C, O and F channels. However, only in the target-like particle channels: Li, Be, B, and C , were isotropic distributions of ${d\sigma / d\theta_{c.m.}}$ observed for all the reactions. Alternatively, the N, O and F products, associated with projectile-like particles, present forward-peaked angular distributions with ``life-angles" in the range $25^o < \alpha < 50^o$ (see Fig. 6), suggesting that a more peripheral reaction mechanism dominates the small-angle yields in these channels. The velocity distributions of the emitted fragments suggest the occurrence of a binary process. This is shown for the $^{18}$O+ $^{11}$B reaction at $E_{lab}$=63 MeV in Fig. 27 from ref. \cite{aacf94}. Although the singles data offer compelling evidence of binary nature of the fully energy-damped yields, the experimental confirmation of such a nature requires the detection of both fragments in coincidence. Such coincidence measurements were performed for the $^{18}$O+ $^{11}$B and $^{17}$O+ $^{11}$B reactions at $E_{lab}$=53 MeV, with the observation that most of the yields correspond to $Z_1+Z_2=Z_{CN}$. This indicates that the effect of secondary light-particles emission following (or preceding) scission is negligible for these reactions at the measured energy. Coincidences between heavy fragments and $Z<3$ particles were associated with evaporation residues. The degree of inelasticity of the binary yields associated with distributions of constant ${d\sigma / d\theta_{c.m.}}$ is indicated by the total kinetic energy $<TKE>$ values displayed in Fig. 28 for these yields. Compared to the experimental results are lines showing the expected values in the case of totally relaxed processes in which the outgoing particles carry essentially the barrier energy. Although these results were calculated assuming spherical fragments, similar results are found using the transition-state model for fission where more realistic shapes are assumed for the compound system configuration \cite{aacf94}. Support for the picture of a statistically equilibrated compound nucleus is found in the data obtained for the $^{17}$O + $^{11}$B, $^{18}$O + $^{10}$B and $^{19}$F + $^9$Be reactions, each of which populates the $^{28}$Al compound system. Figure 29 presents excitation functions for intermediate mass fragment cross sections for the three systems. The close agreement among the exit yields for the various entrance channels shows that the Bohr Hypothesis \cite{bw79} is satisfied. This hypothesis states that the exit channel observables for compound-nucleus reactions should be independent of the entrance channel except for such conserved quantities as angular momentum and total energy. A similar conclusion is reached by comparing the ratio of yields for different exit channels as a function of the excitation energy of the emitted fragments. Figure 30 presents the ratio $R= \sigma_C / \sigma_B$ of yield for carbon products $(\sigma_C)$ compared to the yield for boron $(\sigma_B)$, for the three entrance channels populating the same compound nucleus. Again, entrance channel independence of the cross section ratios is observed. Although it is not possible, in general, to form a compound nucleus with the same excitation energy and angular momentum distribution using two different entrance channels, the three reactions reaching $^{28}$Al achieve similar conditions because of the small variation in the entrance channel mass asymmetry. Predictions of the transition-state model for the fission charge distributions are also compared to the experimental data in Fig. 29. In general, model predictions are found to be satisfactory in reproducing the charge, mass and bombarding energy dependence of the observed yields, thus further supporting the idea that these yields have a fusion-fission origin. It can also be noted that for these reactions, where the statistical fission description seems to work well, the normalized Number of Open Channels (N/F) for the composite system decay are relatively high---some two to four orders of magnitude larger than the ones obtained for the $ ^{12}$C+ $^{12}$C or $^{12}$C+ $^{16}$O resonant systems, as shown in Fig. 23. Bhattacharya {\it et al.} \cite{bbb91,bbm95,3b96} have looked for a possible entrance channel behavior for the fully energy damped yields from the $^{31}$P compound system as populated through the $^{12}$C+ $^{19}$F, $^7$Li+ $^{nat}$Mg, and $\alpha$+ $^{27}$Al reactions. In this case there is a very large difference in the mass asymmetries of the respective reactions. Again, in a comparison of the damped binary yields for the three systems, no prominent entrance channel behavior is observed beyond what would be expected from the very different partial cross-section distributions. At the respective bombarding energies, the normalized NOC for the grazing partial wave is large for each of these systems. \subsection{$32 \leq A_{CN} \leq 44$} \label{sec:a32to44} The first observation of fully energy damped reaction yields in light systems was reported by Shapira {\it et al.} \cite{sfgsd79} in an investigation of $^{20}$Ne+ $^{12}$C inelastic scattering at backward angles. Large cross sections were found when summing the yields over many unresolved mutual excitations at high excitation energies. The resulting angular distributions were found to show most of the characteristics, such as a ${d \sigma / d\Omega} \propto{1 / \sin\theta_{c.m.}}$ angular dependence, of a long-lived, orbiting, $^{20}$Ne+ $^{12}$C dinuclear complex. This initial measurement was followed by a very detailed study of the same system \cite{sfg82} where resonant-like behavior was found in the excitation functions for several outgoing channels, reminiscent to that observed for the well-known, resonant, symmetric $^{16}$O+ $^{16}$O system. This quasi-molecular resonance behavior extends to $Q$-value regions where the total kinetic energy in the exit channel is consistent with an orbiting-type mechanism. It was noted by Shapira {\it et al.} \cite{sfg82} that the small number of open reaction channels for $^{20}$Ne+ $^{12}$C scattering might be related to the observation of the resonance behavior. In a study of the $^{20}$Ne+ $^{16}$O reaction, Shapira {\it et al.} \cite{sdc83} again found evidence of an ``orbiting-like" component in the large angle $^{12}$C and $^{16}$O yields. Subsequent to this, the strongly energy damped component in the $^{24}$Mg+ $^{12}$C reaction, leading to the same $^{36}$Ar composite system, was investigated by the Munich group through a series of measurements \cite{gdbh90,dgh88}. In the study by Glaesner {\it et al.} \cite{gdhk86}, cross section fluctuations for fully damped yields were observed for the first time in such a heavy system, allowing for an Ericson-type fluctuation analysis \cite{emk66}. Excitation functions were generated for different $Q$-value ranges, as shown for the Z=6 channel in Fig. 31, using the data of Ref. \cite{gdhk86}. The observed structure is less pronounced than that observed in discrete-level, low $Q$-value channels because of the averaging over a large number of states. The coherence width obtained in this measurement of approximately 300-400 keV corresponds to a mean rotation angle for an orbiting-like configuration of 180$^o$ to 360$^o$ . In further support of an orbiting-like picture, Konnerth {\it et al.} \cite{kdetc88} have deduced spin alignments for the $^{24}$Mg+ $^{12}$C system by measuring the out-of-plane $\gamma$-ray anisotropy. The large positive values of the observed spin alignments suggest the geometry of a sticking dinuclear complex in a stretched configuration. Perhaps the most striking example of orbiting behavior in a light nuclear system can be seen in the $^{28}$Si+$^{12}$C reaction yields. This system has been extensively studied over the past two decades for a number of reasons. In addition to showing a very strong orbiting-like behavior \cite{s82,ssf84,rsh91,rlv86} (see Fig. 22), it also demonstrates strong resonant structures in its elastic and quasi-elastic yields \cite{blb79,cfo78,ocd79}. Fortunately, the reaction is favorable for experimental study since intense $^{28}$Si beams are readily available at tandem accelerator facilities and $^{12}$C makes a very good and relatively contaminant free target. The energy spectra for the most intense C, N and O fragment exit channels for the $^{28}$Si+ $^{12}$C reaction have been measured at backward angles for a large number of incident energies in the range $29.5\leq E_{c.m.}\leq 54 MeV$. At lower bombarding energies the excitation spectra for the $^{12}$C fragments are dominated by single and mutual excitations of the $^{12}$C and $^{28}$Si fragments. At higher bombarding energies, however, the dominant strength for all three channels shifts to higher excitation energies \cite{ssf84}. For these higher energy spectra the most probable $Q$-values are found to be independent of detection angle. The corresponding, energy-integrated angular distributions are found to follow a $1/\sin\theta_{c.m.}$ angular dependence near 180$^o$\cite{sfg82}. These characteristics suggest a long lived, orbiting-like mechanism where energy equilibration has been achieved. Many of the salient features of the orbiting yields, such as their inelasticity and anisotropy, are indistinguishable, however, from those expected for a compound nucleus fusion-fission mechanism. Early on, the fission mechanism was considered as a possible explanation for the energy damped $^{28}$Si+$^{12}$C yields, but abandoned because the observed cross sections were much greater than model predictions based on the standard rotating liquid drop model. The possibility of a significant fission contribution was again suggested after it was shown that finite-nuclear range effects can lead to significantly reduced fission barriers in lighter systems \cite{s91,skb87,skb88}. However, as shown in Fig. 22, even though the newer fission calculations may be able to account for the observed cross sections in the nitrogen and oxygen channels, the calculations significantly understate the cross sections measured in the carbon channel \cite{s91}. Good agreement with the observed fully damped cross sections, evaporation residue cross sections, and average total kinetic energy values for the damped yields has been obtained using the equilibrium model for fusion and orbiting \cite{sah87,ass88}, as discussed in Sec. \ref{sec:eqorb}. In particular, the observed saturation trend of the TKE values as a function of the incident energy is well described by this latter model, as shown in Fig. 21. In the equilibrium orbiting model, saturation occurs because a value of the orbital angular momentum is reached, after dissipation, beyond which the formation of a dinuclear complex is not allowed because of the centrifugal repulsion. Additional studies support the idea that an orbiting-like mechanism dominates that component of the carbon yield characterized by an angular dependence with ${d\sigma / d\Omega}\propto {1 / \sin\theta_{c.m.}}$. The entrance-channel dependence of the process was demonstrated by Ray {\it et al.} \cite{rk85} by forming the $^{40}$Ca nucleus with both the $^{24}$Mg+ $^{16}$O and $^{28}$Si+ $^{12}$C reactions at closely matched excitation energies and angular momenta. Figure 32 shows the observed ratio of the oxygen to carbon cross section for these reactions as a function of excitation energy. If compound-nucleus fission dominated the cross sections, this ratio would be expected to be very similar for the two reactions. The observation of a strong entrance-channel dependence of the ratios suggests a non-compound mechanism. It is interesting to note that the entrance channel effect becomes smaller at larger excitation energies. This might suggest stronger fission competition for the more strongly damped yields \cite{cnrs94}. Shiva Kumar {\it et al.} \cite{kbg92} have extended the entrance channel studies by comparing the carbon to oxygen cross section ratios over an extended range of $^{40}$Ca excitation energies. This study suggests an approach to an equilibrated compound nucleus ratio as the beam energy increases. In another study that supports an orbiting-like explanation for the less damped yields, the population of magnetic substates of $^{12}$C$(2_1^+)$ and $^{28}$Si states for the $^{12}$C( $^{28}$Si, $^{12}$C) $^{28}$Si reaction at 180$^o$ has been measured by Ray {\it et al.} \cite{rlv86} using $\gamma$-ray angular correlation techniques. Qualitatively, the observed selective population of the m=0 magnetic substate with respect to the beam axis agrees well with a simple dinuclear sticking picture. A subsequent, more complete and detailed experiment has been performed to measure density matrices of the $^{12}$C and $^{28}$Si excited states \cite{rsh91}. The data are found consistent with a dinuclear picture in which bending and wriggling motions are the dominant spin carrying modes. It is interesting to note that these $\gamma$-ray angular correlation experiments using the $^{28}$Si+ $^{12}$C orbiting reaction have been used to illuminate certain conceptual aspects of quantum mechanics \cite{ray93}. In general, orbiting processes and resonant-like structures have been found to coexist in collisions for which surface transparency is expected based on the small number of open reaction channels \cite{abe95}. The $^{28}$Si+$^{12}$C and $^{24}$Mg+$^{16}$O reactions, for example, both show large ``orbiting" yields as well as strong resonances at backward angles in their elastic, inelastic \cite{blb79,kovar85}, and transfer exit channels \cite{kovar85}. As seen in Fig. 24, the normalized number of open reaction channels for both of these reactions is very small. The mass-symmetric $^{20}$Ne+ $^{20}$Ne reaction, which leads to the same $^{40}$Ca composite system as $^{28}$Si+ $^{12}$C and $^{24}$Mg+ $^{16}$O, has also been investigated \cite{sdc83} and found to exhibit strong orbiting-type behavior. Again, this system is found to have a small number of open reaction channels. However, to our knowledge, this is the only ``orbiting" system to show almost structureless elastic excitation functions \cite{bzm95}. The $^{28}$Si+$^{14}$N reaction is an important test case for exploring how surface transparency, as reflected in the number of open channel calculations, is related to the occurrence of orbiting and molecular resonance behavior. Although this system is close in mass to the $^{28}$Si+ $^{12}$C system, the phase space available to reaction channels is very different for the two: whereas the $^{28}$Si+ $^{12}$C reaction has only a few exit channels, the $^{28}$Si+ $^{14}$N reaction is characterized by a large value of N/F (see Fig. 24). The observed orbiting-like cross sections for $^{28}$Si+ $^{14}$N reaction \cite{ssa88,sss86} are significantly smaller than for the $^{28}$Si+ $^{12}$C reaction. The equilibrium model for orbiting \cite{sah87} has been found to give a reasonably good description \cite{ass88} of the ensemble of the experimental data \cite{ssa88,sss86} for this reaction, although fusion-fission calculations performed at $E_{c.m.} = 40 MeV$ \cite{s91} are found to reproduce the overall experimental behavior with comparable success. As a consequence of, or at least coincident with, the increased reaction phase space, the damped binary breakup yield behave more like ``normal" fission. Pronounced backward angle yields have also been observed in the $^{27}$Al+ $^{16}$O reaction at several $^{27}$Al bombarding energies by Shapira {\it et al.} \cite{sfg80}. This is also a system where the number of open reaction channels is large and the observed cross sections for the damped yields are consistent with a fission interpretation. The symmetric and nearly symmetric mass binary decay of $^{44}$Ti has been studied in inclusive and exclusive measurements of the $^{32}$S+ $^{12}$C reaction at 280 MeV bombarding energy \cite{pbb86}. Although substantial post-scission evaporation occurs at this high energy, it was possible to extract the energy-damped reaction cross sections and $\langle TKE \rangle$ values for this system. The analysis of these data is one of the earlier attempts to describe the damped binary yields in a light system in terms of a fusion followed by fission picture. However, the experimental set-up was designed for the detection of two comparable mass fragments, rather than the unequal mass fragments expected to dominate the fission cross sections for a system of fissility below the Businaro-Gallone point \cite{bg55}. Similar studies \cite{gmp84} focusing on symmetric mass breakup yields have shown the existence of fission-like yields for three other light systems in this mass region: $^6$Li+ $^{40}$Ca, $^9$Be+ $^{40}$Ca and $^{12}$C+ $^{40}$Ca, leading to the $^{46}$V, $^{49}$Cr and $^{52}$Fe compound nuclei, respectively. In a study of the $^{28}$Si+$^{16}$O reaction at $E_{c.m.}$=39.1 and 50.5 MeV, Oliveira {\it et al.} \cite{oliv96} found fully energy damped yields which were attributed to a deep inelastic scattering mechanism. These data have also been analyzed in terms of the transition state model and found consistent with a fusion-fission mechanism \cite{sdt97}. In another study of the $^{44}$Ti compound system, Barrow {\it et al.} \cite{bzb95} have found evidence of correlated resonance phenomena in excitation functions of binary channels from the $^{24}$Mg+ $^{20}$Ne reaction. The data suggest that the observed resonances can be characterized by angular momenta close to that of the grazing angular momentum in the entrance channel. This is taken to suggest a different origin for these structures than the very pronounced resonances seen in elastic and inelastic scattering yields for the $^{24}$Mg+ $^{24}$Mg reaction \cite{zkbsh83}, which seem to arise from spins higher than the corresponding grazing angular momentum \cite{qzkp90}. The $^{24}$Mg+ $^{20}$Ne study focuses on low-lying excitations and does not address the question of whether this resonance system is also found to exhibit enhanced orbiting-like yields in the more strongly energy-damped channels. A similar resonance behavior to that seen for the $^{24}$Mg+$^{20}$Ne reaction is also observed for the $^{28}$Si+ $^{20}$Ne reaction, as also studied by Barrow {\it et al.} \cite{bzb95}. Again, the possibility of an orbiting-like component has not been explored. \subsection{$44 < A_{CN} \leq 56$} \label{sec:exp44to56} For reactions populating compound nuclear masses in the range $44 < A_{CN} \leq 56$ there is relatively little evidence for the pronounced orbiting-like yields observed in some lighter systems. In general, the experimental cross sections for the strongly damped binary yields are in good agreement with expectations based on fission-model calculations. There is still, however, evidence for heavy-ion resonance behavior for some of the systems studied in this mass range. To study the possible competition between the fission and orbiting mechanisms, the population and decay of the $^{47}$V compound system has been extensively studied through three different entrance channels: $^{35}$Cl+ $^{12}$C; $^{31}$P+ $^{16}$O and $^{23}$Na+ $^{24}$Mg \cite{bdhd89,bdhf92,bmetc95,alot96}. These systems cover a wide range of entrance-channel mass asymmetries and therefore allow for a strong test of the decoupling of the observed binary yields from the entrance channel, one of the signatures that can be used to differentiate between the fission and orbiting mechanism. The binary decay properties of the $^{47}$V nucleus, produced in the $^{35}$Cl+$^{12}$C reaction, have been investigated between 150 and 280 MeV by means of a kinematics coincidence technique \cite{bdhd89,bdhf92,bmetc95}. The angular distributions of the lightest fragments are found to follow a $1/\sin\theta_{c.m.}$ angular dependence. This is shown in Fig. 33 for fragments with $5\leq Z\leq 11$ from measurements at $E_{lab}=180$ MeV and 200 MeV. Distributions of $d\sigma/d\theta_{c.m.}$ are shown. The distributions are found to be independent of the scattering angle for each exit channels indicating that the lifetime of the dinuclear complex is comparable to or longer than the rotational period. The binary nature of the reaction products has been clearly established with the coincidence measurement. Complete energy relaxation of the fragments with $5\leq Z \leq 11$ is evident from the angle independence of their observed TKE values, as shown in Fig. 33 \cite{bdhf92}. Although these results are obtained using singles data, equivalent results have been obtained in the coincidence measurements. The averaged TKE values for all detected fragments vary little with incident energy and the TKE value corresponding to a symmetric mass breakup is close to the prediction of the revised Viola systematics \cite{vkw85}, as shown in Fig. 8. To test the entrance-channel independence of the damped reaction yield for the $^{47}$V system, back-angle $^{12}$C and $^{16}$O yields have been measured in the $^{31}$P+ $^{16}$O \cite{rsc91} and $^{35}$Cl+ $^{12}$C reactions \cite{bdhd89} at energies leading to the same compound-nucleus excitation energy of $E^*_{CN}= 59.0$ MeV and very comparable angular momenta. The observed $^{12}$C and $^{16}$O cross sections are comparable for the two systems and much smaller than those predicted by the equilibrium orbiting model \cite{sah87,ass88}. Also, the ratio of carbon to oxygen cross section, as shown in Fig. 34, has no significant entrance channel effect and is in general agreement with the predictions of the transition-state model calculations \cite{s91}. A similar comparison has been done with the $^{35}$Cl+ $^{12}$C and $^{23}$Na+ $^{24}$Mg \cite{bdhf93} reactions, populating the $^{47}$V compound nucleus at $E^*_{CN}= 64.1$ MeV. Again, as shown in Fig. 34, the observed behavior is in reasonable agreement with the expectations of the transition-state description of fission \cite{bdhf93}. The elemental cross sections for the $^{23}$Na+ $^{24}$Mg reaction are also in agreement with expectations based on the fission picture, as shown in Fig. 19. One of the most striking phenomena observed in heavy-ion reaction studies is the pronounced resonance structures observed in elastic and inelastic scattering of the $^{24}$Mg+$^{24}$Mg system \cite{zkbsh83}. Narrow structures which are correlated in many channels and extending to high excitation energy suggest that a very special configuration of the $^{48}$Cr compound system is formed in this reaction. By measuring the $\gamma$-ray correlations with the $^{24}$Mg fragments, it has been possible to deduce a resonance spin for at least one of the observed structures that is greater than that of the grazing angular momentum \cite{qzkp90,wzk87}. This is taken to suggest a very prolate deformed configuration of the compound system leading to the resonance. In earlier measurements of the resonance behavior, peaks were also observed in excitation-energy spectra for the $^{24}$Mg($^{24}$Mg, $^{24}$Mg) $^{24}$Mg reaction up to energies where secondary $\alpha$-particle evaporation from the fragments obscures any spectroscopic details. This raises the question as to whether the structure observed at higher excitation energies is somehow related to the resonance phenomenon or, instead, is a feature of the fission decay of the compound nucleus. As discussed in Sec. \ref{sec:ternary}, a possible ternary fission mode for the $^{24}$Mg+$^{24}$Mg reaction, as suggested by several model calculations, has also been sought for but not observed. To explore the relationship of the different reaction mechanisms influencing the binary decay yields of the $^{48}$Cr compound system, the energy-damped yields of the $^{36}$Ar(E$_{lab}$=187.7 MeV)+ $^{12}$C \cite{fsd96}, $^{20}$Ne(E$_{lab}$=78.0 MeV)+ $^{28}$Si \cite{fsd96}, and $^{24}$Mg(E$_{lab}$=88.8 MeV)+ $^{24}$Mg \cite{hsfp94} reactions have been studied. Each of these reactions populates the $^{48}$Cr compound nucleus at an excitation energy of about 59.5 MeV. In each case, the outgoing fragments were identified by measuring both fragments in a kinematic coincidence arrangement. The calculated mass distribution of the fully energy damped yields with $6\leq A\leq 24 $ is in excellent agreement with the experimental results for the $^{24}$Mg+ $^{24}$Mg reaction \cite{hsfp94}, with no significant evidence for an excess yield in the entrance channel that might suggest an additional, orbiting mechanism. The agreement is also reasonable good for the $^{36}$Ar+ $^{12}$C entrance channel, although in this case the experimental results show a somewhat greater mass asymmetry of the fission fragments than predicted. The use of the kinematic coincidence technique allows for very good $Q$-value resolution in the final channels, as seen in Fig. 35 where the excitation-energy spectra for the $^{24}$Mg+ $^{24}$Mg channel is shown for each of the three entrance channels. The experimental results are indicated by the thick-line histograms. It is evident from this figure that the structure observed at higher excitation energy is correlated for the three entrance channels, making it improbable that this structure is an artifact of the resonance behavior. Rather, the structure seems to be related to the detailed level structure of the final nuclei. The thin-line histograms in Fig. 35 are obtained using the transition-state model with the saddle-point method and applying the same procedure as discussed for Fig. 16 to associate the flux at the saddle point with specific mutual excitations of the fragments \cite{shp94,fsd96}. The fission picture can account for most of the observed structures, although with some significant discrepancies observed between the calculated and measured yields to low-lying excitations populated through the $^{24}$Mg+ $^{24}$Mg and $^{20}$Ne+ $^{28}$Si channels. This must be expected since both of these reactions show evidence of resonance behavior \cite{zkbsh83,bzb95} which can not be described by the transition-state picture. The energies corresponding to single and mutual excitations of yrast levels and ground-state band members are shown at the bottom of Fig. 35. Although these excitations are found to contribute to the observed structures, they do not appear to dominate the spectra. Instead, the calculations suggest that random groups of high channel spin excitations account for the general appearance of the spectra. Instead of using the saddle-point calculation to predict, {\it a priori}, the excitation energy spectra, Farrar {\it et al.} \cite{fsd96} have also used this method to explore the compound-nucleus spin distribution leading to the observed fission yields. In this analysis, it is found that the average spin obtained from the fitted distribution is comparable to that obtained from the {\it a priori} calculation for the $^{36}$Ar+ $^{12}$C and $^{20}$Ne+ $^{28}$Si reactions, but is smaller than the systematics would suggest for the $^{24}$Mg+ $^{24}$Mg entrance channel. The difference in the average fitted and calculated spin values is even greater when a ``resonance-subtracted" energy spectrum is used for the $^{24}$Mg+$^{24}$Mg reaction. This suggests that there may be direct competition between the heavy-ion resonance and compound-nucleus fission mechanisms for near grazing partial waves of this entrance channel. The $^{56}$Ni compound system has also been explored through multiple entrance channels, using the $^{16}$O($E_{lab}$=69-87 MeV)+ $^{40}$Ca \cite{sba86}, $^{28}$Si($E_{lab}$=85-150 MeV)+ $^{28}$Si \cite{betts81}, and $^{32}$S($E_{lab}$=121 MeV and 142 MeV) + $^{24}$Mg reactions \cite{setc89}. Excitation functions of the elastic and inelastic excitations of the symmetric $^{28}$Si+ $^{28}$Si channel are found to exhibit correlated resonance behavior \cite{bdp81}, although somewhat more weakly than seen in the $^{24}$Mg+ $^{24}$Mg system. Each of the three entrance channels is found to result in fully energy-damped yields that are consistent with a fusion-fission reaction mechanism. Excitation-energy spectra for the $^{24}$Mg( $^{32}$S, $^{28}$Si) $^{28}$Si reaction at $E_{lab}$=121 MeV and 142 MeV \cite{setc89} are shown in Fig. 17 and are found to be well reproduced by the transition-state model using the extension to the saddle-point method \cite{shp94} to calculate these spectra. To further confirm the predicted population of final mutual excitation, $\gamma$ rays were measured in coincidence with the $^{28}$Si fission fragments \cite{shp94} in the excitation energy range $7.6\leq E_x\leq 16.7$ MeV. In general, the observed and predicted transition rates for specific states in the $^{28}$Si fragments were found to be in good agreement. This is shown in Figs. 36 and 37. In Fig. 36 it is seen that the model calculations well reproduce the relative strength of the yrast $2^+\rightarrow 0^+$ and $4^+\rightarrow 2^+$ transitions. Since higher spin states tend to feed the $4^+$ level, this agreement suggests that the population of these higher spin states is being reasonable well described. Figure 37 shows the calculated and measured $\gamma$-ray spectra on an expanded scale. Again, reasonably good agreement is found between the predicted and observed transition strengths. A quantitative comparison, however, reveals evidence of greater population of members of the $K^\pi=0^+_3$ band in $^{28}$Si (taken to include levels at 6691 keV, 7381 keV, 9165 keV, and 11509 keV) than predicted. This band is believed to have a strongly prolate deformed nature \cite{gbs81} and an enhanced population might result from the relatively deformed shapes expected for the nascent fission fragments at the saddle point. It should be noted, however, that there is some debate as to whether the 11509 keV $6^+$ level should be associated with this band \cite{cshs82}. The mass distributions of the fission-like yields for the $^{32}$S+ $^{24}$Mg reaction at $E_{lab}$=121 and 142 MeV \cite{setc89}, obtained by fitting experimental angular distributions assuming a ${d\sigma / d\Omega} \propto {1 / \sin\theta_{c.m.}}$ angular dependence, are shown in Fig. 38 by the open histograms. The corresponding predicted mass distributions based on the transition-state model calculations, shown by the solid histograms, are found to be in excellent agreement with the data. One of the results of the model calculations which has yet to be confirmed in any of the systems studied is that significant fission yield is expected in the $^8$Be channel. By bridging the light-particle evaporation and fission mass ranges, the predicted strength in the $^8$Be channel highlights the idea, as proposed by Moretto \cite{lgm75}, that light-particle evaporation and fission yields have a common origin and should be viewed in terms of their respective decay barriers. In an experiment exploring the energy and spin sharing between fission fragments in the $^{24}$Mg( $^{32}$S, $^{12}$C) $^{44}$Ti reaction at $E_{lab}$=140 MeV, $\gamma$-ray spectra were obtained in coincidence with the $^{12}$C fragments \cite{sbjk90}. The results also indicate a statistical decay process consistent with the predictions of the transition-state model. Moreover, no evidence was found for the spin alignment of the $^{12}$C fragments, contrary to what might be expected for a deep-inelastic scattering origin of the fully energy-damped yields. \subsection{$56 < A \leq 80$} The reaction products from the $^{35}$Cl+ $^{24}$Mg system have been investigated at a bombarding energy $E = 8$ MeV/nucleon with both inclusive \cite{cbd95} and exclusive measurements \cite{nbm96}. The inclusive data provide information on the properties of both the evaporation residues and the binary-decay fragments. The binary process yields are, for instance, successfully described by statistical models based on either the saddle point picture \cite{s91} or the scission point picture \cite{mbnm97}. The similar good agreement with theory that is found for the energy spectra, the angular distributions, and the $\langle TKE \rangle$ values makes the hypothesis that fully energy-damped fragments result from a fusion-fission process quite reasonable for the $^{59}$Cu compound system, in accordance with findings for other, equivalent systems as shown in this report. No evidence is seen in the coincidence data for the occurrence of three-body processes in the $^{35}$Cl+ $^{24}$Mg reaction. This result can be contrasted to the situation reported for somewhat heavier mass systems, where significant three-body breakup yields are evident for the $^{32}$S+ $^{45}$Sc \cite{bmv93} and $^{32}$S+ $^{59}$Co \cite{vml88} reactions, both measured with $E_{lab}$($^{32}$S)=180 MeV (5.6 MeV/u). The nuclear-charge deficits from the compound-nucleus charge found in the $^{35}$Cl+ $^{24}$Mg exclusive measurement, however, can be fully accounted for by the sequential evaporation of light charge particles (LCP), in agreement with the systematics established for a large number of reactions studied at bombarding energies below 15 MeV/nucleon \cite{bsdt96}. The question of whether a small part of the binary reactions come from ternary processes is still an open question and difficult to answer. In general the measured charge-deficit values and other experimental observables (such as cross sections, energy- and angular-distributions or mean TKE values) are very well described by a complete Extended Hauser-Feshbach statistical-model calculation \cite{mbnm97} which takes into account the post-scission LCP and neutron evaporation. Although the fission picture is seen to work well in the $^{35}$Cl+ $^{24}$Mg reaction, a very different conclusion is reached by Yokota {\it et al.} \cite{yno89} in a study of two systems ($^{37}$Cl+ $^{27}$Al and $ ^{16}$O+ $^{48}$Ti) leading to the somewhat heavier $^{64}$Zn compound system. These reactions populate $^{64}$Zn at comparable excitation energies and spins. The components of the reaction yields corresponding to a ${d\sigma / d\Omega}\propto {1 / \sin\theta_{c.m.}}$ angular dependence result in very different $Z$ distributions for the two reactions, contrary to the expectations of the statistical decay of a compound nucleus. It is possible that these systems are again displaying the strong orbiting signature that has been found in several lighter systems. If so, these results could provide an interesting challenge for the number of open channels calculations performed for systems of mass $A_{CN} > 60$. The transition region where it appears that the asymmetric-mass fission observed in lighter systems may change into the symmetric fission behavior characteristic of heavier systems appears to occur around mass $A\approx 80$. Evans {\it et al}. have studied the $^{40}$Ca+ $^{40}$Ca reaction at $E_{c.m.}$=197 MeV and 231 MeV \cite{esp91} and the $^{28}$Si+ $^{50}$Cr reaction at $E_{lab}$($^{28}$Sr)=150 MeV \cite{esps89}. The resulting mass distributions for the lower energy $^{40}$Ca+ $^{40}$Ca reaction and the $^{28}$Si+ $^{50}$Cr reaction are shown in Fig. 39 by the open circles. The mass distributions for the two reactions are found to be quite different, even though the fissility of the two systems is quite similar. This behavior was initially thought to indicate a fast-fission mechanism accounting for the asymmetry dependence of the $^{28}$Si+ $^{50}$Cr yields \cite{esps89}. However, in exploring the possibility of fission competition in these systems, it has been found that for the spin values near the critical angular momentum for fusion, where most of the fission yields is expected to originate, the mass asymmetry dependent fission barriers are actually quite different for the two systems. The $^{80}$Zr compound system, populated through the $^{40}$Ca+$^{40}$Ca reaction, is found to have either a very flat distribution of barrier energies as a function of mass asymmetry or slightly lower barriers for the symmetric mass configuration. For the $^{78}$Sr compound system, populated through the $^{28}$Si+ $^{50}$Cr reaction, however, the distribution favors asymmetric mass breakup. The bold-line histograms in Fig. 39 show the predicted primary fission mass distributions using the transition-state model and the thin-line histogram show the corresponding mass distributions after secondary light-particle emission from the fission fragments is taken into account. For the two systems, the overall trend of the data seems to be reproduced by the calculations, indicating that these two systems may straddle the Businaro-Gallone \cite{bg57} transition from asymmetric- to symmetric-mass fission. \section{Open Problems} \subsection{Time scale} Although the process of binary decay from equilibrated compound nuclei has been clearly identified experimentally and successfully described in terms of phase-space models, more thought is needed on the dynamics through which systems evolve from their entrance channels to deformed and statistically equilibrated compound nuclei that subsequently undergo scission. The occurrence of orbiting and resonance behavior in some systems indicates the superposition of compound nucleus decay products (properly described by the transition state model) and faster direct processes (interpreted on the basis of polarization potentials, entrance channel resonances or DIC orbiting processes). Therefore, the investigation of the dynamics involved in heavy ion collisions leading to totally energy-damped binary exit channels may lead to a better understanding of the competition among the different reaction mechanisms. Within this scenario two questions arise: a) the time scale of the processes and b) the shape evolution of the system. It has been suggested by Thoennessen {\it et al.} \cite{tbb93} that systems with different entrance channel mass asymmetries may evolve towards their compound nucleus configurations with different time scales such that the most asymmetric one reaches full equilibrium faster. This finding suggests that the mass asymmetry may affect not only the angular momentum distribution but also the competition with faster direct processes. In the case of heavier systems, $\gamma$ rays from the decay of the giant dipole resonance (GDR) built on highly excited compound-nucleus states has been shown to be sensitive (through the shape of the energy spectrum) to the time scale of the process as well as to the deformation of the nuclear system. As we go to lighter systems, where the energy of the GDR is quite high, severe experimental problems are expected with such GDR measurements. However, these measurements should be extended to systems that are as light as possible. Interferometry measurements based on the detection of emitted pairs of light charged particles has been employed as a technique to probe time scales and nuclear dimensions \cite{koon77,dgx89}. For heavier systems, these studies have generally involved the use of small angle correlations of protons emitted from a hot source in the determination of emission time scales and sources radii. However, if we go to lighter and equilibrated systems, this technique can be borrowed to obtain estimates of reaction time scales. If a hot composite systems decays into a binary channel and a proton or alpha particle evaporates from one of the fragments, the proximity of the other fragment may distort the kinematics correlation because of the strong Coulomb repulsion. Such a distortion will depend on the time scale for fission and the sequential secondary evaporation. In cases where there is good $Q$-value resolution, the experimental $Q$-value spectra present a structured behavior even at high excitation energies. This is seen, for example, in Fig. 35 for the three entrance channels populating the $^{48}$Cr compound nucleus. These structures can be associated with selectively populated clusters of high-spin mutual excitations. This situation allows for a fluctuation analysis (see, for example, ref. \cite{gdhk86} and Sec. \ref{sec:a32to44}) to obtain the coherence width of the intermediate system and, hence, its lifetime (with $\tau=\hbar / \Gamma$). It is important to assure that the number of levels included in the energy bin is very low, requiring experiments using very thin targets and beams of good energy and spatial resolution. \subsection{ Relationship to heavy-ion resonance behavior-superdeformed minima?} Several of the results presented in this report suggest that a coherent framework may exist which connects the topics of heavy-ion molecular resonances \cite{eb85,gps95}, superdeformation effects as observed in medium mass $\gamma$-ray studies (see the most recent reports quoted in \cite{jk91}), and fission shape isomerism in the actinides \cite{vh73}. The shape of the ``normal" saddle-point configuration in light systems is very similar to two, touching, prolate-deformed spheroids in a neck-to-neck configuration. The shape of the system found in a conjectured, secondary well in the potential energy surface is likely to be similar, although involving greater deformation of the nascent fragments. In calculations of the shape-dependent potential-energy surfaces at high angular momenta (16-40$\hbar$) for the $^{48}$Cr nucleus \cite{aberg90}, a strong superdeformed configuration is predicted that corresponds to an aligned arrangement of two touching and highly deformed $^{24}$Mg nuclei. This superdeformed configuration is a candidate to become yrast at around spin 34$\hbar$, in the high excitation energy region which corresponds to where the quasi-molecular resonances have been observed. Indeed, from spin alignment measurements \cite{qzkp90} of two strong resonances in the $^{24}$Mg+ $^{24}$Mg scattering reaction \cite{zkbsh83}, the deduced spin assignments were found to be comparable to or a few units larger than expected for grazing collisions, leading to the same conclusion that the resonance configurations correspond to a shape of two prolate $^{24}$Mg nuclei placed pole to pole. This observation has been further supported by theoretical calculations of a molecular model \cite{ua93} which indicate a dinuclear nature of the observed resonances and suggest the presence of such a stabilized configuration in $^{48}$Cr at high spins. The conjectured isomeric configuration constituting this aligned, pole-to-pole arrangement of two $^{24}$Mg clusters has a large probability for breakup into two $^{24}$Mg fragments---a situation which is similar to that expected for the symmetric-fission saddle point, suggesting a relationship between the fission mechanism and that responsible for the resonance behavior. The relative strength of the various statistical and non-statistical processes observed in the binary yields of light systems is found to be related to the number of available open channels for the near-grazing partial waves \cite{baa94,bmetc95}. The resonant, non-statistical mode of the $^{24}$Mg+ $^{24}$Mg reaction leading to the $^{48}$Cr compound system emphasizes the dominance of partial waves near or slightly above the grazing angular momentum value \cite{bdp81,zkbsh83}. The fission mechanism has also been found to play a significant role at these spins \cite{hsfp94}, however, indicating that the nuclear configuration leading to the resonance behavior is only slightly more extended than that expected for the nuclear saddle point. The coexistence of fission and a separate reaction mechanism corresponding to heavy-ion resonance behavior has been analyzed in detail \cite{hsfp94} for the binary breakup of the $^{48}$Cr compound system populated with the $^{24}$Mg+ $^{24}$Mg, $^{20}$Ne+ $^{28}$Si and $^{36}$Ar+$^{12}$C reactions (see Sec. 5.3). The conclusion drawn from inspection of the energy spectra shown in Fig. 35 was that a significant fraction of the yield observed in the $^{24}$Mg+$^{24}$Mg exit channel arises from a statistical fission mechanism, with the resonance mechanism primarily influencing the lower excitation energy region of the spectra due to the more symmetric entrance channels. The influence of the fragment structure and the relationship between the fission mechanism and that responsible for the resonance behavior needs to be investigated further with detailed particle-$\gamma$ coincidence measurements of the $^{24}$Mg+ $^{24}$Mg system, including excitation functions measurements in the vicinity of the well-known resonance energies. \section{Conclusions} In this review we have summarized the results and conclusions of many investigators who have studied the fully energy-damped, binary yields arising from reactions involving lighter nuclear systems with $A_{CN} \leq 80$. The experimental and theoretical techniques used in these investigations have been presented and illustrated with experimental results. The general systematics that have been developed for these yields have been reviewed. In general, the data lend support to the newer macroscopic energy calculations based on the finite-range, rotating liquid drop model. Fission like yields have been observed in all of the systems studied. Moreover, the experimental systematics support the expectation, based on model calculations, that fission should favor a mass asymmetric breakup of the compound nucleus in these light systems. Further support for the fission picture comes for the measured total kinetic energy values of the fragments which are found to reflect the expected deformation of the compound nucleus at the point of scission. It is shown that various model calculations that share the premise that the final mass and energy distributions can be described by phase space constraints all lead to comparable predictions of the damped binary yields. These include the transition-state models based on counting the available states at the saddle- or scission-points, the equilibrium orbiting model, and the dynamical breakup model. The strength of the transition-state model using the saddle-point method is found in its ability to describe the general fission behavior over the entire region of mass covered by this report in a relatively ``parameter-free" manner. This general success is believed to be related to the ability of the finite-range rotating liquid drop model to correctly calculate the shape and energy of the saddle-point barrier. The similarity of the saddle- and scission-point configuration in these light systems, however, allows for very similar behavior being predicted when the scission point is used as the ``transition state" or when equilibrium orbiting is considered. The general success of the statistical model calculations allows us to now establish a reference for what is the ``expected" behavior for the damped binary yields and to search for deviations from this behavior. It has become clear that in some systems there is an additional ``orbiting" component that is of much larger cross section than can be accounted for by the fission calculations. Systems where this additional component is present also tend to manifest resonance-like behavior in excitation functions of their elastic, inelastic, and transfer channels. The occurrence ``orbiting/resonance" behavior is found to be strongly correlated with the number of open reaction channels which, in turn, is believed to be associated with the degree of absorption in the grazing partial waves. The precise mechanism(s) involved in the orbiting and resonance behavior is still unknown. The model calculations also make definite predictions of compound nucleus lifetimes and the shapes corresponding to the fission saddle-point. Measurements aimed at confirming these predictions are likely to require triple coincidences of three outgoing particles---the two resulting heavy fragments from the reaction and either the $\gamma$ ray or light particle emitted from one of the two primary fragments. Such measurements are still in their infancy. \leftline{\bf Acknowledgements :} This work was supported by the Conselho Nacional de Desemvolvimento Cientifico e Tecnologico (CNPq), Brazil, the U.S. Department of Energy, Nuclear Physics Division, under Contract No. DE- FG03-96ER40981, the Centre National de la Recherche Scientifique of France within the CNRS/CNPq Collaboration program 910100/94-0, and the U.S. National Science Foundation under the U.S.-France and U.S.-Brazil International Programs offices. \newpage
\section{##1}} } \renewcommand{\theequation}{\thesection.\arabic{equation}} \newcommand{\lrsection}[1]{\section{#1} \setcounter{equation}{0}} \let\orgsection\section \def\lrsection #1{\setcounter{equation}{0} \section{#1}} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \renewcommand{\baselinestretch}{1.5} \begin{document} \hbadness=10000 \newpage \title{BOSON INTERFEROMETRY \\ in \\ HIGH ENERGY PHYSICS\thanks{To appear in Physics Reports, 1999}} \author{{R. M. Weiner\thanks{E.Mail: <EMAIL>} \date{Physics Department, University of Marburg,\\ Wieselacker 8, 35041 Marburg, Germany \\and \\ Laboratoire de Physique Th\'eorique et Hautes \'Energies, Univ. Paris-Sud,\\ 177 rue de Lourmel, 75015 Paris, France}} } \maketitle \begin{abstract} Intensity interferometry and in particular that due to Bose Einstein correlations (BEC) constitutes at present the only direct experimental method for the determination of sizes and lifetimes of sources in particle and nuclear physics. The measurement of these is essential for an understanding of the dynamics of strong interactions which are responsible for the existence and properties of atomic nuclei. Moreover a new state of matter, quark matter, in which the ultimate constituents of matter move freely, is within the reach of present accelerators or those under construction. The confirmation of the existence of this new state is intimately linked with the determination of its space-time properties. Furthermore BEC provides information about quantum coherence which lies at the basis of the phenomenon of Bose-Einstein condensation seen in many chapters of physics. Coherence and the associated classical fields are essential ingredients in modern theories of particle physics including the standard model. Last but not least besides this ``applicative" aspect of BEC, this effect has implications for the foundations of quantum mechanics including the understanding of the concept of ``identical particles". Recent theoretical developments in BEC are reviewed and their application in high energy particle and heavy-ion reactions is analyzed. The treated topics include: a) a comparison between the wave-function approach and the space-time approach based on classical currents, which predicts ``surprising" particle anti-particle BEC b) the study of final state interactions c) the use of hydrodynamics d) the relation between correlations and multiplicity distributions. \noindent \end{abstract} \newpage \tableofcontents \newpage \section{Introduction} The method of photon intensity interferometry was invented in the mid fifties by Hanbury-Brown and Twiss for the measurement of stellar dimensions and is sometimes called the HBT method. In 1959-1960 G.Goldhaber, S.Goldhaber,W.Lee and A.Pais discovered that identical charged pions produced in $\bar p-p$ annihilation are correlated (the GGLP effect). Both the HBT and the GGLP effects are based on Bose-Einstein correlations (BEC). Subsequently also Fermi-Dirac correlations for nucleons have been observed. Loosely speaking both these correlation effects can be viewed as a consequence of the symmetry (antisymmetry) properties of the wave function with respect to permutation of two identical particles with integer (half-integer) spin and are thus intrinsic quantum phenomena. At a higher level, these symmetry properties of identical particles are expressed by the commutation relations of the creation and annihilation operators of particles in the second quantization (quantum field theory). The quantum field approach is the more general approach as it contains the possibility to deal with creation and annihilation of particles and certain phenomena like the correlation between particles and antiparticles can be properly described only within this formalism. Furthermore, at high energies, because of the large number of particles produced, not all particles can be detected in a given reaction and therefore one measures usually only inclusive cross sections. For these reactions the wave function formalism is impractical. Related to this is the fact that the second quantization provides through the density matrix a transparent link between correlations and multiplicity distributions. This last topic has been in the center of interest of multiparticle dynamics for the last 20 years (we refer among other things to Koba-Nielsen-Olesen (KNO) scaling and ``intermittency"). Furthermore one of the most important properties of systems made of identical bosons and which is responsible for the phenomenon of {\em lasing} is quantum statistical coherence. This feature is also not accessible to a theoretical treatment except in field theory. The present review is restricted to Bose Einstein correlations which constitute by far the majority of correlations papers both of theoretical and experimental nature. This is due to the fact that BEC present important heuristic and methodological advantages as compared to Fermi-Dirac correlations. Among the first we mention the fact that quantum coherence appears only in BEC. Among the second, one should recall that pions are the most abundantly produced secondaries in high energy reactions. In the last years there has been a considerable surge of interest in boson interferometry. This can be judged by the fact that at present there is no meeting on multiparticle production where numerous contributions to this subject are not presented. Moreover since 1990 \cite{camp} meetings dedicated exclusively to this topic are organized; this reflects the realization that BEC are an important subject in its own. This development is due in part to the fact that at present intensity interferometry constitutes the only direct experimental method for the determination of sizes and lifetimes of sources in particle and nuclear physics. Since soft strong interactions which are responsible for multiparticle production processes cannot be treated by perturbative QCD, phenomenological approaches have to be used in this domain and space-time concepts are essential elements in these approaches. That is why intensity interferometry has become an indispensable tool in the investigation of the dynamics of high energy reactions. However this alone could not yet explain the explosion of interest in BEC if it were not the search for quark-gluon plasma which has mobilized the attention of most of the nuclear physics and of an appreciable part of the particle physics community. For several reasons the space-time properties of ``fireballs" produced in heavy ion reactions are an essential key in the process of undertanding whether quark matter has been formed. The main emphasis of the present review will be on theoretical developments which took place after 1989-1990. Experimental results will be mentioned only in so far as they illustrate the theoretical concepts. For a review of older references cf. e.g. the paper by Boal, Gelbke, and Jennings \cite{boal}. In the nineties the single most important theoretical event was in our view the space-time generalization of the classical current formalism. For a detailed presentation of this generalization and its applications up to 1993 the reader is referred to ref.\cite{APW}. For a review of experimental results in $e^+-e^-$ reactions cf. \cite{Hay}. Finally a more pedagogical and more complete treatment of the theory of BEC can be found in \cite{book}. There are two categories of papers not mentioned: (i) Those which the reviewer was unaware off; he apologizes to the authors of these papers for this. (ii) Those which he considered as irrelevant or as repetitions of previous work\footnote{As a matter of fact, ``rediscoveries" are not rare in this field and it would be desirable that journals and referees should do a better job in trying to avoid such a situation, not to mention the prevention of publication of misleading statements and/or wrong results.}. The large number of papers quoted despite these restrictions shows that an exhaustive listing of references on BEC is not trivial. \setcounter{equation}{0} \section{The GGLP effect} In the period 1954-1960 Hanbury-Brown and Twiss developped the method of optical intensity interferometry for the determination of (angular sizes) of stars (cf. e.g.\cite{HBT}). The particle physics equivalent of the Hanbury-Brown Twiss (HBT) effect in optics is the Goldhaber, Goldhaber, Lee and Pais (GGLP) effect \cite{GGLP1}, \cite{GGLP} which we shall describe schematically below. However when going over from optics to particle physics the following point has to be considered: In particle physics one does not measure distances $r$ in order to deduce (differences of) momenta $k$ and thus angular sizes, but one measures rather momenta in order to deduce distances. This explains why GGLP were not aware of the HBT method \footnote{For a comparison of optical and particle physics intensity interferometry cf. e.g.\cite{book}.}. Up to a certain point there are two ways of approaching the relationship between optics and particle physics: the wave function approach and the field theoretical approach. Although the first one is only a particular case of the second one it is still useful because it allows sometimes a more intuitive understanding of certain concepts and in particular that of the distinction between boson and fermion correlations. We shall start therefore with the wave function approach. \subsection{The wave function approach} Let us consider for the beginning a source which consists of a number of discrete emission points $i$ each of which is characterized by a probability amplitude \begin{equation} F_{i}({\bf r})=F_{i}\delta({\bf r}-{\bf r}_{i}). \label{eq:emampl} \end{equation} Let $\psi_{{\bf k}}({\bf r})$ be the wave function of an emitted particle. The total probability $P({\bf k})$ to observe the emission of one particle with momentum ${\bf k}$ from this source is obtained by summing the contributions of all points $i$. If this summation is done {\em incoherently} the single particle probability reads \begin{equation} P({\bf k})=\sum_i|F_i\psi(r_i)|^2 \label{eq:incoh} \end{equation} Instead of discrete emission points consider now a source the emission points of which are continuously distributed in space and assume for simplicity that the wave functions are plane waves $\psi _{{\bf k}}({\bf r})\sim exp(i{\bf kr})$. The sum over $x_i$ will now be replaced by an integral and we have \begin{equation} P({\bf k})=\int|F({\bf r})|^2d^3r. \label{eq:Richard7} \end{equation} Similarly the probability to observe two particles with momenta ${\bf k}_1,{\bf k}_2$ is \begin{equation} P({\bf k}_1{\bf k}_2)=\int|\psi_{1,2}|^2f({\bf r}_1)f({\bf r}_2)d^3r_1d^3r_2 \label{eq:prob} \end{equation} Here $\psi_{1,2}= \psi_{1,2}({\bf k}_1,{\bf k}_2;{\bf r}_1, {\bf r}_2)$ is the two-particle wave function and we have introduced the density distribution $f= |F|^2$. Suppose now that the two particles are identical. Then the two-particle wave function has to be symmetrized or antisymmetrized depending on whether we deal with identical bosons or fermions. Assuming plane waves we have \begin{equation} \psi_{1,2}=\frac{1}{\sqrt{2}}\left[e^{i({\bf k}_1{\bf r}_1 +{\bf k}_2{\bf r}_2)}\pm e^{i({\bf k}_1{\bf r}_2 +{\bf k}_2{\bf r}_1)}\right] \label{eq:wf} \end{equation} with the plus sign for bosons and the minus sign for fermions. With this wave function one obtains \begin{equation} P({\bf k}_1,{\bf k}_2)= |\tilde f_I(0)|^2\pm|\tilde f_I({\bf k}_1-{\bf k}_2)|^2 \label{eq:P2inc} \end{equation} The incoherent summation corresponds to random fluctuations of the amplitudes $F_i$. Following the GGLP experiment \cite{GGLP} consider the space points ${\bf r}_{1}$ and ${\bf r}_{2}$ within a source, so that each point emits two identical particles (equally charged pions in the case of GGLP) with momenta ${\bf k}_{1}$ and ${\bf k}_{2}$. These particles are detected in the registration points $ x_1$ and $x_2$ so that in $x_1$ only particles of momentum $k_1$ and in $x_2$ only particles of momentum $k_2$ are registered (cf.Fig.1). Because of the identity of particles one cannot decide which particle pair originates in $r_1$ and which originates in $r_2$. Assuming that the individual emission points of the source act incoherently GGLP derived eq.(\ref{eq:P2inc}) which for bosons leads to the second order correlation function \begin{equation} C_{2}({\bf k}_1,{\bf k}_2)\equiv P_2({\bf k}_1,{\bf k}_2)/P_1({\bf k}_1)P_1({\bf k}_2) =1+|\tilde{f}({\bf q})|^2 \label{eq:Fourier} \end{equation} where $\tilde{f}$ is the Fourier transform of $f$ and ${\bf q}$ the momentum difference ${\bf k}_{1}-{\bf k}_{2}$. Eq.(\ref{eq:Fourier}) shows quite clearly how in particle physics momentum (correlation) measurements can yield information about the space-time structure of the source In the case of coherent summation one gets instead $C_2 = 1$ (cf. e.g.\cite{boal}). We retain herefrom (cf. also below) that coherence reduces the correlation and that a purely coherent source has a correlation function which does not depend on its geometry\footnote{Cf. also \cite{MP2},\cite{Pratt2} for an attempt to approach the issue of coherence and chaos by using wave packets.}. \subsubsection{Newer correlation measurements in $\bar{p}-p $ annihilation} The original GGLP experiment \cite{GGLP1} measured the correlation function in terms of the opening angle \footnote{It may be amusing to note that the use of the opening angle as a kinematical variable in BEC studies was rediscovered in 1994 in ref. \cite{Chu} where it was recommended as a tool for the investigation of final state interactions.} of a pion pair. The GGLP experiment has been repeated in the last years at LEAR (cf.e.g. \cite{Angel} where also older references are quoted \footnote{For a reanalysis of the results of an older annihilation experiment cf. \cite{Gaspero}.}). In these newer experiments the three momenta of particles were measured, although one continued to use as a variable for the correlation function the invariant momentum difference $Q$, as suggested already in \cite{GGLP}. One of the remarkable observations made in all annihilation reactions is the fact that the intercept of the second order correlation function $C_2(k,k)$ appears consistently to exceed the canonical value of $2$ reaching values up to $4$. (This effect was possibly not seen in the original experiment \cite{GGLP1} because of the averaging over the magnitudes of the momenta.) In \cite{Song} this effect was attributed to resonances while subsequently \cite{Amado},\cite{Angel} a (non-chaotic) Skyrmion type superposition of coherent states was proposed as an alternative explanation. Another possible explanation of this intriguing observation may be that in annihilation processes squeezed states (cf. section 2.2) are produced, while this is not the case in other processes. Indeed, it is known that squeezed states can lead to overbunching effects (cf. below). Furthermore, as shown in ref.(\cite{sqIgor}), squeezed states may be preferentially produced in rapid reactions. Or, according to some authors \cite{Amado}, annihilation is a more rapid process than other reactions occuring at higher energies. Given the importance of squeezed states, further experimental and theoretical studies of this issue are highly desirable \footnote{In ref. \cite{sqIgor} the effect of chaotic superpositions of squeezed states in BEC was also studied. It is shown that such a superposition leads always to {\em overbunching}, while pure squeezed states can lead also to antibunching.}. \subsubsection{Resonances, apparent coherence and other experimental problems; the $\lambda$ factor} It is known that in multiparticle production processes, an appreciable fraction of pions originate from resonance decays. Resonances act in two opposite ways on the correlation function of pion pairs. On one hand the interference between pions originating from short lived resonances and ``directly" produced pions leads to a narrow peak in the second order correlation function at small values of $q$ \cite{Grishin}, \cite{Grass}. On the other hand long lived resonances give rise to pions which are beyond the range of detectors and this leads to an apparent decrease of the intercept of the second order correlation function $C_{2}(k,k)$. These modifications of the intercept are important among other things, because their understanding is essential in the search for coherence through the intercept criterion. As mentioned already one of the most immediate consequences of coherence is the decrease of the correlation function at small $q$. Because of the large number of different resonances produced in high energy reactions a quantitative estimate of their effects is possible only via numerical techniques. In sections 4.9.1 and 5.1.3 we will present a more detailed discussion of the influence of resonances on BEC within the Wigner function formalism. For older references on this topic cf. also \cite{boal}, \cite{Ledn}, \cite{Thomas}. Another experimental difficulty is that, because of limited statistics or certain technical problems, sometimes not all degrees of freedom can be measured in a given expriment. This leads to an effective averaging over the non-measured degrees of freedom and hence also to an effective reduction of the correlation function. As a matter of fact BEC experiments have shown from the very beginning that the extrapolation of the correlation function to $q=0$ almost never led to the maximum value $C_2(0)=2$ permitted by eq.(\ref{eq:Fourier}). To take into account empirically this effect experimentalists introduced into the correlation function a correction factor $\lambda$. Thus eq.(\ref{eq:Fourier}) e.g. was modified into \begin{equation} C_{2}({\bf k}_1,{\bf k}_2)=1+\lambda |\tilde{f}({\bf q})|^2 \label{eq:lambda} \end{equation} Because one initially thought that formally this generalization offers also the possibility to describe partially coherent sources, the corresponding parameter $\lambda$ was postulated to be limited by (0,1) and called the ``incoherence" factor: indeed $\lambda = 0$ leads to a totally coherent source and $\lambda= 1$ to a totally chaotic one. Unfortunately this nomenclature is not quite correct, as will be explained below. It should also be mentioned that there exists a strong correlation between the empirical values of $\lambda$ and and those of the ``radius" which enters $\tilde{f}$ (cf. e.g. \cite{Peitz} for a special study of this issue.) \subsubsection{The limitations of the wave function formalism} The wave function formalism presented in the previous subsection has severe shortcomings: a) The correlation function (\ref{eq:Fourier}) depends only on the difference of momenta $q$ and not also on the sum ${\bf k}_1+{\bf k}_2$, in contradiction with experimental data. Although we will see in section 4.3 that this limitation disappears automatically in the quantum statistical (field theoretical) approach, it can be remedied also within the first quantization formalism by using the Wigner function approach. In this approach the source function is from the beginning a function both of coordinates and momenta and therefore the correlation function depends on $k_1$ and $k_2$ separately. There is of course a price to pay for this procedure as it involves a semiclassical approximation \footnote{ String models \cite{stringbowl}\cite{stringander} also use a Wigner type formalism. Here it is postulated that there exists a ``formation" time $\tau$ and therefore the particle production points are distributed around $t^2-x^2=\tau^2$. This implies among other things a correlation between particle production points and momenta.}. b) The wave function formalism may be useful when exclusive reactions are considered as was the case e.g. in the GGLP paper. Indeed $\psi_{1,2}$ in eq.(\ref{eq:wf}) is just the wave function of the two-boson system i.e. the assumption is made that two and only two bosons are produced. At low energies, i.e. low average multiplicities, this condition can be satisfied. However at high energies the identification of all particles is very difficult and up to now has not been done. Therefore one measures practically always inclusive cross sections. This means that instead of (\ref{eq:wf}) one would have to use a wave function which describes the two bosons in the presence of all other produced particles. To obtain such a wave function one would have to solve the Schr\"odinger equation of the many body, strongly interacting system, which is not a very practical proposal. Related to this is the difficulty to treat higher order correlations within the wave function formalism. c) The correlation function $C_2$ in the wave function formalism is independent of isospin and is thus the same for charged and neutral particles. We shall see in section 4.4 that in a more correct quantum field theoretical approach, this is not the case. This will affect among other things the bounds of the correlations and will lead to quantum statistical particle-antiparticle correlations which are not expected in the wave function formalism. d) Coherence cannot be treated adequately (cf. below). The correlation functions derived in this subsection refer in general to incoherent sources and attempts to introduce coherence within the wave function formalism are rather ad-hoc parametrizations. However coherence is the most characteristic and important property of Bose-Einstein correlations among other things because it is the basis of the phenomenon of Bose Einstein {\em condensates} found in many chapters of physics, like superconductors, superfluids, lasers, and the recently discovered atomic condensates \cite{Atomcond97}. It would be very surprising if coherence would not be found also in particle physics given the fact that the wave lengths of the emitted particles are of the same order as that of the sources. Furthermore as pointed out in this connection in \cite{FW} modern particle physics is based on spontaneously broken symmetries. The associated fields are coherent. That is why one of the main motivations of BEC research should be the measurement of the amount of coherence in strong interactions. For this purpose the formalism of BEC has to be generalized to include the presence of (partial) coherence and this again can be done correctly only within quantum statistics i.e. quantum field theory. We conclude this subsection with the observation that the wave function or wave packet approach may nevertheless be useful in BEC for the investigation of final state interactions (cf. below) or for the construction of event generators, where phases or quantum amplitudes are ignored anyway. Also for correlations between fermions where coherence is absent the wave function formalism may be an adequate substitute, although here too a field theoretical approach is possible. \subsection{Quantum optical methods in BEC} In high energy processes in which the pion multiplicity is large, we may in general expect the methods of quantum statistics (QS) \footnote{By QS we understand in the following the density matrix formalism within second quantization.} to be useful. For BEC in particular they turn out to be indispensable. These methods have been applied with great success particularly in quantum optics (QO), superfluidity, supercoductivity etc. What distinguishes optical phenomena from those in particle physics are conservation laws and final state interactions which are present in hadron physics. At high energies and high multiplicities the first are unimportant. Neglecting for the moment also the final state interactions, QS reduces then to QO and we may take over the formalism of QO to interpret the data on multipion production at high energies, provided we consider identical pions. Given the general validity of QS (or QO), it is then clear that any model of multiparticle production must satisfy the laws of quantum statistics and this has far reaching consequences, independent of the particular dynamical mechanism which governs the production process. The main tools in the QO formalism are defined below \footnote{A review of the applications of quantum optical methods to multiparticle production up to 1988 can be found in \cite{applic}.}. \paragraph{Coherent states and squeezed coherent states} Coherent states $|\alpha>$ are eigenstates of the (one-particle) annihilation operator $a$ \begin{equation} a|\alpha>=\alpha|\alpha>. \label{eq:coherent states} \end{equation} Squeezed coherent states are eigenstates $|\beta>_s$ of the two-particle annihilation operator \begin{equation} b=\mu a+\nu a^{\dagger} \label{eq:2.66} \end{equation} with \begin{equation} |\mu|^2-|\nu|^2=1 \end{equation} so that \begin{equation} b|\beta>_s=\beta|\beta>_s. \label{eq:2.69} \end{equation} One of the remarkable properties of these states which explains also their name is that for them the uncertainty in one variable can be {\em squeezed} at the expense of the other so that \begin{equation} (\Delta q)^2_{s}\le\frac{1}{2\omega},\;(\Delta p)^2_{s}\ge\frac{\omega}{2}, \label{eq:Walls 2.54 sq} \end{equation} or vice versa. The importance of this remarkable property lies among other things in the possibility to reduce quantum fluctuations and this explains the great expectations associated with them in communication and measurement technology as well as their interest from a heuristic point of view. It has been found recently in \cite{sqIgor} that squeezed states appear naturally when one deals with rapid phase transitions (explosions). Indeed consider the transition from a system $a$ to a system $b$ and assume that it proceeds rapidly enough so that the relation between the creation and annihilation operators and the corresponding fields in the two ``phases" remains unchanged. Mathematically this process will be described by postulationg at the moment of this transition the following relations between the generalized coordinate $q$ and the generalized momentum $p$ of the field: \begin{eqnarray} q &=& \frac{1}{\sqrt{2E_b}} (b^{\dagger} + b) = \frac{1}{\sqrt{2E_a}} (a^{\dagger} + a) \nonumber\\ p &=& i \sqrt{\frac{E_b}{2}} (b^{\dagger} - b) = i \sqrt{\frac{E_a}{2}} (a^{\dagger} - a) \label{eq:qp} \end{eqnarray} $a^{\dagger},a$ are the free field creation and annihilation operators in the ``phase $a$" and $b^{\dagger},b$ the corresponding operators in the ``phase $b$" . Eq. (\ref{eq:qp}) holds for each mode $p$. Then we get immediately a connection between the $a$ and $b$ operators, \begin{eqnarray} a &=& b \; \cosh \; r \; + \; b^{\dagger} \; \sinh \; r \; , \nonumber\\ a^{\dagger} &=& b \; \sinh \; r \; + \; b^{\dagger} \; \cosh \; r \; \label{eq:aa} \end{eqnarray} with \begin{equation} r = r ({\bf p}) = \frac{1}{2} \, \log \, (E_a/E_b) \, . \label{eq:rr} \end{equation} The transformation (\ref{eq:aa}) is just the squeezing transformation (\ref{eq:2.66}) with \begin{equation} \mu = \cosh r; \qquad \nu = \sinh r \label{eq:expl} \end{equation} which proves the statement made above. The observation of squeezed states in BEC may thus serve as a signal for such rapid transitions. Furthermore the existence of isospin induces in hadronic BEC certain effects which are specific for squeezed states. This topic will be discussed in section 4.4. >From the point of view of BEC what distinguishes ordinary coherent states from squeezed states is the following: for coherent states the intercept $C_2(k,k) = 1$ while for squeezed states it can take arbitrary values. In Fig.2 from ref.\cite{VW} one can see such an example. \paragraph{Expansions in terms of coherent states} Coherent states form an (over)complete set so that an arbitrary state $|f>$ can be expanded in a unique way in terms of these states. Of particular use is the expansion of the density matrix $\rho$ in terms of coherent states. For a pure coherent state the density operator reads \begin{equation} \rho=|\alpha><\alpha|. \label{eq:7.1} \end{equation} For an arbitrary density matrix case we have \begin{equation} \rho=\int{\cal P}(\alpha)|\alpha><\alpha|d^2\alpha. \label{eq:7.6} \end{equation} Here ${\cal P}$ is a weight function which usually, but not always, has the meaning of a probability. The normalization condition for the density operator translates in terms of the ${\cal P}$ representation as \begin{equation} tr\rho=\int{\cal P}(\alpha)d^2\alpha=1. \label{eq:norm1} \end{equation} Eq.(\ref{eq:7.6}) is called also the Glauber-Sudarshan representation. The knowledge of ${\cal P}(\alpha)$ is (almost) equivalent to the knowledge of the density matrix. However in most cases the exact form of ${\cal P}(\alpha)$ is not accessible and one has to content oneself with certain approximations of it. Among these approximations the Gaussian form is priviledged because: (i) one can prove that ${\cal P}(\alpha )$ is of Gaussian form for a certain physical situation which is frequently met in many body physics. (ii) its use introduces an enormous mathematical simplification. Proposition (i) is the subject of the central limit theorem which states that if 1. the number of sources becomes large; 2. they are stationary in the sense that their weight function ${\cal P}(\alpha)$ depends only on the absolute value $|\alpha|$; 3. they act independently, then ${\cal P}(\alpha)$ is Gaussian. These conditions are known to be fulfilled in most cases of optics and presumably also in high energy physics. Chaotic fields and in particular systems in thermal equilibrium are described by a Gaussian density matrix. One of the reasons why the Gaussian form for ${\cal P}$ plays such an important part in correlation studies is the fact that for a Gaussian ${\cal P}(\alpha)$ all higher order correlations can be expressed in terms of the first two correlation functions. On the other hand the coherent state representation is particularly important for correlation studies because in this representation all correlation functions can be expressed in terms of the creation and annihilation operators $a$ and $a^{\dagger}$ of the fields (particles). This follows from the Fourier expansion of an arbitrary field in second quantization \begin{equation} \pi(x)=\sum_k [a_ke^{-ikx} +a^{\dagger}_ke^{ikx}]. \label{eq:2.12} \end{equation} This property will be used extensively in section 4.2 within the classical current formalism. \paragraph{Correlation functions} The first order correlation function reads \begin{equation} G^{(1)}(x,x^{'})\equiv Tr[\rho\pi^{\dagger}(x)\pi(x^{'})] \label{eq:corr1} \end{equation} Higher (n-th) order correlation functions are defined analogously by \begin{equation} G^{(n)}(x_1...x_n,x_{n+1}...x_{2n})\equiv Tr[\rho\pi^ {\dagger} (x_1)...\pi^{\dagger}(x_n)\pi(x_{n+1})...\pi(x_{2n})] \label{eq:corrn} \end{equation} In quantum field theory because of the mathematical complexity of the problem exact solutions of the field equations are available only in special cases. One such case will be discussed later on. However for strong interactions \footnote{Strong interactions are present not only in hadronic physics but also in quantum optics.} even for this case one has to use phenomenological parametrizations of the correlation functions and determine the parameters (which have a definite physical meaning) by comparing with experiment. In optics for stationary chaotic fields two particular parametrizations are used:\\ 1.Lorentzian spectrum: \begin{equation} G^{(1)}(x_1,x_2)=<n_{ch}>e^{-|x_1-x_2|/\xi} \label{eq:Lorentz} \end{equation} 2.Gaussian spectrum: \begin{equation} G^{(1)}(x_1,x_2)=<n_{ch}>e^{-|x_1-x_2|^2/\xi^2} \label{eq:Gaussdef} \end{equation} $\xi$ is the {\em coherence length} in x-space and $<n_{ch}>$ is the mean number of particles associated with the chaotic fields. In \cite{FW} it was proposed to use the analogy between time and rapidity in applying the methods of quantum optics to particle physics. Indeed in optics processes are usually stationary in time while in particle physics the corresponding stationary variable (in the rapidity plateau region) is rapidity \footnote{This property does not hold for other variables.}. Pure coherent or pure chaotic fields are just extreme cases. In general one expects {\em partial coherence} i.e. a superposition of coherent and chaotic fields \begin{equation} \pi = \pi_{coherent} + \pi_{chaotic}. \label{eq:supergen} \end{equation} This leads for the Lorentzian case e.g.to a second order correlation function of the form \begin{equation} C_{2}(x,x')=1+2p(1-p)e^{-|x-x'|/\xi}+p^{2}e^{-2|x-x'|/\xi} \label{eq:super} \end{equation} where $p$ is the chaoticity, which varies between $0$ (for purely coherent sources) and $1$ (for totally chaotic sources). Eq. (\ref{eq:lambda}) is seen to be a particular form of the above equation for $\lambda =p=1$ and it is clear herefrom that $\lambda$ does not describe (partial) coherence as its name would imply. The presence of coherence introduces a new term into $C_2$. However it is remarkable that the number of free parameters in eq.(\ref{eq:super}) is the same as in eq.(\ref{eq:lambda}). Formally it appears as if there would act two sources rather than one, but the ``weights" and the space-time characteristics of these two sources are in a well defined relationship. This circumstance had been forgotten up to 1989 \cite{revisit} both by experimentalists and theorists. The reason for this collective amnesia is the fact that during the eighties the wave function formalism was dominating the BEC literature, especially the experimental one. >From the foregoing discussion it should be clear that there are various reasons besides coherence why the bunching effect in BEC is reduced. However it should also be clear that the empirical description of this state of affairs through the $\lambda$ factor is possible only for totally chaotic sources. Since in an experiment this is never known a priori, this implies that the fitting of data with a formula of this type is misleading and should be avoided, the more so that the correct formula (\ref{eq:super}) does not contain more free parameters than eq.(\ref{eq:lambda}). In the example presented above these free parameters are $p$ and $\xi$ for eq. (\ref{eq:super}) and $\lambda$ and the effective radius (which enters in $\tilde{f}$) for eq.(\ref{eq:lambda}) respectively. Using the rapidity-time analogy of ref.\cite{FW} for a partially chaotic field the second order correlation function in rapidity is then given by eq.(\ref{eq:super}), with $x$ being replaced by rapidity $y$. The fact that the last two terms in eq.(\ref{eq:super}) are in a well defined relationship and depend in a characteristic way on the two parameters $p$ and $\xi$ is a consequence of the superposition of the two {\em fields} (coherent and chaotic) and distinguishes a partially coherent source from a source which is a superposition of two independent {\em chaotic intensities}. Because of this the form (\ref{eq:super}) was proposed in \cite{revisit} to be used as a signal for detection of coherence in BEC \footnote{Eq. (\ref{eq:super}) is a special case of superposition of coherent and chaotic fields; it can be considered as corresponding to point-like coherent and chaotic sources and a momentum independent chaoticity; superpositions of more general finite-size sources are considered in section 4.4 (eq.\ref{eq:c2plus1}) and section 4.8.}. As a matter of fact, an attempt in this direction was done in an experimental study by Kulka and L\"orstad \cite{KL}. In this analysis BEC data from $pp$ and $\bar{p}p$ reactions at $\sqrt{s}=53 GeV$ were used to compare various forms of correlation functions. Among other things one considered formulae of QO type for rapidity \begin{eqnarray} C_{2}=1+2p(1-p)e^{-|y_1-y_2|/\xi}+p^{2}e^ {-2|y_1-y_2|/\xi} \label{eq:superlor} \end{eqnarray} (corresponding to a Lorentzian spectrum) and \begin{eqnarray} C_{2}=1+2p(1-p)e^{-|y_1-y_2|^{2}/\xi^{2}}+p^{2}e^ {-2|y_1-y_2|^{2}/\xi^2} \label{eq:superpgauy} \end{eqnarray} (corresponding to a Gaussian spectrum) as well as arbitrary superpositions of two chaotic sources of exponential or Gaussian form respectively. \begin{eqnarray} C_{2}=1+\lambda_{1}e^{-|y_1-y_2|/\xi}+\lambda_{2}e^ {-2|y_1-y_2|/\xi} \label{eq:superlory} \end{eqnarray} \begin{eqnarray} C_{2}=1+\lambda_{1}e^{-|y_1-y_2|^2/\xi^2}+ \lambda_{2}e^{-2|y_1-y_2|^2/\xi^2} \end{eqnarray} Here $\lambda_{1}$ and $\lambda_{2}$ represent arbitrary weights of the two chaotic sources. Because of the limited statistics no conclusion could be drawn as to the preference of the QO form versus the two-source form. Similar inconclusive results were obtained when one replaced in the above equations $y_1-y_2$ by the invariant momentum difference $Q^2=(k_1-k_2)^2$. \subsubsection{Higher order correlations} We have mentioned above that a characteristic property of the Gaussian form of density matrix (not to be confused with the Gaussian form of the correlator or the Gaussian form of the space-time distribution) is the fact that all higher order correlation functions are determined just by the first two correlation functions. Since all BEC studies in particle physics performed so far assume a Gaussian density matrix, the reader may wonder why it is necessary to measure higher order correlation functions. There are at least three reasons for this: (i) The conditions of the applicability of the above theorem and in particular the postulate that the number of sources is infinite and that they act independently can never be fulfilled exactly. (ii) In the absence of a theory which determines from first principles the first two correlation functions, models for these quantities are used, which are only approximations. The errors introduced by these phenomenological parametrizations manifest themselves differently in each order and thus violate the above theorem even if (i) would not apply. (iii) In experiments, because of limited statistics and sometimes also because of theoretical biases not all physical observables are determined, but rather averages over certain variables are performed, which again introduce errors which propagate (and are amplified) from lower to higher correlations. Conversely, by comparing correlation functions of different order one can test the applicability of the theorem quoted above and pin down more precisely the parameters which determine the first two correlation functions (e.g. the chaoticity $p$ and the correlation length $\xi$ in eqs. (\ref{eq:superlor}), (\ref{eq:superpgauy})), which is essentially the purpose of particle interferometry. The phenomenological application of these considerations will be discussed in the following as well as in section 6.3 for the particular case of the invariant $Q$ variable, but the arguments (i),(ii),and (iii) have general validity. It would be a worthwhile research project to compare the deviations introduced in the relation between lower and higher correlation functions, due to (i) with those introduced by (ii) and (iii). The simplification brought by the variable $Q$ can be enhanced by a further approximation proposed by Biyajima et al. \cite{Biya90}. With the notation $Q_{ij}=k_i-k_j$ the analogue of eq. (\ref{eq:superpgauy}) can be written \begin{equation} C_2 = 1 + 2p(1-p)\exp(-R^{2}Q_{12}^{2}+p^{2}\exp(-2Q_{12}^ {2}R^{2}) \label{eq:C2Gauss} \end{equation} For the third order correlation function one obtains \begin{eqnarray} C_3 &=& 1+2p(1-p)[\exp\{-R^2Q^2_{12}\}+\exp\{-R^2Q^2_{13}\} +\exp\{-R^2Q^2_{23}\}]\nonumber\\ & & +p^2[\exp\{-2R^2Q^2_{12}\}+\exp\{-2R^2Q^2_{13}\}+\exp \{-2R^2Q^2_{23}\}]\nonumber\\ & & +2p^2(1-p)[\exp\{-R^2(Q^2_{12}+Q^2_{23})\}+\exp \{-R^2(Q^2_{13}+Q^2_{23})\}\nonumber\\ & &+\exp\{-R^2(Q^2_{12}+Q^2_{13})\}]+2p^3[\exp\{-R^2(Q^2_{12} +Q^2_{13}+Q^2_{23})\}] \label{eq:Biya11} \end{eqnarray} Ref.\cite{Biya90} proposed to use symmetrical configurations for all two-particle momentum differences i.e. to consider $Q_{ij}$ independent of $(i,j)$. For $C_2$ this assumption does not of course introduce any modifications. However for higher order the simplification is important. Thus e.g. for the third order correlation with $Q_{12}=Q_{13}=Q_{23}$ and with the definition $Q^{2}_{three} = Q^{2}_{12}+Q^{2}_{13}+Q^{2}_{23}$ eq.(\ref{eq:Biya11}) becomes \begin{eqnarray} C_3 &=& 1+6p(1-p)\exp\left\{-\frac{1}{3}R^2 Q^2_{three}\right\}\nonumber\\ & & +3p^2(3-2p)\exp\left\{-\frac{2}{3}R^2Q^2_{three}\right\} +2p^3\exp\{-R^2Q^2_{three}\} \label{eq:Biya9} \end{eqnarray} In ref.\cite{Biya90} similar expressions for $C_4$ and $C_5$, again for a Gaussian correlator, were given. These relations for higher order BEC were subjected to an experimental test in ref.\cite{UA1}, using the UA1 data for $\bar{p}p$ reactions at $\sqrt s = 630$ and $900 GeV$. For reasons which will become clear immediately, we discuss here this topic in some detail. The procedure used in \cite{UA1} for this test consisted in determining $R$ and $p$ separately for each order $q$ of the correlation and comparing these values for different $q$. It was found that a Gaussian correlator did not fit the data. Next in \cite{UA1} one tried to replace the Gaussian correlator by an exponential (cf. eq.(\ref{eq:superlor}). To do this one substituted simply in the expressions for the correlation functions of ref.\cite{Biya90} the factor $\exp(-R^{2}Q^{2})$ with $\exp(-RQ)$. Such a procedure was at hand given the fact that for $C_2$ the QO formulae both for an exponential correlator and a Gaussian correlator were known \cite{revisit} and their comparison suggested just this substitution, as can be seen from eq.(\ref{eq:superpgauy}). In \cite{UA1} one used then for the exponential correlator the relations \begin{equation} C_{2}^{empirical} = 1 + 2p(1-p)\exp(-RQ_{12})+p^{2}\exp(-2Q_ {12}R) \label{eq:C2exp} \end{equation} and \begin{eqnarray} C^{empirical}_{3} &=& 1+6p(1-p)\exp\left(-\frac{1}{3} RQ_{three}\right)\nonumber\\ & &+3p^2(3-2p)\exp\left(-\frac{2}{3}RQ_{three}\right)+ 2p^3\exp\left(-RQ_{three}\right) \label{eq:C3wrong} \end{eqnarray} With these modified formulae one still could not find in \cite{UA1} a unique set of values $p$ and $R$ for all orders of correlation functions. However now a clearer picture of the ``disagreement" between the QO formalism and the data emerged. It seemed that while the parameter $p$ was more or less independent of $q$, the radius $R$ increased with the order $q$ in a way which could be approximated by the relation \footnote{This reminds one of Polonius's bewilderment in Shakespeare's Hamlet: ``Though this be madness, yet there is method in't".} \begin{equation} R_q=R\sqrt {\frac{1}{2}q(q-1)}. \label{eq:emp} \end{equation} Indeed in \cite{mad} it was shown that the findings of \cite{UA1} and in particular eq.(\ref{eq:emp}) not only did not contradict QS but on the contrary constituted a confirmation of it. While eq.(\ref{eq:C2exp}) for the second order correlation function coincides with that derived in quantum optics for an exponential spectrum, this is not the case with the expressions for higher order correlations $C_q^{empirical}$ (eq.(\ref{eq:C3wrong})). The formulae for $C_3$, $C_4$ and $C_5$ corresponding in QS to an exponential correlator and derived in \cite{mad} differ from the empirical ones used in \cite{UA1}. As an example we quote \begin{eqnarray} C_3 &=& 1+6p(1-p)\exp\left(-\frac{1}{\sqrt{3}}RQ_{three} \right)\nonumber\\ & & +3p^2(3-2p)\exp\left(-\frac{2}{\sqrt{3}}RQ_{three} \right)+2p^3\exp(-\sqrt{3}RQ_{three}) \label{eq:C3correct} \end{eqnarray} As observed in \cite{mad} and as one easily can check by comparing eq.(\ref{eq:C3wrong}) with eq.(\ref{eq:C3correct}), one can make coincide the empirical formulae for $C_q$ used in \cite{UA1} with the correct ones, by replacing the parameter $R$ with a scaled parameter $R_s$ and the relation between $R$ and $R_s$, is, nothing else but $R_s=R_q$, where $R_q$ is given by eq.(\ref{eq:emp}). The fact that this happens for three different orders, i.e. for $C_3, C_4$, and $C_5$ makes a coinicidence quite improbable. By empirically modifying the formulae of higher order BEC for the exponential case, paper \cite{UA1} had explicitely violated QS and the ``phenomenological" relationship (\ref{eq:emp}) between $R$ and $q$ just compensated this violation. \begin{footnotesize} The fact that this compensation and the final agreement between theory and experiment was not perfect is not surprising and is discussed in \cite{mad}. Besides the reasons (i), (ii), (iii) mentioned above, one has to take into account that the QO formalism on which eqs.(\ref{eq:C2exp}),(\ref{eq:Biya11}) are based assumes stationarity in $Q$, i.e. assumes that the corrrelator depends only on the difference of momenta $k_1-k_2$ and not also on their sum. As mentioned already this condition is in general not fulfilled in BEC. Furthermore, the parameter $p$, if it is related to chaoticity, is in general momentum dependent (cf. section 4.8). Also, the symmetry assumption, $Q(i,j)$ independent of $i,j$, may be too strong. Besides these theoretical caveats, there are also experimental problems, related to the fact that the $UA1$ experiment is not a dedicated BEC experiment and thus suffers from specific diseases, which are common to almost all particle physics BEC experiments performed so far. Among other things, there is no identification of particles (only $85\%$ of the tracks recorded are pions), and the normalization of correlation functions is the ``conventional" one, i.e. not based on the single inclusive cross sections as the definition of correlation functions demands (cf.(\ref{eq:Fourier}) and section 4.11), but rather uses an empirically determined ``background" ensemble.\end{footnotesize} The very fact that QS was effectively ``reinvented" by an unbiased experimental team is in our view overwhelmingly convincing. It may enter the history of physics as a classical and quite unique example of the intimate relationship between theory and experiment. Ref.\cite{mad} concludes that experiment \cite{UA1} supports QS in general and the standard form of the density matrix in particular. From an epistemological point of view, the message of this ``conspiracy" is highly significant: {\em QS is robust enough to resist attempts of falsification}. In the mean time further theoretical and experimental developments took place. On the theoretical side a new space-time approach to BEC was developped \cite{APW}, \cite{aw} which is more appropiate to particle physics and which contains as a special case the QO formalism. In particular the two exponential feature of the correlation function is recovered. On the experimental side a new technique for the study of higher order correlations was developed, the method of correlation integrals which was applied \cite{UA1new} to a subset of the UA1 data in order to test the above quoted QO fromalism. The fits were restricted to second and third order cumulants only. Again it was found that by extracting the parameters $p$ and $R$ from the second order data, the ``predicted" third order correlation, this time by using a correct QO formula, differed significantly from the measured one. If confirmed, such a result could indicate that the QO formalism provides only a rough description of the data and that higher precision data demand also more realistic theoretical tools. Such tools are the QS space-time approach to BEC presented in section 4.3. A further, but more remote possibility would be to look for deviations from the Gaussian form of the density matrix. However it seems premature to speculate along these lines given the fact that the procedure used to test the relation between the second and third order correlation functions may have to be qualified. Indeed in \cite{UA1new} one did not perform a {\em simultaneous} fit of second and third order data to check the QO formalism. Such a simultaneous fit appears necessary before drawing conclusions, because as mentioned above (cf.(ii) and (iii)), the errors involved in ``guessing" the form of the correlator, and the fact that the variable $Q$ does not characterize completely the two-particle correlation, limit the applicability of the theorem which reduces higher order correlations to first and second ones. As a matter of fact, it was found \cite{Nelly} (cf. also section 6.3) in a comparison of the QS space-time approach with higher order correlation data, that the second order correlation data is quite insensitive to the values of the parameters which enter the correlator, while once higher order data are used in a simultaneous fit, a strong delimitation of the acceptable parameter values results. Thus there are several possible solutions if one restricts the fit to the second order correlation and the correct one among those can be found only by fitting {\em simultaneously} all correlations. If by accident one chooses in a lower correlation the wrong parameter set, then the higher correlations cannot be fitted anymore\footnote{The fact that in \cite{UA1new} the correlations were normalized by mixing events rather than by comparing with the single inclusive cross sections in the same event, as prescribed by the definition of the correlation functions may also influence the applicability of the central limit theorem.}. Before ending these phenomenological considerations in which the variable $Q$ played a major part, a few remarks about its use may be in order. \paragraph{The invariant momentum difference $Q$} BEC studies in particle physics use often a priviledged variable namely the squared momentum difference \begin{equation} Q^2=(k_1-k_2)^2 = ({\bf k}_1-{\bf k}_2)^2-(E_1-E_2)^2. \label{eq:Q} \end{equation} It owes its special role to the fact that it is a relativistic invariant and was used already in the pioneering paper by GGLP \cite{GGLP}. It has also the advantage that it involves all four components of the momenta $k$ simultaneously so that the intercept of the correlation function $C_2({\bf k},{\bf k})$ coincides with $C_2(Q=0)$. Thus by measuring $C_2$ as a function of one single scalar quantity $Q$ one gets automatically the intercept. This is not the case with other single scalar quantities used in BEC like $y_1-y_2$ or $k_{\bot,1}-k_{\bot,2}$ which characterize only partially the intercept. On the other hand, $Q$ suffers from certain serious diseases which make its use for practical interferometrical purposes questionable. The first and most important deficiency of $Q$ is the fact that it mixes time and space coordinates: the associated quantity $R$ in the conventional parametrization of the correlation function $C_2= 1+\lambda \exp(-R^2Q^2)$ is neither a radius nor a lifetime, but a combination of these, which cannot be easily disentangled. Another deficiency of $Q$, which is common to all single scalar quantities is the circumstance that it does not fully characterize the correlation function. Indeed the second order correlation function $C_2$ is in general a function of six independent quantities which cannot be replaced by a single variable. An improvement on $Q$ was proposed by Cramer \cite{Cramer91} with the introduction of {\em coalescence} variables which constitute a set of three boost invariant variables to replace, for a pair of identical particles, the single variable $Q$. They are related to $Q$ by $Q^2 = 2m^2(C^2_{L}+C^2_{T}+C^2_{R})$, where $C_L, C_T, C_R$ denote longitudinal, tranverse and radial coalescence respectively, and $m$ the mass of the particle. They have the properties that $C_L=0$ when $y_1=y_2, C_T=0$ when $\phi_1=\phi_2$ and $C_R=0$ when either $m_1=m_2$ or $k_1=k_2$. Here $m_i$ is the transverse mass and $\phi$ the azimuthal angle in the transverse plane. It is shown in \cite{Cramer91} that with these new variables a Lorentz invariant separation of the space-like and time-like characteristics of the source is possible, within the kinematical assumptions involved by the particular choice of the coalescence variables. This separation is however rather involved. In \cite{Cramer91} the coalescence variables are used for the introduction of Coulomb corrections into second and higher order correlation functions. Another way to compensate in part for the fact that one single variable does not characterize completely the two-particle system, but which maintains the use of $Q$ is, as was explained above, to consider higher order correlations. \section{Final state interactions of hadronic bosons} One of the most important differences between the HBT effect in optics and the corresponding GGLP effect in particle physics is the fact that in the first case we deal with photons while in the second case with hadrons. While photons in a first approximation do not interact, hadrons do interact. This interaction has two effects: (i) it influences the correlation between identical hadrons and (ii) it leads to correlations also between non-identical hadrons. This review deals only with correlations due to the identity of particles and in particular with Bose-Einstein correlations. Therefore only effect (i) will be discussed \footnote{The reader interested in correlations between non-identical particles is refered to \cite{boal} for the period up to 1990; for more recent literature cf. \cite{Ard}, where correlations in low energy heavy ion reactions are reviewed.}. Effect (i) is usually described in terms of final state interactions. In some theoretical studies (cf. e.g. \cite{AHR} and references quoted there) emission of particles at different times is also treated as an effective final state interaction. >From the BEC point of view the final states interaction constitutes in general an unwanted background, which has to be substracted in order to obtain the ``true" quantum statistical effect on which interferometry measurements are based. That this is not always a trivial task will be shown in the following. There are two types of final state interactions in hadronic interferometry: Electromagnetic, traded under the generic name of Coulomb interactions, and strong. Furthermore one distinguishes between one-body final state interactions and many body final state interactions. \subsection{Electromagnetic final state interactions} The plane wave two-body function used in the considerations above (cf. eq.(\ref{eq:wf})) applies of course only for non-interacting particles. As a first step towards a more general treatment consider charged particle interferometry. As a matter of fact the vast majority of BEC studies, both of experimental and theoretical nature, refer to charged pions. For two-particle correlations, we will have to consider the interaction of each member of a pair with the charge of the source and the Coulomb interaction between the two particles constituting the pair. The first effect will affect primarily the single particle probabilities and is not expected to depend on the momentum difference $q$. Attention has been paid so far mostly to the second effect, i.e the modification of the two-particle wave function due to the Coulomb interaction between the two particles. While initially, having in mind the Gamow formula, it was assumed that this effect is (for small $q$ values) quite important, at present serious doubts about these estimates have arisen. The model dependence of corrections for this effect makes it almost imperative that experimental data should be presented also without Coulomb corrections, so that it should be left to the reader the possibility of introducing (or not introducing) corrections according to her/his own prejudice. \subsubsection{Coulomb correction and ``overcorrection"} As usual one separates the center of mass motion from the relative motion. For the last one the scattering wave function reads: \begin{equation} u({\bf r})\longrightarrow_{r \to \infty}e^{ikz}+r^{-1}g (\theta,\phi)e^{ikr} \label{eq:scatt} \end{equation} where the relative position vector ${\bf r}$ has polar coordinates $r, \theta,\phi$. The form of the function $g$ depends on the scattering potential. In the Coulomb case the corresponding Schr\"odinger equation can be solved exactly and the correction to the two-particle wave function and the correlation function can be calculated. So far three more and more sophisticated procedures were used for this purpose. In \cite{GKW} the value of the square modulus of the wave function $u$ in the origin $r=0$ was proposed as a correction term $G$ to the correlation function $C_2$. Up to non-interesting factors this is the Gamow factor which reads \begin{equation} |u(0)|^2=\frac{2\pi\eta}{\exp(2\pi\eta)-1}=G(\eta). \label{eq:Bowl2} \end{equation} Here \begin{equation} \eta=\alpha m_{\pi}/q \label{eq:Gamov} \end{equation} and $=|k_{1} -k_{2}|$. $m_{\pi}$ is the pion mass. However as pointed out by Bowler \cite{Bowler} and subsequently also by others, in BEC this approximation may be questionable. Indeed in a typical $e^{+}-e^{-}$ reaction e.g. the source which gives rise to BEC has a size of the order of $1 fm$ which is a large number compared with a typical ``Coulomb length" $r_{t}$ defined as the classical turning point where the kinetic energy balances the potential Coulomb energy: \begin{equation} \tilde{q}^{2}/2m_{red}=e^2/r_t \label{eq:turn} \end{equation} where $\tilde{q}=q/2$ and $m_{red}$ is the reduced mass of the pion pair. For a typical BEC momentum difference of $q=100 MeV$, one gets from (\ref{eq:turn}) $r_{t}=0.08 fm$. This suggests that by taking the value of the wave function at $r=0$ one overstimates the Coulomb correction. More recently Biyajima and collaborators \cite{BMOW1} (cf. also \cite{BMOW2}) have considered a further correction to the correction proposed by Bowler, which decreases even more the Coulomb effect and which is also of heuristic interest. In \cite{BMOW1} it is pointed out that the wave function (\ref{eq:scatt}) used by Bowler does not yet take into account the symmetry of the two-particle system. It has to be supplemented by an exchange term so that the rhs of (\ref{eq:scatt}) becomes \begin{equation} (e^{iqz}+e^{-iqz})+[f(\theta,\phi)+f(\pi-\theta,\phi+\pi)] r^{-1}e^{iqr} \label{eq:scattsym} \end{equation} The corrections due to this new effect are of the same order as those found in \cite{Bowler} and go in the same direction. Another approach to the Coulomb correction in BEC has been suggested by Baym and Braun-Munzinger\cite{Baym}. Starting from the observation of \cite{Bowler} about the classical turning point these authors propose the use for heavy ion reactions of a {\em classical} Coulomb correction factor arising in the assumption that the Coulomb effect of the pair is negligible for separations less than an initial radius $r_0$. This model is tested by comparing its results with experimental data on $\pi^{+}\pi^{-}, \pi^{-}p,\pi^{+}p$ correlations in heavy ion collisions\footnote{The measured $\pi^{+}\pi^{-}$ correlations were also used in two recent experimental papers \cite{1Na35}, \cite{2Na49} to estimate the Coulomb correction. Why such a procedure is questionable is explained below.} (Au-Au at AGS energies). The assumption behind this comparison is that the observed correlations are {\em solely} due to the Coulomb effect \footnote{This assumption has to be qualified among other things because final state strong interactions effects due to resonances can also influence these correlations. Furthermore there exists also a quantum statistical correlation for the $\pi^{+}\pi^{-}$ system (cf. sections 4.4, 4.7 and 4.8) which, however may be weak.}. Indeed qualitatively this seems to be the case: thus the data for the $\pi^{+}\pi^{-}$ and $\pi^{-}p$ correlations show a bunching effect characteristic for an attractive interaction while the data for the $\pi^{+}p$ correlation show an antibunching effect, characteristic for repulsion. After this test the authors compare their correction with the Gamow correction and find that the last one is much stronger. Herefrom they also conclude that the Gamow factor overestimates quite appreciably the Coulomb effect \footnote{Cf. however also ref.\cite{Hardtke} where rescattering is added to the classical Coulomb effect and where somewhat different results are obtained. It is unclear whether the strong position-momentum correlations implicit in this rescattering model do not violate quantum mechanics.}. An even stronger conclusion is reached by Merlitz and Pelte \cite{MP3} from the solution of the time dependent Schr\"odinger equation for two identical charged scalar bosons in terms of wave packets. These authors find that the ``expected" Coulomb effect in the correlation function is obliterated by the dispersion of the localized states and is thus unobservable. This makes the interpretation of experimentally observed $\pi^{+}\pi^{-}$ correlations in terms of Coulomb effects even more doubtful. The theoretical studies of the Coulomb effect in BEC quoted so far are based on the solution of the Schr\"odinger equation and apply in fact only for the non-relativistic case. While one might argue that the relative motion of two mesons in BEC is for small $q$ non-relativistic, this is not true for the single particle distributions (cf. below). Therefore in principle one should replace the solution of the Schr\"odinger equation in the Coulomb field used above by the corresponding solution of the Klein-Gordon equation. This apparently has not yet been done, with the exception of of a calculation by Barz\cite{Barz} who investigated the influence of the Coulomb correction on the measured values of radii. He found an important change of these radii due to the Coulomb field only for momenta $\leq$ 200 MeV. \begin{footnotesize}Finally, one must mention that the corrections of the wave function described above do not take into account the fact that the charge distribution of a meson is not point-like, but has a finite extension of the order of 1 fm. This means that in principle the Schr\"odinger (or the Klein-Gordon) equation has to be solved with a Coulomb potential modified by this finite size effect. Given the great sensitivity of BEC on small corrections in the wave function, this might be a worthwile enterprise for future research. As a matter of fact it is known from atomic physics (isotopic and isomeric shifts and hyperfine structure) that these finite size effects lead to observable consequences.\end{footnotesize} Besides the wave function effect which influences the BEC due to directly produced particles or those originating from short lived resonances, one has to consider \cite{Bowler} the Coulomb overcorrection applied to pairs of which one particle is a daughter of a long lived state. This effect may bias the correction by up to 20\%. \paragraph{Coulomb correction for higher order correlations} The Coulomb corrections discussed above were limited to single and two-body interactions. In present high energy heavy ion reactions we have already events with hundreds of particles and in the very near future at the relativistic heavy ion collider at Brookhaven (RHIC) this number will increase by an order of magnitude. Then many-body final state interactions may become important. Unfortunately the theory of many-body interactions even for such a ``simple" potential as the Coulomb one is apparently still unmanageable. The long-range nature of the electromagetic interaction does not make this task simpler. Bowler \cite{Bowler} sketched a scenario for Coulomb screening based on the string model. In such a model particles are ordered in space-time, so that e.g. at least one $\pi^+$ must be situated between the members of a $\pi^{-}\pi^{-}$ pair. While the net influence of this effect on $\pi^{+}\pi^{-}$ is expected to be small, for a $\pi^{-}\pi^{-}$ pair the situation is different, because instead of repulsion one obtains attraction. In the case of long range interactions such an effect may become important if the $\pi^{+}_{1}$ propagates together with the $\pi^{-}_{2}\pi^{-}_3$ pair. This happens if $Q_{12}\sim Q_{13}\sim Q_{23}$. To take care of this effect Bowler suggests the replacement \begin{equation} C(Q_{23})\to C(Q_{23})<C^{-1}(Q_{12})C^{-1}(Q_{13})>_{k3} \label{eq:corrlr1} \end{equation} for the Coulomb $\pi^{-}\pi^{-}$ correction and \begin{equation} C(Q_{12})\to C(Q_{12})<C^{-1}(Q_{23})C(Q_{13})>_{k3} \label{eq:corrlr2} \end{equation} for the Coulomb $\pi^{+}\pi^{-}$ correction. Here $< ...>_{k_3}$ symbolizes averaging with respect to the momentum of particle $3$. While on the average \begin{equation} <C^{-1}(Q_{23})C(Q_{13})>_{k3} \label{eq:Bowle} \end{equation} is unity, at small $Q$ the function $C(Q)$ oscillates rapidly and therefore the factor \begin{equation} <C^{-1}(Q_{12})C^{-1}(Q_{13})>_{k3} \label{eq:Bowlf} \end{equation} is sensitive to the distribution of $Q_{12}, Q_{13}$ associated with the local source. According to \cite{Bowler} this last factor for $\pi^{+}\pi^{-}$ pairs does not exceed $0.5\%$ but for like sign pairs no estimate is provided. \paragraph{Coulomb and resonance effect in single inclusive cross sections} At a first look one might be tempted to believe that for single inclusive cross sections in heavy ion reactions the estimate of Coulomb effects is straightforward. Unfortunately this is not the case and so far there is no reliable theoretical estimate of this effect. This is so because the produced charged secondaries move not simply in the electromagnetic field of the colliding nuclei but interact at the same time with all the other secondaries. In view of this situation, recently an attempt has been made to put in evidence experimentally the Coulomb effect in the single inclusive cross section of pions in heavy ion reactions \cite{Na44coul}. In this experiment an excess of negative pions over positive pions in $Pb-Pb$ reactions at 158 AGeV was observed which the authors of \cite{Na44coul} atrributed to the Coulomb interaction of produced pions with the nuclear fireball. However this interpretation has been challenged in \cite{Nellypm} where, in a detailed hydrodynamical simulation it was shown that a similar excess in the $\pi^-/\pi^+$ ratio is expected as a consequence of resonance (especially hyperon) decays. This qualification goes in the same direction as that mentioned above with respect to the exaggeration of the effects of Coulomb interaction in BEC. It also illustrates the complexity of the many body problem of heavy ion reactions even for weak interactions like the electromagnetic one, which in principle are well known. \subsection{Strong final state interactions.} This is a very complex problem because we are dealing with non-perturbative aspects of quantum chromodynamics. That there is no fully satisfactory solution to this problem can be seen from the very fact that we have at least three different approaches to it. As will become clear from the following, these different approaches must not be used simultaneously, as this would constitute double (or triple) counting. That is also why a fourth solution proposed here and which is of heuristic nature will appear in many cases more appealing and more efficient. \subsubsection{Final state interactions through resonances} The majority of secondaries produced in high energy collisions are pions out of which a large fraction (between 40 and 80$\%$) arise from resonances\footnote{For experimental estimates of resonances cf. refs. \cite{expres}.}. Since the resonances have finite lifetimes and momenta, their decay products are created in general outside the production region of the ``direct'' pions (i.e., pions produced directly from the source) and that of the resonances. As a consequence, the two-particle correlation function of pions reflects not only the geometry of the (primary) source but also the momentum spectra and lifetimes of resonances \cite{boal}. Kaons are much less affected by this circumstance \cite{gpkaon}, however correlation experiments with kaons are much more difficult because of the low statistics. For a more detailed discussion of kaon BEC cf. section 5.1.3. The (known) resonances have been taken into account explicitely within the wave function formalism (cf. e.g.\cite{Grishin},\cite{Pod72},\ \cite{Grass}), the string model (\cite{stringbowl}, \cite{stringander}) or within other variants of the Wigner function formalism (cf. e.g. \cite{Gyu89}, \cite{Pra90}, \cite{Ledn}, \cite{Bolz}). The drawback of this explicit approach is that it is rather complicated, it is usually applicable only at small momenta differences $q$ and it presupposes a detailed knowledge of resonance characteristics, including their weights, which, with few exceptions cannot be measured directly and have to be obtained from event generators\footnote{That event generators are not a reliable source of information for this purpose was demonstrated in the case of $e^{+}-e^{-}$ reactions in \cite{DeWolf}, \cite{Verb}.} or other models\footnote{In ref.\cite{Bolz} the weights were determined from thermodynamical considerations.}. With these essential caveats in mind one finds that the distortion of the two-particle correlation function due to resonance decay leads to two obvious effects: (a) the effective radius of the source increases, i.e., the width of the correlation function decreases, and (b) due to the finite experimental resolution in the momentum difference the presence of very long-lived resonances leads to an apparent decrease of the intercept of the correlation function. Effect (b) is particularly important if one wants to draw conclusions from the intercept about a possible contribution of a coherent component in multiparticle production. In hydrodynamical studies of multiparticle production processes one considers resonances within the Wigner function approach (cf. section 5.1.3). \subsubsection{Density matrix approach} In ref. \cite{Stelte} one describes the effect of strong final state interactions by constructing a density matrix based on an effective Lagrangian of Landau-Ginzburg form as used in statistical physics and quantum optics (cf. also \cite{Perugia}). This is a more theoretical approach as it allows to study the effect of the strength of the interaction $g$ on BEC, albeit in an effective Lagrangian description. One writes the density matrix \begin{equation} \rho=Z^{-1}\int\delta\pi|\pi>e^{-F(\pi)}<\pi|\qquad;\qquad Z=\int\delta\pi e^{-F(\pi)} \label{eq:74} \end{equation} where $F$ is the analogue of the Landau-Ginzburg free energy and the integrals are functional integrals over the field $\pi$. This field is written as a superposition of coherent $\pi_c$ and chaotic fields $\pi_{ch}$ \begin{equation} \pi = \pi_{c}(y) + \pi_{ch}(y) \label{eq:39} \end{equation} The variable y refers in particular to rapidity. The total mean multiplicity $<n>$ is related to the field $\pi$ by \begin{equation} <n>=<|\pi ^{2}|>; \label{eq:mean} \end{equation} similar relations hold for the coherent and chaotic parts of $\pi$. One assumes stationarity in y, i.e. the field correlator $G(y,y')=<\pi (y)\pi (y')>$ depends only on the difference $y-y'\equiv \Delta y$. One writes the Landau-Ginzburg form for $F$ as \begin{equation} F(\pi)=\int^y_0dy\left[a\pi(y)+b\left|\frac{\partial\pi(y )}{\partial y}\right|^2+g|\pi(y)|^4\right] \label{eq:77} \end{equation} where $a,b$ and $g$ are constants. The strong interaction coupling is represented by $g$. The constants $a,b$ can be expresed in term of $<n_{ch}>$ and the ``coherence length" $\xi$ which is defined through the correlator $G$ (cf. eq.(\ref{eq:Lorentz}). The main result of these rather involved calculations is the fact that while the interaction does not play any significant role in the value of the intercept $C_2(0)$ it plays an important part in $(C_2(\Delta y\neq 0)$. This situation is ilustrated in Figs.3,4. Thus it is seen in Fig.3 that all $C_2$ curves for various $g$ coincide in the origin. The effect of the interaction in this approach is similar to that of (short lived) resonances, i.e. it leads to a decrease of the width of the correlation function. The sensitivity of $C_2$ on $g$ suggests that the shape of the correlation function can in principle be used for the experimental determination of $g$. One finds furthermore that there is no $g$-dependence for purely chaotic or purely coherent sources. This observation suggests that for a strongly coherent or chaotic field the final state interaction does not manage to disturb the correlation. Note that the curves in Fig. 4 intersect at some $\Delta y$ which depends on the chaoticity. This is characteristic for correlations treated by quantum statistics, when one has a superposition of coherent and chaotic fields, and is a manifestation of the appearance of two (different) functions in the correlation function . We recall (cf. section 2.2) that for the particular case of a Lorentzian spectrum one gets, {\em in the absence of final state interactions}, \begin{equation} C_{2}(y,y')=1+2p(1-p)e^{-|y-y'|/\xi}+p^{2}e^{-2|y-y'|/\xi} \label{eq:41} \end{equation} instead of the empirical relation \begin{equation} C_{2}(y,y')= 1+ \lambda \exp(-|y-y'|/\xi). \label{eq:Deutschmann} \end{equation} As shown in Fig.5 the parametrization (\ref{eq:Deutschmann}) leads to parallel curves for various values of $\lambda$, while the more correct parametrization of ref.\cite{Stelte} leads to intersecting curves. This is due to the fact that the relation for $C_2$ derived in \cite{Stelte} contains as a particular case (for $g=0$) eq.(\ref{eq:41}) and retains the essential feature of eq.(\ref{eq:41}), which consists in the superposition of two exponentials. \subsubsection{Phase shifts} For charged pions the strong final state interactions can be described also by phase shifts. It is known that for an isospin $I=2$ state (this the isopsin of a system of two identically charged pions) the corresponding strong interaction is repulsive. However it was suggested \cite{Bowl1} that the range of strong interactions is smaller than the size of the hadronic source and therefore the correlation should be essentially unaffected by this effect\footnote{The separation between resonances and phase shifts is of course not rigorous because phase shifts reflect also the effect of resonances; however as long as phase shifts constitute a small effect, this should not matter.}. Even for particle reactions like hadron-hadron or $e^{+}-e^{-}$ the size of the source is of the order of $1 fm$ while the range of interaction is only $0.2 fm$. On the other hand the effective size quoted above arises because of the joint contribution to BEC of direct pairs and resonances. So it is interesting to analyze these two contributions separately. Two identically charged pions are practically always produced together with a third pion of opposite charge . Then according to Bowler \cite{Bowl1} one has also to consider the $I=0$ attractive interaction between the oppositely charged pions and this compensates largely the $I=2$ state interaction. Thus it appears that also in particle reactions only resonances play an important role in final state strong interactions. The considerations about final state interactions made in this subsection treat separately Coulomb and strong interactions. This is permitted as long as we deal with small effects or when the ranges of the two types of interactions do not overlap. For very small distances this is not anymore the case. Furthermore the Gamow and the phase shift corrections are based on the wave function formalism which ignores the possibility of creating particles. However when entering the non-classical region the well known difficulties of the wave function formalism become visible (Klein paradox). To consider this effect, in ref.(\cite{Anish} the joint contribution of the strong interaction potential and the Coulomb potential are analyzed in a version of the Bethe-Salpeter equation for spinless particles. It is found that as expected also from the considerations presented above that strong interaction diminishes appreciably the Gamow correction. \subsubsection{Effective currents} As will be explained below the most satisfactory approach to BEC is at present the classical current approach, based on quantum field theory. There three types of source chracteristics appear: the chaoticity, the correlation lengths/times and the space-time dimensions. It is obvious that these quantities contain already information about the nature of the interaction and therefore it is quite natural to consider them as effective parameters which describe all effects of strong final state interactions. This approach to strong final state interactions is probably the most recommendable at the present stage for BEC in reactions induced by hadrons or leptons because: 1. it is simpler; 2. it avoids double counting; 3. it avoids the use of poorly known resonance characteristics; 4. it avoids the use of the ill defined concept of final state interactions for strong interactions. For heavy ion reactions, when hydrodynamical methods are used the explicit consideration of resonances can be practiced up to a certain point without major difficulties and then the strong final state interactions can be taken into account through these resonances (cf. section 5.1.3). To conclude this discussion of final state interactions in BEC, it is interesting to note that in boson condensates the final state interactions might be different than in normal hadronic sources. In a condensate the Bose field becomes long range in configuration space. This can be understood as a consequence of the fact that in a condensate the effective mass of the field carrier vanishes. Indeed a calculation \cite{2Wheeler} based on the chiral sigma model shows that the effective range of the pion field can increase several times due to this effect. \setcounter{equation}{0} \section{Currents} \subsection{Classical versus quantum currents} In section 2.2 we were concerned mainly with the properties of fields and did not ask the question where these fields come from. In the present chapter we shall put and try to answer this question. We start by recalling the definition of correlation functions within quantum field theory. Let $a^{\dagger}_{i} ({\bf k})$ and $a_{i}({\bf k})$ be the creation and annihilation operator of a particle of momentum ${\bf k}$, where the index $i$ labels internal degrees of freedom such as spin, isospin, strangeness, etc. The $n$-particle inclusive distribution is \begin{equation} \frac{1}{\sigma} \frac{d^n \sigma^{i_1 ... i_n}}{d\omega_1 ... d\omega_n} \ =\ (2\pi)^{3n} \prod^n_{j=1} 2E_j Tr \left( \rho_f \ a^\dagger_{i_1} ({\bf k}_1) ... a^\dagger_{i_n} ({\bf k}_n) a_{i_n}({\bf k}_n) ... a_{i_1} ({\bf k}_1) \right) \label{eq:ae} \end{equation} where \begin{equation} d\omega_i = \frac{d^3k_i}{(2\pi)^3 2E_i} \label{eq:di} \end{equation} is the invariant volume element in momentum space. With the notation \begin{equation} G_n^{i_1 \cdots i_n} ({\bf k}_1, ... , {\bf k}_n) \ \equiv \ \frac{1}{\sigma}\frac{d^n \sigma^{i_1 ... i_n}} {d\omega_1 ... d\omega_n} \end{equation} the general $n$-particle correlation function is defined as \begin{equation} C_n^{i_1 ... i_n} ({\bf k}_1, ... ,{\bf k}_n ) \ =\ \frac{G_n ^{i_1 ... i_n} ({\bf k}_1 , ... ,{\bf k}_n)} {G_1^{i_1} ({\bf k}_1) \cdot ... \cdot G_n^{i_n}({\bf k}_n)} \label{eq:cln} \end{equation} The particles which the operators $a^{\dagger}$ and $a$ are associated with are the quanta of the field (which we will denote in general by $\phi$); these particles as well as the density matrix $\rho$ refer to the final state where measurements take place. On the other hand we usually know (or guess) the density matrix only in the initial state. Therefore we will have to transform the above expression so that eventually the density matrix in the final state $\rho_{f}$ is replaced by the density matrix in the initial state $\rho_i$, while the fields will continue to refer to the final state. To emphasize this we wrote in eq.(\ref{eq:ae}) $\rho_f$. \begin{equation} \rho_{f}=S\rho_{i}S^{\dagger} \label{eq:S} \end{equation} so that we have \begin{equation} G_1(\mbold{k}) = (2\pi)(2E_1)Tr\{\rho_i S^{\dagger} a^{\dagger}(\mbold{k}) a(\mbold{k})S\}\\ \end{equation} \begin{eqnarray} G_2(\mbold{k}_1,\mbold{k}_2) =(2\pi)^6(2E_1)(2E_2) Tr\{\rho_iS^{\dagger}a^{\dagger} (\mbold{k}_2) a^{\dagger}(\mbold{k}_1) a(\mbold{k}_1) a(\mbold{k}_2) S\} \label{eq:(3+4)} \end{eqnarray} Thus, if the initial conditions i.e. $\rho_i$ are given, in principle the knowledge of the S matrix suffices to calculate the physical quantities of interest. In one case the $S$ matrix can even be derived without approximations. This happens when the currents are classical and we shall discuss this case in some detail in the present and the following section. Before doing this, we shall consider briefly the more general case when the currents are not necessarily classical. The $S$ matrix is given by the relation \begin{equation} S\ =\ {\cal T}\exp\left\{i\int d^4x L^{int}(x)\right\} \label{eq:S_def} \end{equation} where the interaction lagrangian $L_{int}$ is a functional of the fields $\phi$. $\cal T$ is the chronological time ordering operator; we shall use below also the antichronological time operator $\tilde {\cal T}$. Consider for simplicity a scalar field produced by a current $J$. Then \begin{equation} L^{int}(x)\ \equiv \ J(x)\phi(x). \label{eq:L_def} \end{equation} Equations (\ref{eq:S_def}) and (\ref{eq:L_def}) allow us now to calculate the correlations we are interested in in terms of the currents after eliminating the fields. One obtains thus \begin{eqnarray} P_1(\mbold{k})\;&=&\; Tr\left\{\rho_iJ_H^{\dagger}({\bf k}) J_{H}({\bf k})\right\} \label{eq:P1JH} \\ P_2(\mbold{k}_1\,,\,\mbold{k}_2)\;&=&\; Tr\left\{ \rho_i {\bf \tilde{\cal T}} \left[ J_H^{\dagger}({\bf k}_1)J_H^{\dagger}({\bf k}_2)\right] {\cal T}\left[ J_{H}({\bf k}_1)J_H({\bf k}_2)\right]\right\} \label{eq:P2JH} \end{eqnarray} where the label $H$ stands for the Heisenberg representation. Now the cross sections depend only on the currents and the density matrix in the initial state. The appearance of the time ordering operators $\cal T$ and $\tilde {\cal T}$ in eqs.(\ref{eq:P1JH}),(\ref{eq:P2JH}) is a reminder of the fact that the current $J$ is here an operator. \subsection{Classical currents} Besides the fact that in this case an exact, analytical solution of the field equations is available and that the limits of this approximation are quite clear, the classical current has the important avantage that in it the space-time characteristics of the source are clearly exhibited and thus contact with approaches like the Wigner approach and hydrodynamics are made possible. The assumption that the currents are classical implies that $J$ is a $c$ number and then the order in eq.(\ref{eq:P2JH}) does not matter. This approximation can be used in particle physics when the currents are produced by heavy particles (e.g. nucleons) and/or when the momentum transfer $q$ is small compared with the momentum $K$ of the emitting particles. \footnote{For multiparticle production processes this implies also a constraint on the multiplicity and/or the momenta $k$ of the produced particles}. In section 4.7 a new criterion for the applicability of the classical current assumption in terms of particle-antiparticle correlations will be presented. The classical current formalism was introduced to the field of Bose-Einstein correlations in \cite{shuryak,podgor,GKW}. In this approach, particle sources are treated as external classical currents $J(x)$, the fluctuations of which are described by a probability distribution $P\{ J \}$. >From many points of view like e.g. the understanding the space-time properties of the sources or the isotopic spin dependence of BEC this approach is superior to any other approach. This has become clear only in the last years \cite{aw}, \cite{APW} when a systematic investigation of the independent physical quantities which enter the dynamics of correlation functions has been made (cf.below). The classical current formalism in momentum space is mathematically identical with the coherent state formalism used in quantum statistics and in particular in quantum optics (cf. section 2.2), the classical currents in $k$-space $J(k)$ being proportional to the eigenvalues of the coherent states $|\alpha >$. This explains the importance of the coherent state formalism for applications in particle and nuclear physics. The density matrix is \begin{equation} \rho = \int {\cal D} J \ {\cal P}\{ J \} \ |J><J| \label{eq:what2} \end{equation} where the symbol ${\cal D}J$ denotes an integration over the space of functions $J(x)$, and the statistical weight ${\cal P} \{ J \}$ is normalized to unity, \begin{equation} \int {\cal D}J \ {\cal P} \{ J \}\ =\ 1. \label{eq:dj} \end{equation} (The reader will recognize in eq.(\ref{eq:what2}) the $\cal P$-representation introduced in section 2.2.) Expectation values of field operators can then be expressed as averages over the corresponding functionals of the currents, e.g., \begin{eqnarray} && Tr \left( \rho a^{\dagger}({\bf k}_1)\ ...\ a^{\dagger} ({\bf k}_n) a({\bf k}_{n+1})\ ... \ a({\bf k}_{n+m}) \right)\\ && = \prod^n_{j=1} \frac{(-i)}{\sqrt{(2\pi)^3 \ 2 E_j}} \ \prod^{n+m}_{j=m+1} \frac{i}{\sqrt{(2\pi)^3 2 E_j}} \ <J^*({\bf k}_1)\ ...\ J^*({\bf k}_n) J({\bf k}_{n+1})\ ...\ J({\bf k}_{n+m})>\nonumber \label{eq:trw} \end{eqnarray} In the following we shall discuss the special case when the fluctuations of the currents $J(x)$ are described by a Gaussian distribution ${\cal P}\{J\}$. The reasons for this choice were given in section 2.2. As in quantum optics we write the current $J(x)$ as the sum of a chaotic and a coherent component, \begin{equation} J(x) \ = \ J_{chaotic}(x)\ + \ J_{coherent}(x) \label{eq::jsum} \end{equation} with \begin{eqnarray} J_{coherent}(x) & = & <J(x)>\\ J_{chaotic}(x) & = & J(x) - <J(x)> \label{eq::jchao} \end{eqnarray} By definition, $<J_{chaotic}(x)>=0$. The case $I(x) \not= 0$ corresponds to {\em single particle coherence}. The Gaussian current distribution is completely determined by specifying its first two moments: the first moment coincides, because of eq.(\ref{eq::jchao}), with the coherent component, \begin{equation} I(x) \ \equiv \ <J(x)> \label{eq:a1} \end{equation} and the second moment is given by the 2-current correlator, \begin{eqnarray} D(x,x') & \equiv & <J(x) \ J(x')> - <J(x)> \ <J(x')>\nonumber\\ & = & <J_{chaotic}(x) \ J_{chaotic}(x')> \label{eq:a2} \end{eqnarray} \subsection{Primordial correlator, correlation length and space-time distribution of the source} We come now to a more recent development \cite{aw}, \cite{APW} of the current formalism which has shed new light on both fundamental and applicative aspects of BEC. There are two fundamental and in principle independent aspects of physics which come together in the phenomenon of BEC in particle and nuclear physics. One refers to the {\em geometry} of the source and goes back to the original Hanbury-Brown and Twiss interefence experiment in astronomy. The ``geometry" is characterized by the size of the source e.g.the longitudinal and transverse radius $R_{\parallel}$ and $R_{\perp}$ respectively and the lifetime of the source $R_0$. The second aspect is related to the {\em dynamics} of the source and is expressed through correlation lengths. In the following we will use two correlation lengths $L_{\parallel}$, $L_{\perp}$ and a correlation time $L_0$ \footnote{We refer here to short range correlations. Cf. section 6.1.3 for a distinction between short range and long range correlations.}. As a consequence of the finite space-time size of sources in particle physics one cannot in general separate the geometry from the dynamics in the second (and higher) order correlation function. This separation is possible (cf. below) only by using simultaneously also the single inclusive cross section. This remarkable fact \cite{aw} is not yet realized in most of the theoretical and experimental papers on this subject\footnote{For systems in local equilibrium Makhlin and Sinyukov \cite{MS1} introduced a length scale (called in \cite{MS} ``length of homogeneity") which characterizes the hydrodynamical expansion of the source and can be different from the size of the system. Further references on this topic can be found e.g.in \cite{Akk}, \cite{csorlor}}. One of the main objects of the present review is to clarify this situation. Assuming Gaussian currents, one may take the point of view that the purpose of measuring $n$-particle distributions is to obtain information about the space-time form of the coherent component, $I(x)$, and of the correlator of the chaotic components of the current, $D(x,x')$. In practice, because of limited statistics it is necessary to consider simple parametrizations of these quantities, and to use the information extracted from experi\-mental data to determine the free parameters (such as radii, correlation lengths etc.). The approach considered in \cite{APW} differs in some fundamental aspects from those applications of the density matrix approach in particle physics which are performed in momentum (rapidity) space and are limited usually to one dimension (cf. however \cite{prd44} where rapidity {\em and} transverse momentum were considered). The approach of ref.\cite{APW} is a {\em space-time} approach in which the parameters refer to the space-time characteristics of the source. This approach has important heuristic advantages compared with the momentum (rapidity) space approach as will be explained below. Among other things, in the quantum statistical (QS) space-time approach the parameters of the source as defined above can be considered as effective parameters which contain already the entire information which one is interested in and which one could obtain from experiment, and thus distinguishing between directly produced particles and resonance decay pro\-ducts could amount to double counting. The apparent proliferation of parameters brought about by the QS approach is compensated by this heuristic and practical simplification. Furthermore a new and essential feature of the approach of ref.\cite{APW} as compared with previous applications of the current formalism \cite{podgor,GKW} which assume $L=0$, is the {\em finite} correlation length (time) $L$. This fact has important theoretical and practical consequences. It leads among other things to an effective correlation between momenta and coordinates, so that e.g. the second order correlation functions does not depend only on the difference of momenta $q=k_1-k_2$ but also on the sum $k_1+k_2$. This non-stationarity property, which is observed in experiment, is usually associated with expanding sources and treated within the Wigner function formalism. However from the considerations presented above it follows that expansion is in general not a necessary condition for non-stationarity in $q$. It will be shown how expanding sources can be treated without the Wigner formalism, which restricts unnecessarily the applicability of the results to small $q$ values (cf. section 4.8). The distinction between correlation lengths and radii is possible only in the current formalism; the Wigner formalism provides just a length of homogeneity. The results which follow from the space-time approach \cite{APW} include: (i) The existence of at least 10 independent parameters that enter into the correlation function; (ii new insights into the problem of partial coherence; (iii) Isospin effects: $\pi^0\pi^0$ correlations are different from $\pi^{\pm}\pi^{\pm}$ correlations; there exists a quantum statistical (anti)correlation between particles and antiparticles ($\pi^+\pi^-$ in this case). These effects are associated with the presence of squeezed states in the density matrix, which in itself is a surprising and unexpected feature in conventional strong interaction phenomenology. It turns out that soft pions play an essential role in the experimental investigation of BEC, both with respect to the effect of particle-antiparticle correlations predicted here as well as in the investigation of the coherence of the source. Depending on the relative magnitude of the parameters of the coherent and chaotic component, soft particles can either enhance or suppress the coherence effect. Consider first the case of an infinitely extended source. The correlation of currents at two space-time points $x$ and $y$ is described by a primordial correlator, \begin{equation} <J(x) \ J(y)>_0\ =\ C(x-y) \label{eq:jx} \end{equation} Note that $C$ depends only on the difference $x-y$. The correlator $C(x-y)$ contains some characteristic length (time) scales $L$, the so-called correlation lengths (times) \footnote{ These correspond to the ``coherence lengths" used in the quantum optical literature.}. In the current formalism used in \cite{GKW} $L=0$. $C(x-y)$ is a real and even function of its argument. In the rest frame of the source, it is usually parametrized by an exponential, \begin{equation} C(x-y)\ =\ C_0 \ \exp \left[ -\frac{|x_0-y_0|}{L_0}- \frac{|{\bf x}-{\bf y}|}{L} \right] \label{eq:cxy1} \end{equation} or by a Gaussian, \begin{equation} C(x-y)\ = \ C_0 \ \exp \left[ -\frac{(x_0-y_0)^2}{2L^2_0}- \frac{({\bf x}-{\bf y})^2} {2L^2} \right] \label{eq:cxy} \end{equation} However, it should be clear that in principle any well behaved decreasing function of $|x-y|$ is a priori acceptable, and in practice it is usually up to the experimenter to decide which particular form is more appropriate. The ans\"atze (\ref{eq:cxy1},\ref{eq:cxy}) need to be modified for the case of an expanding source (cf. sections 4.8 and 4.9) where each source element is characterized not only by a correlation length $L^\mu$ but also by a four velocity $u^\mu$. As a matter of fact, the form of the function $C$ is irrelevant as long as one is interested in the general statements of the theoretical quantum statistical (current) formalism. In practical applications, of course, in order to obtain concrete information about the source and the medium (i.e., about $L$) the form of $C$ has to be specified. In principle, a full dynamical theory is expected to determine the functional form of the correlation function, however at present this ``fundamentalist" approach is not yet applicable \footnote{Indeed one may hope that for strongly interacting systems lattice QCD may provide in future the corrrelation length $L$.}. One uses instead a phenomenological approach like that reflected in eqs. (\ref{eq:cxy1}, \ref{eq:cxy}). Effects of the {\em geometry} of the source are taken into account by introducing the space-time distributions of the chaotic and of the coherent component, $f_{ch}(x)$ and $f_c(x)$, re\-spec\-ti\-ve\-ly. The expectation values of the currents, $I(x)$ and $D(x,x')$, take nonzero values only in space-time regions where $f_c$ and $f_{ch}$ are nonzero. Thus, one may write \begin{eqnarray} I(x) &=& f_c(x)\\ D(x,x') &=& f_{ch}(x) \ C(x-x') \ f_{ch}(x') \label{eq:ixh} \end{eqnarray} We turn now to an important new aspect of the current approach \subsection{Production of an isospin multiplet} Following \cite{APW} we generalize the previous results to the case of an isospin multiplet and derive explicit expressions for the single inclusive distributions and correlation functions of particles that form an isotriplet (such as the $\pi^+$,$ \pi^-$ and $\pi^0$-mesons). For the sake of definiteness, we will refer to pions in the discussion below, but it should be understood that the formalism is applicable to an arbitrary isomultiplet. Consider the production of charged and neutral pions, $\pi^+, \pi^- $ and $\pi^0$, by random currents. The initial interaction Lagrangian is written \begin{eqnarray} {\cal L}_{int} &=& J_+(x) \pi^-(x) + J_-(x)\pi^+(x) + J_0(x)\pi^0(x) . \label{eq:lint} \end{eqnarray} The current distribution is completely characterized by its first two moments. They read for the case of an isotriplet \begin{eqnarray} I_i(x) & \equiv & <J_i(x)> \qquad i,i'\ =\ +,-,0\\ D_{ii'}(x,x') & \equiv & <J_i(x) \ J_{i'}(x')> - <J_i(x)> \ <J_{i'}(x')>\nonumber \label{eq:b1} \end{eqnarray} Invariance of the chaotic 2-current correlator under rotation in isospin space implies \begin{eqnarray} D_{00}(x,x') & = & D_{+-}(x,x') \ \equiv D(x,x')\\ D_{++}(x,x') & = & D_{--}(x,x') \ =\ D_{+0}(x,x') \ =\ D_{0-}(x,x') \ =\ 0.\nonumber \label{eq:b35} \end{eqnarray} The corresponding current distribution is \begin{equation} P \{ J_i \} \ = \ \frac{1}{{\cal N}}\ \exp[-{\cal A} \{ J_i \}] \label{eq:b4} \end{equation} where in coordinate representation \begin{equation} {\cal N} \ = \ \int \int \int {\cal D}J_+ {\cal D}J_- {\cal D}J_0 \ \exp[-{\cal A} \{ J_i \}] \label{eq:b45} \end{equation} and \begin{eqnarray} {\cal A} \{ J_i \} & = & \left. \int\int \ d^4x \ d^4y \right[ (J_+(x)-I_+(x)) M(x,y) (J_-(y)- I_-(y))\nonumber\\ & & \qquad \qquad + \left.\frac{1}{2} (J_0(x)-I_0(x)) M(x,y) (J_0(y)-I_0(y))\right] \label{eq:b5} \end{eqnarray} with \begin{equation} M(x,x') \ = \ D^{-1}(x,x') \label{eq:fd2} \end{equation} For the general case of a partially coherent source, the single inclusive distributions of pions of charge $i$ $(i = +,-,0)$ can be expressed as the sum of a chaotic component and a coherent component, \begin{equation} \frac{1}{\sigma} \frac{d\sigma^i}{d\omega} = \left.\frac{1}{\sigma} \frac{d\sigma^i} {d\omega}\right|_{chaotic} + \left.\frac{1}{\sigma} \frac{d\sigma^i}{d\omega} \right|_{coherent} \label{eq:sig1} \end{equation} with \begin{equation} \left.\frac{1}{\sigma} \frac{d\sigma^i} {d\omega}\right|_{chaotic} \ =\ D(k,k) \label{eq:ncha} \end{equation} and \begin{equation} \left.\frac{1}{\sigma} \frac{d\sigma^i}{d\omega} \right|_{coherent} \ =\ |I(k)|^2 \label{eq:nco} \end{equation} \noindent In general, the chaoticity parameter $p$ will be momentum dependent, \begin{equation} p(k) \ =\ \frac{D(k,k)}{D(k,k)+|I(k)|^2} \label{eq:fk} \end{equation} To write down the correlation functions in a concise form, one introduces the normalized current correlators, \begin{equation} d_{rs} = \frac{D(k_r,k_s)}{[D(k_r,k_r)\cdot D(k_s,k_s)]^{\frac{1}{2}}}, \quad \tilde{d}_{rs} = \frac{D(k_r,-k_s)}{[D(k_r,k_r)\cdot D(k_s,k_s)]^{\frac{1}{2}}}, \label{eq:drs} \end{equation} where the indices $r,s$ label the particles. Note that $\tilde{d}$ is the {\em same} function as $d$ but of the change of sign of one of its variables. One may express the correlation functions in terms of the magnitudes and the phases, \begin{eqnarray} T_{rs} &\equiv & T(k_r,k_s) \ =\ |d(k_r,k_s)|\nonumber\\ \tilde{T}_{rs} &\equiv& \tilde{T}(k_r,k_s) \ = \ | \tilde{d}(k_r,k_s)| \nonumber \\ \phi^{ch}_{rs} &\equiv& \phi^{ch}(k_r,k_s)\ =\ {\rm Arg} \ d(k_r,k_s)\\ \tilde{\phi}^{ch}_{rs} &\equiv& \tilde{\phi}^{ch}(k_r,k_s) \ =\ {\rm Arg} \ \tilde{d}(k_r,k_s) \nonumber \label{eq:trs} \end{eqnarray} and the phase of the coherent component, \begin{equation} \phi^c_r \ \equiv \ \phi^c(k_r) \ =\ {\rm Arg}\ I(k_r) \nonumber \, . \label{eq::phic} \end{equation} The same notation will be used for the chaoticity parameter, \begin{equation} p_r \ \equiv \ p(k_r) \label{eq:pr} \end{equation} The two-particle correlation function is \begin{equation} C^{++}_2({\bf k}_1,{\bf k}_2)\ =\ 1 +\ 2 \sqrt{p_1(1-p_1) \cdot p_2(1-p_2)} \ T_{12}\ \cos(\phi^{ch}_{12} - \phi^c_1+\phi^c_2)\ +\ p_1p_2 \ T^2_{12} \label{eq:c2plus1} \end{equation} In \cite{APW} higher order correlation functions up to and including order 5 are given. In the absence of single particle coherence the two-particle correlation functions for different pairs of $\pi^+,\pi^-, \pi^0$ mesons read \begin{eqnarray} C_2^{++}({\bf k_1},{\bf k_2}) &=& 1 + |d_{12}|^2,\nonumber\\ C_2^{+-}({\bf k_1},{\bf k_2}) &=& 1 + |\tilde{d}_{12}|^2, \nonumber\\ C_2^{+0}({\bf k_1},{\bf k_2}) &=& 1, \label{eq:c4}\\ C_2^{00}({\bf k_1},{\bf k_2}) &=& 1 + |d_{12}|^2 + |\tilde{d}_{12}|^2 \nonumber \end{eqnarray} These results \cite{apw} were surprising in that they disagreed with some of the pre\-con\-cei\-ved notions on Bose-Einstein correlations. For instance, it was commonly assumed that without taking into account final state interactions and in the absence of coherence, the maximum of the two-particle correlation of identical pions is 2 (for ${\bf k_1} = {\bf k_2}$). It was also assumed that there are no correlation effects among different kinds of pions because these particles are not identical. (This last assumption is even sometimes used in normalizing the experimental data on $C_2^{\pm\pm}$ with respect to $C_2^{+-}$.) The results (\ref{eq:c4}) show that these assertions are not necessarily true. In particular, looking at the two pion correlations one can see that in addition to the familiar correlations of identical particles (the terms $|d_{12}|^2$) there are particle-antiparticle -- in this case, $\pi^+\pi^-$ -- correlations (the terms $|\tilde{d}_{12}|^2$). The $\pi^0$ has both terms, as it is identical with its antiparticle. Essentially this last fact is the explanation for the appearance of the ``surprising" effects. Obviously it is a specific quantum field effect. It will be shown in section 4.8, that for soft pions and for small lifetimes of the source the terms $|\tilde{d}_{12}|^2$ can in principle become comparable with the conventional terms $|d_{12}|^2$. This implies that the distribution $C_2^{00}({\bf k_1}, {\bf k_2})$ of two neutral pions can be as large as 3, and the maximum value of $C_2^{+-}$ is 2 (instead of 1). The corresponding limit of $C_3^{000}$ is 15 (instead of 6), and that of $C_3^{++-}$ is 6 (instead of 2). However, it should be noted that these are merely upper limits, which for massive particles are not reached except for sources of infinitesimally small lifetimes. For soft photons however the situation is different (cf. below). The ``new" terms, proportional to $<a_\ell(k_1) a_\ell(k_2)>$ are due to the non-stationarity (in $k$ space) of the source. While in quantum optics time-stationarity is the rule, in particle physics this is not the case because of the finite lifetime and finite radius of the sources. The existence of a non-vanishing expectation value of the products $a(k_1) a(k_2)$ is what one would expect (cf. section 2.2) from two-particle coherence (squeezing), just as $<a(k)>\not= 0$ follows from ordinary (one-particle) coherence (note that the latter has not been assumed here). The fact that squeezing which, as mentioned above, is quite an exceptional situation in optics, discovered only recently, is a natural consequence of the formalism for present particle physics, is possibly one of the most startling results obtained recently in BEC. A characteristic feature of squeezed states is the fact that for static sources they lead anti-correlations. Finally, it should be pointed out again that the above results -- in particular, the fact that $C_2^{00}$ in general differs from $C_2^{--}$ -- are consistent with isospin symmetry. In closing this section one should note that the existence of particle-antiparticle correlations is not restricted to pions but applies also to other systems, like e.g. neutral kaons. In principle thus there exist also $K_0\bar{K}_0$ quantum statistical correlations. However since $K_0$ particles cannot be observed except in linear combinations with $\bar{K}_0$ in the form of $K_s$ and $K_l$ the QS particle-antiparticle correlation effect has to be disentangled from the $K_sK_s$ or $K_lK_l$ Bose-Einstein correlation ($K_s$ and $K_l$ are of course bosons and thus subject to BEC) which exist also in the ``old fashioned" wave function formalism which ignores the intrinsic ``new" $K_0\bar{K}_0$ correlation. As a matter of fact $K_sK_s$ correlations have been observed experimentally, however no attempt has been made so far to extract from them the ``surprising" effects \footnote{Because of their larger mass as compared with that of pions, these effects may be even more quenched than in the case of pions, except for sources of very short lifetime (cf. section 4.8).} (For more recent experiments cf. \cite{Opal}) and for a theoretical analysis of the ``old" effects and their possible application in CP violation phenomena cf. \cite{Lipkin}). \footnote{The analysis in the preceding subsections referred to the production of an isotriplet assuming just symmetry between the isospin components (cf. eq.(\ref{eq:lint})) of the current. In principle, for strong interactions the conservation of isospin $I$ has also to be considered. While the chaotic part is not affected by this condition, the coherent component is influenced by conservation of isospin\cite{AB}. In particular this can lead to an additive positive term in the correlation function and thus to an increase of the bounds of BEC for pions. It remains to be seen whether this effect can be distinguished from the effect of long range correlations (cf. section 6.1.3). Moreover in hadronic reactions and in particular those involving nuclei such an effect would be strongly suppressed because of the following circumstance. The initial state has to be averaged over all components of isospin $I$ which is a first ``diluting" factor. Furthermore the effect is appreciable only for low total isospin. This total isospin has to be shared by the chaotic component $I_{ch}$ and the coherent component $I_c$: $I=I_c+I_{ch}$. The first one arises mostly from resonances with different isospin values, so that even if the total isospin takes its minimum value ($I=0$), $I_{ch}$ and therefore $I_c$ can take larger values.} \subsubsection{An illustrative model of uncorrelated point-like random sources} To clarify the origin of different terms in the functions $C_2^{ab}(k_1,k_2)$, let us consider, a source consisting of $N$ point-like random sources \begin{equation} J_a(x) = \sum^N_{i=1} \ j_a(x_i) \ \delta(t-t_i) \ \delta^3 ({\bf x}-{\bf x}_i),\quad a = +,-,0 \label{eq:e1} \end{equation} and assume that the currents $j_a(x_i)$ at different points $x_i$ are mutually independent and have the same statistical properties, i.e. \begin{equation} <j_a^*(x_i) \ j_b(x_j)> = \delta_{ij}<j_a^* j_b>. \label{eq:e2} \end{equation} We also assume in this section \begin{equation} <j_a> = 0, \label{eq:e3} \end{equation} i.e. ignore a possible coherent component $<j_a>$ to make the presentation more transparent. Now the one-particle distribution is \begin{equation} <J_a^*({\bf k}) \ J_a({\bf k})> = N \cdot <j_a^* j_a> \label{eq:e4} \end{equation} and the two-particle distribution takes the form \begin{eqnarray} &<&J^*_a(k_1) \ J_a(k_1) \ J^*_b(k_2) \ J_b(k_2)> = \nonumber\\ &=&\sum^N_{i=1} < j^*_a(x_i)j_a(x_i) j^*_b(x_i)j_b(x_i)> \nonumber\\ &+&\sum^N_{i \not= j} <j^*_a(x_i)j_a(x_i)> \cdot <j^*_b(x_j) j_b(x_j)>\nonumber\\ &+&\sum^N_{i \not= j} <j^*_b(x_i)j_a(x_i)> \cdot <j^*_a(x_j) j_b(x_j)> \cdot e^{i(k_1-k_2)(x_i-x_j)} \nonumber\\ &+&\sum^N_{i \not= j} <j^*_a(x_j)j^*_b(x_j)> \cdot <j_a(x_i) j_b(x_i)> \cdot e^{i(k_1+k_2)(x_i-x_j)} \nonumber\\ \label{eq:e5} \end{eqnarray} Let us consider separately the four different terms on the right-hand side of eq. (\ref{eq:e5}). The first term corresponds to two particles being emitted from a single point (Fig. 6a); it is proportional to the number of emitting points. The second term describes an independent emission of two particles from different points (Fig. 6b). The third term, being non-zero for $a=b$, describes an interference effect of direct and exchange diagrams, characteristic for identical particles, emitted from different points (Fig. 6c). This is the usual BE-correlation effect. The fourth term describes an interference of two-particle emissions from different points (Fig. 6d) (``two-particle sources"). It appears to be non-zero for real currents with $a=b \ (\pi^0\pi^0)$ and for complex currents with $J^*_a = J_b \ (\pi^+\pi^-)$, that is for particle-antiparticle associative emission. In this simple model, it represents the ``surprising" effects discussed above. This diagramatic illustration of the ``surprising" BE-correlations is due to Bowler \cite{bowler}, who found that the ``surprising" effects derived for the first time in ref. \cite{apw} can be understood in terms of these more qualitative considerations. Bowler derived the ``new" effects from the string model. \subsection{Photon interferometry. Upper bounds of BEC.} The advantage of photon BEC resides in the fact that photons are not influenced by final state interactions. Photons present also an interesting subject of theoretical research from the general BEC point of view, since they are spin-one bosons while pions and kaons used in hadronic BEC are scalar particles. We shall see below that this supplementary degree of freedom has specific implications in BEC. Last but not least photon correlations are for various reasons to be discussed below of particular interest for the search of quark matter. We present some of the results of \cite{Leo}, which contain as a special case those of \cite{Neu} and where these topics are discussed. Consider a heavy ion reaction where photons are produced through bremsstrahlung from protons in independent proton-neutron collisions\footnote{Photon emission from proton-proton collisions is suppressed because it is of quadrupole form.}. The corresponding elementary dipole currents are \begin{equation} j^{\lambda}(k)=\frac{ie}{mk^0}{\bf p}.{\bf \epsilon}_{\lambda}(k) \label{eq:Leo1} \end{equation} where ${\bf p} = {\bf p}_i-{\bf p}_f$ is the difference between the initial and the final momentum of the proton, ${\bf \epsilon}_{\lambda}$ is the vector of linear polarization and $k$ the photon momentum; $e$ and $m$ are the charge and mass of the proton respectively. The total current is written \begin{equation} J^{\lambda}(k)=\sum^N_{n=1}e^{ikx_n}j_n^{\lambda}(k). \label{eq:Leo2} \end{equation} For simplicity we will discuss in the following only the case of pure chaotic currents\\ $<J^{\lambda}(k)> = 0$ and refer for coherence effects to the original literature \cite{Leo}, \cite{Neu}. In analogy to the considerations of the previous subsection the index $n$ labels the independent nucleon collisions which take place at different space-time points $x_n$. These points are assumed to be randomly distributed in the space-time volume of the source with a distribution function $f(x)$ for each elementary collision. The current correlator is proportional to products of the form \begin{equation} <J^{\lambda_1}(k_1)J^{\lambda_2}(-k_2)>={\bf {\epsilon}}^{i}_{\lambda_1}(k_1)\left(\sum^N_{n=1}<{\bf p}^{i}_n{\bf p}^j_n>\right){\bf {\epsilon}}^j_{\lambda_2}(k_2) \label{eq:Leo10} \end{equation} For central collisions due to the axial symmetry around the beam direction one has for the momenta the tensor decomposition \begin{equation} <{\bf p}^{i}_n{\bf p}^j_n>=\frac{1}{3}\sigma_n\delta^{ij}+ \delta_nl^{i}l^j, \label{eq:Leo11} \end{equation} where $l$ is the unit vector in the beam direction and $\sigma _n, \delta _n$ are real positive constants. In \cite{Neu} an isotropic distribution of the momenta was assumed. This corresponds to the particular case $\delta _n = 0$. The generalization to the form (\ref{eq:Leo11}) is due to \cite{Leo}. The summation over polarization indexes is performed by using the relations \begin{equation} <({\bf \epsilon}^{i}.{\bf p}_l)({\bf \epsilon}^{j}.{\bf p}_ {l^{'}})>=\frac{1}{3}({\bf \epsilon}^{i}.{\bf \epsilon}^{j}) \delta_{ll'} \label{eq:sum} \end{equation} and \begin{equation} \sum^2_{\lambda=1}{\bf \epsilon}^{i}_{\lambda}(k){\bf \epsilon}^j_{\lambda}(k)=\delta^{ij}-{\bf n}^{i}{\bf n}^j, \label{eq:Leo21} \end{equation} where ${\bf n} = {\bf k}/ {|{\bf k}|}$. We write below the results for the second order correlation function \begin{equation} C_2(k_1,k_2)=\frac{\rho_2(k_1,k_2)}{\rho_1(k_1)\rho_1(k_2)} \label{eq:Leo24} \end{equation} for two extreme cases: (1) Uncorrelated elementary currents (isotropy) ($\sigma \gg \delta$) \begin{equation} C_2(k_1,k_2;\;\sigma\neq0,\;\delta=0)= 1+\frac{1}{4}[1+({\bf n}_{1}.{\bf n}_{2})^2]\left[|\tilde{f} (k_1-k_2)|^2+|\tilde{f}(k_1+k_2)|^2\right], \label{eq:Leo25} \end{equation} leading to an intercept \begin{equation} C_2(k,k)=\frac{3}{2}+\frac{1}{2}|\tilde{f}(2k)|^2 \label{eq:Leo26} \end{equation} limited by the values (3/2,2). (2) Strong anisotropy ($ \sigma \ll \delta $) \begin{equation} C_2(k_1,k_2;\;\sigma=0,\;\delta\neq0)= 1+|\tilde{f}(k_1-k_2)|^2+|\tilde{f}(k_1+k_2)|^2 \label{eq:Leo27} \end{equation} with an intercept \begin{equation} C_2(k,k)=2+|\tilde{f}(2k)|^2 \label{eq:Leo28} \end{equation} limited this time by the values (2,3). Note that due to the form of the photon-current interaction (\ref{eq:Leo1}) in this (strong anisotropy) case the photons emerge practically completely polarized so that the summation over polarizations does not affect the correlation. These results are remarkable among other things because they illustrate the specific effects of photon spin on BEC. Thus while for (pseudo-)scalar pions the intercept is a constant (2 for charged pions and 3 for neutral ones) even for unpolarized photons the intercept is a function of k. For a graphical illustration and explanation of this fact cf. Fig. 7 from ref.\cite{Leo}). It is seen that to perform the summation over polarization implied by eq.(\ref{eq:sum}) only one direction of the linear polarization can be chosen equal for both photons, while the other polarization direction differs by the angle $\theta$ between the two momenta ${\bf k}_1, {\bf k}_2$. One thus finds that, while for a system of charged pions (i.e. a mixture of $50\%$ positive and $50\%$ negative) the maximum value of this intercept $MaxC_2(k,k)$ is 1.5, for photons $MaxC_2(k,k)$ exceeds this value and this excess reflects the space-time properties (represented by $\tilde{f}(k)$) , the degree of (an)isotropy of the source represented by the quantities $\sigma$ and $\delta$, and the supplementary degree of freedom represented by the photon spin. As a consequence of the fact that $\tilde{f}$ is a decreasing function of its argument, in eqs.(\ref{eq:Leo25}),(\ref{eq:Leo27}) the terms with $\tilde{f}(k_1+k_2)$ are in general smaller than the terms with $\tilde{f}(k_1-k_2)$, except for small momenta $k$. The fact that the differences between charged pions and photons are enhanced for soft photons reminds us of a similar effect found with neutral pions (cf. section 4.4). Neutral pions are in general more bunched than identically charged ones and this difference is more pronounced for soft pions. This similarity is not accidental, because photons as well as $\pi^0$ particles are neutral and this circumstance has quantum field theoretical implications which will be mentioned also below. We see thus that photon BEC can provide information both about the space-time form of the source represented by $f$ and the dynamics which are represented by $\delta$ \footnote{Cf. \cite{Marques},\cite{Badala} where photon correlation experiments in low energy (100 MeV/nucleon) heavy ion reactions are reported. For a theoretical discussion of these experiments cf.\cite{Barz96}.} \footnote{The relation between photon interferometry and the formation length of photons is discussed in \cite{PPT}.}. The results on photon correlations presented above refer to the case that the sources are ``static" i.e. not expanding. Expanding sources were considered in \cite{Feld} within a covariant formalism. Some of the results above, in particular eqs. (\ref{eq:Leo25},\ref{eq:Leo26}), which had been initially derived by Neuhauser \cite{Neu}, were challenged by Slotta and Heinz \cite{Heinz}. Among other things, these authors claim that for photon correlations due to a chaotic source ``the only change relative to 2-pion interferometry is a statistical factor $\frac{1}{2}$ for the overall strength of the correlation which results from the experimental averaging over the photon spin". In \cite{Heinz} an intercept $\frac{3}{2}$ is derived which is in contradiction with the results presented above and in particular with eq.(\ref{eq:Leo26}) where besides the factor $\frac{3}{2}$ there appears also the $k$ dependent function $\frac{1}{2}{|\tilde{f}(2k)|^2}$. Similar statements can be found in previous papers \cite{kap}, \cite{kap1}, \cite{kap2}, \cite{Sriv} where more detailed applications concerning heavy ion reactions based on this assertion of \cite{Heinz} are presented. Some of the papers quoted above were criticized immediately after their publication in \cite{bounds}, \cite{Axel} \cite{Feld} and the paper \cite{Heinz} was written with the intention to ``settle" this ``controversy". It should be pointed out here that the reason for the difference between the results of \cite{Neu},\cite{Leo} on the one hand and those of ref.\cite{Heinz} on the other is mainly due to the fact that in \cite{Heinz} a formalism was used which is less general than that used in \cite{Neu} and \cite{Leo} and which is inadequate for the present problem. This implies among other things that unpolarized photons cannot be treated in the way proposed in \cite{Heinz} and that the results of \cite{Neu} and \cite{Leo} are correct, while the results of \cite{Heinz} are not. In \cite{Heinz} the following formula for the second order correlation function is used: \begin{equation} C({\bf k}_1,{\bf k}_2)=1+\frac{\tilde {g}_{\mu \nu}( {\bf q,K})\tilde {g}^{\nu \mu}({\bf -q,K})} {\tilde {g}^{\mu}_{\mu}({\bf 0,k}_1)\tilde {g}^{\mu}_{\mu}( {\bf 0,k}_2)} \label{eq:Heinz23} \end{equation} Here $\tilde{g}$ is the Fourier transform of a source function, ${\bf q}={\bf k}_1- {\bf k}_2$ and $ {\bf K}=\frac{1}{2}({\bf k}_1+{\bf k}_2)$. This formula is a particular case of a more general formula for the second order correlation function derived by Shuryak \cite{shuryak} using a model of uncorrelated sources, when emission of particles from the same space-time point is negligible (cf. section 4.4.1). Since this equation is sometimes used in the recent literature without giving the reader the possibility of evaluating the approximations used in its derivation, we will sketch this derivation in the following. In \cite{shuryak} one starts with the current correlator \begin{equation} <J^{*}_i(x_1)J_j(x_2)>=\delta_{ij}J_i(x,\Delta x), \label{eq:shuryak1} \end{equation} where $J_i(x)$ is the current emitted by point $x$ and \begin{eqnarray} x=(x_1+x_2)/2,\qquad \Delta x=x_1-x_2. \end{eqnarray} Eq. (\ref{eq:shuryak1}) assumes that the individual currents $(i \neq j$) are uncorrelated. With the notation $\tilde{I}_{i}(q,K)$ for the Fourier transform of $I_{i}(x, \Delta x)$ and $\tilde{I}(q,K)\equiv\sum_i\tilde{I}(q,K)$ the inclusive single particle distribution reads \begin{equation} W(k)=\left<\left|\sum_j\int e^{ikx}J_j(x)d^4x\right|\right>= \sum_i\tilde{I}_i(0,k)=\tilde{I}(0,k) \label{eq:shuryak3} \end{equation} and the two particle distribution is given by \begin{equation} W(k_1k_2)=\left<\left|\sum_{i,j}\int(e^{ik_1x_1+ik_2x_2}+ e^{ik_1x_2+ik_2x_1})J_i(x_1)J_j(x_2)dx_1dx_2\right|^2\right> \label{eq:shuryak4} \end{equation} or finally \begin{eqnarray} W(k_1,k_2) &=& \tilde{I}(0,k_1)\tilde{I}(0,k_2)+ |\tilde{I}(q,K)|^2\nonumber\\ & & +\sum_{i}[<J^{*}_{i}(k_1)J^{*}_{i}(k_2)J_{i}(k_1)J_{i} (k_2)> \nonumber\\ & & -\tilde{I}_i(0,k_1)\tilde{I}_i(0,k_2)- |\tilde{I}_i(q,K)|^2] \label{eq:shuryak5} \end{eqnarray} The first term on the rhs of eq.(\ref{eq:shuryak5}) is the product of one-particle distributions and the second term is the conventional interference term, corresponding to figure 6c. By going over to the Wigner source function (cf. also section 4.9) \begin{equation} g_{\mu\nu}(x,K)=\int d^4ye^{-iKy}<J^{*}_{\mu}(x+\frac{1}{2}y)J_{\nu}(x-\frac{1}{2} y)> \label{eq:heinz21} \end{equation} these first two terms result in eq.(\ref{eq:Heinz23}) of \cite{Heinz}. However as is clear from eq.(\ref{eq:shuryak5}) there exists also a third term, neglected in eq.(\ref{eq:Heinz23}) and which corresponds to the simultaneous emission of two particles from a single point ($x_i$) as indicated in figure 6d. While for massive particles this term is in general suppressed, this is not true for massless particles and in particular for soft photons. In \cite{Neu} and \cite{Leo} this additional term had not been neglected as it was done in \cite{Heinz} and therefore it is not surprising that ref.\cite{Heinz} could not recover the results of refs.\cite{Neu} and \cite{Leo}. The neglect of the term corresponding to emission of two particles from the same space-time point is not permitted in the present case. As mentioned in section 4.4.1, in a model of uncorrelated point-like random sources like the present one, emission of particles from the same space-time point corresponds in a first approximation to particle-antiparticle correlations and this type of effect leads also to the difference between BEC for identical charged pions and the BEC for neutral pions. This is so because neutral particles coincide with the corresponding antiparticles. (As a consequence of this circumstance e.g. while for charged pions the maximum of the intercept is 2, for neutral pions it is 3 (cf. section 4.4)). Photons being neutral particles, similar effects like those observed for $\pi^0$-s are expected and indeed found (cf. above). This misapplication of the current formalism invalidates completely the conclusions of ref.\cite{Heinz} and confirms and strengthens the criticism expressed in \cite{bounds}, \cite{Axel}, \cite{Feld} of the papers \cite{kap}-\cite{Sriv}. The fact that for unpolarized photons $MaxC_2(k,k)$ is 2 and not 1.5 as stated in \cite{Heinz}, can be understood by realizing that a system of unpolarized photons consists on the average of $50\%$ photons with the same helicities and $50\%$ photons with opposite helicities. The first ones contribute to the maximum intercept (of the unpolarized system) with a factor of $3$ and the last ones with a factor of 1 (coresponding to unidentical particles). For the sake of clarification it must be mentioned that ref. \cite{Heinz} contains also other incorrect statements. Thus the claim in \cite{Heinz} that the approach by Neuhauser ``does not correctly take into account the constraints from current conservation" is completely unfounded as can be seen from eq.(\ref{eq:sum}) which is an obvious consequence of current conservation (cf.e.g. eq.(7.61) in \cite{BD}). Last but not least the statement that because the tensor structure in eq.(20) of ref.\cite{Feld} is parametrized in terms of $k_1$ and $k_2$ separately ``instead of only in terms of $K$, leading to spurious terms in the tensor structure which eventually result in their spurious momentum-dependent prefactor" has also to be qualified. As mentioned above, eq.(\ref{eq:Heinz23}) to which this observation about the $K$ dependence of \cite{Heinz} refers is not general enough for the problem of photon interferometry. \subsection{Coherence and lower bounds of Bose-Einstein correlations} We mentioned in the previous subsections that the intercepts of the second and higher order correlation functions can deviate from the canonical values derived within the wave function formalism. This effect is important for at least three reasons: (i) it illustrates the limitations of the wave function approach (ii) it can in principle (provided other effects like final state interactions are taken into account) be used for the determination of the degree of coherence. (iii) it can serve as a test of models of BEC, since the value of the intercept follows from very general quantum statistical considerations, in particular the Gaussian nature of the density matrix. In most BEC models the intercept is identical to the maximum of the correlation function and therefore it can be studied by limiting the discussion to chaotic sources as was done in the previous subsection where the upper bounds of correlation functions were investigated. On the other hand the minimum of the correlation functions is determined both by the form of the density matrix and the amount of coherence (cf.ref.(\cite{bounds})) because coherence leads to a decrease of the correlation function. This will be illustrated below by discussing the lower bounds of this function. We will show among other things (cf. \cite{APW}) that in the quite general case of a Gaussian density matrix, for a purely chaotic system the two-particle correlation function must always be greater than one. On the other hand, in the presence of a coherent component the correlation function may take values below unity. Some implications for experimental and theoretical results found in the literature will be discussed here as well as in section 5.1.6. We have seen in section 4.4 that for identically charged bosons (e.g., $\pi^+$) the two- particle correlation function reads \begin{equation} C^{++}_2({\bf k}_1,{\bf k}_2)\ =\ 1 +\ 2 \sqrt{p_1(1-p_1) \cdot p_2(1-p_2)} \ T_{12}\ \cos(\phi^{ch}_{12} - \phi^c_1+\phi^c_2)\ +\ p_1p_2 \ T^2_{12} \label{eq:c2plus} \end{equation} For neutral bosons like photons, or $\pi^0$'s, the terms $\tilde{d}(k_r,k_s)$ also appear (cf. eq:(\ref{eq:drs}) in the BEC function: \begin{eqnarray} C^{00}_2({\bf k}_1,{\bf k}_2) & = & 1 + 2 \sqrt{p_1(1-p_1)\cdot p_2(1-p_2)} \ T_{12}\ \cos(\phi^{ch}_{12} - \phi^c_1+\phi^c_2)\ +\ p_1p_2 \ T^2_{12}\nonumber\\ & & + 2 \sqrt{p_1(1-p_1)\cdot p_2(1-p_2)} \ \tilde{T}_{12}\ \cos(\tilde{\phi}^{ch}_{12} - \phi^c_1- \phi^c_2) \ + \ p_1p_2 \ \tilde{T}^2_{12}\nonumber\\ && \label{eq:cnulnul} \end{eqnarray} Let us first consider the case of a purely chaotic source. Insertion of $p(k) \equiv 1$ in eqs. (\ref{eq:c2plus}) and (\ref{eq:cnulnul}) immediately yields $C_2({\bf k}_1,{\bf k}_2)\geq 1$. In the case of partial coherence, the terms containing cosines come into play and consequently $C_2$ may take values below unity. Eqs. (\ref{eq:c2plus},\ref{eq:cnulnul}) imply that $C^{--}_2({\bf k}_1,{\bf k}_2)\geq 2/3$ and $C^{00}_2({\bf k}_1,{\bf k}_2)\geq 1/3$. Because of the cosine functions in (\ref{eq:c2plus},\ref{eq:cnulnul}) one would expect $C_2$ as a function of the momentum difference $q$ to oscillate between values above and below $1$. \begin{footnotesize}Such a behaviour of the Bose-Einstein correlation function has been observed in high energy $e^{+}-e^{-}$ collision experiments (cf., e.g., ref. \cite{osc}), but apparently not in hadronic reactions. This observation was interpreted as a consequence of final state interactions in ref. \cite{bowler}. If final state interactions determine this effect, it is unclear why the effect is not seen in hadronic reactions. On the other hand, if coherence is responsible for it, this would be easier to understand. Indeed multiplicity distributions of secondaries in $e^{+}e^{-}$ reactions are much narrower (almost Poisson-like) than in $pp$ reactions, which is consistent with the statement that hadronic reactions are more chaotic than $e^{+}-e^{-}$ reactions \cite{fried}\footnote{This statemant is not necessarily in contradiction with the empirical observation that the $\lambda$ factor in $e^{+}-e^{-}$ reactions appears in general to be larger than in $p-p$ reactions, given the fact that $\lambda$ is not a true measure of coherence.}. \end{footnotesize} So far, two methods have been proposed for the detection of coherence in BEC: the intercept criterion \cite{gnf} ($C_2({\bf k},{\bf k})<2$) and the two exponent structure of $C_2$ \cite{revisit}. Both these methods have their difficulties because of statistics problems or other effects. The observation of $C_2({\bf k}_1,{\bf k}_2)<1$ could constitute a third criterion for coherence. In \cite{kap},\cite{kap1} the two-particle correlation function has been calculated for photons emitted from a longitudinally expanding system of hot and dense hadronic matter created in ultra\-rela\-ti\-vist\-ic nuclear collisions . For such a system, the particles are emitted from a large number of independent source elements (fluid elements), and consequently one would expect the multiparticle final state to be described by a Gaussian density matrix. However, although the system is assumed to be purely chaotic the correlation function calculated in \cite{kap} is found to take values significantly below unity. Clearly, this is in contradiction with the general result derived above from quantum statistics ($C_2\geq 1$ for a chaotic system). The reason for this violation of the general bounds derived for a purely chaotic source is in this concrete case the use of an inadequate approximation in the evaluation of the space-time integrals. However as pointed out in \cite{Axel} the expression for the two-particle inclusive distribution used in ref. \cite{kap} (equation (3) of that paper), which in our notation takes the form \begin{equation} P_2({\bf k}_1,{\bf k}_2) \ =\ \int d^4x_1 \int d^4x_2\ g\left(x_1,k_1\right)\ g\left(x_2,k_2\right)\ [1\ +\ \cos((k_1-k_2)(x_1-x_2))], \label{eq:3} \end{equation} is also unsatisfactory \footnote{This formula appears apparently for the first time in \cite{Yano} and was criticized (for other reasons) already in \cite{PaGyGa}. It is nevertheless used in certain event generators for heavy ion reactions (cf. section 4.10).} because for certain physical situations it can lead to values below unity for the two-particle correlation function even if the integrations are performed exactly. To see this, consider, e.g., the simple ansatz \begin{equation} g(x,k)\ =\ const. \ \exp[-\alpha({\bf x}-\beta {\bf k})^2]\ \delta(t-t_0) \end{equation} where $\alpha$ and $\beta$ are free parameters. The expression for $P_2({\bf k}_1,{\bf k}_2)$ used in ref.\cite{kap} then yields \begin{equation} C_2({\bf k}_1,{\bf k}_2) \ =\ 1 \ +\ \exp\left[-\frac {{\bf q}^{\ 2}} {2\alpha}\right] \ \cos[\beta {\bf q}^{\ 2}] \end{equation} Clearly, if $\beta$ exceeds $\alpha^{-1}$ the above expression will oscillate and take values below unity. On the other hand, in the current formalism (cf. below) one obtains with the same ansatz for $g$ \begin{equation} C_2({\bf k}_1,{\bf k}_2) \ =\ 1 \ +\ \exp\left[-\frac {{\bf q}^{\ 2}}{2\alpha}\right] \ \geq \ 1. \end{equation} Thus, eq. (\ref{eq:3}) can lead to values $C_2<1$ if there is a strong correlation between the momentum of a particle and the space-time coordinate of the source element from which it is emitted. Such correlations between $x$ and $k$ can occur in the case of an expanding source. The reason for this pathological behaviour is the fact that the simultaneous specification of coordinates and momentum as implied by eq.(\ref{eq:3}) is constrained in quantum mechanics by the Heisenberg incertainty relation and any violation of this constraint leads necessarily to a violation of quantum mechanics. This violation manifests itself sometimes, as in the present case, through a violation of the conservation of probability. This phenomenon is also met when using the Wigner function, which for this reason cannot always be associated with a bona fide probability amplitude. We will discuss this problem also in section 5.1.6. The above considerations concerning bounds for the BEC functions refer to the case of a Gaussian density matrix. In general, a different form of the density matrix may yield correlation functions that are not constrained by the bounds derived here. For instance we have seen in section 2.2 that for a system of squeezed states $C_2$ can take arbitrary positive values. Moreover, for particles produced in high energy hadronic or nuclear collisions, the fluctuations of quantities such as impact parameter or inelasticity may introduce additional correlations which may also affect the bounds of the BEC functions. \subsection{Quantum currents} The results derived in the previous section, in particular the isospin dependence of BEC, were obtained in the assumption that the currents were classical. The question arises up to what point these conclusions survive in a fully quantum treatment of the problem. It would also be important to get a more precise criterium for the phenomenological applicabilbity of the classical assumption, besides the no-recoil prescription. This question was discussed in \cite{LeoQFT} where it was found that the ``surprising" effects not only persist when the currents are quantum, but that they can serve as an experimental estimate of the size of the quantum corrections. We shall sketch briefly in the following the results of ref.(\cite{LeoQFT}). As in the classical current case one starts with the interaction Lagrangian \begin{equation} L_{int}(x)\ \equiv \ J_{(+)}(x)\pi^{(-)}(x)\ +\ J_{(-)}(x)\pi^{(+)}(x) \ +\ J_0(x)\pi^0(x) \label{eq:L_def1} \end{equation} The currents $J^{(+)},\; J^{(-)},\; J^{0}$ are {\em operators} which we assume again for simplicity not to depend on the $\pi^{(\pm)}$ and $\pi^0$ fields \mbox{$([J,\pi]=0)$}. Taking into account the different isospin components in eqs.(\ref{eq:P1JH}) and (\ref{eq:P2JH}) we find that as in the classical current case the single and double inclusive cross sections depend on these components, e.g.\footnote{For reasons of notational simplicity we have replaced $J^{\dagger}(k)$ by $J(-k)$.} \begin{eqnarray} G^{(-)}_1(\mbold{k})\;&=&\; Tr\left\{\rho_iJ_H^{(+)}(-k) J_H^{(-)}(k)\right\} \label{eq:P1JH_m} \\ G^{(-+)}_2(\mbold{k}_1\,,\,\mbold{k}_2)\;&=&\; Tr\left\{ \rho_i {\bf \tilde{\cal T}} \left[ J_H^{(+)}(-k_1)J_H^{(-)}(-k_2)\right] {\cal T}\left[ J_H^{(-)}(k_1)J_H^{(+)}(k_2)\right]\right\} \label{P2JH_mp} \end{eqnarray} >From now on we shall omit the label $H$ and assume that all operators are written in the Heisenberg representation. Assuming a Gaussian density matrix one gets \begin{eqnarray} G^{(0)}_1(\mbold{k})\ &=&\ F^n(k,k) \label{P0:F} \\ G^{(-)}_1(\mbold{k})\ &=&\ G^{(+)}_1(\mbold{k})\ =\ F^{ch}(k,k) \label{P-:F}\\ G^{(--)}_2(\mbold{k}_1,\mbold{k}_2)\ &=&\ G^{(-)}_1(\mbold{k}_1) G^{(-)}_1(\mbold{k}_2) + |F^{ch}(k_1,k_2)|^2 \label{P--:F}\\ G^{(00)}_2(\mbold{k}_1,\mbold{k}_2)\ &=&\ G^{(0)}_1(\mbold{k}_1) G^{(0)}_1(\mbold{k}_2) + |F^{n}(k_1,k_2)|^2 + |\Phi^{n}(k_1,k_2)|^2 \label{P00:F}\\ G^{(-+)}_2(\mbold{k}_1,\mbold{k}_2)\ &=&\ G^{(-)}_1(\mbold{k}_1) G^{(+)}_1(\mbold{k}_2) + |\Phi^{ch}(-k_1,k_2)|^2 \label{P-+:F} \end{eqnarray} where the functions $F$ and $\Phi$ are defined for charged particles (upper index $ch$) and for neutral ones (upper index $n$) as follows: \begin{eqnarray} F^{ch}(k_1,k_2)\ &\equiv&\ \ll\!\! {\cal T}_c \left\{ J^{(+)}_{\oplus}(-k_1)J^{(-)}_{\ominus}(k_2) \right\} \!\!\gg \ =\ \ll\!\! J^{(+)}(-k_1)J^{(-)}(k_2) \!\!\gg \label{Fch} \\ \Phi^{ch}(k_1,k_2)\ &\equiv&\ \ll\!\! {\cal T}_c \left\{ J^{(+)}_{\oplus}(-k_1)J^{(-)}_{\oplus}(k_2) \right\} \!\!\gg \ =\ \ll\!\! {\bf \tilde{\cal T}} \left\{ J^{(+)}(-k_1)J^{(-)}(k_2) \right\} \!\!\gg \label{Fch:t} \\ F^n(k_1,k_2)\ &\equiv&\ \ll\!\! {\cal T}_c \left\{ J^{(0)}_{\oplus}(-k_1)J^{(0)}_{\ominus}(k_2) \right\} \!\!\gg \ =\ \ll\!\! J^{(0)}(-k_1)J^{(0)}(k_2) \!\!\gg \label{Fn} \\ \Phi^{n}(k_1,k_2)\ &\equiv&\ \ll\!\! {\cal T}_c \left\{ J^{(0)}_{\oplus}(-k_1)J^{(0)}_{\oplus}(k_2) \right\} \!\!\gg \ =\ \ll\!\! {\bf \tilde{\cal T}} \left\{ J^{(0)}(-k_1)J^{(0)}(k_2) \right\} \!\!\gg \label{Fn:t} \end{eqnarray} In contrast to the classical current approach \cite{APW} which deals with only one type of two-current correlator we have here two different kinds of two-current correlators depending on their ordering prescriptions. Moreover and most remarkably, the difference between these two correlators is reflected by the difference between $--$ and $+-$ correlations. It thus follows from \cite{LeoQFT} that the ``surprising'' effects found in \cite{apw} i.e. the presence of particle-antiparticle Bose-Einstein type correlations and a new term in the Bose-Einstein correlation function for neutral particles are reobtained, but under a more general form which contains also the quantum corrections. These equations also prove that the above effects are not an artifact of the classical current formalism but have general validity. Morover and most remarkably, from the above equations follows that {\em the difference between the effects of the classical and quantum currents resides in just these ``new effects" and in particular in the difference between $00$ and $--$ correlations i.e. in the $+-$ correlations}. This result can serve as an estimate of the importance of quantum corrections to the classical current formalism of BEC. Since $+-$ correlations are in general small, it follows that the classical current approach is a good approximation, except for very short lived sources, where the $+-$ correlations become comparable to the $--$ correlations. It also follows that the experimental measurement of $+-$ correlations is a highly rewarding task, since they are a rather unique tool for the investigation of two very interesting effects in BEC, namely squeezed states and quantum corrections. \subsection{Space-time form of sources in the classical current formalism} In \cite{APW} two types of sources were considered, a ``static" one which corresponds to a source in rest and an expanding one. We will present below some of the results, as they exemplify certain important features of the space-time approach within the classical current formalism. \paragraph{A static source} The space-time distributions of static sources, as well as the primordial correlator, are parametrized as Gaussians: \begin{eqnarray} f_{ch}(x) & = & \exp\left(-\frac{x_0^2}{R_{ch,0}^2} -\frac{x_{\parallel}^2}{R_{ch,\parallel}^2} -\frac{x_{\perp}^2}{R_{ch,\perp}^2}\right) \label{eq:fstat1}\\ f_{c}(x) & = & \exp\left(-\frac{x_0^2}{R_{c,0}^2} -\frac{x_{\parallel}^2}{R_{c,\parallel}^2} -\frac{x_{\perp}^2}{R_{c,\perp}^2}\right) \label{eq:fstat} \end{eqnarray} and \begin{equation} C(x-y)\ =\ \exp\left[-\frac{(x_0-y_0)^2}{2L_0^2} -\frac{(x_{\parallel}-y_{\parallel})^2}{2L_{\parallel}^2} -\frac{({\bf x}_\perp-{\bf y}_\perp)^2}{2L_{\perp}^2}\right] \label{eq:cxmy} \end{equation} Note that the term {\em static} here does not imply time independence but rather a specific time dependence defined by eqs. (\ref{eq:fstat1}), (\ref{eq:fstat}) corresponding to source elements being at rest. This is to be contrasted to the {\em expanding} source, discussed in the next section, which explicitly contains velocities of source elements. The main justification for this particular form of parametrization is mathematical convenience, because, as will be shown below, for this case the correlation functions in momentum space can be calculated analytically and the physical implications can be read off immediately. In eqs.(\ref{eq:fstat1}-\ref{eq:cxmy}), $R_{ch,\alpha}$ and $R_{c,\alpha} (\alpha = 0,\perp, \parallel)$ are the lifetimes, transverse radii and longitudinal radii of the chaotic source and of the coherent source, respectively, and $L_{\alpha} (\alpha = 0,\perp, \parallel)$ are the correlation time and the corresponding correlation lengths in transverse and in longitudinal direction. The relative contributions of the chaotic and the coherent component are determined by fixing the value of the (momentum dependent) chaoticity parameter $p$ at some arbitrary scale (in this case, at $k=0$): \begin{equation} p_0 \ \equiv \ p(k=0) \label{eq:pop} \end{equation} The model contains 10 independent parameters: the radii and lifetimes of the chaotic and of the coherent source, the correlation lengths in space and time, and the chaoticity $p_0$. In \cite{APW} it is assumed that $L_{\parallel} = L_{\perp}\equiv L$, i.e., that the medium is isotropic, which leaves us with 9 independent parameters. With the definitions \begin{equation} R_{\alpha L}^2\ =\ \frac{R_{ch,\alpha}^2 L_{\alpha}^2} {R_{ch,\alpha}^2 + L_{\alpha}^2} \qquad (\alpha = 0, \perp , \parallel ) \label{eq:rperp} \end{equation} one may write the single inclusive distribution in the form \begin{equation} E \frac{1}{\sigma} \frac{d^3 \sigma}{d^3k}\ = \ \left. E\frac{1}{\sigma}\frac {d^3\ \sigma}{d^3 k} \right| _{k=0} \left( p_0\ s_{ch}(k)\ +\ (1-p_0)\ s_c(k) \right) \label{eq:eins} \end{equation} where \begin{equation} s_{ch}(k) \ = \ \exp \left[-\frac{E^2 R^2_{0L}}{2} -\frac{k_{\parallel}^2 R^2_{\parallel L}}{2} - \frac{{\bf k}_{\perp}^2 R^2_{\perp L}}{2}\right] \label{eq:schak} \end{equation} and \begin{equation} s_{c}(k) \ = \ \exp \left[-\frac{E^2 R^2_{c,0}}{2} -\frac{k_{\parallel}^2 R^2_{c,\parallel}}{2} - \frac{{\bf k}_{\perp}^2 R^2_{c,\perp}}{2}\right] \label{eq:schuk} \end{equation} The scales which determine the mean energy-momentum of the coherently produced particles are given by the inverse lifetime and radii, $R^{-1}_{c, \alpha}$, of the coherent source. For the chaotically produced particles, these scales are given by the inverse of a combination of correlation lengths and dimensions of the chaotic source, $R^{-1}_{\alpha L}$. Eq.(\ref{eq:rperp}) implies that $R_{\alpha L} \leq R_{ch,\alpha}$. The radius of the chaotic source enters the single inclusive distribution only in combination with the correlation length $L$. This feature which occurs also for higher order correlations leads to the important consequence that experimental measurements of BEC do not provide separately information about radii (lifetimes) of sources, nor about correlation-lengths (-times), but rather about the combination of these quantities as given by eq. (\ref{eq:rperp}). On the other hand, by measuring both the single and the double inclusive distribution one can determine radii and correlation lengths separately. It follows from eqs.(\ref{eq:eins}, \ref{eq:schak},\ref{eq:schuk}) that in the presence of partial coherence in general (i.e., unless $R_{\alpha L} = R_{c,\alpha}$) the single inclusive distribution is a superposition of two Gaussians of different widths. If the geometry of the coherent source is the same as that of the chaotic source, one has $R_{c, \alpha} = R_{ch, \alpha} > R_{L \alpha}$, which would imply that coherently produced particles can be observed predominantly in the soft regime. However, if the coherent radii are small compared to the chaotic ones, this situation is reversed. As a next step, consider the correlation functions. The correlation function of two negatively charged pions is \begin{equation} C^{--}_2(k_1,k_2)\ =\ 1 +\ 2\sqrt{p_1(1-p_1) \cdot p_2(1-p_2)} \ T_{12}\ \cos(\phi^{ch}_{12} - \phi^c_1+\phi^c_2)\ +\ p_1p_2\ T^2_{12} \label{eq:non} \end{equation} For the Gaussian parametrizations all phases in the second order correlation function disappear \footnote{This is not the case anymore for an expanding source e.g. (cf. below) or in general for higher order correlations}, \begin{equation} \phi_{12}^{ch}\ =\ \tilde{\phi}_{12}^{ch}\ =\ \phi_{j}^{c}\ =\ \tilde{\phi}_{j}^{c}\ =\ 0 \label{eq:fis} \end{equation} and \begin{eqnarray} T_{12} & = & \exp \left[-\frac{(E_1-E_2)^2(R^2_{ch,0}-R^2_{0L})}{8} -\frac{(k_{1,\parallel}-k_{2,\parallel})^2(R_{ch,\parallel}^2-R^2_{\parallel L} )} {8}\right.\nonumber\\ && \qquad \left. - \frac{({\bf k}_{1\perp}-{\bf k}_{2\perp})^2(R^2_{ch,\perp}- R^2_{\perp L})}{8}\right] \label{eq:fiss} \end{eqnarray} \begin{eqnarray} \tilde{T}_{12} & = & \exp \left[-\frac{(E_1+E_2)^2(R^2_ {ch,0}-R^2_{0L})}{8} -\frac{(k_{1,\parallel}+k_{2,\parallel})^2(R_{ch,\parallel}^2- R^2_{\parallel L}) }{8} \right. \nonumber\\ && \qquad \left. - \frac{({\bf k}_{1\perp}+{\bf k}_{2\perp})^2 (R^2_{ch,\perp}-R^2_{\perp L})}{8}\right] \label{eq:fiat} \end{eqnarray} The two particle correlation function $C_2^{--}$ is the sum of a purely chaotic term $(\propto \ T^2_{12})$ and a mixed term $(\propto \ T_{12})$. The momentum dependence of the chaoticity parameter, $p=p(k)$, implies a momentum dependence of the contribution of the mixed term relative to that of the purely chaotic term. To see how this affects the interplay between the two terms (i.e., the interplay between the two Gaussians), it is useful to explicitly insert the momentum dependence of the chaoticity parameter by writing \begin{equation} p_r \ =\ p(k_r)\ =\ \frac{p_0}{A_r} \qquad (r=1,2) \label{eq:pr1} \end{equation} with \begin{equation} A_r \ \equiv \ A(k_r) \ =\ p_0 + (1-p_0) \ S_{rr} \qquad (r=1,2) \label{eq:akr} \end{equation} and \begin{eqnarray} S_{rs} & = & \exp \left[-\frac{(E_r^2+E_{s}^2)(R^2_{c,0}-R^2_{0L})}{4} -\frac{(k_{r\parallel}^2+k_{s\parallel}^2) (R_{c,\parallel}^2-R^2_{\parallel L} )} {4}\right. \nonumber\\ && \qquad \left. - \frac{({\bf k}_{r\perp}^2+{\bf k}_ {s \perp}^2)(R^2_{c,\perp}-R^2_{\perp L})}{4}\right] \label{eq:swelve} \end{eqnarray} With this, $C_2^{--}$ takes the form \begin{equation} C^{--}_2(k_1,k_2)\ =\ 1 +\ \frac{2p_0(1-p_0)S_{12}}{A_1 A_2} \ T_{12}\ +\ \frac{p_0^2}{A_1 A_2}\ T^2_{12} \label{eq:cett} \end{equation} The momentum-dependence of the relative contributions of the purely chaotic and of the mixed term is reflected in the factor $S_{12}$. Depending on the sign of the combinations $R^2_{c,\alpha} - R^2_{\alpha L}$, $\ S_{12}$ may act either as a suppression factor or as an enhancement factor of the mixed term relative to the chaotic term. This is a consequence of the fact that, in contrast with the case of the correlation function $C_2$ derived within the wave function formalism, where the $C_2$ depends only on the difference of momenta $k_1-k_2$ now the correlation function depends also on $k_1+k_2$. \footnote{Up to recently this desirable physical property, which is observed in most experimental data on BEC, was considered to be a consequence of the {\em expansion} of the source and used to be derived within the Wigner function formalism, which is also a particular case of the classical current formalism.) As shown in the example treated above (cf.\cite{aw}) it can be considered also a consequence of the (partial) coherence of a non-expanding source.} It is instructive to discuss the tilde terms that give rise to the particle-antiparticle correlations for the parametrization (\ref{eq:fstat1}-\ref{eq:cxmy}) of a static source. For the sake of transparency, consider only the purely chaotic case, $p_0 = 0$. The correlation functions of like and unlike charged pions then take the form \begin{eqnarray} C_2^{--} (k_1,k_2) &=& 1 + T^2_{12}\\ C_2^{+-} (k_1,k_2) &=& 1 + \tilde{T}^2_{12} \label{eq:tills} \end{eqnarray} >From (\ref{eq:fiss}) and (\ref{eq:fiat}) it can be seen that the ``new" $\tilde{T}_{12}$ terms that appear in the particle-antiparticle correlations are in general small compared to the ``ordinary" $T_{12}$ terms that determine the particle-particle correlations. The term $\tilde{T}_{12}$ gives rise to an anticorrelation effect due to the factor in eq. (\ref{eq:fiat}) containing the sum ${\bf k}_1 + {\bf k}_2$, if the first factor, containing $E_1+E_2$, is not too small. The latter is possible, if the time duration of the pion emission process and/or the pion energies are small enough. We thus expect an enhanced contribution of the ``new" terms for soft pions. The appearance of anti-correlations \footnote{Some authors have recently called them ``back to back" correlations.} is, as mentioned above, a general property of squeezed states, which are present in the space-time formalism of \cite{APW}. The tilde terms arise as a consequence of the non-stationarity of the source. In the limit of a stationary source, $R_0 \rightarrow \infty$, and $\tilde{T}_{12} \rightarrow 0$. An upper limit is given by \begin{equation} C_2^{+-} \ = \ 1 + |\tilde{T}_{12}|^2 \ \leq \ 1 + \exp\left[ -(R_0^2 - R^2_{0L}) m^2_\pi\right] \label{eq:rol} \end{equation} In the limit $L_0 >> R_0$ on the other hand, $R_0 \simeq R_{0L}$ and $C_2^{+-}$ reaches its maximum value $(C_2^{+-})_{max} = 2$. We observe in the above equation that the contribution of the ``tilde" terms increases with decreasing mass of the particles and reaches its maximum of $2$ for massless particles at fixed and non-vanishing $R_0-R_{0L}$. This is related to the observation made in the case of photon BEC where we saw that for unpolarized photons the maximum of $C_2$ is also $2$. Indeed the role of the charge degrees of freedom ($(+,-$) for pions is for unpolarized photons taken over by the spin. From this mass dependence or more general from the energy dependence of the anticorrelation follows that if the factor $E_1+E_2$ in eq.(\ref{eq:fiat}) could be decreased, an enhancement of the ``new" terms would emerge. A possible mechanism for this could be the sudden transition mechanism considered in \cite{sqIgor}. Indeed as shown in this reference, for a chaotic source the correlator $<a({\bf k}_1)a({\bf k}_2)>$ characteristic for the ``new" terms turns out to be an increasing function of the parameter $r=\frac{1}{2}\log(E_a/E_b)$ where $E_a, E_b$ are the energies of the particle in the vacuum and medium respectively. Thus by allowing for medium effects, which in a certain sense is equivalent to an effective change of mass \footnote{This particular possibility was suggested in \cite{Asakawa1}. Andreev \cite{IgorMatra} suggested a time evolution scenario for the medium effect, which involves two modes $k$ and $-k$.}, one can possibly enhance the anticorrelation effect \footnote{Anticorrelations in disoriented chiral condensates are considered in \cite{Hideaki}.}in BEC. \paragraph{Expanding source} High energy multiparticle dynamics suggests that the sources of produced particles are expanding. This property is reflected in particular in hydrodynamical models and also in string models. In terms of the current formalism this means that the correlators are velocity dependent. While many of the studies of Bose-Einstein correlations for expanding sources have followed, with slight variations, a Wigner function type of approach, the use of the Wigner approach is in general too restrictive and is recommendable only in the case when a full-fledged hydrodynamical description of the system is performed. In the present section, following \cite{APW}, we shall therefore start with a more general discussion of the expanding source which is based on the space-time current correlator and the space-time form of the coherent component and which is not affected by the semi-classical and small $q$ approximations inherent in the Wigner function approach. We introduce the variables $\tau$, $\eta$ and $x_{\parallel}$, with \begin{equation} \tau \ =\ \sqrt{x_0^2-x_{\parallel}^2}, \qquad \eta \ =\ \frac{1}{2} \ln \frac{x_0+x_{\parallel}}{x_0-x_{\parallel}} \end{equation} Here $\tau$ is the proper time, $x_{\parallel}$ the coordinate in the longitudinal direction (e.g. the collision axis in p-p reactions or the jet axis in $e^{+}-e^{-}$ reactions) and $\eta$ the space-time rapidity. An ansatz which is invariant under boosts of the coordinate frame in longitudinal direction will be considered (Fig.8). Physically this ansatz is motivated by the prejudice that the single inclusive distribution in rapidity is flat. The space-time distributions of the chaotic and of the coherent source and the correlator are then parametrized as \begin{eqnarray} f_{ch}(x) & \sim & \exp\left(-\frac{(\tau-\tau_{0,ch})^2} {(\delta \tau_{ch})^2}\right) \ \exp\left( -\frac{x_{\perp}^2}{R_{ch}^2}\right) \label{eq:chaos1}\\ f_{c}(x) & \sim & \exp\left(-\frac{(\tau-\tau_{0,c})^2} {(\delta \tau_{c})^2}\right) \ \exp\left( -\frac{x_{\perp}^2}{R_{c}^2}\right) \label{eq:coh1} \end{eqnarray} \begin{eqnarray} C(\tau_1-\tau_2,\eta_1-\eta_2,x_{\perp,1}- x_{\perp,2}) & = & \exp \left[-\frac{(\tau_{1}-\tau_{2})^2} {2L_{\tau}^2} -\frac{2 \tau_1 \tau_2}{L_{\eta}^2} \sinh^2\left(\frac{\eta_1-\eta_2} {2}\right) \right. \nonumber\\ & & \ \qquad \left.-\frac{({\bf x}_{\perp,1}-{\bf x}_{\perp,2})^2} {2L_{\perp}^2}\right] \end{eqnarray} The model contains again 10 independent parameters: The proper time coordinates of the chaotic and the coherent source, $\tau_{0,ch}$, $\tau_{0,c}$, their widths in proper time, $\delta \tau_{ch}$ and $\delta \tau_{c}$, the transverse radii, $R_{ch}$ and $R_{c}$, the correlation lengths $L_\tau$, $L_{\perp}$ and $L_{\eta}$, and the chaoticity parameter $p_0$. In order to be able to obtain explicit expressions for the single inclusive distribution and the correlation functions, a further simplifying assumption is made, namely, that $\delta \tau_{ch} = \delta \tau_{c} =0$. Eqs.(\ref{eq:chaos1},\ref{eq:coh1}) then take the form \begin{eqnarray} f_{ch}(x) & \sim & \delta(\tau-\tau_{0,ch})\ \exp\left( -\frac{x_{\perp}^2}{R_{ch}^2}\right)\\ f_{c}(x) & \sim & \delta(\tau-\tau_{0,c})\ \exp\left( -\frac{x_{\perp}^2}{R_{c}^2}\right) \end{eqnarray} Now the results no longer depend on the correlation length $L_\tau$, and one is left with 7 independent parameters: $\tau_{0,ch}$, $\tau_{0,c}$, $R_{ch}$, $R_{c}$, $L_{\perp}$, $L_{\eta}$ and $p_0$. The Fourier integrations necessary to obtain $D(k_1,k_2)$ and $I(k)$ can be performed by doing a saddle point expansion; this should provide a good approximation if \begin{equation} a_i \ \equiv \ \frac{m_{i\perp} \tau_{0,ch}}{2} \gg 1, \qquad b \ \equiv \ \frac{\tau_{0,ch}^2}{2L_{\eta}^2} \gg 1 \label{eq:ab} \end{equation} With the definitions: \begin{equation} R_{L}^2\ \equiv\ \frac{R_{ch}^2 L_{\perp}^2} {R_{ch}^2 + L_{\perp}^2} \label{eq:rlexp} \end{equation} \begin{equation} \gamma_{12} \ \equiv \ \frac{\tau_{0,ch}(m_{1\perp}-m_{2\perp})} {L_{\eta}^2 m_{1\perp}m_{2\perp}} \qquad \tilde{\gamma}_{12} \ \equiv \ \frac{\tau_{0,ch}(m_{1\perp}+m_{2\perp})} {L_{\eta}^2 m_{1\perp}m_{2\perp}} \label{eq:gamma12} \end{equation} the single inclusive distribution can be written as the sum of a chaotic and a coherent term, \begin{equation} E \frac{1}{\sigma} \frac{d^3 \sigma}{d^3k}\ = \left( p_0\ s_{ch}(k)\ +\ (1-p_0)\ s_c(k) \right) \left. E \frac{1}{\sigma}\frac{d^3 \sigma}{d^3 k} \right|_{k=0} \label{eq:eins1} \end{equation} with \begin{eqnarray} s_{ch}(k) & = & \frac{m_{\pi}}{m_{\perp}} \ \exp \left[ - \frac{{\bf k}_{\perp}^2 R^2_{L}}{2}\right]\\ s_{c}(k) & = & \frac{m_{\pi}}{m_{\perp}} \ \exp \left[ - \frac{{\bf k}_{\perp}^2 R^2_{c}}{2}\right] \end{eqnarray} where $m_{\perp}$ is the transverse mass of the pions emitted. The momentum dependence of the chaoticity parameter takes the form \begin{equation} p_r \ =\ p(k_r)\ =\ \frac{p_0}{A_r} \qquad (r=1,2) \label{eq:pr11} \end{equation} with \begin{equation} A_r \ \equiv \ A(k_r) \ =\ p_0 + (1-p_0) \ S_{rr} \qquad (r=1,2) \label{eq:akr1} \end{equation} and \begin{equation} S_{rs} \ = \ \exp \left[- \frac{({\bf k}_{r \perp}^2+{\bf k}_{s \perp}^2) (R^2_{c}-R^2_{L})}{4}\right] \label{eq:sxp12} \end{equation} Unless $R_c=R_L$, the transverse momentum distribution is a superposition of two Gaussians of different widths. The rapidity distribution is uniform, $dN/dy =const.$, as a result of boost-invariance. In opposition to what is assumed usually in simplified pseudo-hydrodynamical treatments, the transverse radius of the chaotic source, $R_{ch}$, cannot be determined independently by measuring only the single inclusive distribution, as the quantity $R_L$ which sets the scale for the mean transverse momentum of the chaotically produced particles is a combination of $R_{ch}$ and the correlation length $L_{\perp}$. We recall that the second order correlation functions \begin{equation} C^{++}_2({\bf k}_1,{\bf k}_2)\ =\ 1 +\ 2 \sqrt{p_1(1-p_1) \cdot p_2(1-p_2)} \ T_{12}\ \cos(\phi^{ch}_{12} - \phi^c_1+\phi^c_2)\ +\ p_1p_2 \ T^2_{12} \label{eq:c2plus11} \end{equation} are defined in terms of the magnitudes and phases of $d_{rs}$ and $\tilde{d}_{re}$, $T_{rs}$, $\tilde{T}_{rs}$, $\phi^{ch}_{rs}$ and $\tilde{\phi}^{ch}_{rs}$ of the chaotic source as well the phases of the coherent component, $\phi^{c}_{r}$. The expressions for these quantities read for an expanding source: \begin{eqnarray} T_{12} & = & \left(1+\gamma_{12}^2\right)^{-1/4} \exp \left[-\frac{b}{1+\gamma_{12}^2}(y_1-y_2)^2 - \frac{({\bf k}_{1\perp}-{\bf k}_{2\perp})^2(R^2_{ch}-R^2_ {L})}{8}\right]\nonumber\\ \tilde{T}_{12} & = & \left(1+\tilde{\gamma}_{12}^2\right)^ {-1/4}\exp \left[-\frac{b}{1+\tilde{\gamma} _{12}^2}(y_1-y_2)^2 - \frac{({\bf k}_{1\perp}-{\bf k}_{2\perp})^2(R^2_{ch}-R^2_ {L})}{8}\right] \end{eqnarray} and \begin{eqnarray} \phi_{12}^{ch} & = & \frac{b \gamma_{12}} {1+\gamma_{12}^2}(y_1-y_2)^2\ -\ \tau_{0,ch} (m_{1\perp}-m_{2\perp})\ - \ \frac{1}{2}\arctan \gamma_{12}\\ \tilde{\phi}_{12}^{ch} & = & \frac{b \tilde{\gamma}_{12}}{1+\tilde{\gamma}_{12}^2}(y_1-y_2)^2\ -\ \tau_{0,ch} (m_{1\perp}+m_{2\perp})\ - \ \frac{1}{2}\arctan \tilde{\gamma}_{12}\\ \phi_{j}^{c} & = & -\ \tau_{0,c} m_{j\perp} \end{eqnarray} One thus finds again that the correlation functions do not depend separately on the geometrical radii $R$ or on the correlation lengths $L$ but rather on the combination $R_L$ defined in (\ref{eq:rlexp}). This expression reduces in the limit $R_{ch}\gg L$ to $L$ and in the limit. $R_{ch} \ll L$ to $R$. The model considered in \cite{GKW} is thus a particular case of the space-time approach \cite{APW} for $L=0$. As in the static case the tilde terms give rise to the particle-antiparticle correlations. For a purely chaotic system the intercept of the $\pi^+\pi^-$ correlation function is \begin{equation} C^{+-}_2(k,k) \ = \ 1 \ + \ \left(1+4\left(\frac{b}{a}\right)^2\right)^{-1/2} \ = \ 1 \ + \ \left(1+4\left(\frac{\tau_{0,ch}} {m_{\perp}L_{\eta}^2}\right)^2\right)^{-1/2} \label{eq:rol1} \end{equation} We conclude this section with the observation that in \cite{APW} a correspondence between the correlation length $L$ in the primordial correlator $C(x-y)$ and the temperature $T$ for a pion source that exhibits thermal equilibrium was established. In the limit of large volume $V\propto R^3$ and lifetime $R_0$ of the system, it reads \begin{equation} L \ \sim \ T^{-1} \label{eq:A.11} \end{equation} \subsection{The Wigner function approach} As mentioned previously, the experimental observation of the fact that the two particle correlation function depends not only on the difference of momenta $q=k_1-k_2$ but also on the sum $k_1+k_2$ led to the introduction and the use \cite{Pratt} of a ``source" function within the well known Wigner function formalism of quantum mechanics \footnote{An attempt to consider the correlation between coordinates and momentum was also performed earlier within the ordinary wave function formalism by Yano and Koonin \cite{Yano} who proposed a formula for the second order correlation function of form (\ref{eq:3}). However this form turned out subsequently to have pathological features as it leads in some cases to a violation of the lower bounds of the correlation function (cf. section 5.1.6). The reason for this mis-behaviour was mentioned in section 4.6 and will also be discussed in the following.}. While it turned out later that this property of the correlation function can be derived within the current formalism without the approximations involved by the Wigner formalism, this formalism is still useful when applied within a hydrodynamical context. On the other hand it should be clear that there is no justification for using this formalism without a full hydrodynamical apparatus (cf. also below). This message has apparently not yet come through, as many phenomenological and experimental papers have continued to use this restrictive formalism. The Wigner function approach for BEC was proposed in a non-relativistic form in Ref.\cite{Pratt} and subsequently generalized in \cite{Schlei}, \cite{APW}(cf. also \cite{shuryak},\cite{PaGyGa}). The Wigner function called also source function, $g(x,k)$, may be regarded as the quantum analogue of the density of particles of momentum $k$ at space-time point $x$ in classical statistical physics. It is defined within the wave function formalism as \begin {eqnarray} g({\bf x},{\bf k},t)& =&\int d^3x^{'}\psi^*\left({\bf x}+\frac{1}{2}{\bf x^{'}},t\right)\psi\left({\bf x}-\frac{1}{2}{\bf x^{'}},t\right)e^{i{\bf kx{'}}}\nonumber\\ &=&\int d^3k^{'}\psi^*\left({\bf k}+\frac{1}{2}{\bf k{'}},t\right) \psi\left({\bf k}-\frac{1}{2}{\bf k{'}},t\right)e^{-i{\bf k{'}x}} \label{eq:Wigner} \end{eqnarray} and is related to the coordinate and momentum densities by the relations \begin{equation} n({\bf x},t)=\int d^3kg({\bf x},{\bf k},t) \label{eq:(B)} \end{equation} and \begin{equation} n({\bf k},t)=\int d^3xg({\bf x},{\bf k},t) \label{eq:(C)} \end{equation} respectively. Due to its quantum nature the function $g(x,k)$ takes real but not necessarily positive values. Although eq.(\ref{eq:Wigner}) is nothing but a definition which does not imply any approximation, its form suggests that it might be useful when simultaneous information about coordinates and momenta are desirable, provided of course that the limits imposed by uncertainty relations are not violated. As a matter of fact as will be shown below, the Wigner function is useful for BEC only if a more stringent condition is fulfilled, namely that the difference of momenta $q$ of the pair is small, as compared with the individual momenta of the produced particles. It is thus clear that its applicability is more restricted than that of the classical current approach, where only the ``no recoil" condition, i.e. small total momentum of produced particles, as compared with the momentum of incident particles, must be respected. This circumstance is often overlooked when comparing theoretical predictions based on the Wigner approach with experimental data. In particular herefrom also follows that the application of the Wigner formalism to data has necessarily to take into account from the beginning resonances which dominate the small $q$ region. It turns out that the use of the Wigner function for BEC is justified only in special cases as e.g. when a coherent hydrodynamical study is performed, i.e. when the observables are related to an equation of state and when simultaneously single and higher order inclusive dstributions are investigated. Unfortunately only very few papers, where the Wigner function formalism is used, are bona fide hydrodynamical studies. The majority of ``theoretical" papers in this context are ``pseudo-hydrodynamical" (cf. sections 5.1.5 and 5.1.6) in the sense that the form of the source function is expressed in terms of {\em effective} physical variables like temperature or velocity, which are not related by an equation of state. In this case the application of the Wigner approach is a ``luxury" which is not justified. This is a fortiori true since, as will be shown in the following, the Wigner approach is mathematically not simpler that the classical current approach, of which it is a particular case. Thus the space-time model \cite{APW} presented above (cf. section 4.8) is more general than the Wigner approach, albeit it is not more complicated and has not more independent parameters In second quantization $g(x,k)$ is defined in terms of the correlator $< a^\dagger ( {\bf k}_i ) a ( {\bf k}_j )>$ by the relation \begin{equation} < a^\dagger ( {\bf k}_i ) a ( {\bf k}_j ) > = \int d^4 x \exp [- i x_\mu ( k_i{}^\mu - k_j{}^\mu ) ] \cdot g [ x , \textstyle{\frac{1}{2}} ( k_i + k_j ) ] \label{eq:wigdef} \end{equation} This is a natural generalization of (\ref{eq:(C)}) to which it reduces in the limit $k_{i}=k_{j}$. Accordingly, for the second order correlation function one writes \begin{equation} P_2({\bf k}_1,{\bf k}_2) \ =\ \int d^4x_1 \int d^4x_2 \left[g\left(x_1,k_1\right) g\left(x_2,k_2\right)\ +\ g\left(x_1,K\right) g\left(x_2,K\right) \ \exp\left[iq_\mu(x^\mu_1-x^\mu_2)\right] \right] \label{eq:P2} \end{equation} where $K^\mu=(k_1^\mu+k_2^\mu)/2$ and $q^\mu=k_1^\mu-k_2^\mu$ are the mean momentum and momentum difference of the pair \footnote{For neutral particles, there are additional contributions to $P_2({\bf k}_1,{\bf k}_2)$ which play a role for soft particles and which will be neglected here.}. The relation between this Wigner approach and the classical current approach is established by expressing the rhs of eq.(\ref{eq:wigdef}) in terms of the currents. One has \begin{equation} g(x,k)\ =\ \frac{1}{2 \sqrt{E_{i}E_{j}}(2\pi)^{3}} \ \int d^4z \ <J\left(x+\frac{z}{2}\right) J\left(x-\frac{z}{2}\right)> \ \exp\left[-ik^\mu z_\mu\right] \label{eq:wigcur} \end{equation} The derivation of the Wigner formalism from the classical current formalism has the important advantage that it avoids violations of quantum mechanical bounds as those mentioned previously. Note that in the rhs of eq.(\ref{eq:P2}) enters the off-mass shell average momentum $\frac{1}{2}(k_1+k_2)$ which is not equal to the on-mass shell average $K=\frac{1}{2}\sqrt{E^2-m_{1}^2+m_{2}^2}$ where $E$ is the total energy of the pair (1,2). This means among other things that in this approach it is not enough to postulate the source function $g$ in order to determine the second (and higher order) correlation function $C_2$, but further assumptions are necessary. Usually oe neglects the off-mass shellness i.e. one approximates $E$ by the sum $E_1+E_2$ where $E_i$ are the on-shell energies of the particles (1,2), which means that one neglects quantum corrections \footnote{That these corrections can be important has also been shown in \cite{Bertsch94}.} which is permitted as long as $k_1-k_2=q$ is small \footnote{It is sometimes argued that the relevant $q$ range in BEC is given by $D^{-1}$ where $D$ is a typical length scale of the source and therefore for heavy ion reactions this should be allowed. This is not quite correct, because the {\em shape} of the correlation function from which one determines the physical parameters of the source is not given just by the values of the correlation function near the origin, but depends also on its values at large $q$.}. As mentioned already, the use of the Wigner formalism is worthwhile within a true hydrodynamical approach when the relation with the equation of state is exploited. In this case the probability to produce a particle of momentum $k$ from the space-time point $x$ then depends on the fluid velocity, $u^\mu(x)$, and the temperature, $T(x)$, at this point, and one has \begin{eqnarray} &&\sqrt{E_i E_j} < a^\dagger ( {\bf k}_i ) a ( {\bf k}_j ) > \nonumber\\ &&= \frac{1}{(2\pi)^3} \int\limits_\Sigma \frac{\textstyle{\frac{1}{2}} (k_i{}^\mu + k_j{}^\mu) d \sigma_\mu (x_\mu)}{\exp \left[\displaystyle{ \frac{\textstyle{\frac{1}{2}} (k_i{}^\mu + k_j{}^\mu) u_\mu (x_\mu)}{T_f (x_\mu)} } \right] - 1 } \cdot \exp [- i x_\mu (k_i{}^\mu - k_j{}^\mu)] \label{eq:hydro} \end{eqnarray} Here, $d\sigma^\mu$ is the volume element on the freeze-out hypersurface $\Sigma$ where the final state particles are produced. We will discuss applications of this approach in section 5. \subsubsection{Resonances in the Wigner Formalism} For a purely chaotic source, the formalism to take into account the effects of resonance decays on the Bose-Einstein correlation function can be found, e.g. in \cite{Grass,Gyu89}. An extension of this approach is due to \cite{Bolz} and \cite{Schlei} which allows to consider also the effect of coherence and provides rather detailed and subsequently, apparently, confirmed predictions for heavy ion reactions. It is based on the Wigner function formalism. The correlation function of two identical particles of momenta ${\bf k}_1$ and ${\bf k}_2$ can be written as \begin{equation} C_2( {\bf k}_1 , {\bf k}_2 ) \:=\: 1 + \frac{A_{12}\:A_{21}}{A_{11}\:A_{22}} \label{eq:001} \end{equation} where the matrix elements $A_{ij}$ are given in terms of source functions $g(x,k)$ as follows \begin{equation} A_{ij}\:=\:\sqrt{E_i E_j}<a^\dagger({\bf k}_i) a({\bf k}_j)> \:=\:\int d^4x\: g(x_\mu,k^\mu)\: e^{\textstyle{i q^\mu x_\mu}} \label{eq:002} \end{equation} A typical source function reads \begin{equation} g(x_\mu,p^\mu)\: =\: g_\pi^{dir}(x_\mu,p^\mu) \:+ \sum_{res=\rho,\omega,\eta,...} g_{res \rightarrow \pi}(x_\mu,p^\mu) \label{eq:003} \end{equation} where the labels $dir$ and $res \rightarrow \pi$ refer to direct pions and to pions which are produced through the decay of resonances (such as $\rho$, $\omega$, $\eta$, ... etc.), respectively. The contribution from a particular resonance decay is estimated in \cite{Bolz},\cite{Schlei} using kinematical and phase space considerations as well as the source function of that resonance. The source distribution for the direct production of pions and resonances is calculated assuming local thermodynamcal and chemical equilibrium as is appropiate for a hydrodynamical treatment. \begin{equation} g_{\alpha}^{dir}(x_\mu,p^\mu)\: =\: \frac{2J+1}{(2\pi)^3} \int_{\Sigma} \: \frac{p^\mu d\sigma_\mu(x'_\mu) \: \delta^4(x_\mu-x'_\mu)} {\exp \left[ \displaystyle{\frac{p^\mu u_\mu(x'_\mu) -B_{\alpha}\mu_B(x'_\mu)-S_{\alpha}\mu_S(x'_\mu)} {T_f(x'_\mu)}} \right] - 1} \label{eq:00x} \end{equation} Here $\alpha$ denotes the particular resonance and $d\sigma^\mu$ is the differential volume element and the integration is performed over the freeze-out hypersurface $\Sigma$. $u^\mu(x)$ and $T_f$ are the four velocity of the fluid element at point $x$ and the freeze-out temperature, respectively. $B$ and $S$ are the baryon number and the strangeness of the particle species labeled $\alpha$, respectively, and $\mu_B$ and $\mu_S$ are the corresponding chemical potentials. $J$ is the spin of the particle. This approach is then extended \cite{Schlei} to include also a coherent component rsulting in a second order correlation function of the form \begin{equation} C_2( {\bf k}_1 , {\bf k}_2 )= 1 + 2\:p_{eff}\:(1-p_{eff})\: Re\:d_{12} + p^2_{eff}\:|d_{12}|^2 \label{eq:def} \end{equation} where $p_{eff}$ is an effective chaoticity related to the true chaoticity $p_{dir}$ \footnote{It is assumed that only directly produced particles have a coherent component.} via \begin{equation} p_{eff}\: =\: p_{dir} \:(1 - f^{res})\: +\: f^{res}. \label{eq:017} \end{equation} $f^{res}$ is the fraction of particles arising from resonances. The form of this equation us the same as that derived previously for a partially coherent source within the current formalism and manifests the characteristic two component structure. The sensitivity of the correlation function on the chaoticity parameter $p_{dir}$ can be estimated e.g. from the intercept (cf. eq. (\ref{eq:def})) \begin{equation} I_o\: =\: C_2({\bf k},{\bf k})\: =\: 1 + 2\: p_{eff} - p_{eff}^2 \label{eq:018} \end{equation} Defining the fractions of pions produced directly (chaotically and coherently) and from resonances, \begin{equation} f_{ch}^{dir} = \frac{p_{dir}\: A_{ii}^\pi}{A_{ii}}\:, \quad f_{co}^{dir} = \frac{(1-p_{dir})\:A_{ii}^\pi}{A_{ii}}\:, \quad f^{res} = \frac{\Sigma_{res=\rho,\omega,\eta,...}\: A_{ii}^{res}}{A_{ii}} \label{eq:014} \end{equation} with \begin{eqnarray} f_{ch}^{dir} + f_{co}^{dir} + f^{res} = 1 \nonumber \label{eq:015} \end{eqnarray} In fig. 9 the intercept of the correlation function is shown as a function of $p_{dir}$ and $f^{res}$. In order to read off the fraction of {\it direct} chaotically produced particles, $p_{dir}$, from the intercept of the correlation function, one has to extract the effective chaoticity $p_{eff}$ according to eq. (\ref{eq:018}) and then correct for the fraction of pions from resonance decays. Note that $p_{eff}<p_{dir}$. In particular, if a large fraction of pions arise from resonance decays, $p_{eff}\rightarrow 1$ and it will take very precise measurements of the two-particle correlation function at small ${\bf q}$ to determine the true chaoticity, $p_{dir}$. A further complication arises if a fraction of particles are the decay products of long-lived resonances \footnote{In some papers \cite{Cs1996} pions originating from long lived resonances are associated with a ``halo" while those coming form short lived resonances or directly produced are related to a ``core". Then it is claimed among other things that the ``core" parameters of the source (like radius and $\lambda$ factor) can be obtained from the data just by eliminating the small $Q$ points and fitting only the remaining points. Even if such a separation would be clear cut (there are doubts about this because of the $\omega$ resonance), it would be of course dependent on the resolution of the detector.}. This topic as well as the problem of misidentification are discussed in \cite{Bolz}. \subsection{Dynamical models of multiparticle production and event generators} Due to the lack of a full fledged theory of multiparticle production in strong interactions different models of multiparticle dynamics were proposed. Bose-Einstein correlations measurements have been used either to test a particular model or/and to determine some of its parameters. Among other things these models might be used to predict the dependence of the chaoticity on the type of reaction. In the following we will sketch the main theoretical ideas on which these models are based and mention briefly their relation to data. One of the first models of particle production from which definite predictions on BEC can be derived is the Schwinger model \cite{Schwinger} for $e^{+}-e^{-}$ reactions. It visualizes the source as an one-dimensional string in a coherent state and thus predicts the absence of any bunching effect. A similar prediction follows from the bremsstrahlung model \cite{Gunion}. Recoilless bremsstrahlung can be described by a classical current which also corresponds to a coherent state. Given the fact that in all hadron production processes BEC have been seen, it follows that the above two models are ruled out by experiment. More complex predictions follow from a {\em dual topological} model due to Giovannini and Veneziano \cite{GV} which associates the processes $e^{+}-e^{-} \rightarrow hadrons$ to a unitarity cut in one plane, reactions due to Pomeron exchange to a cut in two planes, and annihilation reactions $\bar{p}-p$ to a cut in three planes. This model predicts then among other things that for $\pi^-\pi^-$ BEC the intercepts $C_2(k,k)$ of the second order correlation functions for the above reactions should satisfy the following relation: \begin{equation} [C_2^{e^+e^-}(k,k)-2]/[C_2^{\pi p}(k,k)-2]/[C_2^{ann}(k,k)-2]=1/\frac{1}{2}/\frac{1}{3}. \label{eq:GV} \end{equation} (A similar, but quantitatively different relationship is predicted for $\pi^+\pi^-$ correlations.) Despite the fact that since the publication of this paper in 1977 many experimental BEC studies of these reactions were performed, the above predictions could not yet be tested quantitatively in a convincing manner. This is due among other things to experimental difficulties (cf. below, in particular factors (a),(b)) and illustrates the unsatisfactory status of experimental BEC investigations. A qualitative remark can however be made: The expectation that the annihilation reaction leads to more bunching than other reactions is apparently confirmed (cf. e.g. ref.\cite{Angel} and section 2.1.1). As to the difference between $e^{+}-e^{-}$ reactions and hadronic reactions the experimental situation is rather confused (cf. also below). A somewhat related dynamical model based on Reggeon theory, was proposed already in \cite{LRT}. A straightforward extension of of this formalism to heavy ion reactions does not work as it predicts that the longitudinal radius is of ``hadronic" size \cite{CK}. A different approach to BEC based on the classical current formalism is proposed in \cite{Casado}. The currents are associated with the chains of the dual parton model, and contrary to what is assumed in other applications of the classical current formalism, all the phases of these elementary currents are fixed, so that the source is essentially coherent. This is a special case of the classical current approach presented in sections 4.2, 4.3 and 4.8 where allowance is made both for a chaotic and coherent component. The model is intended to work for $p-p$ reactions where the authors state that resonances do not play an important role. It explains, according to the authors, the dependence of the $\lambda$ parameter in the empirical formula for the second order correlation function \begin{equation} C_2= 1 + \lambda \exp (-R^2 q^2) \label{eq:lambda1} \end{equation} on the multiplicity and energy. Unfortunately the claim that in $p-p$ reactions pions are only directly produced is unfounded. Furthermore there are other factors which influence the multiplicity dependence of $\lambda$ (cf. section 6.2) which are not considered in\cite{Casado} and which are of more general nature. An orthogonal point of view for the interpetation of the same $\lambda$ factor (also for directly produced pions, only,) is due to \cite{BSWW}. In this approach the source is made up of totally chaotic elementary emitting cells which are occupied by identical particles subject to Bose-Einstein statistics. Different cells are independent so that correlations between particles in different cells lead to $\lambda=0$, while correlations between particles in the same cell are characterized by $\lambda=1$. From the interplay of these two types of correlations, one obtains with an appropiate weighting, large $\lambda$ values in $e^{+}-e^{-}$ reactions and small $\lambda$ values in $p-p$ reactions, as in \cite{Casado}, but within a completely different approach. We conclude the discussion of these two approaches by the following remarks. Besides the reservations about the role of directly produced pions in BEC expressed above and which presents the two approaches in a rather academic light, it is unclear whether the $\lambda$ factor in $e^{+}-e^{-}$ reactions is larger than in $p-p$ reactions as assumed in \cite{BSWW}. This issue awaits a critical analysis of the specific experimental set-ups. The fact that quite different approaches lead to similar conclusions about the $\lambda$ factor confirms that the parametrization of the second order correlation function in the form (\ref{eq:lambda1}) is, as pointed out already in section 2.2, inadequate. We discuss now other two, closely related, approaches, which make more detailed predictions about the form of the correlation function in $e^{+}-e^{-}$ reactions: \cite{stringbowl}, \cite{bowler} on the one hand, and \cite{stringander},\cite{AR1}, \cite{AR2} on the other. Both approaches are based on a variant of the string model (for a more extended review of this topic cf. e.g. \cite{Bow0}). Such a string represents a coloured field formed between a quark $q$ and an antiquark $\bar{q}$, which tend to separate. Because of confinement the breakup of the string can be materialized only through creation of new $q\bar{q}$ pairs, which are the mesons produced in the reaction. The difference between the Schwinger model of confinement on the one hand and the models of Bowler and Andersson-Hofmann-Ringn\'er on the other is that in the former the field couples directly and locally to a meson, while in the latter ones the quarks, which constitute the meson, are created at different points. This feature destroys the coherence inherent in the Schwinger model and makes possible the Bose-Einstein bunching effect. For massless quarks the second order correlation function can be approximated by the relation \cite{stringander} \begin{equation} C_2= 1+<\cos(\kappa\Delta A)/\cosh(b\Delta A/2)> \label{eq:AH} \end{equation} where $\Delta A$ denotes the difference between the space-time areas of coloured fields spanned by the two particles, $\kappa$ is the string tension and $b$ a parameter characterizing the decay probability of the string. For massive quarks the formulae become more involved and were approximated analytically in \cite{stringbowl} or calculated numerically in \cite{stringander},\cite{AR1}, \cite{AR2}. In this model the correlation function depends both on the difference of momenta $k_1-k_2$ as well as on their sum $k_1+k_2$, reflecting the correlation between the momentum of the particle and the coordinate of its production point. This is a consequence of the fact that string models use a Wigner function type approach. From the above equations follows that there are two length scales in the problem, one associated with $\kappa$ and another with $b$. Phenomenologically these correspond to $q_{\parallel}$ and $q_{\bot}$. Both these lengths are correlation lengths rather than geometrical radii. (As a matter of fact there is no geometrical radius in the string model.) Their magnitudes are quite different. In both string approaches \cite{bowler}, \cite{stringander} one obtains a difference between BEC for identically charged and neutral pions as found in \cite{apw}. However while in \cite{stringbowl} there is room for coherence, this is apparently not the case for \cite{stringander},\cite{AR1}, \cite{AR2}, which predict a totally chaotic source. Furthermore in \cite{stringbowl} an energy dependence of the BEC is predicted (the correlation function is expected to shrink with increasing energy), while in \cite{stringander},\cite{AR1},\cite{AR2} the correlation function does not depend on energy. \footnote{To make the model more realistic in ref.\cite{stringander} resonances were included according to the variant of the Lund model (JETSET) in use at that time (1986) and agreement with $e^{+}-e^{-}$ data was found. Subsequently however it was pointed out in \cite{Verb} that some resonance weights used in \cite{stringander} were incorrect, so that the agreement mentioned above is probably accidental.} A rather discordant note in this string concerto \cite{stringbowl}, \cite{stringander} is represented by the paper by Scholten and Wu \cite{Scholten}. These authors, using a different hadronization mechanism conclude that dynamical correlations, at least in $e^{+}-e^{-}$ reactions dominate over BEC correlations so that BEC cannot be used to infer information about the size and lifetime of the source \footnote{What concerns $e^{+}-e^{-}$ reactions similar skepticism was expressed by Haywood \cite{Hay}.}. This point of view seems too extreme, as it is contradicted by some simple empirical observations: in $e^{+}-e^{-}$ reactions, as well as in all other reactions, correlations between identical particle are observed which are much stronger than those of non-identical ones, the correlation functions are (in general) monotoneously decreasing functions of the momentum difference $q$ and in nuclear reactions the ``radii" obtained from identical particole correlations increase with the mass number of the participating nuclei. All these observations are in agreement with what one would expect from BEC, which suggests that dynamical correlations cannot distort too much this picture. However a thorough comparison of BEC in different reactions, using the same experimental techniques, appears highly desirable. \paragraph{Event generators} The model \cite{stringander} was implemented by Sjostrand \cite{Sjos} into JETSET under the name LUBOEI by modifying a posteriori the momenta of produced pions so that identical pairs of pions are bunched according to \cite{stringander}. This manipulation ``by hand" violates energy-momentum conservation which was imposed at the beginning in JETSET. To compenate for this, the momenta are rescaled so that energy-momentum conservation is restored. However this rescaling introduces spurious long-range correlations, which bias the BEC. Nevertheless, in general this program leads to a reasonable description of the bunching effect in the second order correlation function \footnote{Cf. however ref. \cite{FW97}}. More refined features of BEC which reflect the quantum mechanical essence of the effect, cannot be obtained of course. One reason for this, of rather technical nature, is due to the fact that the ad-hoc modification of two-particle correlations does not yet include many-body correlations, reflected in the symmetrization (or antisymmetrization) of the entire wave-function. Another reason of fundamental character is the fact that event generators like any Monte Carlo algorithm deal in general with probabilities \footnote{For heavy ion reactions the "Quantum Molecular Dynamics" (QMD) model \cite{QMD} attempts to surpass this deficiency by using wave functions rather than probabilities as input. However this model also neglects (anti)symmetrization effects and cannot be used for interferometry studies.} and therefore cannot account for quantum effects, which are based on phases of amplitudes \footnote{It is interesting to mention that for {\em one} string Andersson and Hofmann (\cite{stringander})proposed a formulation of the BEC effect in terms of {\em amplitudes}. However this procedure cannot be used to generate events.}. The event generator JETSET was further developped by L\"onnblad and Sj\"ostrand , (\cite{SL1}, \cite{SL2}) and used to estimate the influence of BEC on the determination of the mass of W in $e^{+}-e^{-}$ reactions, a subject of high current interest for the standard model and in particular for the search of the Higgs particle. This effect was also studied using different event generators in \cite{Jadach} and \cite{Kart}. The argument of L\"onnblad and Sj\"ostrand is the following. Consider the reaction $e^{+}-e^{-}\rightarrow W^+W^- \rightarrow q_1\bar{q}_2q_3\bar{q}_4$ when both $W$-s decay into hadrons. Then according to \cite{SK} the typical space-time separation of the decay vertices of the $W^+$ and the $W^-$ is less than 0.1 fm (at LEP 2 energies) and thus much smaller than a typical hadronic radius ($\sim 0.5$ fm. There will thus be a Bose Einstein interference between a pion from the $W^+$ and a pion (with the same charge) from the $W^-$ and one cannot establish unambigously the ``parenthood" of these pions. This prevents then in this model a precise determination of the invariant mass of the $W$-s. In \cite{SL1} algorithms for the inclusion of this effect into the determination of the mass of the $W$ are proposed and for certain scenarios mass corrections of the order of 100 MeV at 170 GeV c.m. energy are obtained. However, as emphasized in \cite{SL1} other scenarios with less or no effect of BEC on the mass determination of the $W$ are possible. Thus in \cite{Jadach} and \cite{Kart} effects of the order of only 20 MeV are found. For more details we refer the reader to the original literature. This aspect of BEC is interesting in itself as it illustrates the possible applications of this effect in electroweak interactions, a domain which is beyond the usual application domain of BEC, i.e. that of strong interactions. The Lund model was applied also to heavy ion reactions and then extended to include BEC (e.g. the SPACER version \cite{Pra90} of the Lund model). The topic of event generators for heavy ion reactions is of current interest because of the ongoing search for quark gluon plasma. Padula, Gyulassy and Gavin \cite{PaGyGa} suggested to use for this purpose the Wigner function formalism in order to take into account explicitly the correlation between momenta and coordinates, as implied by the inside-outside cascade approach. This is evidently another way of expressing the non-stationarity of the correlation function mentioned above. (An explicit introducton of momentum-coordinate correlations in particle physics event generators like JETSET/LUBOEI is not necessary, because the non-stationarity is delivered ``free house" by the string model used in the LUND generator.) On the other hand the Wigner formalism may present also another advantage as emphasized more recently by Bialas and Krzywicki \cite{BK}. This has to do with the important difficulty mentioned also above and which is inherent in all event generators, namely the probabilistic nature of Monte Carlo methods. The Wigner function has in certain limits the meaning of a wave function and thus provides quantum amplitudes. The proposal of Bialas and Krzywicki consists then in starting from the single particle distribution $\Omega_0({\bf k})$ constructed from non-symmetrized particle wave functions as produced by conventional event generators and writing the Wigner function \begin{equation} g({\bf k};{\bf x}) = \Omega_0({\bf k}) w({\bf k};{\bf x}) \label{eq:BK8} \end{equation} where w({\bf k};{\bf x}) is the conditional probability that given that the particles with momenta $k_1,, k_2, ...k_n$ are present in the final state, they are produced at the points $x_1, ...x_n$. Then the art of the model builder consists in guessing the probability $w({\bf k};{\bf x})$. This may be easier than guessing from the beginning the exact Wigner function. For example a simple ansatz would be to assume that the likelyhood to produce a particle from a given space point is statistically independent of what happens to other particles. This means that $w({\bf k};{\bf x})$ can be factorized in terms of the individual particles. Implementations of this scheme were discussed in \cite{Wosiek}, \cite{FWIT} When using the Wigner formalism or any model (like those used in event generators) which specifies momenta and coordinates simultaneously, one must of course watch that the correlations between momenta and coordinates do not become too strong. This apparently has not always been done.\footnote{E.g. it is unclear to us whether the ``pure" multiple scattering approach of \cite{Humanic} satisfies the above constraint.} That such a procedure is dangerous since it can lead to unphysical {\em antibunching} effects, i.e. to violation of unitarity was already mentioned in \cite{Axel} (cf. section 4.6). This point has been reiterated recently e.g. in \cite{Pratt97},\cite{WFKMSZ},\cite{Martin}. Concluding this section one should emphasize that event generators are just an experimental tool, sometimes useful in the design of detectors or for getting rudimentary information about experimentally unaccessible phase. Often they are however ab-used, e.g. to search for ``new" phenomena: if agreement between data and event generators is found, one states that no ``new " physics was found. Such a procedure is injustified, because agreement with a model or an event generator is often accidental. Furthermore, for the reasons mentioned above event generators cannot be used to obtain the ``true" correlation function, i.e. they are no substitute for a bona fide HBT experiment (cf. also \cite{Aich} for a critical analysis of transport models from the point of view of interferometry). \subsection{Experimental problems} The confrontation of model predictions with experimental BEC data has been hampered by two major facts: (i) most models are idealizations, i.e. they use assumptions which are too strong. Examples of such assumptions are: neglect of final state interactions, boost invariance, particular analytical forms of the correlation functions; (ii) for various reasons in almost no BEC experiment so far a ``true" or ``complete" correlation function was measured, i.e. a correlation function as is defined by \begin{equation} C_2({\bf k}_1,{\bf k}_2) = \frac{P_2({\bf k}_1,{\bf k}_2)}{P_1({\bf k}_1)P_1({\bf k}_2)}. \label{eq:defC2} \end{equation} Here $P_2$ and $P_1$ are the double and single inclusive cross sections respectively. Note in particular that the single inclusive cross sections in the denominator are defined in terms of the same density matrix as the double inclusive in the numerator. This means that these cross sections and the corresponding ratios have to be measured for each event separately and only after that one is allowed to average over events. What is measured usually instead is a function which differs from eq.(\ref{eq:defC2}) in two respects: the normalization of $C_2$ is not done in terms the product of single inclusive cross sections $P_1$ but in terms of a ``background" double inclusive cross section, which is obtained either by considering pairs of (identically charged) particles which come from different events, or by considering oppositely charged particles, or by simulating $P_2$ with an event generator which does not contain BEC. This means that the structure of the event is falsified and therefore this kind of normalization biases the results. Furthermore, one does not (yet) measure the full correlation function $C_2$ in terms of its six independent variables, but rather projections of it in terms of single variables like the momentum difference $q$, rapidity difference $y_1-y_2$ etc. Last but not least, the intercept of the correlation function, which contains important information about the amount of coherence, cannot be really be measured at present because (a) one does not yet control sufficiently well the final state interactions which contribute to the intercept; (b) its experimental determination implies an extrapolation to $q=0$. Such an extrapolation can be performed only if the analytical form of the correlation function at $q\geq 0$ is known, which is not the case. For these reasons at present it is difficult to test quantitatively a given model, except when its predictions are very clear-cut. This circumstance limits certainly the usefulness of BEC as a tool in determining the exact dynamics of a reaction. \setcounter{equation}{0} \section{Applications to ultrarelativistic nucleus-nucleus collisions} \subsection{BEC, hydrodynamics and the search for quark-gluon plasma} The use of BEC in the search for quark-gluon plasma is in most cases based on hydrodynamics. This is so because the space-time evolution of the system is given by the equations of hydrodynamics the solutions of which are different depending whether a QGP is formed or not. In this way hydrodynamics also provides information about the equation of state (EOS). QGP being a (new) {\em phase} it is described by a specific EOS which is different from that of ordinary hadronic matter. The proof that this phase has been seen must include information about its EOS and thus the combination of hydrodynamics with BEC constitutes the only consistent way through which the formation of QGP can be tested. QCD predicts that the phase transition from hadronic matter to QGP takes place only when a critical energy density is exceeded. To measure this density we need to know the initial volume of the system. While via photon interferometry (cf. section 4.5 and 4.6) one can in principle measure the dimensions (and thus the energy density) of the initial state, hadron interferometry yields information only about the final freeze-out stage when hadrons are created. To obtain information about the initial state with hadronic probes, hydrodynamical models have to be used in order to extrapolate backwards from the final freeze-out stage where hadrons are created, to the interesting initial stage. The lifetime of the system as given by BEC is also an important piece of information for QGP search. Indeed, in order to decide whether we have seen the new phase, we have to measure the lifetime of the system. Only lifetimes exceeding significantly typical hadronic lifetimes ($10^{-23} sec.$) could prove the establishment of QGP (cf. below) \subsubsection{General remarks about the hydrodynamical approach} Besides the main advantages of hydrodynamics related to the information about initial conditions, freeze-out and equation of state, for the study of BEC in particular hydrodynamics is very useful because it provides also the single inclusive distributions which are intimately connected with higher order distributions as well as the weights and the space-time and momentum distributions of resonances, which influence strongly the correlations. The phenomenological applications of the hydrodynamical approach to data are however hampered by two circumstances. (i) While the ultimate goal of BEC is the extraction of the minimum set of parameters which include radii and coherence lengths both for the chaotic and coherent components of the source, in practice, mainly because of limited statistics (but also because of an inadequate analysis of the data) one had to limit oneself to the determination of a reduced number of parameters, which we will call in the following ``effective radii" $R_{eff}$ and ``effective chaoticity" $p_{eff}$. In reality $R_{eff}$ is a combination of correlation lengths\footnote{In the following the concept of length refers to space-time.} $L$ and geometrical lengths $R$ as introduced in sections 4.3 and 4.8. Only in the particular case that one length scale is much smaller than the other, one can assume that one measures a ``pure" radius or a ``pure" correlation length. For simplicity in the following we shall assume that this is the case and in particular we will assume that $R \gg L$ so that, $R_{eff}$ reduces to $L$. This limit might perhaps correspond to what is seen in experiment, if one considers the expansion of the system in the hadronic phase. (In the high temperature limit $L\approx T^{-1}$ (cf. section 4.8). (ii) The presentation of the data is yet biased by theoretical prejudices. Instead of a consistent hydrodynamical analysis, much simplified models are used (cf. section 5.1.5 where these models are presented under the generic name of pseudo-hydrodynamics) for this presentation and therefore to obtain the real physical quantities, one would have to solve a complicated mathematical ``inverse" problem, i.e. one would have to reconstruct the raw data from those presented in the experimental papers and then apply the correct theoretical analysis to these. This has not been done so far and even if the statistics would be sufficient for this purpose, the outcome is questionable because of the difficulties implied by the numerics. (That is why it would be desirable that experimentalists and theorists perform a joint analysis of the data or at least that the data should be presented in ``raw" form.) The nearest approximation to the solution of the ``inverse" problem found in the literature, is that of \cite{Schlei97} based on the application of the HYLANDER code by the Marburg group: It consists in fitting the results of the hydrodynamical calculation to the Gaussian form used by experimentalists, \begin{equation} C_2( {\bf k}_1 , {\bf k}_2 ) = 1 +\lambda \exp [ -\textstyle{\frac{1}{2}} q_\parallel^2 R_\parallel^2 -\textstyle{\frac{1}{2}} q_{out}^2 R_{out}^2 -\textstyle{\frac{1}{2}} q_{side}^2 R_{side}^2 ] \label{eq:015a} \end{equation} and comparing with the inverse width of the correlation function as presented in the experimental papers \footnote{In \cite{chapman} it was recommended that experimentalists should use the more complete formula \begin{equation} C_2( {\bf k}_1 , {\bf k}_2 ) = 1 +\lambda \exp [ -\textstyle{\frac{1}{2}} q_\parallel^2 R_\parallel^2 -\textstyle{\frac{1}{2}} q_{out}^2 R_{out}^2 -\textstyle{\frac{1}{2}} q_{side}^2 R_{side}^2-2q_{out}q_{\parallel}R^2_{{out\parallel}}] \label{eq:015bis} \end{equation} where $2q_{out}q_{\parallel}R^2_{{out\parallel}}$ is called ``cross" term. For a more detailed discussion of its meaning and dependence on the coordinate system cf. ref. \cite{SSX}. Of course, in view of the defficiencies of the entire phenomenological procedure outlined above and discussed in greater length in section 5.1.5, these details are of limited importance. In particular they do not affect the conclusions discussed here.}. Here $q^\mu \equiv p_1{}^\mu - p_2{}^\mu $, $K^\mu \equiv \frac{1}{2} (p_1{}^\mu + p_2{}^\mu)$ and $q_\parallel$ and $K_\parallel$ denote the components of ${\bf q}$ and ${\bf K}$ in beam direction, and $q_\perp$ and $K_\perp$ the components transverse to that direction; $q_{out}$ is the projection of the transverse momentum difference, ${\bf q}_{\perp}$ on the transverse momen\-tum of the pair, $2 {\bf K}_\perp$, and $q_{side}$ the component perpendicular to ${\bf K}_\perp$. (For a source with cylindrical symmetry, the two-particle correlation function can be expressed in terms of the 5 quantities $K_\parallel$, $K_\perp$, $q_\parallel$, $q_{side}$ and $q_{out}$.) $R_\parallel$, $R_{side}$, $R_{out}$ are effective parameters, associated via eq.(\ref{eq:015a}) to the corresponding $q$ components. Eq.(\ref{eq:015a}) is equivalent to an expansion of the correlation function $C({\bf q},{\bf K})$ for small $q$. The use of eq.(\ref{eq:015a}) for the representation of correlations data implies then that one does not measure the geometrical radius of the system but the length of homogeneity, which means that energy density determinations based on BEC are an overestimate \footnote{In \cite{SinHot}, \cite{AkkSi} a distinction is made between the ``local" length of homogeneity $L_{h}(x,k)$ and the ``hydrodynamical" length $\L_{h}(x)$ which is the ensemble average of the former.}. To take into account the fact that the correlation function depends in general not only on the momentum difference $q=k_1-k_2$ but also on the sum $K=\frac{1}{2}(k_1+k_2)$ , the parameters $R$ and $\lambda$ are assumed to be functions of $K$ and rapidity $\frac{1}{2}(y_1+y_2)$. Hadron BEC refer to the freeze-out stage. This stage is usually described by the Cooper-Frye formula \cite{cfrye}: \begin{equation} E\frac{dN}{d{\bf k}}=\frac{g_i}{(2\pi)^3} \int_\sigma \frac{p_{\mu} d\sigma^{\mu}}{\exp(\frac{p_{\mu} u^{\mu} - \mu_s-\mu_b} {T_f}) - 1}, \label{eq:single} \end{equation} \\ which describes the distribution of particles with degeneracy factor $g_i$ and 4-momentum $p^{\mu}$ emitted from a hypersurface element $d\sigma^\mu$ with 4-velocity $u^{\mu}$ \footnote {In most applications particles produced with momenta $p_\mu$ pointing into the interior of the emitting isotherm $(p_\mu d\sigma^\mu < 0)$ were assumed to be absorbed and therefore their contribution to the total particle number was neglected. In ref \cite{latp} this effect was indeed estimated to be negligible and recent attempts to reconsider it could not change this conclusion. Another effect is the interaction of the freeze-out system with the rest of the fluid. This effect can be estimated by comparing the evolution of the fluid with and without the frozen-out part. This is done by equating the frozen-out part with that corresponding in the equation of hydrodynamics to the case $p_\mu=0$. The fluid parameters are modified by this procedure at a level not exceeding $10\%$ \cite{ornik}. The influence of the freeze-out mechanism on the determination of radii via BEC has been discussed recently in several papers; cf.e.g. \cite{HV} and references quoted there.}. After the cascading of the resonances we obtain the final observable spectra. \subsubsection{Transverse and longitudinal expansion} The equations of hydrodynamics are non-linear and therefore good for surprises. An illustration of this situation is represented by the realization, to be described in more detail below, that the naive intuition about the role of transverse expansion in the determination of the transverse and longitudinal radius may be completely misleading. Only a systematic analysis based on 3+1 dimensional hydrodynamics could clarify this issue. In the present section we will discuss Bose-Einstein correlations of pions and kaons produced in nuclear collisions at SPS energies in the framework of relativistic hydrodynamics. Concrete applications were done for the symmetric reactions $S+S$ and $Pb+Pb$ at 200 AGeV. Many of the theoretical results were predictions at the time they were performed. These predictions were subsequently confirmed in experiment. In \cite{Hum} it was found that the transverse radius extracted from data on Bose-Einstein correlations (BEC) for O$+$Au at $200$ AGev reached in the central rapidity region a value of about $8 fm$. It was then natural to conjecture that this could be an indication of transverse flow\cite{Hei},\cite{Lee},\cite{Alb}. In the meantime the experimental observation in itself has been qualified \cite{Fer} and it now appears that the transverse radius obtained from the BEC data does not exceed a value of $4$-$5 fm$ (see however ref.\cite{Alb}). Motivated by this situation in ref. \cite{Schlei97} an investigation \footnote{In ref.\cite{KagMin} the dependence of the effective transverse radius on the transverse velocity field was investigated for a fixed freeze-out hypersurface. The important effects of transverse expansion on the shape and position of the hypersurface are not considered there.} of the role of $3$-dimensional hydrodynamical expansion on the space-time extension of the source was performed and compared with a 1+1 d calculation. Contrary to what one might have expected it was found that {\em transverse flow does not increase the transverse radius}. On the other hand, a strong dependence of the longitudinal radius on the transverse expansion was estblished. Fig.10 shows two typical examples of the Bose-Einstein correlations as functions of $q_\parallel$ and $q_\perp$ for the $1$- and the $3$-dimensional hydrodynamical solution. The dependence of $C_2 (q_\parallel)$ on transverse expansion agrees qualitatively with what one would expect. For a purely longi\-tu\-dinal expansion, the effective longi\-tu\-di\-nal radius of the source is larger than in the case of 3-dimensional expansion, which is reflected in a decrease of the width of the correlation function (see also in Fig.11 below). On the other hand, the results for $C_2(q_\perp)$ were at a first glance rather surprising. Naively one might have expected that the transverse flow would lead to an increase of the transverse radius, i.e., to a narrower correlation function $C_2(q_\perp)$. However, in Fig.10 the curves that describe the 1-dimensional and the 3-dimensional results are almost identical. If anything, one would conclude that the effective transverse radius is {\it smaller} in the presence of transverse expansion. This effect can however be explained if one takes a closer look at the details of the hydrodynamic expansion process as investigated in ref.\cite{Schlei97}. Due to the strong correlation between the space-time point where a particle is emitted, and its energy-momentum, the effective radii obtained from Bose-Einstein correlation data present a characteristic dependence on the average momentum of the pair, $K^\mu$. Fig.11 shows the dependence of the effective radii $R_\parallel$, $R_{side}$ and $R_{out}$ on the rapidity $y_K$ and the mean transverse momentum of the pair $K_\perp$, both for the 1-dimensional and for the 3-dimensional calculation. The longitudinal radius $R_\parallel$ becomes considerably smaller (by a factor of 2-3) if transverse expansion is taken into account. For the 1-dimensional case, an approximate analytic expression has been derived for the $y_K$- and the $K_\perp$-dependence of the longitudinal `radii`" in Refs.\cite{LorSin}) \footnote{In this reference eq.(\ref{eq:016}) is used to disprove the applicability of hydrodynamics to $p-p$ reactions in the ISR energy range ($\sqrt{s}=53 GeV$). Such a conclusion seems dangerous given the approximations involved both in the derivation of this formula as well as in the interpretation of the BEC measurements in the above reactions.},\cite{Sin} (cf. also \cite{PaGyGa}): \begin{equation} R_\parallel = \sqrt{\frac{2T_f}{m_\perp}} \frac{\tau_o} {\cosh(y_K)} \label{eq:016} \end{equation} where $m_\perp = (m_{\perp 1}+m_{\perp 2})/2$ is the average transverse mass of the two particles, $T_f$ is the freeze-out temperature and $\tau_o =(\partial u_\parallel / \partial x)^{-1}$ is the inverse gradient of the longitudinal component of the $4$-velocity in the center (at $x=0$) \footnote{The expression (\ref{eq:016}) for $R_\parallel$ denotes in fact the length of homogeneity $L_h$ mentioned above. It refers to the region within which the variation of Wigner function is small. By definition $L_h \leq R$ where $R$ is the geometrical radius.}. In \cite{Schlei97} one finds that this approximate expression describes $R_\parallel (K_\perp,y_K)$ for $S+S$ reactions quite well, both for the 1-dimensional and the 3-dimensional case (see Fig.11). However for $Pb+Pb$ reactions the same formula fails to account for the data\cite{Pbpl}. This is not surprising, because eq.(\ref{eq:016}) is based on the assumption of boost invariance, i.e. no stopping. This assumption is not justified at SPS energies where there is considerable stopping. The inelasticity increases with atomic number and this may explain the breakdown of the above formula. This exemplifies the limitations of the boost-invariance assumption, an assumption which must not be taken for granted but in special circumstances. In \cite{jan} it was proposed to use the information obtained from fitting the single inclusive distribution to constrain the parameters that enter into the hydrodynamic description, and then to calculate the transverse radius directly. Indeed let $\Sigma $ denote the hypersurface in Minkowski space on which hadrons are produced. Then one can define e.g. a transverse radius \begin{equation} R_{\perp}=\frac{\int_{\Sigma}R\frac{p^{\mu}d\sigma_{\mu}} {\exp[(p^{\mu}u_{\mu}-\mu)/T]-1}} {\int_{\Sigma }\frac{p^{\mu}d\sigma_{\mu}} {\exp[(p^{\mu}u_{\mu}-\mu)/T]-1}} \label{eq:BOW10} \end{equation} where $u_{\mu}, T$ and $\mu$ denote the four-velocity, temperature, and chemical potential on the hypersurface $\Sigma$, respectively as in eq.(\ref{eq:single}). It is interesting to note that this method for the determination of transverse radii based on the single inclusive cross sections provides a geometrical radius while the use of the second order correlation function provides a coherence length (length of homogeneity). \begin{footnotesize} Comparing the effective transverse radius $R_\perp$ extracted from the Bose-Einstein correlation function to the mean transverse radius as calculated directly in \cite{jan} according to eq. (\ref{eq:single}) , one finds in \cite{Schlei97} that the two results agree to an accuracy of about 10$\%$. This conclusion is confirmed and strengthened in a more recent study by Schlei \cite{Schlei97} for kaon correlations.\end{footnotesize} Of course, this approach can be used only if a solution of the equations of hydrodynamics is available; with pseudo-hydodynamical methods this is not possible. \subsubsection{Role of resonances and coherence in the hydrodynamical approach to BEC} This problem was investigated using an exact 3+1 d numerical solution of hydrodynamics in \cite{bernd3}. The source distribution $g(x,k)$ was determined from a 3-dimensional solution of the relativistic hydrodynamic equations. Fig. 12 illustrates the effect of successively adding the contributions from $\rho$, $\omega$, $\Delta$ and $\eta$ decays to the the BEC correlation functions of directly produced (thermal) $\pi^-$ (dotted lines), in longitudinal and in transverse direction. The width of the correlation progressively decreases as the decays of resonances with longer lifetimes are taken into account, and the correlation looses its Gaussian shape. The long-lived $\eta$ leads to a decrease of the intercept. \paragraph{Pion versus kaon interferometry} Ideally, a comparison of pion and kaon interferometry should lead to conclusions concerning possible differences in the space-time regions where these particle decouple from the hot and dense matter. It was proposed that kaons may decouple (freeze out) at earlier times and higher temperatures than pions \cite{heinz}. Indeed, preliminary results had indicated that the effective longitudinal and transverse source radii extracted from $\pi\pi$ correlations were significantly larger than those obtained from $KK$ correlations \cite{NA44}. However, as we have seen in fig. 12, the BEC of pions are strongly distorted by the contributions from resonance decay. It was pointed out in ref. \cite{gpkaon} in a study based on the Lund string model that such distortions are not present for the BEC of kaons, and that consequently for the effective transverse radii one expects $R_\perp(K^\pm) < R_\perp(\pi^\pm)$, even in the absence of any difference in the freeze-out geometry of {\it directly produced} pions and kaons. These conclusions were confirmed in \cite{bernd3} within the hydrodynamical approach and one found furthermore that this effect is even more pronounced if one considers longitudinal rather than transverse radii. Furthermore the interplay between coherence and resonance production which was not considered in \cite{gpkaon} was studied in \cite{bernd3}. There are also some striking differences between \cite{gpkaon} and \cite{bernd3} in the resonance production cross sections used. In fig. 13, BEC functions of $\pi^-$ (solid lines) and of $K^-$ (dashed lines) are compared, at $k_\perp=0$ and $k_\perp=1$ GeV/c, respectively. The dotted lines correspond to the BEC function of thermally produced $\pi^-$. It can be seen that the distortion due to the decay contributions from long-lived resonances disappear only at large $k_\perp$. Fig. 14 shows the effective radii $R_{||}$, $R_{side}$ and $R_{out}$ as functions of rapidity and transverse momentum of the pair, both for $\pi^-\pi^-$ (solid lines) and for $K^-K^-$ pairs (dashed lines). For comparison, the curves for thermally produced pions (dotted lines) are also included. The effective longitudinal radii extracted from $\pi^-\pi^-$ correlations are considerably larger than those obtained from $K^-K^-$ correlations. In the central region the two values for $R_{||}$ differ by a factor of $\sim 2$. For the transverse radii, the factor is $\sim 1.3$. A comparison between results for $K^-$ and thermal $\pi^-$ shows that part of this effect can be accounted for by kinematics (the pion mass being smaller than the kaon mass; see also eq. (\ref{eq:016})). Nevertheless, the large difference between the widths of pion and kaon correlation functions is mainly due to the fact that pion correlations are strongly affected by resonance decays, which is not the case for the kaon correlations. In the hydrodynamic scenario of ref.\cite{Bolz}, about $50\%$ of the pions in the central rapidity region are the decay products of resonances \cite{jan}, while less than $10\%$ of the kaons are created in resonance decays ($K^\star \rightarrow K\pi$ dominates, contributing with about $5\%$). In ref.\cite{bernd3} the problem of coherence within the hydrodynamical approach to BEC was also investigated. Fig. 15 shows the $\pi^-\pi^-$ correlation functions in the presence of partial coherence. In order to extract effective radii from Bose-Einstein correlation functions in the presence of partial coherence, eq.(\ref{eq:015a}) must be replaced by the more general form \begin{eqnarray} C_2({\bf k}_1,{\bf k}_2)\: = &1& +\:\: \lambda \cdot \:p_{eff}^2\: \exp \left[-\: \frac{1}{2}\: \sum (qR)^2\: \right] \nonumber\\ &&+ \:\: \sqrt{\lambda} \cdot \: 2\:p_{eff}\:(1\:-\:p_{eff})\: \exp \left[-\: \frac{1}{4}\: \sum (qR)^2\: \right] \label{eq:024} \end{eqnarray} \subsubsection{Comparison with experimental data} Some of the predictions made in \cite{Bolz} for $S+S$ reactions could be checked experimentally in refs.\cite{Alber}\cite{Ferenc}, in particular the rapidity and transverse momentum dependence of radii and remarkable agreement was found. In \cite{OrnikPR96} the hydrodynamical calculations were extended to $Pb+Pb$ reactions and compared with $S+S$ reactions and, where data were available, with experiment. The calculation of Bose-Einstein correlations (BEC) was performed using the formalism outlined in refs. \cite{bernd2,bernd3,Schlei97} including the decay of resonances. The hadron source was assumed to be fully chaotic. Figs.16 and 17 show the calculations for the effective radii $R_\parallel$, $R_{side}$ and $R_{out}$ as functions of rapidity $y_K$ and transverse momentum $K_\perp$ of the pion pair compared to the corresponding NA35 and preliminary NA49 data \cite{alber,QM95}, respectively. All these calculations, which in the case of $S+S$ had been true predictions, agree surprisingly well with the data\footnote {The EOS used in the hydrodynamical studies quoted above included a phase transition from QGP to hadronic matter. How critical this assumption is for the agreement with data is yet unclear and deserves a more detailed investigation. On the other hand the very use of hydrodynamics is based on local equilibrium, and this equilibrium is favoured by the large member of degrees of freedom due to a QGP.}. This suggests that our understanding of BEC in heavy ion reactions has made progress and confirms the usefulness of the Wigner approach when coupled with full fledged hydrodynamics. An important issue in comparing data with theory is the detector acceptance of a given experiment. This is also discussed in detail in \cite{OrnikPR96}. {\footnotesize Another application of hydrodynamics to the QGP search in heavy ion reactions is due to Rischke and Gyulassy \cite{Gyu96} who investigate the ratio $r=R_{out}/R_{side}$. Based on considerations due to Pratt \cite{Pratt}, this quantitiy had been proposed by Bertsch and collaborators\cite{Bertsch} as a signal of QGP. Under certain circumstances one could expect that for a long lived QGP phase $r$ should exceed unity, while for a hadronic system, due to final state interactions, the out and side sizes should be comparable. The authors of \cite{Gyu96} performed a quantitative hydrodynamical study of $r$ in order to check whether this signal survives a more realistic investigation, albeit they did not take into account resonances. For directly produced pions it is found that $r$ indeed reflects the lifetime of an intermediate (QGP) phase. However we have seen from \cite{Bolz}, \cite{OrnikPR96} that for pion BEC, when resonances are considered, hydrodynamics with an EOS containing a long lived QGP phase, leads (in agreement with experiment) to values of $r$ of order unity. To avoid this complication, in \cite{Risch} it was proposed to consider kaons at large $k_{\bot}$. However even this proposal may have to be qualified, besides the fact that it will be very difficult to do kaon BEC at large $k_{\perp}$. Firstly one has to recall that the entire formalism on which the $r$ signal is based and in particular the parametrization (\ref{eq:015a}) is questionable. Secondly it remains to be proven that this signal survives if one imposes simultanoeusly the essential constraint due to the single inclusive distribution\footnote{I am indebted to B. R. Schlei for this remark.}. Furthermore it is unclear up to what values of $k_{\perp}$ the Wigner formalism, (which is a particular case of the classical current formalism) on which the theory is based, is applicable. For these reasons the determination of the lifetime of the system via ``pure" hydrodynamical considerations is certainly an alternative which deserves to be considered seriously, despite its own difficulties.} \subsubsection{Bose-Einstein correlations and pseudohydrodynamics} As mentioned already, the initial motivation for proposing the Wigner function formalism for BEC was to explain why the experimentally observed second order correlation functions $C_2$ were depending not only on the momentum difference $q=k_1-k_2$ but also on their sum $K=k_1+k_2$. However in the mean time it was shown \cite{aw},\cite{APW} that this feature follows from the proper application of the space-time approach in the current formalism even without assuming expansion. Furthermore the Wigner formalism is useful only at small $q$ and cannot be applied in the case of strong correlations between positions and momenta while the current formalism is not limited by these constraints. As explained above the use of the Wigner formalism can be defended if combined with bona-fide hydrodynamics and an equation of state. This notwithstanding, besides a few real, albeit numerical, hydrodynamical calculations, most phenomenological papers on BEC in heavy ion reactions (cf. e.g. \cite{MS},\cite{Bert1},\cite{Pratt3}, \\ \cite{Bertsch94}, \cite{Chao94}-\cite{Bert2},\cite{Akk},\cite{csorlor}, \cite{Csorg4}-\cite{Wu98} have used the Wigner formalism without a proper hydrodynamical treatment, i.e. without solving the equations of hydrodynamics; hydrodynamical concepts like velocity and temperature were used just to parametrize the Wigner source function. While such a procedure may be acceptable as a theoretical exercise, it is certainly no substitute for a professional analysis of heavy ion reactions. This is a fortiori true when real data have to be interpreted\footnote{A recent experimental paper \cite{2Na49} where such a procedure is used is a good illustration of the limits of pseudo-hydrodynamical models. Despite the fact that the statistics are so rich that ``the statistical errors on the correlation functions are negligible", the outcome of the analysis is merely the resolution of the ambiguity between temperature and transverse expansion velocity of the source. It is clear that such an ambiguity is specific to pseudo-hydrodynamics and is from the beginning absent in a correct hydrdynamical traetment. Moreover even this result is questionable given some doubtful assumptions which underly this analysis. To quote just two: (i) The assumption of boost invariance made in \cite{2Na49} decouples the longitudinal expansion from the transverse one. This not only affects the conclusions drawn in this analysis but prevents the (simultaneous) interpretation of the experimental rapidity distribution. (ii) The neglect of long lived resonances which strongly influences the $\lambda$ factor and thus also the extracted radii. Of course, despite the claimed richness of the data, no attempt to relate the observations to an equation of state can be made within this approach.}. As exemplified in previous sections such a procedure is unsatisfactory, among other things because it can lead to wrong results. The use of this ``pseudo-hydrodynamical" approach is even more surprising if one realizes the fact that the Wigner formalism not only is not simpler than the more general current \footnote{We remind that the classical current formalism, in opposition to the Wigner function formalism, does not apply only to small $q$ and to semiclassical situations. Furthermore it allows also for coherence and new phenomena like particle-antiparticle correlations.} formalism but it is also less economical. The number of independent parameters necessary to characterize the BEC within the Wigner formalism of ref.\cite{Nix} is \footnote{For each value of ${\bf K}$ a Gaussian ellipsoid is described by three spatial extensions, one temporal extension, three components of the velocity in the local rest frame and the three Euler angles of orientation.} $10$, i.e. it is as large as that in the current formalism. However the $10$ parameters of \cite{Nix} describe a very particular source \footnote{To consider such an approach as ``model independent" \cite{Heavy}, \cite{HeinzNP} is misleading.}, as compared with that of the current formalism: besides the fact that the correlation function source is assumed to be Gaussian, it is completely chaotic and it can provide only the length of homogeneity $L_h$ \cite{MS}. For the search of quark gluon plasma, however, the geometrical radius $R$ is relevant, because the energy density is defined in terms of $R$ and the use of $L_h$ instead of $R$ leads to an overestimate of the energy density \footnote{One may argue that the ``length of homogeneity" \cite{MS} $L_h$ defined in terms of the Wigner function is a particular case of the correlation length $L$ defined in terms of the correlator (cf. section 4.3). While $L_h$ is always limited from above by $R$, $L$ can be either smaller or larger than $R$.}. Furthermore the physical significance of the parameters of the Wigner source is unclear if the Gaussian assumption does not hold\footnote{Even if a Gaussian form would hold for directly produced pions, resonances would spoil it \cite{Bolz}, \cite{OrnikPR96}, \cite{Schlei97}.}. Not only is there no a priori reason for a Gaussian form, but on the contrary, both in nuclear and particle physics as well as in quantum optics, there exists experimental evidence that in many cases an exponential function in $|q|$ is at small $q$ a better approximation for $C_2$ than a Gaussian. Furthermore, in the presence of coherence, no single simple analytical function, and in particular no single Gaussian is expected to describe $C_2$. This is a straightforward consequence of quantum statistics. Given the fact that good experimental BEC data are expensive both in terms of accelerator running time and man-power, the use of inappropiate theoretical tools, when more appropiate ones are available, is a waste which has to be avoided. For the reasons quoted above we will not discuss in more detail the numerous and sometimes unnecessarily long papers which use pseudo-hydrodynamical methods. \paragraph{Concluding remarks.} In a consistent treatment of single and double inclusive cross sections for identical pions via a realistic hydrodynamical model, resonances play a major role leading to an increase of effective radii of sources. Effective longitudinal radii are more sensitive to the presence of resonances than transverse ones. >From the hydrodynamical treatment we learn that the hadron source (the real fireball) is represented by a very complex freeze-out hypersurface (cf. ref. \cite{OrnikPR96}). The longitudinal and transverse extensions of the fireball change dynamically as a function of time, rather than show up in static effective radii. Thus, the interpretation of BEC measurements is also complicated. This is a theoretical task and if only for this reason experimentalists would be well advised to include in their teams also qualified theorists. For heuristic applications, when no quantitative comparison with data is intended, besides the current formalism, analytical approximations of the equations of hydrodynamics can be useful, because they allow a better qualitative understanding of hydrodynamical expansion. However when a quantitative interpretaion of experiments is intended and in particular a connection with the equation of state is looked for, the only recommendable method is full-fledged hydrodynamics. \subsubsection{Photon correlations, pseudo-hydrodynamics and pseudo-Wigner formalism} Photon correlations have been investigated within the context of quark-gluon plasma search, since they present certain methodological advantages as compared with hadrons. While experimentally genuine photon BEC in high energy heavy ion reactions have not yet been unambigously identified, because of the strong $\pi{^0}$ background and the small cross sections for photon production, there are several theoretical studies devoted to this topic. The advantage of photon BEC resides in the fact that, while correlations between hadrons are influenced by final state interactions, photon correlations are ``clean" from this point of view. For high energy physics photons present another important advantage due to the fact that they can provide direct information about the early stages of the interaction where quarks and gluons dominate and hadrons have not yet been created. In particular photon BEC contain information about the lifetime of the quark-gluon plasma \cite{kap}-\cite{Sriv}, \cite{Sriv5}, \cite{Axel}. Among other things it was argued e.g. in \cite{Sriv5} and confirmed in \cite{Axel} using a more correct formalism, (cf. below), that the correlation function $C_2$ in the transverse direction can serve as a signal for QGP as it is sensitive to the existence of a mixed phase. Unfortunately, some of these studies (refs.\cite{kap}-\cite{Sriv}, \cite{Sriv5}) besides suffering from the more general disease of pseudo-hydrodynamics, use an input formula for the second order BEC, which is essentially incorrect even within the Wigner-formalism (cf. section 4.6). Besides this, some approximations made in \cite{kap} e.g. are inadequate. This question was analyzed in more detail in \cite{Axel} where it was found that only some of the results of ref. (\cite{kap}) survive a more critical analysis. The formula for the two-particle inclusive probability used in (\cite{kap}-\cite{Sriv},\cite{Sriv5}) reads \begin{equation} P_2({\bf k}_1,{\bf k}_2) \ =\ \int d^4x_1 \int d^4x_2\ g\left(x_1,k_1\right)\ g\left(x_2,k_2\right)\ [1\ +\ \cos((k_1-k_2)(x_1-x_2))], \label{eq:4} \end{equation} while ref.(\cite{Axel} uses the more correct formula \begin{equation} P_2({\bf k}_1,{\bf k}_2) \ =\ \int d^4x_1 \int d^4x_2\ g\left(x_1,\frac{k_1+k_2}{2}\right)\ g\left(x_2,\frac{k_1+k_2}{2}\right)\ [1\ +\ \cos((k_1-k_2) (x_1-x_2))], \label{eq:3a} \end{equation} (From eq.(\ref{eq:3a}) one gets for the second order correlation function eq.(\ref{eq:Heinz23})). It is found in \cite{Axel} that two rather surprising properties of the two-photon correlation function presented in \cite{kap} are artifacts of inappropriate approximations in the evaluation of space-time integrals. In \cite{kap}, it was claimed that the BEC function in the longitudinal direction (a) oscillates and (b) takes values below unity. As property (b) is inconsistent with general statistical bounds, it was important to clarify the origin of this discrepancy. On the other hand, it was confirmed in \cite{Axel} that the correlation function in the transverse direction does exibit oscillatory behaviour in the out component of the momentum difference. Furthermore, in this reference a change of the BEC function in $\Delta y$ from Gaussian to a two-component shape with decreasing transverse photon momentum was found which may serve as evidence for the presence of a mixed phase and, hence, as a QGP signature. However even after correcting the wrong input BEC formula it is questionanble whether the other approximations made in refs.(\cite{kap}-\cite{Sriv},\cite{Sriv5},\cite{Axel}) may not invalidate the above result. Besides the use of a simplified hydrodynamical solution one has to recall that (i) the Wigner formalism like the more general classical current formalism, is limited to small momenta $k$ of produced particles (no recoil approximation) (ii) Besides this general limitation to small $k$ the Wigner formalism is specifically limited to small differences of momenta $q$. (iii) Although Eq. (\ref{eq:Heinz23}) does not suffer from the violation of unitarity disease mentioned in section 4.6 of chapter 3, it is based on an approximation which makes it sometimes inapplicable for photons (cf. section 4.5). >From the above discussion one may conclude that the experimental problems of photon BEC are matched by yet unsolved theoretical problems. \subsection{Pion condensates} One of the most interesting phenomena related to Bose-Einstein correlations is the effect of Bose condensates. The remarkable thing about this effect is that it is not specific for particle or nuclear physics, but occurs in various other chapters of physics, in particular in condensed matter physics like superconductivity and superfluidity. Moreover, recently not only the condensation of a gas of atoms has been experimentally achieved \cite{Atomcond95} but the quantum statistical coherence of these systems has been experimentally proven through Bose-Einstein correlations \cite{Atomcond97}. The proposal to use BEC for the detection of condensates was made a long time ago \cite{Stelte}. The more recent developments in heavy ion reactions have made this subject of current interest. Already in experiments at the SPS (e.g., $Pb+Pb$ at $E_{beam}=160$ AGeV) secondary particles are formed at high number densities in rapidity space \cite{pbpb} and in future experiments at RHIC and the LHC one expects to obtain even higher multiplicities on the order of a few thousand particles per unit rapidity. If local thermal (but not chemical) equilibrium is established and the number densities are sufficiently large, the pions may accumulate in their ground state and a Bose condensate may be formed\footnote{Recently a different type of pion condensate, the disordered chiral condensate, has received much of attention in the literature. It has been argued \cite{kogan} that such a condensate would lead to the creation of squeezed states.}. A specific scenario for the formation of a Bose condensate, namely, the decay of short-lived resonances, was discussed in Ref. \cite{bo-co} where conditions necessary for the formation of a Bose condensate in a heavy ion collision were investigated. In Ref. \cite{bo-co} it was found that if a pionic Bose condensate is formed at any stage of the collision, it can be expected to survive until pions decouple from the dense matter, and thus it can affect the spectra and correlations of final state pions. In ref.\cite{cond97} one investigated the influence of such a condensate on the single inclusive cross section and on the second order correlation function of identically charged pions (Bose-Einstein correlations BEC) in hadronic reactions for {\em expanding} sources. A hydrodynamical approach was used based on the HYLANDER routine. The Bose-condensate affects the single inclusive momentum distributions $EdN/d^3k$, the momentum-dependent chaoticities $p$ and the Bose-Einstein correlation functions $C_2$ only over a limited momentum range. This is due to the fact that in a condensate there exists a maximum velocity (which implies also a maximum momentum difference $q_{max}$) and leads to a very characteristic structure in single and double inclusive spectra. In Fig. 18 the results of the numerical evaluations of the Bose-Einstein correlation functions $C_2$ are shown for a spherically and for a longitudinally expanding source. The presence of a Bose-condensate of only $1\%$ results in a decrease of the intercept by about 15\%. Furthermore due to a limited value of $q_{max}$ a part of the tail of the two-particle correlation functions is not affected by the pionic Bose-condensate and a peak appears. To what extent such peaks can be observed in experimental data depends among other things on the size of the source, the details of the freeze-out, width of the momentum distribution in the bosonic ground state, and detector acceptance. \paragraph{Plasma droplets?} If the phase transition from hadronic matter to QGP and \\viceversa is of first order then one could expect the formation of a mixed phase, in which QGP and hadronic matter coexist. Such a mixed phase manifests itself in the hydrodynamical evolution of the system \cite{mixed} and it influences among other things the transverse momentum distribution of photons as we have seen in section 5.1.6. It was suggested by Seibert \cite{Seibert} that the mixed phase could also lead to a granular structure which might be seen in the fluctuations of the velocity distributions of secondaries produced in the hadronization stage. Pratt, Siemens and Vischer \cite{droplet1} (cf. also \cite{droplet2} proposed subsequently that in Bose-Einstein correlations, too, one might see a signature of this granularity. \setcounter{equation}{0} \section{Correlations and Multiplicity Distributions} \subsection{From correlations to multiplicity distributions} An important physical observable in multiparticle production is the multiplicity distribution\footnote{The multiplicity distribution will be denoted sometimes in the following also by MD} $P(n)$, i.e., the probability to produce in a given event $n$ particles. The link between the multiplicity distribution $P(n)$ and BEC is represented by the density matrix $\rho$ , since it is the same $\rho$ which appears in the definition of $P(n)$ \begin{equation} P(n) \equiv <n|\rho|n> \label{eq:rho} \end{equation} and the definition of correlation functions (cf.e.g. eq.(\ref{eq:cln})). Usually one expresses $\rho$ in terms of the $\cal P(\alpha)$ representation (cf. section 2.2 ) and by using for $\cal P(\alpha)$ simple analytical expressions one is able to derive the most characterisitc forms of $P(n)$ in an analytical form, like the Poisson or the negative binomial representation (cf. e.g. \cite{book}). Sometimes there exist however physically interesting cases where no analytical expression for the multiplicity distribution $P(n)$ exists, but instead the moments of $P$ are given. >From the phenomenological point of view the approach to MD via moments presents sometimes important advantages because it allows the construction of an {\em effective} density matrix from the knwoledge of a few physical quantities like the correlation lengths and mean multiplicity, which in turn can be obtained from experiment (cf.e.g. \cite{Botke}). In the following we will address this aspect of the problem, the more so that in this way the link between correlations and MD becomes clearer. We start by recalling some definitions. Besides the normal moments of the MD given by \begin{equation} <n^q> \ \equiv \ \sum_n \ P(n) n^q \label{eq:normal} \end{equation} one uses frequently the factorial moments \begin{equation} \Phi_q \ \equiv \ \sum_n \ P(n)\ \left(n(n-1)\cdot...\cdot(n-q+1)\right) \label{eq:deffq} \end{equation} These can be expressed in terms of the inclusive correlation functions $\rho_q$ through the relation \begin{equation} \Phi_q \ =\ \int_\Omega d\omega_1 ...\ \int_\Omega d\omega_q \ \ \rho_q({\bf k}_1 , ... , {\bf k}_q) \ = \ \langle \frac{n!}{(n-q)!} \rangle \label{eq:facmom} \end{equation} Eqs. (\ref{eq:deffq}) and (\ref{eq:facmom}) illustrate the fundamental fact that the inclusive cross sections $\rho_q$ and thus the correlation functions determine the moments of the MD, which are nothing else but the integrals of $\rho_q$. Although this relation between moments of MD and correlation functions is a straightforward aspect of multiparticle dynamics, the connection between MD and correlations was often overlooked. This is in part due to the fact that measurements of correlations are, for reasons of statistics and other technical considerations, frequently performed in different (narrower) regions of phase space than measurements of MD \footnote{This is e.g the case when MD are measured with no proper identification of the particles, while BEC refer of course to identical particles. Thus the UA5 experiment \cite{UA5}, which discovered the violation of KNO scaling in MD-s, measured only {\em charged} particles, without distinguishing between positive and negative charges. However the rapidity region accessible in this experiment was much broader than the corresponding region in the UA-1 experiment \cite{bu} where correlation measurements were performed and where a distinction between positive and negative charges could be made.}. However the importance of the use of the relationship between MD and correlations can hardly be overemphasized, just because of the different experimental methods used in the investigation of these two observables. In the absence of a theory of multipaticle production, the form of the correlators and the amount of chaoticity are unknown and have therefore to be parametrized and then determined experimentally in correlation experiments. Both the parametrization and the measurements are affected by errors. Similar considerations apply for MD, but because of the different experimental conditions under which correlation measurements and MD measurements are made, the corresponding errors are different. (Cf. also the discussion of the importance of higher order correlation, section 2.2.1). Therefore from a phenomenological and practical point of view, MD and correlations are rather complimentary and have to be interpreted together. In the following we will exemplify the usefulness of this point of view. \subsubsection{Rapidity dependence of MD in the stationary case} The dependence of moments of multiplicity distributions $P(n)$ on the width of the bins in momentum or rapidity space has been in the center of multiparticle production studies for the last 15 years. It got much attention after: 1) the experimental observation \cite{UA5} by the UA5 collaboration that the normalized moments of $P(n)$ in the rapidity plateau region increase with the width of the rapidity window $\Delta y$; 2) the proposal by Bialas and Peschanski \cite{bipesh} that this behaviour, which at a first look was power-like may reflect ``intermittency", i.e. the absence of a fixed scale in the problem, which could imply that self-similar phenomena play a role in multiparticle production. However soon after this proposal was published it was pointed out in \cite{car} that the quantum statistical approach, presented in section 2.2 and which implies a {\em fixed} scale \footnote{This scale is the correlation length $\xi$ of eqs. (\ref{eq:Lorentz}) or (\ref{eq:Gaussdef}).}, predicts a similar functional relationship between the moments of MD and $\Delta y$. In Fig. 19 from \cite{car} some examples of this behaviour are plotted and compared with experimental data. For small $-\delta y$ the (semilogarithmic) plot can be approximated by a power function as indicated by the data. Recalling that the QS formalism applies to identical particles, it follows that BEC could be at the origin of the so called intermittency effect. This point of view was corraborated subsequently also in \cite{gyul} and \cite{cap} and was confirmed experimentally by the observation that the ``intermittency" effect is strongly enhanced when identical particles are considered \footnote{The UA5 data refer to a mixture of equal numbers of positive and negative particles; this dilutes the BEC effect.} and /or when studied in more than one dimension \footnote{This last observation also supports the idea that BEC is the determining factor in intermittency because the integration over transverse momentum implied by a a one dimensional $y$ investigation diminishes the BEC.}. For a more recent very clear confirmation of this point of view cf. the studies by Tannenbaum \cite{tan}. Further developments related to ``intermittency" will be discussed in section 6.3. \subsubsection{Rapidity dependence of MD in the non-stationary case} The assumption of translational invariance in rapidity permitted to apply the quantum optical formalism, in which time has the analogous property, to MD and led to a simple interpretation of the observed broadening of the MD with the decrease of the width of the rapidity window in high energy reactions. However stationarity in rapidity is expected to hold only in the central region (and only at high energies). Indeed experimental data on proton(antiproton)-proton collisions in the energy range $52 < \sqrt{s} < 540 GeV$ show that if one considers shifted rapidity bins along the rapidity axis the MD in these bins depend on the position of the bin: in the central region the MD is broader and can be decribed by a negative binomial distribution while in the fragmentation region it is narrower and can be described by a Poisson MD \cite{2comp}. The mean multiplicity in the central region increases faster (approximately like $s^{1/4}$) with the energy than that in the fragmentation region. This was interpreted in \cite{2comp} as possible evidence for the existence of two sources, one of chaotic nature localized in the central region and another coherent in the fragmentation region. This interpretation is in line with the folklore that gluons which interact stronger (than quarks) form a central blob which may be equilibrated, while the fragmentation region is populated by througoing quarks associated with the leading particles. \footnote{For a microscopic interpretation of this effect in terms of a partonic stochastic model, cf. ref.\cite{stoch}.}. A few years later the NA35-collaboration \cite{Na35} measured BEC in $^{16}O-Au $ reactions at 200 GeV/nucleon in a relatively broad y region and found evidence for a larger and more chaotic source in the central rapidity region, and a smaller and more coherent source in the fragmentation region. It was then natural to correlate (cf. ref.\cite{revisit}) the two observations, i.e. that refering to MD and that refering to BEC. Taken together the credibility of the conjecture made in ref.\cite{2comp} is strongly enhanced, just because we face here different physical observables and different experiments, each with its own specific corrections and biases. Moreover it was pointed out in \cite{revisit} that experiment \cite{Na35} did not necessarily imply that the two sources were independent, but could also be interpeted as due to a single partially coherent source. Indeed consider in a simplified approach as used e.g. in quantum optics a superposition $\pi$ of two fields one coherent denoted by $\pi_c$ and another chaotic denoted by $\pi_{ch}$ so that $\pi = \pi_{c}(k_{\perp}^{(1)} +\pi_{ch}(k_{\perp}^{(2)})$. whre $Q_{\perp} =k_{\perp}^{(1)} -k_{\perp}^{(2)}$. Assuming boost invariance the correlator depends only on $k_{\perp}$ and we have for the second order correlation function \begin{equation} C_2=1+2p(1-p)e^{-Q^2_{\perp}R^2_{\perp}/2}+p^2e^{-Q^2_ {\perp}R^2_{\perp}} \label{eq:super1} \end{equation} where the transverse ``radius" $R_{\perp}$ plays the role of the correlation length and $p=\frac{<|\pi_{ch}>|^2>}{|\pi_c|^2 + <|\pi_{ch}>|^2>}$. Assume now that the chaoticity $p$ is rapidity dependent so that in one rapidity region, denoted by (A), $p(A) \approx 1$. In that region then the third term in the equation (\ref{eq:super1}) dominates, i.e. \begin{equation} C_2 \approx 1 + p^{2}(A)e^{-Q^{2}_{\perp}R^2_{\perp}(A)} \label{eq:chaotic} \end{equation} In the parametrization used in \cite{Na35}, according to which we have two independent sources, this suggests that the effective radius of source A is $R^{A}_{\perp}$. Conversely, for the more coherent region denoted by B, $p(B)$ is small and $C_2$ reads \begin{equation} C_2 \approx 1+2p(B)(1-p(B))e^{-Q^{2}_{\perp}R^{2}_{\perp}(B) /2} \label{eq:coherent} \end{equation} with an effective radius $R_{\perp}(B)/\sqrt{2}< R_{\perp}(A)$. This corresponds qualitatively to the observations made in ref.\cite{Na35}. Unfortunately these observations have not yet been confirmed by another, independent experiment so that the reader should view these considerations with prudence \footnote{For various caveats concerning the analysis of BEC and MD data cf. refs.(\cite{Na35},\cite{UA5},\cite{2comp},\cite{revisit}, \cite{Perugia}, \cite{QM88}).}. In any case they prove the usefulness of a global analysis which incorporates both BEC and MD. On the other hand, a dedicated simultaneous investigation of the rapidity dependence of these two observables appears very desirable. \subsubsection{Energy dependence of MD and its implications for BEC; long range fluctuations in BEC and MD} This subject has been discussed recently in \cite{lrc}. Besides the rapidity dependence, the dependence of MD on the center of mass energy of the collision, $\sqrt{s}$, constitutes an important topic in the study of high energy multiparticle production processes. This energy dependence is usually discussed in terms of the violation \cite{UA5} of KNO scaling \cite{KNO}. KNO scaling implies that the normalized moments $\langle n^m \rangle /\langle n \rangle^m$ are constant as a function of $s$ (for high energies, i.e. large $\langle n \rangle$, these moments coincide with the normalized factorial moments). For charged particles it turned out that while KNO scaling is approximately satisfied over the range of ISR energies ($20$ $GeV\leq \sqrt{s} \leq 60$ $GeV$), it is violated if one goes to SPS-Collider energies ($200$ $GeV\leq \sqrt{s} \leq 900$ $GeV$), i.e. one finds a considerable increase of multiplicity fluctuations with increasing energy. In the following we will show how BEC can be used to to understand the origin of this $s$-dependence. To do this, one needs to distinguish between long range (dynamical) correlations (LRC) and short range correlations (SRC)\footnote{The importance of making this distinction was pointed out, among other things, in \cite{foa,capella}. In \cite{faes} rapidity correlations were measured for events at fixed multiplicity in order to get rid of the effect of LRC.}. For identical bosons, one important type of SRC are BEC which reflect quantum statistical interference. In addition, there exist dynamical SRC like final state interactions, which however are quite difficult to be separated from BEC. Up to 1994 one usually assumed that LRC do not play an important part in BEC measurements (see however \cite{gyul},\cite{cap}) in the sense that for $Q>1 \ GeV$ the two-particle correlation functions do not significantly exceed unity. However in ref.\cite{lrc} evidence, based on observational data, was presented showing that this is not the case and that new and important information about LRC is contained in the BEC data obtained by the UA1-Minimum-Bias Collaboration \cite{bu}. In principle, the observed increase of multiplicity fluctuations with $\sqrt{s}$ could be due to a change of the SRC as seen in BEC, i.e. of the chaoticity and radii/lifetimes. This possibility was discussed in \cite{fowler} but could not be tested because of the lack of identical particle data for multiplicity distributions at Collider energies. At that time only the UA5 data \cite{UA5} for multiplicity distributions of charged particles were available. Furthermore, up to this point, the effect of LRC had only been studied in terms of two-particle correlations as a function of rapidity difference, i.e. in one dimension. With the advent of the newly analyzed UA1 data \cite{bu} for identical particles in three dimensions (essentially in $Q^2=-(k_1-k_2)^2$) this situation changed. The analysis of \cite{lrc} led to the conclusion of the existence in BEC data of long range fluctuations in the momentum space density of secondaries and to the realization that the increase with energy of multiplicity fluctuations is to a great extent due to an increase of the asymptotic values of the $m$-particle correlation functions $C_m^{asympt.}$, i.e. their values in the limit of large momentum differences where BEC do not play a role \footnote{The effect of LRC on BEC was discussed in ref.\cite{gyu}, where several models specific for nucleus-nucleus collisions were considered, but at that time no evidence for this effect could be found.}. In ref.\cite{bu}, the UA1 collaboration presents the two-particle correlation of negatively charged secondaries as a function of the invariant momentum difference squared $Q^2=({\bf k}_1-{\bf k}_2)^2- (E_1-E_2)^2$. The data (cf. Fig. 20) have two unusual features: (I) at large $Q^2$ the correlation function saturates above unity, and (II) at small $Q^2$ it takes on values above $2$. The higher order correlation functions also exhibit property (I) \cite{bu}\footnote{For higher order correlations the equivalent of property (II) is $C_m>m!$. (cf. also below). The values of $C_m({\bf k},...,{\bf k})$ for $m>2$ are apparently not yet available.}. By comparing the asymptotic values of the correlation functions at large momentum differences $C_m^{asympt.}$ ($m=2,...,5$), with the normalized factorial moments, $\phi_m \equiv \langle n (n-1)\cdot ... (n-m+1)\rangle / \langle n \rangle^m$ in the momentum space region$ |y| \leq 3, k_{\perp} > 0.15$ $GeV$ one finds that the contribution of the BE interference peak to the moments is negligible for such large rapidity windows. Herefrom one concludes in \cite{lrc} that (I) indicates the presence of LRC in the momentum space density of secondary particles and that it is quite plausible (cf. below) that (II) has to a great extent the same explanation. We sketch here the arguments of ref.\cite{lrc}. In general, LRC may be related to fluctuations in impact parameter or inelasticity, or fluctuations in the number of sources. In what follows, let us label these fluctuations by a parameter $\alpha$. The $m$-particle Bose-Einstein correlation function at a fixed value of $\alpha$ is given by \begin{equation} C_m({\bf k}_1,...,{\bf k}_m|\alpha) \ =\ \frac {\rho_m({\bf k}_1,...,{\bf k}_m|\alpha)} {\rho_1({\bf k}_1|\alpha) \cdot ... \cdot \rho_1({\bf k}_m|\alpha)} \end{equation} where $\rho_\ell({\bf k}_1,...,{\bf k}_\ell|\alpha)$ are the $\ell$-particle inclusive distributions. The fluctuations in $\alpha$ are described by a probability distribution $h(\alpha )$ with \begin{equation} \int d\alpha \ h(\alpha) \ = \ 1. \end{equation} If the experiment does not select events at fixed $\alpha$, the measured inclusive distributions are \begin{equation} \rho_m({\bf k}_1,...,{\bf k}_m) \ =\ \langle\rho_m({\bf k}_1,...,{\bf k}_m|\alpha)\rangle\:, \end{equation} where the symbol $\langle ... \rangle$ denotes an average over the fluctuating parameter $\alpha$, i.e. \begin{equation} \langle X(\alpha ) \rangle \ \equiv \ \int d\alpha \ h(\alpha) \ X(\alpha ) \end{equation} The $m$-particle correlation function at the intercept reads \begin{equation} C_m({\bf k},...,{\bf k}) \ =\ m! \ \frac{\langle\alpha^m \rangle}{\langle \alpha \rangle^m}, \end{equation} where the symbols $<>$ refer to averaging with respect to $h(\alpha)$. At large momentum differences one has \begin{eqnarray} &&C_m({\bf k}_1,...,{\bf k}_m) \ \rightarrow \ \frac{\langle \alpha^m\rangle }{\langle \alpha \rangle^m} = C_m^{asympt.} \qquad {\rm for} \quad |{\bf k}_i-{\bf k}_j|\rightarrow \infty \nonumber\\ &&(i\neq j, \ i,j=1,...m), \label{eq:cal} \end{eqnarray} i.e., the $m$-particle correlation functions can have intercepts above $m!$ and saturate at values above unity for large momentum differences. The most obvious candidate for the function $h(\alpha)$ is the inelasticity distribution which describes the event to event fluctuations of the inelasticity $K$. With the identification $\alpha \equiv \langle n(K)\rangle$, where $\langle n(K)\rangle$ is the mean multiplicity at inelasticity $K$, one obtains from the above considerations a first ``experimental" information about this important physical quantity at collider energies. Previous experimental information about this distribution was derived in \cite{fow1} from the data of ref. \cite{brick} at $\sqrt s \simeq 17$ $GeV$. The conclusions obtained in \cite{lrc} about LRC are based among other things on the different normalizations used in different experiments. Some tests related to these conclusions are proposed: \begin{itemize} \item Analysis of BEC at lower energies (NA22 range) as well as at $\sqrt s =1800$ $GeV$ with the same normalization as that used by the UA1 Collaboration. The values of $C_2$ at $Q^2 > 1$ $GeV^2$ and possibly also at very small $Q^2$ obtained in this way should exceed those obtained with the fixed multiplicity normalization in the same experiments. The following inequalities for $C_2(Q^2 > 1$ $(GeV)^2)$, should be observed if this normalization is used:\\ $C_2({\rm NA22}) <C_2({\rm UA1})<C_2({\rm Tevatron})$. \item Analysis of BEC at UA1 energies with the same normalization as that used so far by the NA22 and Tevatron groups (fixed multiplicity). The enhancement of $C_2$ at large $Q$ (and possibly also at small $Q$) observed so far, should disappear to a great extent.\\ \end{itemize} \subsection{Multiplicity dependence of Bose-Einstein correlations} The operators for the field (intensity) and number of particles do not commute. This means that measurements of ``ideal" BEC can be performed only when no restriction on the multiplicity $n$, which fluctuates from event to event, is made. In practice, however, very often such restrictions are imposed, either because of technical reasons or because of theoretical prejudices. To the last category belong considerations imposed by the search for QGP in high energy heavy ion reactions. Thus one expects that by selecting events with $n\geq n_{min}$, where $n_{min}$ is in general an energy dependent quantity, one gets information about the interesting "central" collisions. Another reason why multiplicity constraints are of practical importance for QGP experiments is the need to compare various QGP signals in a given event and at the same time determine for that event the radius, lifetime and chaoticity of the source, among other things in order to be able to estimate the energy density achieved in that event. This means that for QGP search it is interesting to perform interferometry measurements for single events, which of course have a given multiplicity. For these reasons the investigation of the multiplicity dependence of BEC constitutes an important enterprise which we shall address in the following section. \subsubsection{The quantum statistical formalism} Correlation functions defined in quantum statistics and used in quantum optics refer to ensemble averages of intensities \footnote{Herefrom the name ``intensity interferometry".} $I$ of fields $\pi$, where \begin{equation} I({\bf k}_{\perp},y)=|\pi({\bf k}_{\perp},y)|^2. \label{eq:Shih1} \end{equation} The total multiplicity $n$ of {\em identical particles} over a given phase space region is given by \begin{equation} n=\int d{\bf k}^2_{\perp}\int dy|\pi({\bf k}_{\perp},y)|^2 \label{eq:Shih2} \end{equation} Both the field $\pi({\bf k}_{\perp},y)$ and the intensity $I({\bf k}_{\perp},y)$ are stochastic variables. Averaging over an appropriate ensemble, we get the mean total multiplicity \begin{equation} <n>=\int d{\bf k}^2_{\perp}\int dy<|\pi({\bf k}_{\perp},y)|^2>. \label{eq:Shih2a} \end{equation} In \cite{Shih} the dependence of BEC within the QS formalism on the total multiplicity $n$ and on $n_{min}$ was investigated and it was found that the size of this effect is (especially at low $<n>$) surprisingly large and must not be ignored, as had been done before. Both the $n_{min}$ constraint and the $<n>$ constraint lead to a decrease of the correlation function $C_2$ at fixed $y_g = y_1-y_2$, i.e. to an antibunching effect. The last effect can be approximated, except for very large $n$ and small $y_g$ by a simple analytical formula \begin{equation} C^{(n)}_{(2)}(y_g)\approx C_{(2)}(y_g)\frac{n-1}{nf_2} \label{eq:Shih424} \end{equation} where $C_{2}^{n}$ denotes the correlation function at fixed $n$ and $f_2$ is the reduced factorial moment. These results show that BEC parameters like radii, lifetimes, and chaoticity do depend on the particular experimental conditions under which the measurements are performed. {\footnotesize It is worth mentioning that a multiplicity dependence of BEC was observed experimentally for $p-p$ and $\alpha-\alpha$ reactions at $E_{cm}= 53$ GeV and $31$ GeV respectively already in \cite{Axial}. There it was found that the transverse radius increases with the multiplicity of charged particles $n_{charged}$. This effect was interpreted by Barshay \cite{Barshay} to be a consequence of the impact parameter dependence. The same effect was seen in heavy ion reactions \cite{Na35} and got a similar interpretation in \cite{Kag} \footnote{The approach of \cite{Kag} combines simplified (1-d) hydrodynamics with a multiple scattering model, which also exploits the impact parameter dependence.}. The interpretation in terms of impact parameter dependence could be checked directly in heavy ion reactions since here the impact parameter can be determined on an event by event basis. Another mechanism for the increase of radius with multiplicity was proposed by Ryskin \cite{Ryskin88}. He pointed out that in high multiplicity events, which are associated by many authors to large transverse momenta of partons and thus to a regime where perturbative QCD applies, one expects that the size of the hadronization region should increase with the multiplicity like $\sqrt{n}$.} A related topic is the dependence of BEC on the rapidity density $d=\Delta n/\Delta y$, which has also been observed experimentally in $\bar p-p$ reactions at the CERN SPS collider \cite{Wheeler} and at the Fermilab tevatron \cite{Alex}. Using a parametrization \begin{equation} C_2=1+\lambda \exp(-R^{2}Q^2) \label{eq:simple} \end{equation} where $Q$ is the invariant momentun transfer, it was found that $R$ increased with $d$ while $\lambda$ decreased with $d$. This last observation is compatible with the results from \cite{Shih}; a more quantitative comparison would be possible only if, among other things, the data were parametrized in a way more consistent with quantum statistics. Thus the four momentum difference $Q$ is not an ideal variable for BEC (cf. section 2.1.5) and the coherence effect has to be taken into account as outlined in section 2.2 and not by the simple empirical $\lambda$ factor. \subsubsection{The wave function formalism; ``pasers"?} The dependence on total multiplicity of BEC was investigated also within the wave function formalism. In section 2.1 where the GGLP theory was presented it was pointed out that the wave function formalism may be useful for exclusive processes or for event generators. Indeed, in a first approximation, the wave function $\psi_n$ of a system of $n$ identical bosons e.g. can be obtained from the product of $n$ single particle wave functions $\psi_1$ by symmetrization. Then the calculation of $C_{(2)}^{(n)}$ is in principle straightforward and follows the lines of GGLP. However when the multiplicities $n$ become large (say $n>20$) the explicit symmetrization of the wave function formalism becomes difficult. This lead Zajc \cite{Zajc} to use numerical Monte Carlo techniques for estimating $n$ particle symmetrized probabibilities, which he then applied to calculate two-particle BEC. He was thus able to study the question of the dependence of BEC parameters on the multiplicity. For an application of this approach to Bevalac heavy ion reactions, cf. \cite{WNZhang}. Using as input a second order BEC function paramerized in the form (\ref{eq:simple}) Zajc found, (and we have seen above that this was confirmed in \cite{Shih},) that the ``incoherence" parameter $\lambda$ decreased with increasing $n$ \footnote{In \cite{Zajc} the clumping in phase space due to Bose symmetry was also illustrated;}. However Zajc did not consider that this effect means that events with higher pion multiplicities are denser and more coherent. On the contrary he warned against such an interpretation and concluded that his results have to be used in order to eliminate the {\em bias} introduced by this effect into experimental observations \footnote{The same interpretation of the multiplicity dependence of BEC was given in \cite{Shih}. In this reference the nature of the ``fake" coherence induced by fixing the multiplicity is even clearer, as one studies there explicitly coherence in a consistent quantum statistical formalism.}. This warning apparently did not deter the authors of \cite{Pratt1} and \cite{Chao} to do just that. Ref.\cite{Pratt1} went even so far as to deduce the possible existence of pionic lasers from considerations of this type \footnote{The philosophy of ref.\cite{Pratt1} is reflected e.g. in statements like ``multiparticle interference can enhance the emission of pions". This language is apparently motivated by the practice of event generators where BEC (or in general quantum statistical effects) are ``added" to the usual schemes at the end of the calculation. That this does not reflect the true sequence of the physical processes considered is obvious.}. Ref. \cite{Pratt1} starts by proposing an algorithm for symmetrizing the wave functions which presents the advantages that it reduces very much the computing time when using numerical techniques, which is applicable also for Wigner type source functions and not only plane wave functions, and which for Gaussian sources provides even analytical results. Subsequently in ref.\cite{Zimanyi} wave packets were symmetrized and in special cases the matrix density at fixed and arbitrary $n$ was derived in analytical form. This algorithm was then applied to calculate the influence of symmetrization on BEC and multiplicity distributions. As in \cite{Zajc} it is found in \cite{Pratt1} that the symmetrization produces an effective decrease of the radius of the source, a broadening of the multiplicity distribution $P(n)$ and an increase of the mean multiplicity as compared to the non-symmetrized case. What is new in \cite{Pratt1} is (besides the algorithm) mainly the meaning the author attributes to these results. In a concrete example Pratt considers a non-relativistic source distribution $S$ in the absence of symmetrization effects: \begin{equation} S(k,x)=\frac{1}{(2\pi R^2mT)^{3/2}}\exp\left(-\frac{k_0}{T}- \frac{x^2}{2R^2}\right)\delta(x_0) \label{eq:Pratt8} \end{equation} where \begin{eqnarray} k_0/T=k^2/2{\Delta}^2 \end{eqnarray} Here $T$ is an effective temperature, $R$ an effective radius, $m$ the pion mass and $\Delta$ a constant with dimensions of momentum. Let $\eta_0$ and $\eta $ be the number densities before and after symmetrization, respectively. In terms of $S(k,x)$ we have \begin{equation} \eta_0= \int S(k,x)d^4kd^4x \label{eq:dens} \end{equation} and a corresponding expression for $\eta $ with $S$ replaced by the source function after symmetrization. Then one finds \cite{Pratt1} that $\eta $ increases with $\eta_0$ and above a certain crtitical density $\eta^{crit}_0 $, $\eta $ diverges. This is interpreted by Pratt as ``pasing". The reader may be rightly puzzled by the fact that while $\eta $ has a clear physical significance the number density $\eta_0 $ and a fortiori its critical value have no physical significance, because in nature there does not exist a system of bosons the wave function of which is not symmetrized. Thus contrary to what is alluded to in ref.\cite{Pratt1}, this paper does not does address really the question how a condensate is reached. Indeed the physical factors which induce condensation are, for systems in (local) thermal and chemical equilibrium \footnote{For lasers the determining dynamical factor is among other things the inversion.}, pressure and temperature and the symmetrization is contained automatically in the distribution function \begin{equation} f=\frac{1}{\exp[(E-\mu]/T]- 1} \label{eq:f} \end{equation} in the term -1 in the denominator; $E$ is the energy and $\mu$ the chemical potential. To realize what is going on it is useful to observe that the increase of $\eta_0 $ can be achieved by decreasing $R$ and/or $T$. Thus $\eta_0$ can be substituted by one or both of these two physical quantities. Then the blow-up of the number density $\eta $ can be thought of as occuring due to a decrease of $T$ and/or $R$. However this is nothing but the well known Bose-Einstein condensation phenomenon and does not represent anything new. While from a purely mathematical point of view the condensation effect can be achieved also by starting with a non-symmetrized wave function and symmetrizing it afterwards ``by hand" , the causal i.e. physical relationship is different: one starts with a bosonic i.e symmetrized system and obtains condensation by decreasing the temperature or by increasing the density of this {\em bosonic} system. Another confusing interpretation in \cite{Pratt1} relates to the observation made also in \cite{Zajc} that the symmetrization produces a broadening of the multiplicity distribution (MD). In particular starting with a Poisson MD for the non-symmetrized wf one ends up after symmetrization with a negative binomial. While Zajc correctly considers this as a simple consequence of Bose statistics, ref.\cite{Pratt1} goes further and associates this with the so called pasing effect. That such an interpretation is incorrect is obvious from the fact that for true lasers the opposite effect takes place. Before ``condensing" i.e. below threshold their MD is in general broad and of negative binomial form corresponding to a chaotic (thermal) distribution while above threshold the laser condensate is produced and as such corresponds to a coherent state and therefore is characterized by a Poisson MD. Last but not least the fact that this broadening increases with $n$ is not due, as one might be mislead to believe from \cite{Pratt1},\cite{Chao}, to the approach to ``lasing criticality", but simply to the fact that the larger $n$ the larger the number of independent emitters is and the better the central limit theorem applies. This theorem states (cf. section 2.2) that the field produced by a large number of independent sources is chaotic. Finally a terminological remark appears necessary here. We believe that names like ``paser" or pionic laser used in the papers quoted above are unjustified and misleading. The only characteristic which the systems considerd in these papers possibly share with lasers is the condensate property i.e. the bunching of particles in a given (momentum) state. However lasers are much more than just condensates; one of the main properties of lasers which distinguishes them from other condensates is the directionality, a problem which is not even mentioned in the ``paser" literature. \footnote{For a model of directional coherence, not necessarily related to pion condensates, cf.\cite{FW85}; experimental hints of this effect have possibly been seen in \cite{Zajc84}.} To conclude this excursion into what in our opinion is a surrealistic interpretation of multiparticle correlations, it is hard to find new physics in the ``paser" papers, although they probably represent progress in the mathematical problem of symmetrization of wave functions or wave packets\footnote{For another investigation of the effect of symmetrization on the single inclusive cross section cf. \cite{Ray} and \cite{Wied88}. In \cite{Wied88} second order correlation functions are also considered.(Cf. also ref.\cite{RH} for this topic.)}. \subsection{The invariant $Q$ variable in the space-time approach; higher order correlations; ``intermittency" in BEC?} The issue of apparent power-like rapidity dependence of moments of MD was discussed in section 6.1.1 where it was pointed out that this dependence could in principle be understood within the QS formalism without invoking the idea of intermittency. However this was not the end of the story because:1) it was observed \cite{bu} that the power like behaviour extends also to BEC data (in the invariant variable $Q$). This was surprising because up to that moment BEC data could usually be fitted by a Gaussian or exponential function, albeit these data did not extend to such small $Q$ values as those measured in \cite{bu}; 2) Bialas \cite{bi} (cf. also \cite{bizi}, \cite{zi}) proposed that the source itself has no fixed size, but is fluctuating from event to event with a power distribution of sizes. Since the measurements made in \cite{bu} were in the invariant variable $ Q$ rather than $\Delta y$, one had to understand whether the $QS$ approach which implies fixed scales does not lead to a similar behaviour in $Q$. Refs. \cite{intpl}, \cite{intpr} addressed this question and proved that, indeed, by starting from a space-time correlator with a fixed correlation length and a source distribution with a fixed radius, one gets after integrations over the unobserved variables a correlation function which is power-like in a limited $Q$ range. In the following we shall sketch how this happens. The two-particle Bose-Einstein correlation function is defined as \begin{equation} C_2({\bf k}_1,{\bf k}_2) = \frac{\rho_2({\bf k}_1,{\bf k}_2)} {\rho_1({\bf k}_1) \cdot \rho_1({\bf k}_2)} \label{eq:c2rho} \end{equation} where $\rho_1({\bf k}^{})$ and $\rho_2({\bf k}_1,{\bf k}_2)$ are the one- and two-particle particle inclusive spectra, res\-pec\-tive\-ly. The two-particle correlation function projected on $Q^2$ is \begin{equation} C_2(Q^2) = 1 + \frac{{\cal I}_2(Q^2)}{{\cal I}_{11}(Q^2)} \label{eq:c2rat} \end{equation} with the integrals \begin{eqnarray} {\cal I}_{11}(Q^2) &=& \int d\omega_1 \int d\omega_2 \: \ \delta \left[ Q^2+(k_1^\mu-k_2^\mu)^2 \right] \rho_1({\bf k}_1) \rho_1({\bf k}_2) \/ \nonumber\\ {\cal I}_2(Q^2) &=& \int d\omega_1 \int d\omega_2 \: \ \delta \left[ Q^2+(k_1^\mu-k_2^\mu)^2 \right] \left(\rho_2({\bf k}_1,{\bf k}_2) \ - \ \rho_1({\bf k}_1) \rho_1({\bf k}_2)\right) \label{eq:i112} \end{eqnarray} where $d\omega \equiv d^3k/(2\pi)^32E$ is the invariant phase space volume element. Two types of sources, a static one and an expanding one, were considered. The parameters employed are physically meaningful quantities in the sense that they give the lifetime, radii and correlation lengths of the source. This is not the case for the ad-hoc parametrization of $C_2(Q^2)$, e.g., with a Gaussian \begin{equation} C_2(Q^2)\:=\:1\:+\:\lambda_Q\:e^{-R_Q^2 Q^2} \label{eq:c2qmi1} \end{equation} To illustrate the behaviour of the correlation function as a function of $Q^2$, one applied the formalism to describe UA1 data in the phase space region $| y | \leq 3.0$ and $k_\perp \geq 150 MeV$. In \cite{intpl},\cite{intpr} $C_2(Q^2)$ was calculated for the static and for the expanding source by Monte Carlo integration. (for the static source, approximate analytical results could be obtained only for $|y|>1.5$, where they agree with the numerical results \cite{intpr}). Fig. 21 shows the results of fits to the UA1 data\footnote{The $C_2(Q^2)$ data of \cite{bu} were normalized so that at large $Q^2$, $C_2(Q^2) \approx 1$. An explanation for the experimental observation of \cite{bu} that $C_2(Q^2)$ exceeds by a multiplicative factor of $\sim 1.3$ both the upper and lower ``conventional'' limits of 2 and 1, respectively, is discussed in ref. \cite{lrc}.} for the static source (dashed line) and for the expanding source (solid line), which were obtained under the assumption of a purely chaotic source, $p_0=1$ (the data are consistent with an amount of $\leq 10 \%$ coherence\footnote{The older UA-1 data \cite{UA1} were limited to larger $Q^2$ values. Furthermore the quantum statistical interpretation \cite{mad} of these data is different from that in \cite{intpl}, which is based on space-time concepts. This explains the different values of chaoticity obtained in \cite{intpl}, \cite{intpr} on the one hand and in refs. \cite{UA1},\cite{mad} on te other hand}, but the sensitivity to $p_0$ was not sufficient to further constrain the degree of coherence within these limits). For comparison, the result of a power law fit (dotted line) as suggested in ref. \cite{bu} \begin{equation} C_2(Q^2)\:=\:a\:+\:b\cdot \left(\frac{Q^2} {1 GeV^2}\right)^{-\phi} \label{eq:power} \end{equation} is also plotted. One finds that the data can be well described both with the static and the expanding source model with reasonable values for the radii, lifetime and correlation length. The results of above show that the power-like behaviour of $C_2$ (the same holds for higher order correlations) can be reproduced by assuming a conventional space-time source with fixed parameters, i.e., without invoking ``intermittency" \footnote{A similar point of view is expressed in \cite{PPT1} for the particular case of bremsstrahlung photons.}. This conclusion of \cite{intpl} is strengthened by an explicit consideration of resonances in \cite{intpr}. The advantages of the QS formalism as compared with the wave function formalism emerge clearly also in the problem of higher order correlations. In the QS formalism higher order correlations are treated on the same footing as lower order ones and emerge just as consequences of the form of the density matrix. Therefore questions found sometimes in the literature like ``what is the influence of higher order correlations on the lower one" do not even arise in QS and in fact do not make sense. We emphasized in section 2.2.1 the importance of higher order correlations for the phenomenological determination of the form of the correlation function and this applies in particular when a single variable like $Q$ is used instead of the six independent degrees of freedom inherent in the correlator. For this reason the space-time integration at fixed $Q$ has been used recently \cite{Nelly} also in the study of higher order correlations and applied to the NA22 data \cite{Na22}. It was found among other things that an expanding source with fixed parameters as defined in chapter 4 can account for the data up to and including the fourth order, confirming in a first approximation the Gaussian form of the density matrix. The fact that previous attempts in this direction like \cite{Na22} and \cite{UA1new} met with difficulties may be due to the fact that in these two experimental studies the QO formalism in momentum space was used and possibly also because no simultaneous fit of all orders of correlations was performed, as was the case in \cite{Nelly}. \section{Critical discussion and outlook} For historical reasons related to the fact that the GGLP effect was observed for the first time in annihilation at rest when (almost) exclusive reactions were studied, the theory of BEC was initially based on the wave function formalism. This formalism is not appropiate for inclusive reactions in high energy physics, among other things because it yields a correlation function which depends only on the momentum difference $q$ and not also on the sum of momenta, it does not take into account isospin, and cannot treat adequately coherence. This last property is essential as it leads to one of the most important applications of BEC, i.e. to condensates. Furthermore this property strongly affects another essential application of BEC, namely the determination of sizes, lifetimes, correlation lengths and correlation times of sources. The adhoc parametrization of the correlation function under the form (\ref{eq:lambda}) where $\lambda$ is supposed to take into account (in)coherence is unsatisfactory. An improvement based on an analogy with quantum optics leads to an additional term (with the same number of parameters). A more complete and more correct treatment of BEC is provided by the space-time formalism of classical currents. Almost all studies of BEC assume a Gaussian density matrix. Possible deviations from this form could and should be be looked for by studying higher order correlations. The classical current approach is based on an exact solution of the equations of quantum field theory and it can be considered at present the most advanced and complete description of BEC. It introduces a primordial correlator of currents which is characterized by finite correlation lengths and correlation times. The geometry of the source is an independent property of the system and it is here that the traditional radii and lifetimes enter. The form of the geometry and of the correlator is not given by the theory and it is up to experiment to determine these. For a Gaussian density matrix the space-time approach within the classical current formalism leads to a minimum of ten independent parameters; these include geometrical and dynamical scales as well as the chaoticity. Phenomenologically these scales can be separated only by considering simultaneously single and double inclusive cross sections. Experimentally this separation as well as the determination of all parameters has not yet been performed and constitutes an important task for the future. {\em This should be done not only for particle reactions but also for heavy ion reactions, as an alternative to the pseudo-hydrodynamical approach which cannot provide this separation.} It would also be desirable to extend the parametrization for an expanding source by dropping the assumption of boost invariance. Besides the conventional $++$ or $--$ pion correlations there exist also $+-$ correlations, which become important for sources of small lifetimes. These "surprising" field theoretical effects represent squeezed states, which unlike what happens in optics, appear in particle physics ``for free". These effects can be used to investigate the difference between classical and quantum currents. Their detection constitute one of the most important challenges for future experiments. BEC are influenced by final state interactions. Coulomb final state interactions do not play a major part except at very small $q$. These small values have apparently not yet been reached in experiment and it is questionable whether they will be reached in the foreseable future. In heavy ion reactions this is due to the large values of radii which multiply $q$ and even for typical scales of $1$ fm the present resolution of detectors is not sufficient to make this effect very important. Nevertheless since Coulomb corrections have been studied so far only within the wave function formalism and usually by applying the Schr\"odinger equation, it would be desirable to extend this study by considering the Klein-Gordon equation which is more appropiate for mesons. Even more interesting would be to study Coulomb corrections within the classical current formalism. Resonances play an essential part in final state interactions and progress has been achieved in their understanding, in particular in heavy ion reactions, where their influence has been investigated using solutions of the equations of hydrodynamics within the Wigner function formalism BEC have been investigated in $e^{+}-e^{-}$, hadron-hadron and heavy ion reactions, however a systematic comparison of results using the same paramterization and normalization awaits still to be done\footnote{At present even for the same type of reaction different normalizations are frequently used and this can lead to {\em apparently} different results, as exemplified in the case of $NA22$ and $UA1$ data. Tests for a better understanding and elimination of these discrepancies have been proposed (cf. section 6.1.3)}. Correlations are intimately related to multiplicity distributions which can serve as complementary tools in the determination of the parameters of sources. Therefore a systematic investigation of BEC and multiplicity distributions in the same phase space region is desirable. How useful this can be has been shown by proving in this way the influence of long range correlations on BEC. BEC can be useful in heavy ion reactions and in particular for the search of quark-gluon plasma if one of the two conditions are satisfied: (A) the investigation is based on full-fledged hydrodynamics, impying the solution of the equations of hydrodynamics with explicit consideration of the equation of state. (B) the investigation is based on the classical current space-time approach. Case (A) has the advantage that the dependence of the equation of state on the phase transition may be reflected also in single inclusive cross sections and in the BEC. (So far, however, the sensitivity of BEC on the equation of state could not yet been proven with present data \cite{SchleiHung}.) It has however the disadvantage that its applicability is restricted because it is based on the Wigner function, which is a particular case of the classical current formalism, for small $q$ and not too strong correlations between momenta $k$ and coordinates $x$. Case (B) has the disadvantage that there is no contact with the equation of state. However it has the advantage it is not restricted to small $q$ and weak correlations between $k$ and $x$. Unfortunately many of the ``theoretical" papers on BEC in heavy ion reactions do not satisfy either condition (A) or condition (B). This is the case with most of the pseudo-hydrodynamical papers which use a parametrization of the source function based on qualitative hydrodynamical considerations without the use of an equation of state and without solving the equations of hydrodynamics. This pseudo-hydrodynamical approach has the disadvantages of (A) and (B) but none of their advantages\footnote{Pseudo-hydrodynamics as compared with hydrodynamics has also the supplimentary disadvantage that it does not allow a separation between geometrical radii and correlation lengths. Moreover it does not have even the excuse of simplicity, since the number of free parameters in the psudo-hydrodynamical approach is as large as that in the classical current space-time approach.}. >From the above considerations it is seen that there are many usolved problems in the investigation of BEC, some of them of theoretical, but most of them of experimental nature. While an analysis of the experimental BEC deserves a special review, some of the obvious reasons for this unsatisfactory experimental situation are: 1) Most of the BEC experiments performed so far use inadequate detectors, because they are not dedicated experiments but rather byproducts of experiments planned for other purposes. What is needed amng other things is track by track detection and improved identification of particles. 2) Insufficient statistics. An improvement of statistics especially at small $q$ by at least one order of magnitude is necessary to address some of the problems enumerated above. 3) Incorrect or incomplete parametrizations of the correlation functions. Very often and in part because of 2) not all six independent variables of $C_2$ are measured, but projection of these. Very popular among these projections is the relativistically invariant variable $Q$. This is not a good variable for BEC studies, because among other things it mixes the space and time variables in an uncontrollable way. Furthermore, in most parametrizations, coherence is not (or inadequately) considered. 4) Inadequate normalizations. Practically all BEC experiments use a normalization procedure of the correlation function which does not correspond to its definiton. This definition refers to the single inclusive cross sections obtained in the same event as the double inclusive cross section. The use of an ``uncorrelated background", instead, biases the results. The solution of the problems mentioned above will make of boson interferometry what it is supposed to be: a reliable method for the determination of sizes, lifetimes, correlation lengths, and coherence of sources in subatomic physics. A more pedagogical presentation of the theory of Bose-Einstein correlations, which discusses also its quantum optical context, including a comparison between the HBT and the GGLP effects and between photon and hadron intensity interferometry can be found in the book by the author \cite{book}. Some of the most representative theoretical and experimental papers on BEC and which are frequently quoted within the present review have been reprinted in a single volume in ref. \cite{reprints}. {\bf Acknowledgements} I am indebted to D. Strottman for a careful reading of the manuscript and for many helpful comments. \newpage \section{Figure captions} \begin{description} \item[Fig. 1] The GGLP experiment schematically. \item[Fig. 2] Second order correlation function $C_2(k,k)\equiv g^{(2)}$ as a function of the squeezing parameter $r$ for pure squeezed states (from. ref.\cite{VW}). \item[Fig. 3] Second order correlation function as given by the quantum statistical formalism of ref. \cite{Stelte} for various values of the coupling constant $g$ for $\kappa =0.5$ (from \cite{Stelte}). The parameter $\kappa$ is related to the chaoticity $p$ via the relation $\kappa =<n_c>/<n_{ch}> =1/p-1$ where $p=<n_{ch}/n>$ and $n = n_{c} +n_{ch}$ is the total multiplicity. $Y$ is the maximum rapidity. \item[Fig. 4] The same for various values of $\kappa$ at $g = 0$ (from \cite{Stelte}). \item[Fig. 5] Second order correlation function for various values of the ``incoherence" parameter $\lambda$ as given by the phenomenological equation (\ref{eq:Deutschmann}). \item[Fig. 6] Diagrams contributing to different terms of the two-particle correlator in the point-like random source model (from ref.\cite{APW}). \item[Fig. 7] The linear polarization vectors $\epsilon_{lambda}(k)$ for two photons with momenta ${\bf k}_1$ and ${\bf k}_2$ (from ref.\cite{Feld}). \item[Fig 8] Geometry of the boost-invariant source. \item[Fig. 9] Intercept of two-particle correlation function in the presence of coherence and resonances (from ref. \cite{Bolz}). \item[Fig. 10] Bose-Einstein correlation functions in longitudinal and transverse direction, for the 3-dimensional (solid lines) and the 1-dimensional calculations (dashed lines) (from ref. \cite{Schlei97}). \item[Fig. 11] Dependence of the longitudinal and transverse radii extracted from Bose-Einstein correlation functions on the rapidity $y_K$ and average momentum $K_\perp$ of the pair. As before, solid lines correspond to the 3-dimensional and dashed lines to the 1-dimensional results. The open circles indicate values of $R_\parallel$ obtained from \mbox{eq.(\ref{eq:016})} (from ref. \cite{Schlei97}). \item[Fig. 12] Bose-Einstein correlation functions of negatively charged pions, in longitudinal and transverse direction. The separate contributions from resonances are successively added to the correlation function of direct (thermal) $\pi^-$ (dotted line). The solid line describes the correlation function of all $\pi^-$ (from ref. \cite{bernd3}). \item[Fig. 13] Correlation functions of all $\pi^-$ (solid lines), thermal $\pi^-$ (dotted lines) and all $K^-$ (dashed lines), for $k_\perp=0$ and for $k_\perp=1$ GeV/c (from ref. \cite{bernd3}). \item[Fig. 14] Dependence of the longitudinal and transverse radii extracted from Bose-Einstein correlation functions on the rapidity $y_k$ and average momentum $k_\perp$ of the pair, for all $\pi^-$ (solid lines), thermal $\pi^-$ (dotted lines) and all $K^-$ (dashed lines). The full circles were obtained by substituting the value $<t_f(z=0,r_{\perp})> = 2$ fm/c for the average lifetime of the system (calculated directly from hydrodynamics by averaging over the hypersurface) into eq. (\ref{eq:016}), with \mbox{$T_f=0.139$ GeV} (from ref. \cite{bernd3}). \item[Fig. 15] $\pi^-\pi^-$ Bose-Einstein correlation functions in the presence of partial coherence (from ref. \cite{bernd3}). \item[Fig. 16] Effective radii extracted from Bose-Einstein correlation functions as a function of the rapidity $y_k$ of the pair and the transverse average momentum $K_{\perp}$ of the pair for all pions ( from ref. \cite{OrnikPR96}). \item[Fig. 17] Effective radii extracted from Bose-Einstein correlation functions as a function of the transverse average momentum $K_{\perp}$ of the pair for all pions compared with data (from ref. \cite{OrnikPR96}). \item[Fig. 18] Two-particle BE correlation functions for a spherically and a longitudinally expanding source. The different line styles correspond to different condensate densities $n_{co}$ compared to the thermal number densities $n_{th}$ (from ref. \cite{cond97}). \item[Fig. 19] Normalized factorial moments $\Phi_q$ of order $q$ in finite (pseudo) rapidity windows of width $\delta y$ around $y_{CMS} = 0$, plotted against $\delta \eta $ (from ref. \cite{car}). \item[Fig. 20] Second order correlation function for negative particles at $\sqrt s = 630 GeV$ (from ref. \cite{bu}). \item[Fig. 21] $C_2(Q^2)$ for the static source model (dashed line) and for the expanding source model (solid line) compared to the UA1 data \cite{bu}. The dotted line shows a power law fit. \end{description} \newpage \section{References}
\section{Introduction} The Coulomb gas picture of rational conformal field theory, and its various generalizations, has amply proved to be a very efficient tool for the computation of conformal blocks, structure constants and braiding properties in these theories \cite{FF,DF,FFK,PRY}. Interestingly, Coulomb gas techniques have turned out to describe (certain sectors of) irrational conformal field theories as well, such as WZNW theory and Liouville/Toda theory. Within the Gervais-Neveu quantization of Liouville theory, an efficient formulation of screened vertex operators, avoiding completely the manipulation of contours, was introduced long ago in \cite{GN832,Sch90,GS93}. The historical development of the exploration of Liouville theory in the operator framework was in fact such that for a long time, only primary fields corresponding to the Kac table were considered, in close analogy to the situation for minimal models. The next step of generalization \cite{GS94,GR94} consisted in formulating observables with arbitrary conformal dimensions, which however involved only a {\em non-negative} integer number of screenings. It was pointed out in ref. \cite{Sch96} that while those observables (the general Liouville exponentials) are formally described by an infinite sum over positive integer powers of screening operators, these infinite sums do not permit any naive evaluation even in the simple context of three-point functions. Depending on the three-point function considered, negative screening powers can arise as a non-perturbative effect, establishing contact in this way with the Goulian-Li procedure \cite{GL}. The analysis of the three-point functions furthermore suggests that there should exist a conformal algebra, closed under fusion and braiding, involving chiral vertex operators with both positive and negative integer powers of screenings. This would constitute a non-trivial generalization of the standard Coulomb gas picture and open very interesting perspectives towards an operator formulation of a new class of irrational conformal field theories. A great virtue of the Gervais-Neveu approach consists in the fact that negative integer powers of screenings are just as well defined as positive ones so that their introduction does not require any analytic continuation procedure. Negative and positive screening powers are in fact related by a Weyl reflection \cite{BG89,ANPS94}, exchanging one of the two equivalent free fields of the Gervais-Neveu approach with the other. However, except for the cases considered in \cite{GR94}, so far very little has been known about the braiding and fusion algebras of vertex operators involving negative powers of screenings. It is clear that the $R$-matrix and the fusion matrix should be given by an appropriate analytic continuation of $q$-deformed $6j$-symbols, as was the case already in the generalization from the Kac table to arbitrary continuous spins \cite{GS93}. The objective of the present paper is to establish the braiding algebra of chiral vertex operators with {\em arbitrary} integer screenings and continuous spins. The operator product is then determined as well by the general proportionality relation of Moore and Seiberg \cite{MS} between fusion and braiding matrices. The remaining part of the paper is organized as follows: In Section 2 we introduce our notation and provide some well known background material. In Section 3 we present our proposal (ansatz) for the analytic continuation of the $R$-matrix to negative screening numbers. The continuation procedure is remarkably simple, and uses no more than a well-known transformation formula for (truncating) $q$-hypergeometric sums of type $_4F_3$, which constitute the essential part of the analytic expression for the braiding matrix. Depending on the signs of the screening numbers of the vertex operators to be braided, the standard braiding matrix for positive integer screenings of refs. \cite{GS93,GS94} can always be brought into a form such that the continuation corresponds to a simple substitution of positive by negative screening numbers. In Section 4 we derive, as a simple consequence of the polynomial equations of Moore and Seiberg \cite{MS}, a system of determining equations for $R$-matrices with a mix of both positive and negative ``ingoing'' screening numbers. We verify that our proposal fulfills these equations, starting with the case where only one of the vertex operators to be braided has negative screening. In the process we need several new relations for $q$-hypergeometric functions, in particular of the type $_4F_3$ and generalizations thereof. Proofs are outlined and will appear in more detail elsewhere \cite{RS}. We also discuss the delicate issue of uniqueness of the solution. In Section 5 we turn to the class of $R$-matrices where both ingoing screening numbers are negative. In order to verify our analytic continuation proposal for this case, we will exploit the fact that it can be reduced to the ones treated in Section 4 by means of a concatenation procedure. Substantial evidence in favour of our proposal is then provided by checking explicitly the resulting identity for several non-trivial classes of examples. In Section 6 we discuss the connection of our extended $R$-matrices with $6j$-symbols. Section 7 describes the trivial generalization from only one type of screening charge to the case where the conjugate screening charge is included. Finally, Section 8 is devoted to concluding remarks and a speculative outlook. We discuss in particular the non-chiral case in the context of Liouville theory, including the strong coupling regime. Appendix A contains some useful observations on certain transformations of $q$-hyper- \break geometric functions of type ${}_4F_3$, whereas Appendix B is devoted to further considerations on the uniqueness of our proposal for the braiding matrix. \section{Coulomb Gas Picture and Liouville Theory} We start by introducing our notation and by recalling some elementary facts about the Coulomb gas picture as used in refs. \cite{Sch90,GS93,GS94,Sch96}. As usual, Coulomb gas vertex operators are written as a product of a free field vertex operator and some power of screening operators. A particularity of the formulation chosen in the above references is that the integration contour for the screenings is fixed once and for all, and does not depend on the correlator in question. Concretely, for a vertex operator of spin $J$ and $U(1)$ charge $m$ we have \ben U_m^{(J)}(\sigma)\equiv V^{(J)}_{-J}(\sigma)S^{J+m}(\sigma) \een with \ben V^{(J)}_{-J}(\sigma)=e^{\alpha_- J X}(\sigma), \qquad S(\sigma)=e^{2ih(\varpi+1)}\int_0^\sigma dx V_1^{(-1)}(x) +\int_\sigma^{2\pi}dx V_1^{(-1)}(x) \label{vertexdef} \een Here, $X(\sigma)$ is a canonical free field and $\alpha_-$ is the ``semi-classical'' screening charge, related to the central charge $c$ of the theory in the usual way: \ben c=1+{12\over \alpha_-^2}(1+{\alpha_-^2\over 2})^2 \een Note that $c$ is arbitrary continuous, and $c>25$ for $\alpha_-$ real. In terms of the deformation parameter $h$ ($q=e^{ih}$ is the deformation parameter relevant for the quantum group interpretation \cite{G90,CGR942,CGS97}) we have \ben \alpha_-=\sqrt{2h\over \pi}, \qquad c=1+{6\pi\over h}\left(1+{h \over \pi}\right)^2 \een Moreover, $\varpi$ is essentially $i$ times the momentum zero mode of the free field, and is taken to be real in the following (this corresponds to the elliptic sector of Liouville theory \cite{GN82}, and is also the choice appropriate for instance for the description of minimal models in this framework \cite{G93}). Explicitly, one has \ben X(\sigma)=q_0+p_0\sigma+i\sum_{n\ne 0} \ e^{-in\sigma}{p_n\over n} \een and $\varpi=ip_0\sqrt{2\pi\over h}$. The $U_m^{(J)}$ operators are characterized by their conformal weight $\Delta_J$, their $U(1)$ charge (momentum shift) $m$, and their normalization $I_m^{(J)}(\varpi)\equiv\bra{\varpi}U^{(J)}_m(\sigma=0)\ket{\varpi+2m}$: \bea \Delta_J&=&-J-{h\over\pi}J(J+1)\nn U_m^{(J)}\varpi&=&(\varpi+2m)U_m^{(J)}\nn I_m^{(J)}(\varpi)&=&\left( 2 \pi \Gamma(1+{h\over\pi}) \right )^{J+m} e^{ ih(J+m)(\varpi-J+m)}\nn &\cdot&\prod_{\ell=1}^{J+m}\frac{\Gamma[1+(2J-\ell+1)h/\pi]}{ \Gamma[1+\ell h/\pi]\Gamma[1-(\varpi+2m-\ell)h/\pi] \Gamma[1+(\varpi+\ell)h/\pi]} \label{3.6} \eea The braiding properties and the operator product of the $U_m^{(J)}$ operators were derived in ref. \cite{CGR941} in the case of degenerate operators ($2J=0,1,2,...$ and $m=-J,-J+1,...,J$) and generalized in refs. \cite{GS93,GR94} to the case of arbitrary $J$ and integer positive screening numbers $n=J+m$. Their algebra fulfills the Moore-Seiberg equations \cite{MS} of conformal field theory \cite{GR94}. For later use, we introduce a special notation for the leading order operator product (fusion) of the two operators $U_m^{(J)}$ and $U_{m'}^{(J')}$: \ben U^{(J)}_m(\sigma)\odot U^{(J')}_{m'}(\sigma):= \lim_{\sigma'\to\sigma} \frac{U^{(J)}_m(\sigma)\cdot U^{(J')}_{m'}(\sigma')}{ (1-e^{i(\sigma'-\sigma)})^{-\Delta_J-\Delta_{J'}+\Delta_{J+J'}}} \label{odotdef} \een (as usual, $\sigma'$ has to be given a small positive imaginary part to make the above expression well defined). One has the simple multiplication law \ben U_{m}^{(J)}(\sigma)\odot U_{m'}^{(J')}(\sigma)=q^{2J'(J+m)}U_{m+m'}^{(J+J')} (\sigma) \label{odot} \een for any (positive or negative) integer values of the screenings. Local observables (in the Liouville context they are the Liouville exponentials) can be constructed as bilinear combinations of the $U_m^{(J)}$ and their right moving counterparts ${\overline U}_m^{(J)}$, with $\varpi$-dependent coefficients \cite{Sch90,G93,GS93}. Locality is a simple consequence of the orthogonality relations obeyed by the $q$-deformed $6j$-symbols that constitute the essential part of the chiral braiding matrices. In the present paper, we shall concentrate on the algebra of the chiral vertex operators; the application of the extended formalism with negative screenings to the construction of local observables, as considered in \cite{Sch96}, is left for a future publication. A crucial feature of the Gervais-Neveu approach is the introduction of a second free field, related to the one above by a quantum canonical transformation \cite{GN832}. As pointed out in ref. \cite{Sch96}, this is of direct relevance for the problem of constructing negative (integer) powers of screening operators: Negative powers of screening operators in terms of $X$ are nothing else than positive powers of screening operators in terms of $\tilde X$, the second free field. Therefore, there is no need of any analytic continuation procedure to define negative screenings, and the corresponding ambiguities are absent from the start. Concretely, we have (cf. ref. \cite{Sch96}) \ben S^{-1}(\sigma)=\tilde S(\sigma)\frac{1}{I^{({\hf})}_{\hf}(\varpi+1) I^{({\hf})}_{\hf}(-\varpi-1)} \label{ss} \een where $\tilde S$ is the screening operator constructed from $\tilde X$. We remark here that, $X$ and $\tilde X$ being completely equivalent, the replacement of one by the other must not change any physical observable. {}From the quantum group point of view, this means that observables are invariant not only under infinitesimal $U_q(sl(2))$ transformations, but also under Weyl reflections \cite{BG89,ANPS94}. As discussed in \cite{Sch96}, for Liouville exponentials outside the Kac table ($2J\ne 0,1,2,...$), this is a highly non-trivial condition which is in conflict with naive charge conservation rules familiar from the standard Coulomb gas picture. It serves, in fact, as an important guideline for the correct {\em non-perturbative} evaluation of matrix elements of the Liouville exponentials as constructed within the Gervais-Neveu approach. \section{The Braiding Matrix} The braiding matrix $R$ describes the braiding of the two chiral fields $U_{m}^{(J)}(\sigma)$ and $U_{m'}^{(J')}(\sigma')$ \ben U_{m}^{(J)}(\sigma)U_{m'}^{(J')}(\sigma')=\sum_{n_1,n_2} R(J,J';\varpi)_{n,n'}^{n_2,n_1}U_{m_2}^{(J')}(\sigma')U_{m_1}^{(J)}(\sigma) \label{Rdef} \een The ordering of $\sigma$ and $\sigma'$ is implicit in this definition of $R$ and we shall deal with the case $0<\sigma<\sigma'<2\pi$ explicitly, reserving the notation $\overline{R}$ for the opposite ordering of $\sigma$ and $\sigma'$. The sums extend over the integers \ben n_1=J+m_1\spa n_2=J'+m_2 \een the ranges of which we shall discuss below. The parameters are subject to the condition \ben m_1+m_2=m+m' \een so that there is really just one sum in Eq. (\ref{Rdef}). In general, the combinations \ben n=J+m\spa n'=J'+m' \een are related to the screening numbers and have mainly been treated as {\em non-negative} integers in the literature. In particular, the braiding matrix for non-negative integers $n$ and $n'$ and arbitrary spins is known to be \cite{GS93} \bea &&R(J,J';\varpi)_{n,n'}^{n_2,n_1}=\exp\left\{-i\pi(\D_x+\D_{x+m+m'} -\D_{x+m_2}-\D_{x+m})\right\}q^{2nJ'-2n_2J}\nn &\cdot&\q{2x-2J'+2n_2+1}\binomial{n}{n_1}\nn &\cdot&\frac{\q{2x+n_2+2}_{n_1} \q{2x-2J-2J'+n+n_2+1}_{n'}\q{2J'-n_2+1}_{n-n_1}}{\q{2x-2J'+n_2+1}_{n+n'+1}}\nn &\cdot&{}_4F_3\left(\begin{array}{l}-2J+n,\ -2J'+n_2,\ -n_1,\ -n'\\ -2x-n-n'-1,\ n-n_1+1,\ 2x-2J-2J'+n+n_2+1 \end{array};\ q,\ 1 \right)\nn &=&\exp\left\{-i\pi(\D_x+\D_{x+m+m'} -\D_{x+m_2}-\D_{x+m})\right\}q^{2nJ'-2n_2J}\nn &\cdot&\q{2x-2J'+2n_2+1}\binomial{n}{n_1}\nn &\cdot&\frac{\q{2x+n_2+2}_{n_1} \q{2x-2J-2J'+n+n_2+1}_{n'}\q{2J'-n_2+1}_{n-n_1}}{\q{2x-2J'+n_2+1}_{n+n'+1}}\nn &\cdot&\sum_{l=0}^{\infty}\frac{\q{-2J+n}_l\q{-2J'+n_2}_l\q{-n_1}_l\q{-n'}_l}{ \q{l}!\q{-2x-n-n'-1}_l\q{n-n_1+1}_l\q{2x-2J-2J'+n+n_2+1}_l} \label{R} \eea with $x$ defined by $\varpi=2x+1+{\pi\over h}$. Here, the standard notation for the $q$-hypergeometric function has been used \bea {}_4F_3\left(\begin{array}{llll}a,&b,&c,&d\\ e,&f,&g&{} \end{array} ;\ q,\ \rho \right)&=&\sum_{n=0}^\infty\frac{\q{a}_n\q{b}_n\q{c}_n\q{d}_n}{ \q{e}_n\q{f}_n\q{g}_n\q{n}!}\rho^n\nn \q{a}_0=1\spa\q{a}_n&=&\q{a}\q{a+1}...\q{a+n-1} \label{F} \eea It should be noted that in the standard case ($n$ and $n'$ non-negative integers) the summation in (\ref{R}) truncates after a {\em finite} number of terms. We shall use the convention $\q{y}=\sin(yh)/\sin(h)$. A $q$-hypergeometric function (\ref{F}) is balanced if $g=a+b+c+d-e-f+1$ and Saalschutzian if furthermore $d$ (or equivalently $a$, $b$ or $c$) is a non-positive integer. A $q$-deformed Saalschutzian ${}_4F_3$ hypergeometric function satisfies the standard transformation (ST) rule \cite{GR94} \bea {}_4F_3\left(\begin{array}{l}a,\ b,\ c,\ d\\ e,\ f,\ g\end{array};\ q,\ 1\right) &=&\frac{\q{f-c}_{-d}\q{e+f-a-b}_{-d}}{\q{f}_{-d}\q{e+f-a-b-c}_{-d}}\nn &\cdot&{}_4F_3\left(\begin{array}{l}e-a,\ e-b,\ c,\ d\\ e,\ e+f-a-b,\ c+d-f+1 \end{array};\ q,\ 1\right) \label{ST} \eea The present work is devoted to a discussion of the braiding matrix for the case where the ingoing screening numbers are {\em arbitrary integers}. We will assume that the outgoing screening numbers are also integers - see below for further comments on this point. As the braiding matrix for non-negative integer screenings Eq. (\ref{R}) is expressed in terms of a $q$-hypergeometric function, the most naive procedure for an extension to negative screenings would be to simply replace one or both of the ingoing screening numbers $n,n'$ by negative integers and simultaneously allow the outgoing screening numbers to be arbitrary integers. However, this would in general lead to $q$-hypergeometric sums that do not truncate, and such sums may well diverge for $|q|=1$ since the denominators of individual terms cannot be bounded away from zero. The remedy is given by employing ST (\ref{ST}) {\em before} continuing the ingoing screening numbers after which a continuation results in a {\em finite} and well defined expression, as will be discussed below. We thus arrive at an explicit proposal for the $R$-matrix, which will subsequently be verified. \subsection{Analytic Continuation} {}From a mathematical point of view, the extension of the explicit expression (\ref{R}) to arbitrary integer screenings is certainly not unique as we are trying to analytically continue a function given only on a discrete set of points. Let us make a division into subcases characterized by the signs of the ingoing screening numbers and treat them one by one. The expressions for the $R$-matrices we obtain should of course only be considered as an ansatz. The evidence for their validity will be given in the following sections. \subsubsection{{\bf I}: Positive-Positive Case, $n,n'\geq0$} It is well known \cite{GS93} that the set of chiral vertex operators corresponding to non-negative screening numbers closes under braiding, ensuring that both of the outgoing operators have non-negative screening numbers. It is easily verified algebraically that precisely then is the $R$-matrix (\ref{R}) non-vanishing and has no poles in the limits where $2J$ or $2J'$ becomes an integer\footnote{Such an analysis is similar in spirit to the discussion of fusion rules in ref. \cite{PRY}.}. The finite summation range for the truncating $q$-hypergeometric function is given explicitly by \ben max(0,n_1-n)\leq l\leq min(n_1,n') \label{selI} \een and $n_1,n_2\geq0$. This analysis suggests that a more natural expression exists for $n_1>n$, which allows to avoid cancellations of the form $\q{-1}!/\q{-1}!$. Let us introduce the notation $R_>^I$ and $R_\leq^I$ for the type {\bf I} $R$-matrix ($n,n'\geq0$) in the cases $n_1>n$ and $n_1\leq n$, respectively. By construction, $R_\leq^I$ is given by (\ref{R}) whereas $R_>^I$ is naturally represented as \bea &&R_>^I(J,J';\varpi)_{n,n'}^{n_2,n_1}=\exp\left\{-i\pi(\D_x+\D_{x+m+m'} -\D_{x+m_2}-\D_{x+m})\right\}q^{2nJ'-2n_2J}\nn &\cdot&\q{2x-2J'+2n_2+1}\binomial{n'}{n_1-n}\nn &\cdot&\frac{\q{2x+n_2+2}_n\q{2x-2J-2J'+n+n'+1}_{n_2}\q{-2J+n}_{n_1-n}}{ \q{2x-2J'+n_2+1}_{n+n'+1}}\nn &\cdot&{}_4F_3\left(\begin{array}{l}-2J+n+n'-n_2,\ -2J'+n',\ -n,\ -n_2\\ 2x-2J-2J'+n+n'+1,\ -2x-n-n_2-1,\ n'-n_2+1 \end{array};\ q,\ 1\right) \label{RI>} \eea This expression may be obtained either by shifting the summation over $n_1-n\leq l\leq n'$ to $0\leq l\leq n_2$, or by a repeated use of ST (\ref{ST}). In Appendix A some useful techniques based on ST are discussed and the present situation is presented as an illustration. \subsubsection{{\bf II}: Negative-Positive Case, $n<0\leq n'$} Inspection of Eqs. (\ref{R}) and (\ref{RI>}) shows that $R^{II}$ may be represented by exactly the same mathematical expressions as $R^I$ when in the latter $n$ has been continued to negative integers: \ben R^{II}_\leq=R^I_\leq\spa R^{II}_>=R^I_> \label{II=I} \een For the second equality to make sense, it is crucial that Eq. (\ref{RI>}) vanishes term by term for $n_2<0$ due to the binomial prefactor. $R^{II}_\leq$ also vanishes in this case. An alternative way to obtain Eq. (\ref{II=I}) is the following: In order to have a well defined and non-vanishing $R$-matrix for $n<0\leq n'$ and $2J$ and $2J'$ non-integer, we find that the summation variable $l$ in Eq. (\ref{R}) must satisfy \ben 0\leq l\leq min(n_1-n-1,n')\ \ {\mbox{or}} \ \ max(0,n_1-n)\leq l\leq n' \een It should be noted that there is no overlap between these two regimes. The first regime is excluded by our second demand that the $R$-matrix must be well defined in the limit where $2J'$ is an integer (analyzing the limit where $2J$ is an integer does not provide new information). This leaves us with \ben max(0,n_1-n)\leq l\leq n'\ \ \rightarrow\ \ 0\leq n_2 \label{selII} \een The last inequality expresses a selection rule preserving the non-negativity of the screening number under braiding (from the left) with a negative screening vertex operator. \subsubsection{{\bf III}: Positive-Negative Case, $n'<0\leq n$} In complete analogy with the analysis of case {\bf II} we find that the type {\bf III} $R$-matrix may be represented by the following single surviving range for the summation variable $l$ (\ref{R}) \ben max(0,n_1-n)\leq l\leq n_1\ \ \rightarrow\ \ 0\leq n_1 \label{selIII} \een Accumulating the information from cases {\bf I}, {\bf II} and {\bf III}, we observe that the property of non-negativity of a screening number is preserved under any braiding; see also Section 4. \subsubsection{{\bf IV}: Negative-Negative Case, $n,n'<0$} In the final case where both $n$ and $n'$ are negative integers our construction goes as follows. We shall initially focus on the $q$-hypergeometric part of $R^I$ which will be denoted $I$. Employing ST once results in the rewriting \bea I&=&{}_4F_3\left(\begin{array}{l}-2J+n,\ -2J'+n_2,\ -n-n'+n_2,\ -n'\\ -2x-n-n'-1,\ -n'+n_2+1,\ 2x-2J-2J'+n+n_2+1 \end{array};\ q,\ 1\right)\nn &=&\frac{\q{-2x+2J'-n-n'-n_2-1}_{n'}\q{-2x+2J-n-n'}_{n'}}{\q{-2x-n-n'-1}_{n'} \q{-2x+2J+2J'-n-n'-n_2}_{n'}}\nn &\cdot&{}_4F_3\left(\begin{array}{l}2J-n-n'+n_2+1,\ n+1,\ -2J'+n_2,\ -n'\\ -n'+n_2+1,\ -2x+2J-n-n',\ 2x-2J'+n+n_2+2 \end{array};\ q,\ 1\right) \label{Ianal} \eea It should be stressed that a possible pole or zero in $I$ which is matched by an appropriate zero or pole, respectively, in the prefactors of $R^I$, must also be present in the right hand side. In this sense, the two expressions are completely equivalent. Now, the right hand side is also well defined after the substitution $n_i \rightarrow-n_i-1$. This is the core of our analytic continuation and we {\em define} the object $I_{-n-1.-n'-1}^{-n_2-1,-n_1-1}$ by the right hand side in (\ref{Ianal}) after the substitution. The full type {\bf IV} $R$-matrix is obtained by multiplying $I_{-n-1.-n'-1}^{-n_2-1,-n_1-1}$ by the original prefactors in $R^I$ in which the substitution $n_i\rightarrow -n_i-1$ has been performed. The shift by $-1$ in the substitution $n_i\rightarrow -n_i-1$ is convenient because $R^{IV}$ is defined only for purely negative ingoing screening numbers. Nevertheless, let us state our proposal in the form \bea &&R^{IV}(J,J';\varpi)_{-n,-n'}^{-n_2,-n_1} =\exp\left\{-i\pi(\D_x+\D_{x+m+m'} -\D_{x+m_2}-\D_{x+m})\right\}q^{-2nJ'+2n_2J}\nn &\cdot&\q{2x-2J'-2n_2+1}\binomial{-n}{-n_1}\nn &\cdot&\frac{\q{2x-n_2+2}_{n_2-1} \q{2x-2J-2J'-2n-n'-n_2+2}_{n-1}\q{2J'+n_2+1}_{n'-n_2}}{ \q{2x-2J-2n-n'+2}_{n+n'-1}}\nn &\cdot&{}_4F_3\left(\begin{array}{l}-2J-n,\ -2J'-n_2,\ 1-n_2,\ 1-n\\ -2x,\ n'-n_2+1,\ 2x-2J-2J'-2n-n'-n_2+2 \end{array};\ q,\ 1 \right) \label{RIV} \eea where an additional ST has been employed. Here $n$ and $n'$ are {\em positive} integers and $-n_i=J_i+m_i$. \\[.2cm] In conclusion, our two analytic continuation procedures result in the same expressions for the braiding matrices. In case {\bf IV}, though, only one approach seems to apply directly. In \cite{RS} we shall return to the question of infinite sum representations of $q$-hypergeometric functions and braiding matrices. We have seen that all four types may be constructed using well defined concatenations of ST followed by trivial continuations of one or two of the ingoing screening numbers. Thus, the approach based on repeated use of ST demonstrates the {\em universality} of the mathematical expression for the original braiding matrix. Of course, this statement pertains to our proposal only. In the following sections we shall argue for the validity of this proposal. \subsection{Weyl Reflection Symmetry} In this section we shall perform a first test on our proposal by comparing it with the braiding matrices obtainable by Weyl reflection symmetry from the well known type {\bf I}. Since Weyl reflections act on the screening numbers as $n_i\rightarrow 2J_i-n_i$, subclasses of all four types are reachable, depending on the signs of $2J-n$ and $2J'-n'$. As we are not addressing the question of non-integer screening numbers we shall restrict ourselves to the case $2J,2J'\in\mbox{$Z\hspace{-2mm}Z$}$. The explicit relation between $\tilde{S}^n$ and $S^{-n}$ is \bea \tilde{S}^n&=&S^{-n}K_n(\varpi)\nn K_n(\varpi)&=&K_1(\varpi+2(n-1))K_1(\varpi+2(n-2))...K_1(\varpi) \eea where $K_1(\varpi)$ is given by the normalization factor in the right hand side of Eq. (\ref{ss}) \ben K_1(\varpi)=I^{(\hf)}_\hf(\varpi+1)I^{(\hf)}_\hf(-\varpi-1) \een Weyl reflection dictates that \bea R(J,J';\varpi)^{-k_2,-k_1}_{-n,-n'}&=& R(J,J';-\varpi)^{k_2,k_1}_{n,n'} \frac{I^{(J)}_J(\varpi)I^{(J')}_{J'}(\varpi+ 2J-2n)}{I^{(J')}_{J'}(\varpi)I^{(J)}_J(\varpi+2J'-2k_2)}\nn &\cdot&\frac{K^{(J')}_{k_2}(\varpi+2J'-2k_2)K_{k_1}(\varpi+2J+2J'-2k_2-2k_1) }{K_n(\varpi+2J-2n)K_{n'}(\varpi+2J+2J'-2n-2n')} \eea and one may show that this reduces to \ben q^{2nJ'-2n_2J}R(J,J';-\varpi)_{2J-n,2J'-n'}^{2J'-n_2,2J-n_1} =q^{-2nJ'+2n_2J}R(J,J';\varpi)_{n,n'}^{n_2,n_1} \label{ref} \een This relation may be seen as boundary conditions on our proposal imposed by the reflection symmetry. Due to the universality of the {\em mathematical} structure of the braiding matrix, it is almost straightforward to verify that our proposal satisfies (\ref{ref}). In principle, one should distinct between the four resulting types of braiding matrices. However, we know from Appendix A that Weyl reflections for $2J,2J'\in\mbox{$Z\hspace{-2mm}Z$}$ and (concatenations of) the STs we employ commute so it is sufficient to consider only one type. In the representation of type {\bf I} we have \bea &&q^{2nJ'-2n_2J}R^I(J,J';-\varpi)_{2J-n,2J'-n'}^{2J'-n_2,2J-n_1}\nn &=&\exp\left\{-i\pi(\D_x+\D_{x+m+m'} -\D_{x+m_2}-\D_{x+m})\right\}\q{2x-2J'+2n_2+1}\binomial{2J-n}{2J-n_1}\nn &\cdot&\frac{\q{2x-2J-2J'+n+n'+1}_{2J-n_1} \q{2x-2J'+n+n'+n_2+2}_{2J'-n'}\q{n_2+1}_{n'-n_2}}{ \q{2x-2J-2J'+n+n'+n_2+1}_{2J+2J'-n-n'+1}}\nn &\cdot&{}_4F_3\left(\begin{array}{l}-2J+n_1,\ -2J'+n',\ -n_2,\ -n\\ -2x-n-n_2-1,\ n'-n_2+1,\ 2x-2J-2J'+n+n'+1 \end{array};\ q,\ 1\right)\nn &=&q^{-2nJ'+2n_2J}R^I(J,J';\varpi)_{n,n'}^{n_2,n_1} \eea where the last equality is due to a concatenation of STs. \section{Double Braiding Relations for the Type II and III $R$-matrices} In this section we shall set up simple relations for the $R$-matrices using one of the polynomial equations of Moore and Seiberg \cite{MS}, namely the commutativity of fusion and braiding. The Moore-Seiberg equations were originally set up within the context of rational conformal field theory, but can be viewed as a set of consistency relations for conformal field theory in general. For the present operator algebra, which extends the standard Coulomb gas in a rather non-trivial fashion, their validity may not seem self-evident - especially since both fusion and braiding matrices involve an infinite number of conformal blocks. We will take the commutativity of fusion and braiding as an axiomatic starting point of our analysis, though in principle our explicit operatorial construction should allow to discuss its validity. In any such discussion, manipulations with infinite sums over conformal operators will necessarily arise and one may speculate that the Moore-Seiberg equations will act as a defining principle for their treatment; clearly, further analysis will be necessary to elucidate this point. There are several possibilities for choosing an appropriate equation system to study. It turns out that a convenient choice is the recursive system to be analyzed in the following. \subsection{Recursive Equation Systems} Let us first consider the product \ben U_{-m_1}^{(-J_1)}(\sigma)\odot U_{m_2}^{(J_2)}(\sigma)U_{m_3}^{(J_3)}(\sigma') \een (cf. Eq. (\ref{odotdef})) and braid through from the right the operator $U_{m_3}^{(J_3)}(\sigma')$. Demanding that the above expression can be evaluated either by braiding first and then fusing the operators at $\sigma$, or vice versa, allows us to derive the double braiding relation \bea &&\sum_{l=0}^{n_2+n_3} \sum_k \, q^{2J_2(n_1-k)} R(J_2,J_3;\varpi-2m_1)_{n_2,n_3}^{n_2+n_3-l,l} R(-J_1,J_3;\varpi)_{-n_1,n_2+n_3-l}^{-n_1+n_2+n_3-l+k,-k}\nn &\cdot&U_{m_3+n_2-n_1-l+k}^{(J_3)}(\sigma')U_{-(J_2-J_1)+l-k}^{(-J_1+J_2)} (\sigma) \nn &=&\sum_iR(-J_1+J_2,J_3;\varpi)_{-n_1+n_2,n_3}^{-n_1+n_2+n_3-i,i} U_{n_2-n_1+m_3-i}^{(J_3)}(\sigma')U_{i-(J_2-J_1)}^{(-J_1+J_2)}(\sigma) \label{n1n2n3} \eea $n_1$ is a positive integer while $n_2$ and $n_3$ are non-negative ones. The sums over $k$ and $i$ could in principle contain non-integer values or even an integration, since they are related to outgoing screening numbers about which we have no a priori knowledge as soon as we leave the safe grounds of the standard Coulomb gas with positive screenings only. In Appendix B, we present an argument involving degenerate field techniques which suggests that non-integer values may actually occur, with the corresponding $R$-matrix elements decoupling from those for integer outgoing screenings. This is quite puzzling, as one would expect our explicit operator construction to select one particular $R$-matrix, removing the ambiguity. However, braiding problems with negative ingoing screenings involve infinite sums over products of outgoing vertex operators, the evaluation of which is expected to be delicate \cite{Sch96}; in particular, several equivalent representations with numerically different $R$-matrices may exist, as is suggested by the argument in Appendix B. These questions, though clearly important, go beyond the scope of this first analysis of the braiding problem with negative screenings. We will therefore discuss here only the $R$-matrix elements for integer outgoing screenings. We consider the case $n_3=0$ in Eq. (\ref{n1n2n3}) for which we may deduce the equation system\\[.2cm] {\bf Triangular Equation System} \bea &&\sum_{l=0}^{n_2}q^{2J_2(n_1+L-l)} R(J_2,J_3;\varpi-2m_1)_{n_2,0}^{n_2-l,l} R(-J_1,J_3;\varpi)_{-n_1,n_2-l}^{-n_1+n_2-L,L-l}\nn &=&R(-J_1+J_2,J_3;\varpi)_{-n_1+n_2,0}^{-n_1+n_2-L,L} \label{tri} \eea by comparison of coefficients of like operators. For integer outgoing screenings, $L$ is integer. Note that in any case there is no coupling between matrix elements for integer and non-integer outgoing screenings. Due to the ``triangular'' structure of this equation system, all braiding matrices of type {\bf II} are determined recursively in terms of $R$-matrices of type {\bf I} and the subclass of type {\bf II} consisting of $R$-matrices of the form $R(J,J';\varpi)_{-n,0}^{n_2,-n_1}$. However, the latter may be determined by an additional and independent equation system and a selection rule. Consider the product \ben U_{n/2}^{(n/2)}(\sigma)\odot S^{-n}(\sigma)V_{-J'}^{(J')}(\sigma') \een and braid through from the right the operator $V_{-J'}^{(J')}(\sigma')$. One obtains the double braiding relation \bea &&\sum_{n_1,n_1'}R(0,J';\varpi+n)_{-n,0}^{n_1-n,-n_1} R(n/2,J';\varpi)_{n,n_1-n}^{n_1-n-n_1',n_1'}\nn &\cdot&V_{-J'}^{(J')}(\sigma') S^{n_1-n-n_1'}(\sigma')V_{-n/2}^{(n/2)}(\sigma)S^{n_1'-n_1}(\sigma)\nn &=&e^{ihnJ'}V_{-J'}^{(J')}(\sigma')V_{-n/2}^{(n/2)}(\sigma) \eea Now we may use that degenerate fields remain degenerate fields after braiding \cite{GN84} and we derive the equation system\\[.2cm] {\bf Initial Equation System} \ben e^{-ihnJ'}\sum_{l=0}^n \, R(n/2,J';\varpi)_{n,K+l-n}^{K,l} R(0,J';\varpi+n)_{-n,0}^{K+l-n,l-K}=\delta_{K,0} \label{ini1} \een which we shall denote the ``initial'' one. This equation system fixes the remaining indeterminacy, if we impose the additional selection rule\\[.2cm] {\bf Selection Rule} \bea R(J_1,J_2;\varpi)_{-n, n'}^{n_2,n'-n-n_2}=0 \ \ \ \ \mbox{for}\ \ n>0,\ n'\geq0,\ n_2<0\nn R(J_1,J_2;\varpi)_{n,-n'}^{n-n'-n_1,n_1}=0 \ \ \ \ \mbox{for}\ \ n\geq0,\ n'>0,\ n_1<0 \label{selrule} \eea All screening numbers are taken to be integers. The selection rule means that positive screenings always remain positive screenings after braiding. This rule is certainly fulfilled by our analytic continuation of the $R$-matrix as already discussed, but we do not have, at present, an a priori derivation. Its validity can be verified in the case where one of the ingoing vertex operators is degenerate, cf. Appendix B. It is an interesting question whether it can be deduced in general by invoking other polynomial equations, or by using more information about the explicit structure of the negative screening vertex operators. If we consider the case of integer $K$ and take into account the selection rule, we can rewrite Eq. (\ref{ini1}) as \ben e^{-ihnJ'}\sum_{l=max(0,K-n)}^KR(n/2,J';\varpi)_{n,l}^{K,n+l-K} R(0,J';\varpi+n)_{-n,0}^{l,-n-l}=\delta_{K,0} \label{ini} \een This equation system now determines recursively all remaining unknowns, if we consider successively $K=0,1,2...$. We will now outline the proof that our explicit proposal is the unique solution of Triangular and Initial equation systems plus selection rule, if only integer outgoing screenings are admitted. The argument is based on certain new identities for $q$-hypergeometric functions and generalizations thereof. Detailed proofs of these identities will be presented elsewhere \cite{RS}. \subsection{Verification of Proposal} Here we shall verify that the analytically continued expression (\ref{R}) subject to (\ref{selII}) provides a solution to the two double braiding relations discussed above. We first consider the triangular equation system (\ref{tri}). Insertion of our proposal yields the equation system \bea &&\sum_{l=0}^{n_2}\q{2x+2J_1-2J_3-2n_1+2n_2-2l+1}\binomial{n_2}{l} \tilde{R}(-J_1,J_3;\varpi)_{-n_1,n_2-l}^{-n_1+n_2-L,-l+L}\nn &\cdot&\frac{\q{2x+2J_1-2n_1+n_2-l+2}_l\q{2J_3-n_2+l+1}_{n_2-l}}{ \q{2x+2J_1-2J_3-2n_1+n_2-l+1}_{n_2+1}}\nn &=&\Theta(-n_1+n_2-L\geq0)\q{2x-2J_3-2n_1+2n_2-2L+1}(-1)^{L} \frac{\q{n_1-n_2}_{L}}{\q{L}!}\nn &\cdot&\frac{\q{2x-n_1+n_2-L+2}_L\q{2J_3+n_1-n_2+L+1}_{-n_1+n_2-L}}{ \q{2x-2J_3-n_1+n_2-L+1}_{-n_1+n_2+1}} \label{n1n20} \eea where \bea &&\tilde{R}(-J_1,J_3;\varpi)_{-n_1,n_2-l}^{-n_1+n_2-L,-l+L}\nn &=& \q{2x-2J_3-2n_1+2n_2-2L+1}(-1)^{L-l}\frac{\q{n_1}_{L-l}}{\q{L-l}!} \q{2x-n_1+n_2-L+2}_{L-l}\nn &\cdot&\frac{\q{2x+2J_1-2J_3-2n_1+n_2-L+1}_{n_2-l} \q{2J_3+n_1-n_2+L+1}_{-n_1-L+l}}{\q{2x-2J_3-n_1+n_2-L+1}_{-n_1+n_2-l+1}}\nn &\cdot&\sum_{i=max(0,n_1+L-l)}^{n_2-l}\frac{\q{2J_1-n_1}_{i}\q{-2J_3-n_1 +n_2-L}_{i}}{\q{i}!\q{-2x+n_1-n_2+l-1}_{i}}\nn &\cdot&\frac{\q{-L+l}_{i}\q{l-n_2}_i}{\q{-n_1-L+l+1}_{i} \q{2x+2J_1-2J_3-2n_1+n_2-L+1}_{i}} \label{hat} \eea $\tilde{R}$ is merely the $R$-matrix (\ref{R}) subject to (\ref{selII}) without the explicit phases and powers of $q$ which cancel out their analogues in the other $R$-matrices. Our proof will be by induction in $n_2$ and it is recalled that $n_1\geq1$. Introduce the integer \ben N=-n_1+n_2-L \een Due to the selection rule, the equation system becomes trivial for $N<0$, so in the following we may assume that $N\geq0$. It is easily verified that the Triangular equation system (\ref{n1n20}) is satisfied for the initial value $n_2=0$. For general $n_2$, the left hand side in (\ref{n1n20}) will be denoted ${\cal L}$ whereas the right hand side will be denoted ${\cal R}$. The finite double summation in ${\cal L}$ may be expressed as \ben {\cal L}=\sum_{j\geq0}S_j \label{sumS} \een where \bea &&S_j\nn &=&(-1)^{n_2}\q{2x-2J_3+2N+1}\frac{\q{n_2}!\q{n_1+N-j-1}! \q{2J_3-n_2-N+j+1}_{n_2+N-j}}{\q{n_2-j}! \q{n_1-1}!\q{j}!\q{N-j}!\q{-2x-N-1}_{n_1+N-j}}\nn &\cdot&\frac{\q{-2J_1+n_1-n_2+j+1}_{n_2-j}}{ \q{-2x-2J_1+2J_3+2n_1-2n_2-1}_{n_2+1} \q{-2x+2J_3+n_1-n_2-N-1}_{-n_1+n_2+1}}\nn &\cdot&\sum_{l=0}^{n_2-j}\q{2x+2J_1-2J_3-2n_1+2n_2-2l+1} \q{-2x-2J_1+2J_3+n_1-n_2-N+l}_j\nn &\cdot&\frac{\q{-n_2+j}_l \q{-2x-2J_1+2n_1-n_2-1}_l}{\q{l}!\q{-2J_1+n_1-n_2+j+1}_l}\nn &\cdot&\frac{\q{-2x-2J_1+2J_3+2n_1-2n_2-1}_l\q{-2x+2J_3+n_1-n_2-N-1}_l}{ \q{2J_3-n_2-N+j+1}_l\q{-2x-2J_1+2J_3+2n_1-n_2}_l}\nn \eea We also introduce \ben S_j(\left\{-p\right\})=S_j(2x+p,2J_1,2J_3-p,n_1,n_2-p,N-p) \label{shiftS} \een where the parameters $n_2,\ N,\ -2x$ and $2J_3$ all have been subtracted $p$ while the remaining parameters $n_1$ and $2J_1$ are left unchanged. ${\cal R}(\left\{-p\right\})$ is defined analogously where the unshifted ${\cal R}$ is given explicitly by \ben {\cal R}=-\frac{\q{2x-2J_3+2N+1}\q{n_1-n_2+N-1}!\q{2J_3-N+1}_N \q{-2x+2J_3-N}_{n_1-n_2-1}}{\q{N}!\q{n_1-n_2-1}!\q{-2x-N-1}_{n_1-n_2+N}} \label{calR} \een We shall make a case distinction and consider $N=0$ first in which case \ben {\cal L}=S_0 \een {\bf Lemma 1} \ben S_0={\cal R}(2x,2J_1,2J_3,n_1,n_2,N=0) \label{1} \een {}\\[.2cm] For $N>0$, let $M_k$ denote the following sum \bea M_k&=&(-1)^{n_2}\q{2x-2J_3+2N+1}\frac{\q{n_1+N-1}! }{\q{n_1-1}!\q{k}!\q{N}\q{N-k-1}!}\nn &\cdot&\frac{\q{2J_3-n_2-N+k+1}_{n_2+N-k}\q{-2J_1+n_1-n_2+k+1}_{n_2-k}}{ \q{-2x-N-1}_{n_1+N-k}\q{-2x-2J_1+2J_3+2n_1-2n_2+k-1}_{n_2-k+1}}\nn &\cdot&\frac{1}{ \q{-2x+2J_3+n_1-n_2-N-1}_{-n_1+n_2+1}}\nn &\cdot&\sum_{l=0}^{n_2-k}\q{2x+2J_1-2J_3-2n_1+2n_2-2l+1-k}\nn &\cdot&\frac{ \q{-n_2+k}_l\q{-2x-2J_1+2n_1-n_2-1}_l}{\q{l}!\q{-2J_1+n_1-n_2+k+1}_l}\nn &\cdot&\frac{\q{-2x-2J_1+2J_3+2n_1-2n_2+k-1}_l\q{-2x+2J_3+n_1-n_2-N-1}_l}{ \q{2J_3-n_2-N+k+1}_l\q{-2x-2J_1+2J_3+2n_1-n_2}_l} \label{Mk} \eea where $M_0=S_0$. The following lemma determines a relation between $S_j$ and $M_k.$\\[.2cm] {\bf Lemma 2}\\[.2cm] For $k<n_2$ we have \ben M_k-M_{k+1}=\sum_{m=0}^{k+1}A_k^mS_m(\left\{-k-1+m\right\}) \label{Mk'} \een where \ben A_k^m=(-1)^{k+m}\frac{\q{n_1-n_2+N-1}_{k+1-m} \q{2J_3-k+m}_{k+1-m}}{\q{k+1-m}!\q{-2x+2J_3-N-k-1+m}_{k+1-m}} \label{A} \een and in particular $A_k^{k+1}=-1$.\\[.2cm] It then follows that \ben {\cal L}=\sum_{k=0}^{min(N,n_2)-1} \sum_{m=0}^kA_k^mS_m(\left\{-k-1+m\right\})+\Theta(n_2<N)M_{n_2} \een from which (\ref{n1n20}) may be deduced, thus completing the proof of the Triangular equation system (\ref{tri}). Now we turn to the Initial equation system (\ref{ini}). Insertion of our proposal yields the equation system \bea &&\delta_{K,0}\nn &=&\q{2x-2J'+2K+1}(-1)^n\frac{\q{-n}_{n-K}\q{2x+K+2}_{n-K} \q{-2J'}_K\q{2x-2J'+2}_{n-1}}{\q{n-K}!\q{2x-2J'+K+1}_{n+1}\q{2x+2}_n}\nn &\cdot&\sum_{l=max(0,K-n)}^K \q{2x-2J'+n+2l+1}\frac{\q{-K}_l\q{2x-2J'+K+1}_l}{\q{l}!\q{n-K+1}_l}\nn &\cdot&\frac{\q{n}_l\q{2x-2J'+n+1}_l}{\q{2x-2J'+n+K+2}_l\q{2x-2J'+2}_l} \label{ini2} \eea It is immediate that the right hand side vanishes for $K<0$ and that it reduces to 1 for $K=0$. We may therefore assume $K>0$. After a case distinction into $0<K\leq n$ and $0\leq n<K$ the identity follows by induction in $K$ and $n$, respectively, thus completing the proof of the Initial equation system (\ref{ini}). Details on the proofs of both the Triangular and the Initial equation systems will appear elsewhere \cite{RS}. \subsection{The Type {\bf III} $R$-matrix} So far we have only verified our proposal for $R$-matrices of type {\bf II}. Type {\bf III} $R$-matrices may be obtained from the latter essentially by hermitian conjugation. We have \bea \left( V^{(J)}_{-J}(\sg)S^n(\sigma)V^{(J')}_{-J'}(\sg') S^{-n'}(\sigma') \right)^\dagger &=&S^{-n'\dagger}(\sg')V^{(J')\dagger}_{-J'}(\sigma') S^{n\dagger}(\sg)V^{(J)\dagger}_{-J}(\sigma)\nn &=&\tilde S^{-n'}(\sg')\tilde V^{(J')}_{-J'}(\sigma')\tilde S^n(\sg) \tilde V^{(J)}_{-J}(\sigma) \eea Here we have used that for real $\varpi$, the two free fields are just hermitian conjugates of each other \cite{GN832} ($\tilde V^{(J')}_{-J'}$ and $\tilde S$ denote operators constructed from the field $\tilde X$). The order of screening operators and free field exponentials can be reverted to the standard form at the expense of simple phase factors, so that \ben \left( V^{(J)}_{-J}(\sg)S^n(\sigma)V^{(J')}_{-J'}(\sg') S^{-n'}(\sigma') \right)^\dagger =q^{2(Jn-J'n')}\tilde V^{(J')}_{-J'}(\sg')\tilde S^{-n'}(\sigma') \tilde V^{(J)}_{-J}(\sg)\tilde S^n(\sigma) \label{conj1} \een The expression on the right hand side is of the type {\bf II} already discussed. Braiding the operators on the left hand side with the unknown type {\bf III} $R$-matrix, we obtain \ben R^*(J,J';\varpi-2n_1+2n_2+2(J+J'))_{n,-n'}^{-n_2,n_1}\ q^{2(Jn_1-J'n_2)} =q^{2(Jn-J'n')}{\overline R} (J',J;-\varpi)_{-n',n}^{n_1,-n_2} \een The bar on the $R$-matrix on the right hand side indicates that this $R$-matrix braids an operator at $\sigma'$ with an operator at $\sigma$ (cf. comment following Eq. (\ref{Rdef})), contrary to the one on the left hand side. Notice also the replacement $\varpi \to -\varpi$ in this $R$-matrix which reflects the fact that it applies to the braiding of ``tilded'' operators. The final result for the type {\bf III} braiding matrix thus becomes \ben R(J,J';\varpi)_{n,-n'}^{-n_2,n_1}=q^{-2[J(n-n_1)-J'(n'-n_2)]} {\overline R}^*(J',J; -\varpi-2n_1+2n_2+2(J+J'))_{-n',n}^{n_1,-n_2} \label{conj2} \een In the case of braiding of positive screening vertex operators, it is well known that ${\overline R}(J,J';\varpi)_{n,n'}^{n_2,n_1}$ differs from $R(J,J';\varpi)_{n,n'}^{n_2,n_1}$ only by a phase factor \cite{CGR941}, and it is evident from our analytic continuation procedure that the type {\bf II} $R$-matrix will keep this property. Therefore, we can write Eq. (\ref{conj2}) as a relation between $R$-matrices corresponding to the same position ordering: \bea R(J,J';\varpi)_{n,-n'}^{-n_2,n_1}&=&q^{-2[J(n-n_1)-J'(n'-n_2)]} e^{-2i\pi(\Delta_x +\Delta_{x+m+{m}'} -\Delta_{x+{m}_2}-\Delta_{x+m})}\nn &\cdot&R^*(J',J; -\varpi-2n_1+2n_2+2(J+J'))_{-n',n}^{n_1,-n_2} \label{conj3} \eea where ${m}'=-n'-J'$ and ${m}_2=-n_2-J'$. Eq. (\ref{conj3}) determines the type {\bf III} $R$-matrix in terms of the type {\bf II} one, which we have already verified. We show that our analytic continuation fulfills Eq. (\ref{conj3}) by considering separately the two cases $n\geq n_1$ and $n<n_1$. From our formula for the type {\bf II} $R$-matrix, we obtain for the right hand side (up to explicit phases and powers of $q$, cf. comment following Eq. (\ref{hat})) \bea &&\tilde{R}(J',J;-\varpi-2n_1+2n_2+2(J+J'))_{-n',n}^{n_1,-n_2}=\q{-2x+2J' +2n_2-1}\binomial{-n'}{-n_2}\nn &\cdot&\frac{\q{-2x+2J+2J'-n_1+2n_2}_{-n_2}\q{-2x-n+n_2-1}_n \q{2J-n_1+1}_{n_2-n'}}{\q{-2x+2J'-n_1+2n_2-1}_{n-n'+1}}\nn &\cdot&\sum_{l=max(0,n'-n_2)}^n\frac{\q{-2J'-n'}_l\q{-2J+n_1}_l \q{n_2}_l\q{-n}_l}{\q{l}!\q{2x-2J-2J'+n_1-n_2+}_l\q{n_2-n'+1}_l \q{-2x-n+n_2}_l} \label{Rstar} \eea In the first case $n\geq n_1$, the summation in (\ref{Rstar}) may be written \bea &&\sum_{l=n'-n_2}^n\ \rightarrow\ \frac{\q{-n}_{n'-n_2}\q{n_2}_{n'-n_2}}{ \q{n'-n_2}!\q{n_2-n'+1}_{n'-n_2}}\nn &\cdot&\frac{\q{-2J'-n'}_{n'-n_2}\q{-2J+n_1}_{n'-n_2}}{\q{2x-2J-2J'+n_1 -n_2+1}_{n'-n_2}\q{-2x-n+n_2-1}_{n'-n_2}}\nn &\cdot&\sum_{l=0}^{n_1}\frac{\q{-2J'-n_2}_l\q{-2J+n}_l\q{n'}_l\q{-n_1}_l}{ \q{l}!\q{n'-n_2+1}_l\q{2x-2J-2J'+n-n_2+1}_l\q{-2x-n+n'-1}_l} \eea and by re-inserting it we recognize the left hand side of Eq. (\ref{conj3}). In the second case $n<n_1$, the summation on the left hand side of Eq. (\ref{conj3}) may be written \bea &&\sum_{l=n_1-n}^{n_1}\ \rightarrow\ \frac{\q{-n_1}_{n_1-n}\q{n'}_{n_1-n}}{\q{n_1-n}! \q{n-n_1+1}_{n_1-n}}\nn &\cdot&\frac{\q{-2J+n}_{n_1-n}\q{-2J'-n_2}_{n_1-n}}{\q{-2x-n+n'-1}_{n_1-n} \q{2x-2J-2J'+n-n_2+1}_{n_1-n}}\nn &\cdot&\sum_{l=0}^n\frac{\q{-2J+n_1}_l\q{-2J'-n'}_l\q{-n}_l\q{n_2}_l}{\q{l}! \q{n_1-n+1}_l\q{-2x+n_2-n-1}_l\q{2x-2J-2J'+n_1-n_2+1}_l} \eea and by re-inserting it we recognize the expression (\ref{Rstar}). This concludes the verification of our proposal for the type {\bf III} $R$-matrix. \section{Connection with $6j$-symbols} The braiding matrix for vertex operators from the Kac table was shown in \cite{CGR941} to coincide with $q$-deformed $6j$-symbols for $U_q(sl(2))$, up to normalization factors that were determined explicitly. This result continues to be true for general positive screening vertex operators \cite{GS93,GR94}. We will not show here that our expression for the $R$-matrix in the presence of negative screenings can indeed be interpreted consistently as a $q$-deformed $6j$-symbol - this would require verifying all of the defining properties of the latter - but rather offer a most natural generalization of the standard relation between $R$-matrix and $6j$-symbol, as given in ref. \cite{GS94}: \ben R(J,J',\varpi)_{n,n'}^{n_2,n_1}= e^{-i\pi (\Delta_c+\Delta_b-\Delta_e-\Delta_f)} {\kappa_{ab}^e \kappa_{de}^c\over \kappa_{db}^f \kappa_{af}^c} \left\{ ^{a}_{d}\,^{b}_{c}\right. \left |^{e}_{f}\right\} \label{6j} \een The arguments of the $q$-deformed $6j$-symbol are \bea a=J,\quad b=x+m+m',\quad c=x\nn d=J',\quad e=x+m_2,\quad f=x+m \label{3.21} \eea and the coefficients $\kappa_{J_1 J_2}^{J_{12}}$ are given by \bea \kappa_{J_1 J_2}^{J_{12}}&=& \left ({ he^{-i(h+\pi)} \over 2\pi \Gamma(1+h/\pi) \sin h}\right )^{J_1+J_2-J_{12}} e^{ih (J_1+J_2-J_{12})( J_1-J_2-J_{12})} \nn &\cdot& \prod_{k=1}^{J_1+J_2-J_{12}} \sqrt{ \lfloor 1+2J_1-k\rfloor \over \lfloor k \rfloor \, \lfloor 1+2J_2-k \rfloor\, \lfloor -(1+2J_{12}+k) \rfloor } \label{3.9} \eea The last equation makes sense for arbitrary $J_1,J_2, J_{12}$ such that $J_1+J_2-J_{12}$ is a non-negative integer, but is readily extended to negative screening numbers if we define products with negative upper limits as usual by \begin{equation} \prod_{j=1}^{-n}f(j):={1\over \prod_{j=1}^n f(1-j)} \label{negprod} \end{equation} As our method provides us directly with the braiding matrix, rather than the $6j$-symbol, Eq. (\ref{6j}) should be read as our {\em definition} of (or proposal for) the $q$-deformed $6j$-symbol, extended to arbitrary integer screening numbers. An extension to the purely negative case where all (ingoing and outgoing) screening numbers are negative, has been provided in ref. \cite{GR94}. Here we shall show that our definition (\ref{6j}) based on our proposal for the braiding matrix, reproduces (and generalizes\footnote{In contrast to ref. \cite{GR94}, we do not assume that the outgoing screenings are negative.}) the result of \cite{GR94} in the purely negative case. The idea is to verify invariance of our proposal under the replacement $J_i\rightarrow-J_i-1$ (implying in particular that screening numbers are transformed according to $n_i\rightarrow-n_i-1$) \ben \left\{ ^{J}_{J'}\,^{x-J-J'+n+n'}_{x}\right. \left |^{x-J'+n_2}_{x-J+n}\right\}= \left\{ ^{-J-1}_{-J-1'}\,^{-x+J+J'-n-n'-1}_{-x-1} \right. \left |^{-x+J'-n_2-1}_{-x+J-n-1}\right\} \label{GR} \een which is the defining property for the extension in ref. \cite{GR94}. In order to do that let us write down explicitly the well known $q$-deformed $6j$-symbol \bea &&\left\{ ^{J}_{J'}\,^{x-J-J'+n+n'}_{x} \right. \left |^{x-J'+n_2}_{x-J+n}\right\}\nn &=&\left(\frac{\q{2J-n+1}_{n_2-n'}\q{2x-2J-2J'+n+n'+n_2+1}_{n-n_2}}{ \q{2J'-n_2+1}_{n_2-n'}\q{2x+n_2+2}_{n-n_2}}\right.\nn &\cdot&\left.\frac{\q{2x-2J'+n_2+1}_{n+n'+1}\q{2x-2J+2n+1}\q{n_2}!\q{n_1}!}{ \q{2x-2J+n+1}_{n+n'+1}\q{2x-2J'+2n_2+1}\q{n}!\q{n'}!}\right)^{\hf}\nn &\cdot&R^I(J,J';\varpi)_{n,n'}^{n_2,n_1} \label{6jexp} \eea and our proposal in the purely negative case \bea &&\left\{ ^{-J-1}_{-J-1'}\,^{-x+J+J'-n-n'-1}_{-x-1} \right. \left |^{-x+J'-n_2-1}_{-x+J-n-1}\right\}\nn &=&\left(\frac{\q{-2J+n}_{n'-n_2}\q{-2x+2J+2J'-n-n'-n_2}_{n_2-n}}{ \q{-2J'+n_2}_{n'-n_2}\q{-2x-n_2-1}_{n_2-n}}\right.\nn &\cdot&\left.\frac{\q{-2x+2J'-n_2}_{-n-n'-1}\q{-2x+2J-2n-1}\q{-n_1-1}! \q{-n_2-1}!}{ \q{-2x+2J-n}_{-n-n'-1}\q{-2x+2J'-2n_2-1}\q{-n-1}!\q{-n'-1}!}\right)^{\hf}\nn &\cdot&R^{IV}(-J-1,-J'-1;-\varpi)_{-n-1,-n'-1}^{-n_2-1,-n_1-1} \eea Employing ST six times, it is now straightforward to verify the invariance (\ref{GR}). In ref. \cite{GR94} it is observed that the $q$-deformed $6j$-symbol possesses the simple symmetry \ben \left\{ ^{J'}_{J}\,^{x-J-J'+n+n'}_{x} \right. \left |^{x-J+n}_{x-J'+n_2}\right\} =\left\{ ^{J}_{J'}\,^{x-J-J'+n+n'}_{x} \right. \left |^{x-J'+n_2}_{x-J+n}\right\} \label{6jsym} \een An explicit proof based on (\ref{6jexp}) is immediate, so let us conclude this section by stating the equivalence of (\ref{6jsym}) in terms of $R$-matrices \bea &&q^{2nJ'-2n_2J}R^I(J',J;\varpi)_{n_2,n_1}^{n,n'}\nn &=&\frac{\q{2x-2J+2n+1}\q{n_2}!\q{n_1}!\q{2x-2J-2J'+n+n'+n_2+1}_{n-n_2}}{ \q{2x-2J'+2n_2+1}\q{n}!\q{n'}!\q{2x+n_2+2}_{n-n_2}}\nn &\cdot&\frac{\q{2J-n+1}_{n_2-n'}\q{2x-2J'+n_2+1}_{n+n'+1}}{ \q{2J'-n_2+1}_{n_2-n'}\q{2x-2J+n+1}_{n+n'+1}} q^{-2nJ'+2n_2J}R^I(J,J';\varpi)_{n,n'}^{n_2,n_1} \eea \section{Construction of the Type {\bf IV} $R$-matrix by Concatenation} In this section we shall argue for the validity of our proposal for the type {\bf IV} $R$-matrix by explicit comparison with an expression obtained by a 3 step concatenation. Step 1 is a direct verification that our proposal for $J=J'=0$ is identical to the result of imposing Weyl reflection symmetry on the corresponding type {\bf I} $R$-matrix. Eq. (\ref{ref}) readily ensures that. Step 2 is then the generalization to $J'=0$ and $J$ generic, using the result of Step 1. Finally, Step 3 completes the generalization to $J$ and $J'$ generic. In both Step 2 and Step 3, we consider the braiding of $\vj\smn(\sigma)$ with $\vjpr\smnpr(\sigma')$, for $n,n'>0$. \subsection{Step 2} The $R$-matrix we want to establish is given by \ben \vj(\sg)\smn(\sg)\smnpr(\sg')=\sum_{n_2}R(J,0;\varpi)^{-n_2,-n_1}_{-n,-n'} S^{-n_2}(\sg')\vj(\sg)S^{-n_1}(\sg) \een where $n_1\equiv n+n'-n_2$. On the other hand we also have ($k_2\equiv n+n'-k_1$) \ben \vj(\sg)\smn(\sg)\smnpr(\sg')=\vj(\sg)\sum_{k_1} \, R(0,0;\varpi)^{-k_2,-k_1}_{-n,-n'}S^{-k_2}(\sg')S^{-k_1}(\sg) \een where $0\le k_1,k_2\le n+n'$ by reflection symmetry (whereby the sum over $k_1$ is seen to be finite). The right hand side can be further rewritten as ($l_1\equiv l_2-k_2$) \ben \sum_{k_1}\, R(0,0;\varpi-2J)^{-k_2,-k_1}_{-n,-n'} \ \sum_{l_2} \, R(J,0;\varpi) ^{-l_2,l_1}_{0,-k_2} \ S^{-l_2}(\sg')\vj(\sg)S^{l_1-k_1}(\sigma) \een leading to the equation\\[.2cm] {\bf Step 2 Equation} \ben R(J,0;\varpi)^{-n_2,n_2-n-n'}_{-n,-n'}=\sum_{k=0}^{min(n_2,n+n')} \, R(0,0;\varpi-2J)^{-k,k-n-n'}_{-n,-n'} \ R(J,0;\varpi)^{-n_2,n_2-k}_{0,-k} \label{step2} \een The right hand side is fully known as the type {\bf III} braiding has been solved. The upper summation limit $n_2$ is due to the selection rule and could also be inferred by a ``fusion rule analysis'' (see Section 3.1) of the left hand side. Though we do not have a complete proof of (\ref{step2}) we do have substantial evidence since we have been able to prove it for $n_2=0,1$ or 2. \subsection{Step 3} It is firstly noted that \ben \vj(\sg)\smn(\sg)\vjpr(\sg')\smnpr(\sg')=q^{2J'n'} \vj(\sg)\smn(\sg)\smnpr(\sg')\vjpr(\sg') \een We will consider this last expression, assuming that our proposal for $R(J,0;\varpi)^{-n_2,n_2-n-n'}_{-n,-n'}$ has already been verified (Step 2). We then have \bea &&\vj(\sg)\smn(\sg)\smnpr(\sg')\vjpr(\sg')\nn &=&\sum_{k_2\ge 0} \, R(J,0;\varpi)^{-k_2,k_2-n-n'}_{-n,-n'}\, S^{-k_2}(\sg')\, \sum_{l_2}\, R(J,J';\varpi)^{l_2,l_1}_{k_2-n-n',0}\nn &\cdot&\vjpr(\sg')S^{l_2}(\sg')\vj(\sg)S^{l_1}(\sg)\nn &=&\sum_{k_2\ge 0}\,q^{-2J'k_2} R(J,0;\varpi)^{-k_2,k_2-n-n'}_{-n,-n'} \ \sum_{l_2}\, R(J,J';\varpi-2k_2)^{l_2,l_1}_{k_2-n-n',0}\nn &\cdot&\vjpr(\sg')S^{l_2-k_2}(\sg')\vj(\sg)S^{l_1}(\sg) \eea in addition to \ben \vj(\sg)\smn(\sg)\vjpr(\sg')\smnpr(\sg')= \sum_{n_2} R(J,J';\varpi)^{-n_2,-n_1}_{-n,-n'} \vjpr(\sg')S^{-n_2}(\sg') \vj(\sg)S^{-n_1}(\sg) \een with $l_1,n_1$ again defined by conservation of total screening number. By comparing the two results for the braiding in question, we find\\[.2cm] {\bf Step 3 Equation} \ben R(J,J';\varpi)^{-n_2,n_2-n-n'}_{-n,-n'}= \sum_{k\ge 0} \, q^{2J'(n'-k)} \, R(J,0;\varpi)^{-k,k-n-n'}_{-n,-n'} R(J,J';\varpi-2k)^{k-n_2,n_2-n-n'}_{k-n-n',0} \label{step3} \een By analyzing the second factor on the right hand side we find that the summation over $k$ is restricted as \bea n_2\leq k&\mbox{for}&n+n'\leq n_2\nn n_2\leq k\leq n+n'-1&\mbox{for}&n_2\leq n+n'-1 \eea An explicit proof of Eq. (\ref{step3}) is quite laborious, but we have checked it explicitly in the case $n=1$ and $n_2\le n'$, thus presenting further evidence in favour of our proposal. A further discussion on the type {\bf IV} $R$-matrix will appear elsewhere \cite{RS}. \section{Case of Two Screening Charges: $\alpha_+$ and $\alpha_-$} Up to now, we have restricted ourselves to vertex operators involving only screenings corresponding to the ``semi-classical'' screening charge $\alpha_-$. However, it is straightforward to include the screenings corresponding to the conjugate charge $\alpha_+$, with $\alpha_+\alpha_-= 2\pi$. The corresponding deformation parameter is ${\widehat h} \equiv {\pi^2\over h}$. Denoting by ${\widehat S}$ the screening operators corresponding to $\alpha_+$, one may consider the general vertex operators $U_{m^e}^{(J^e)}$ given by \begin{equation} U_{m^e}^{(J^e )}:=V^{(J^e{})}_{-J^e{}}S^{n}{\hat S}^{{\widehat n}} \label{doublecharge} \end{equation} with $n,{\widehat n}$ integer and $m^e= n+{\widehat n} {\alpha_+\over\alpha_-}-J^e$. One may then follow exactly the same reasoning as in ref. \cite{GS94} to obtain the braiding matrix for $U_{m{\widehat m}}^{(J{\widehat J})}$ from that of the single screening vertex operators; the essential observation here is that screening operators of opposite types commute, $[S(\sigma), {\widehat S}(\sigma')]=0$.\footnote{It is quite remarkable that the technique used in refs. \cite{GS93,GS94} to derive the braiding matrix continues to work in the presence of negative screenings. We have been able to show that the fundamental equations (2.17) of ref. \cite{GS93} and (5.7) of ref. \cite{GS94} continue to hold, when using the standard continuation (\ref{negprod}) to products with negative upper limits. These equations could in fact have been used as an alternative starting point for our analysis.} One thus obtains, exactly as in the positive screening case, \begin{equation} R(J^e{},{J^e{}}';\varpi)_{ \underline{n},\underline{n}'}^{\underline{n}_2, \underline{n}_1}=q^{-J^e{}{J^e{}}'}{\widehat q}^{-{\Jhat^e}{}{{\Jhat^e}{}}'} R(J^e{},{J^e{}}';\varpi )_{n,n'}^{n_2,n_1} \hat R({\Jhat^e}{},{{\Jhat^e}{}}';{\widehat \varpi} )_{{\widehat n}, {{\widehat n}}'} ^{{\widehat n_2}, {\widehat n_1}} \label{doubleR} \end{equation} The conventions here are the following: $\underline{n}=(n,{\widehat n})$ represents the two screening numbers corresponding to $\alpha_-$ and $\alpha_+$, ${\Jhat^e}{} =J^e{}{h\over\pi}=J^e{}{\alpha_-\over\alpha_+}$, ${\widehat \varpi}=\varpi{h\over\pi}$ and ${\widehat q}\equiv e^{i{\widehat h}}$. Finally, $\hat R$ is the same function of its arguments as $R$, except that $\alpha_-$ is replaced by $\alpha_+$ everywhere (or equivalently $h$ by ${\widehat h}$). \section{Conclusions and Outlook} What have we learnt from the present analysis? First of all, we believe to have presented convincing evidence that there is a closed exchange algebra for a set of generalized Coulomb gas vertex operators involving positive and negative powers of screenings alike, and we have obtained the explicit form of the braiding matrix in this general context, through natural analytic continuations of the braiding matrix for positive screenings. The basic lesson is that the additional structure generated by the inclusion of negative screenings is essentially determined by the algebra for vertex operators with positive screenings alone, through fundamental consistency conditions of conformal field theory. This is because negative screening operators are, in a rather precise sense, inverses of positive screening operators. Our results can be expressed, as in the positive screening case, in terms of $_4F_3$ $q$-hypergeometric functions which truncate. As pointed out in Appendix B, the question of uniqueness of our solution is more delicate than might have been naively expected. Whether there is a deeper meaning to the formal possibility of introducing, in our analysis, $R$-matrix elements for non-integer outgoing screenings merits further investigation. Likewise, the origin of the selection rule of Section 4.1 should be elucidated. The technique employed in ref. \cite{GS93} (cf. footnote in Section 7), which uses detailed input from the operatorial construction of the vertex operators, may allow to shed some light on both of these questions. To complete the description of the conformal algebra, we also need the fusion matrix. While it should be given in general in terms of the braiding matrix through one of the Moore-Seiberg relations, one could also think of an independent derivation along similar lines as for the braiding matrix. It appears that this is in fact possible, and one obtains recursion relations from the associativity (${\cal F}{\cal F}{\cal F}={\cal F}{\cal F}$) relation. We hope to come back to this question in a future publication. The growing body of results thus indicates strongly that on the level of the chiral operator algebra, there exists a sequence of inclusions\footnote{See ref. \cite{GR94} for a discussion on how the Kac table is embedded into the larger shells from the point of view of the polynomial equations.}, each representing a consistent solution of the Moore-Seiberg equations; see Fig. \ref{shellmodel}. \begin{figure} \epsfxsize=10truecm \epsfysize=10truecm \epsffile{shellmodel.eps} \caption[]{``Shell model'' of chiral vertex operator algebra} \label{shellmodel} \end{figure} The situation will, however, be much more complicated on the non-chiral level, as will be discussed below. Another interesting direction for future investigations would be an analysis of the quantum group aspect of the generalized Coulomb gas picture advocated here. It is well known \cite{B88,G90,CGR942} that there exists another basis of chiral vertex operators $\xi^{(J)}_M$ which renders the underlying $U_q(sl(2))$ symmetry of the algebra manifest; braiding and fusion symbols for these vertex operators are given by universal $R$-matrix and $3j$-symbols for $U_q(sl(2))$, respectively. The ``covariant'' vertex operators are related to those of the Coulomb gas picture by a basis transformation with $\varpi$-dependent coefficients $|J,\varpi)_M^m$. In succession of our work ref. \cite{Sch96}, we were able to carry out formally an analysis similar to the one above within the quantum group covariant basis, with the role of the screening number $n$ played by $N=J+M$. It is not hard to obtain the braiding matrix for the case where one or both of $N,N'$ are negative (integer). The analysis is in fact much simpler than in the Coulomb gas basis as the algebraic expression for the $R$-matrix in the covariant basis is essentially just a ratio of $q$-factorials, and the generalized $R$-matrix can essentially be obtained by naive analytic continuation. Similar selection rules as for the Coulomb gas basis hold. However, the covariant vertex operators for generic $2J$ are given by infinite sums over Coulomb gas vertex operators, the precise meaning of which is not a priori clear. In particular, by replacing one free field by the other, one would seem to obtain two different copies of the same quantum group representation. It is interesting to speculate whether non-perturbative mechanisms along the lines of ref. \cite{Sch96} will render these two copies equivalent. We hope to return to the study of these questions in the future. \subsection{Non-chiral Case and Liouville Theory} Our analysis so far has addressed the chiral operator algebra only; let us now make some more speculative remarks on the conformal field theories that could be obtained by combining the two chiralities, and in particular about the relevance of our analysis for Liouville theory. The general rule for constructing the non-chiral theory is that operators of the latter should have crossing-symmetric correlators, and obey a consistent closed fusion and braiding algebra. To find the general answer to this question is, obviously, a formidable task. However, within the operator approach to Liouville theory, two mechanisms for producing local fields have been identified. In the standard ``weak coupling'' ($c>25$) regime, locality arises from the orthogonality relations for the $q$-deformed $6j$-symbols. In fact, one writes the local fields - the Liouville exponentials - as diagonal combinations of left and right moving vertex operators, both in the sense of conformal weight and $U(1)$ quantum number $m$: \begin{equation} e^{-J\alpha_-\Phi}=\sum_{n\ge 0} a_n^{(J)}(\varpi) U_m^{(J)}(u){\overline U}_m^{(J)}(v) \label{Liou} \end{equation} with $u:=\tau+\sigma, \ v:=\tau-\sigma$. The coefficients $a_n^{(J)}(\varpi)$ can actually be absorbed in the normalization of the vertex operators, upon which the braiding of the latter will be given in terms of $q$-deformed $6j$-symbols alone \cite{GS94}. Mutual locality of two Liouville exponentials then becomes the statement that the contraction of left and right moving $R$-matrices over the $m$-indices yields the unit matrix. When written in terms of the quantum group covariant basis, Eq. (\ref{Liou}) simply becomes the singlet formed from left and right moving spin $J$ representations of highest and lowest weight, respectively. Remarkably, this mechanism reproduces in particular the minimal models when one continues to central charges $c<1$; that is, there is a description in terms of Liouville exponentials of the observables of these models \cite{G93}. For $c>1$ (in fact, $c>25$), the closest analog of the minimal models is Liouville theory restricted to the Kac table. The diagonal sums are finite (the $n$ -sum in Eq. (\ref{Liou}) truncates at $n=2J$), and closure of the chiral and non-chiral algebra become essentially equivalent. When we leave the Kac table, the situation changes rather drastically: it is no longer possible to give a {\it unique} representation of the Liouville exponentials in terms of a sum over chiral vertex operators. As pointed out in ref. \cite{Sch96}, this is due to two reasons: First, the formal expansion Eq. (\ref{Liou}) cannot be evaluated term by term, using naive charge conservation rules, when the sum is infinite (see also ref. \cite{RPS98} for a recent discussion on this fact from another point of view). Second, when $2J$ is not an integer, the expansion Eq. (\ref{Liou}) suffers from a multi-valuedness problem as regards its zero mode dependence, and therefore does not represent accurately the ``true'' Liouville exponentials even classically. According to the ideas of ref. \cite{PRY}, adapted to the Liouville context in \cite{Sch96}, this does not mean that such expansions cannot be used; rather, one needs to introduce a new (continuous) parameter $\beta$ which controls the monodromy properties of the expansion with respect to the zero mode. One arrives then at generalized expansions of the form \begin{equation} \sum_{n} a_{m;\beta}^{(J)}(\varpi) V^{(J)}_{-J}{\overline V}^{(J)}_{-J} S^{\beta+n}{\overline S}^{\beta+n}. \label{Liougen} \end{equation} For the three point function, with operators situated at $0,1,\infty$, all $\beta$ parameters are fully determined, up to integers. For the operator at $z=1$, not replaced by a highest weight state, one has to choose $\beta$ such that the total monodromy becomes trivial, i.e. the multi-valuedness disappears. One can in particular consider the set of three point functions where $\beta$ is of the form $n+{\widehat n}{\alpha_+\over\alpha_-}$ ($n,{\widehat n}$ integer), so that the exponential can be described by Eq. (\ref{Liou}) or its generalization to both screenings $\alpha_+,\alpha_-$. This was done in ref. \cite{Sch96}. In this context, one already needs the negative screening operators discussed in the present paper, because they are generated non-perturbatively by a careful evaluation of the expansion Eq. (\ref{Liou}). The non-perturbative contributions can be rendered perturbative in terms of an ``effective'' representation of the Liouville exponentials that can be evaluated using naive charge conservation rules: \begin{equation} e^{-J\alpha_-\Phi}_{\hbox{\tiny{eff}}} =\sum_{n=-\infty}^\infty a_n^{(J)}(\varpi) U_m^{(J)}(u){\overline U}_m^{(J)}(v) \label{Lioueff} \end{equation} It is an interesting question whether Eq. (\ref{Lioueff}) is more than just a recipe for the computation of a class of three point functions. In ref. \cite{GS94} it was shown that Eq. (\ref{Liou}) is compatible with locality in the sense that any two Liouville exponentials of the form (\ref{Liou}) will commute at equal times, order by order in powers of screening operators (i.e. perturbatively). It was also observed that in the same sense, the Liouville equation is fulfilled by the exponential with $J=-1$ (see also \cite{OW86}). It is natural to ask whether the same properties will obtain non-perturbatively, i.e. when taking into account the additional negative screening contributions of Eq. (\ref{Lioueff}). While this is essentially trivial for the equations of motion, non-perturbative locality requires that \ben [e^{-J\alpha_-\Phi}(\tau,\sigma), \widetilde {e^{-J'\alpha_-\Phi}} (\tau,\sigma')]=0 \een Here, both exponentials are given by Eq. (\ref{Liou}) and understood to be treated perturbatively, but the second one is constructed in terms of the free field $\tilde X$. The results of the present paper provide all the necessary tools in order to answer this question. One has to show that there is a new orthogonality relation between $q$-deformed $6j$-symbols corresponding to positive and negative screenings, respectively; work on this problem is in progress. If locality holds both in the perturbative and the non-perturbative sense, at least for integer (positive or negative) $2J$, this would be a hint that there exist consistent conformal field theories described by positive integer, and arbitrary integer screening powers, respectively. Some counter-evidence, however, seems to be provided by the analysis of the ``perturbative'' operator product of two Liouville exponentials. While it formally closes on the set of Liouville exponentials with integer positive screenings, there appear arbitrarily strong short-distance singularities \cite{GS94}\footnote{The formula (\ref{fus5}) corrects a misprint in ref. \cite{GS94}. Note also that we are working here in Euclidean coordinates on the sphere where $z:=e^{\tau+i\sigma}$.} that render the result dubious: \bea &&e^{-J_1\alpha_-\Phi(z_1, {\bar z}_1 )} e^{-J_2\alpha_-\Phi(z_2, {\bar z}_2 )}= \sum_{ J_{12}=-\infty}^{J_1+J_2} \sum _{\{\nu\}, \{\bar \nu\} } e^{-J_{12} \alpha_-\Phi^{\{\nu\}, \{\bar \nu\}}(z_2, {\bar z}_2 )}\nn &\cdot& |z_1-z_2|^{2(\Delta_{J_{12}}-\Delta_{J_1}-\Delta_{J_2})} \langle\varpi_{J_{12}}, {\overline \varpi}_{J_{12}}; \{\nu\}, \{\bar \nu\} | e^{-J_1\alpha_-\Phi(z_1-z_2, {\bar z}_1-{\bar z}_2 )} | \varpi_{J_2}, {\overline \varpi}_{J_2}\rangle \label{fus5} \eea where $\varpi_J:=\varpi_0+2J\equiv 1+{\pi\over h}+2J$, and ${\nu}$ denotes descendant contributions. The matrix element on the right hand side represents the operator product coefficient as well as the descendant contributions to the conformal block given by $J_{12}$; see ref. \cite{GS93} for details. The contributions from very large negative $J_{12}$ are non-vanishing in general and so there appear unwanted short-distance singularities of arbitrarily high order as $\Delta_{J_{12}} \to -\infty$. Even if there is a fully consistent non-chiral algebra involving only integer screenings, we should be very careful as to its interpretation as a restriction of Liouville theory. In fact, the operator product of Liouville exponentials is expected \cite{ZZ96} to depend on all operators entering in the correlator, and to involve ``hyperbolic'' states corresponding to purely imaginary $\varpi$. Crossing symmetry, or locality, should then involve the same intermediate channels, and the interpretation within Liouville theory of locality properties based on integer screenings only is a priori not very clear. On the other hand, general $\beta$-dependent expansions of type Eq. (\ref{Liougen}) would actually allow for the introduction of hyperbolic intermediate states. Consider a four point function with operators situated at $0,1,\infty$ and $z$, represented as \ben \langle\varpi_4|e^{-J_3\alpha_-\Phi}(z,\bar z)e^{-J_2\alpha_-\Phi}(1)| \varpi_1\rangle \een The parameters $\beta_3,\beta_2$ appearing in the representation of the exponentials in the middle are coupled by the monodromy condition. For a given conformal block, they have a unique value. The question of what range to choose for the one free $\beta$ parameter is thus tantamount to the factorization problem, and imaginary $\beta$ will in fact produce hyperbolic intermediate states. While in Liouville theory restricted to the Kac table, locality and conformal invariance are sufficient to fully determine the operator construction of the theory, we see that in the present, much larger context this may well be an illusion. The simple correspondence between chiral and non-chiral algebras seems to be lost, and additional information is needed for the construction of non-chiral correlation functions. A careful analysis of the zero mode dependence of the exponentials, combined with group-theoretical arguments as in refs. \cite{T97,T972}, seems to be a step in the right direction. The presence of a context-dependent, floating $\beta$ parameter as we advocate it here, means that general Liouville exponentials will invoke arbitrary quantum group representations, and not just highest or lowest weight ones as one could have naively expected from the classical picture. This would resemble the situation in affine $SL(2)$ current algebra where continuous representations of intertwining operators appear for non-integrable weights \cite{PRY}. In any case, it seems clear that a deeper understanding of the full Liouville dynamics will have to pass through a study of the algebra of the chiral vertex operators, and it seems desirable to continue the program towards the analysis of arbitrary non-integer screenings. The study of correlation functions of the chiral vertex operators may be interesting in its own right, and is expected to lead to new mathematical functions already on the integer (including negative ones) screening level. \subsection{Application to Strong Coupling Liouville Theory} In a series of works \cite{GN85,G91,GR942+96,GR94}, Gervais and collaborators have proposed a conformal field theory with $1<c<25$ which they interpret as a candidate for Liouville theory in the forbidden, ``strong coupling'' region $1<c<25$. This theory exists for certain discrete values of the central charge ($c=7,13,19$), and on a spectrum of highest weight states specified by the unitary truncation theorems of ref. \cite{G91}. Locality arises in a rather different way; local observables become products of sums of chiral vertex operators rather than sums of products, and each chiral factor is local up to a phase factor. The local operators can be divided into two sets, one with negative and one with positive conformal weights. The latter operators involve negative screenings; they can be written in the form (we write the left-moving factor only) \begin{equation} \chi_+^{(J)}(u)=\sum_{p\in \Nat_0} \, (-1)^{(2-s)p\left[2(p+{x-J\over {\pi\over h}-1}) +{p+1\over 2}\right]} g^{x}_{J,x-J-1+p({\pi\over h}-1)} \ V^{(J)}_{p({\pi\over h}-1)-1-J}(u) \label{chi} \end{equation} Here, $x:={1\over 2}(\varpi-\varpi_0)$ with $\varpi_0:=1+{\pi\over h}$, and $s=0,\pm 1$ determines the central charge ($c=1+6(s+2)$). The operators $V^{(J)}_m$ are nothing but normalized $U_m^{(J)}$ fields, $V^{(J)}_m\equiv {1\over I^{(J)}_m(\varpi)} U_m^{(J)}$. Finally, the coupling constants $g^{J_{12}}_{J_1,J_2}$ appear as prefactors of the $6j$-symbols in the fusion and braiding relations of the $V^{(J)}_m$ operators and were determined in \cite{CGR941}. The values that $\varpi$ is allowed to take are not arbitrary continuous, but discrete: \ben \varpi=({\pi\over h}-1)(l+1+{r\over 2-s}) \een with $l\in \mbox{$Z\hspace{-2mm}Z$}$, $r=0,...,1-s$. Similarly, $J$ is restricted to be \ben J=({\pi\over h}-1)(l'+{r'\over 1-s})-1 \een with $l'\in \mbox{$Z\hspace{-2mm}Z$}$, $r'=0,...,2-s$. The screening number with respect to $\alpha_-$ is $-p-1$, while for $\alpha_+$ it is $p$. Thus indeed, negative screenings are involved, and one can obtain a second analogous set with the roles of $\alpha_+$ and $\alpha_-$ interchanged. In the absence of an algebraic control over vertex operators with negative screenings, the properties of the $\chi$ fields were described using the formal symmetry $J\to -J-1$ of the $q$-deformed $6j$-symbols already mentioned above. The present analysis not only allows to further corroborate this discussion (at least as far as the braiding is concerned) by establishing a direct derivation from the operator construction, but also opens the road towards an investigation of new local operators with negative screenings which are not accessible by the $J\to -J-1$ analytic continuation technique. The results of the present analysis and its possible future extensions should also find direct applications in theories closely related to Liouville, such as $SL(2,\mbox{\hspace{.04mm}\rule{0.2mm}{2.8mm}\hspace{-1.5mm} R})$ or $SL(2,\C)/SU(2)$ WZNW theory \cite{T97}. {}From a larger perspective, this program can be viewed as a kind of bootstrap approach to irrational conformal field theory in the simplest context where the chiral symmetry algebra is just Virasoro. It is an interesting question whether all solutions can be identified either with (subsectors of) weak or strong coupling Liouville theory. \\[.4cm] {\bf Acknowledgment}\\[.2cm] We would like to thank J. Teschner for useful discussions and comments. J.R. gratefully acknowledges the partial financial support from the Danish Natural Science Research Council, contract no. 9700517. He also thanks Laboratoire de Math\'ematiques et Physique Th\'eorique, Universit\'e de Tours and The Niels Bohr Institute, where parts of this paper were written down, for their kind hospitality.
\section{Introduction} The conjecture \cite{mal,KW} on the nexus between Type--IIB superstring theory on $\ads{5}\times{\cal H}$, where $\ads{5} $ is the five-dimensional anti-de~Sitter space and $\cal H$ is a five-dimensional variety, called {\em horizon} in the sequel, and superconformal field theories on the four-dimensional boundary of the five-dimensional anti-de~Sitter space, thought of as a theory of D3--branes, has instigated a new spate of research. The conjecture, known by now as the AdS-CFT correspondence, has already been tested in sundry instances. The tests can be broadly classified into two categories. The first is a consideration of symmetries of the two candidates. Apart from a diagonal $U(1)$ subgroup of the gauge group, corresponding to a free-photon in the gauge theory \cite{mal,Wit-hol,MP}, the group of global isometries of the supergravity background in the presence of suitable four-form fluxes corresponding to the Type--IIB superstring theory is identified with the R-symmetry group of the deemed superconformal theory on the four-dimensional boundary of $\ads{5}$ [see, for example, \cite{FZ}]. The other consists in an identification of operators in the two theories and in the comparison of their correlation functions [see, for example, \cite{GKP,Wit-hol,roma2}]. When the five-dimensional horizon $\cal H$ is the five-sphere, ${\mathbb{S}}^5$, the conformal counterpart on the D3--brane is the ${\cal N} =4$ supersymmetric gauge theory of the boundary of $\ads{5}$. The horizon $\cal H$ is envisaged as being part of the space, namely ${\mathbb{C}}^3$, transverse to the world-volume of the D3--brane. One of the interesting courses of development in the subject is a generalisation of the original conjecture \cite{mal} to models with less than maximal supersymmetry \cite{KW,KS,GRW,MP}. Considering a theory of D3--branes on a Gorenstein canonical singularity of, for example, the type ${\mathbb{C}}^3/\Gamma$, where $\Gamma$ is a discrete subgroup of $SU(3)$, it is possible to break some of the supersymmetries of the field theory of the brane. The corresponding candidate dual theory is then the Type--IIB theory compactified on $\ads{5}\times {\mathbb{S}}^5/\Gamma$ \cite{KS,KW}. For example, if $\Gamma$ is isomorphic to a ${\mathbb{Z}}_2$ subgroup of an $SU(2)$ subgroup of $SU(3)$, then the resulting dual theory has been found to be a four-dimensional ${\cal N} =2$ gauge theory \cite{LNV,GRW,Wit-bar}. Theories of a D3--brane on orbifolds of ${\mathbb{C}}^3$ of the form ${\mathbb{C}}^3/\Gamma$ have been considered earlier, for different groups $\Gamma$. Examples include discrete groups $\Gamma$ isomorphic to ${\mathbb{Z}}_k$, for $k = 3, 5$ \cite{DGM}, 7, 9, 11\cite{muto1}, to ${\mathbb{Z}}_2\times{\mathbb{Z}}_2$ without discrete torsion \cite{brg,subir} and with discrete torsion \cite{md:dis,DF}, as well as some cases where $\Gamma$ is non-abelian \cite{GLR,muto2}. In some of these analyses \cite{DGM,muto1,brg,subir,GLR,muto2} the moduli space of a D3--brane at an orbifold singularity, derived as a solution to the F- and D-flatness conditions of the corresponding supersymmetric gauge theory of the D3--brane, admits a toric geometric description. It is interesting to enquire whether one can also use these other cases to derive an admissible horizon for $\ads{5}$. For an appropriate action of a $U(1)$ group on the horizon $\cal H$, compatible with the complex structure of its metric cone $C({\cal H})$, one can obtain the toric data of the base ${\cal H}_{B} = {\cal H}/U(1)$ from the toric data of the moduli space of the D3--brane on ${\mathbb{C}}^3/\Gamma$ \cite{MP}. It has been shown \cite{keh} that if the space ${\cal H}_B$ admits a K\"ahler-Einstein metric, then one can, from the knowledge of the $U(1)$-action, find the horizon $\cal H$ by considering $U(1)$-bundles of ${\cal H}_B$. This has been illustrated for several examples arising from partial resolutions of a ${\mathbb{C}}^3/({\mathbb{Z}}_2\times{\mathbb{Z}}_2)$ singularity \cite{MP}. In order to obtain the horizon from the moduli space of a D3--brane, one starts from the field theory on the D3--brane on ${\mathbb{C}}^3/\Gamma$. This field theory is the blown-down limit of the theory of a D3--brane on the Calabi--Yau~ singularity. In order for this theory to be dual to the Type--IIB theory on $\ads{5}\times{\cal H}$, the horizon $\cal H$ is taken to be a space on which ${\mathbb{C}}^3/\Gamma$ is a cone. It has been shown that in order to retain some supersymmetry in the quotient theory, the horizon has to be a {Sasaki--Einstein}~ manifold \cite{FigF,keh} and correspondingly, the manifold ${\cal H}_B = {\cal H}/U(1)$ must be K\"ahler--Einstein. If the group-action of $U(1)$ on the horizon $\cal H$ is \emph{regular}, meaning that the orbits of the $U(1)$-action are closed and have the same length, the base ${\cal H}_B$ of this $U(1)$-foliation admits a K\"ahler--Einstein metric \cite{FK,MP}. See \cite{BG} for results with a weaker condition of a quasi-regular $U(1)$-action. If the discrete group $\Gamma$ is isomorphic to ${\mathbb{Z}}_3$, for example, then the group action of $U(1)$ is regular on the horizon \cite{FK,MP}. However, if the $U(1)$-action on the horizon is not regular, the issue of the existence of K\"ahler--Einstein metrics on the corresponding bases needs be settled on a case by case basis. In this note we consider the question of the existence of K\"ahler--Einstein metrics on the base ${\cal H}_B$ of the D3--brane moduli spaces in two examples, namely D3--branes on ${\mathbb{C}}^3/\Gamma$, where $\Gamma$ is a discrete subgroup of $SU(3)$, isomorphic to ${\mathbb{Z}}_3$ and ${\mathbb{Z}}_5$, that is for the orbifolds ${\mathbb{C}}^3/{\mathbb{Z}}_3$ and ${\mathbb{C}}^3/{\mathbb{Z}}_5$. The first corresponds to a horizon that admits a regular $U(1)$-action, while the other does not. The corresponding theories of D3-branes have been studied earlier \cite{DGM}. In both cases we work out the toric data for the base, starting from the toric data of the orbifolds, obtained earlier \cite{DGM}. We then prove the existence of a K\"ahler--Einstein metric on the respective four-dimensional bases. For the base of ${\mathbb{C}}^3/{\mathbb{Z}}_3$, the canonical toric metric is the Fubini-Study metric on $\cp{2}$, which is {K\"ahler--Einstein}. The canonical toric metric on the base of ${\mathbb{C}}^3/{\mathbb{Z}}_5$ is, however, K\"ahler, but not Einstein. We find out the condition for the existence of a K\"ahler--Einstein metric on ${\cal H}_B$ in terms of a deformation of the canonical K\"ahler potential, following \cite{abru,rk}. This leads to a linear partial differential equation whose solution purveys a K\"ahler--Einstein metric on the corresponding base manifold ${\cal H}_B$, thereby establishing the existence of a K\"ahler--Einstein metric on ${\cal H}_B$. The structure of this note is as follows. In \S\ref{cone}, we briefly review the argument establishing the necessity of a K\"ahler--Einstein metric on ${\cal H}_B$, following \cite{keh}. In \S\ref{guill}, we review the necessary information and formulas for the calculations, following \cite{gui,abru,rk}. Finally in \S\ref{zed3} and \S\ref{zed5}, we present the calculation of K\"ahler--Einstein metrics on the bases of ${\mathbb{C}}^3/{\mathbb{Z}}_3$ and ${\mathbb{C}}^3/{\mathbb{Z}}_5$, before concluding in \S\ref{conclusion}. \section{The cone, the horizon \& the base}\label{cone} Given a space $\cal H$, the \emph{metric cone} $C({\cal H })$ on $\cal H$ is defined by the warped product of $\cal H$ and the half-line, ${\mathbb{R}}^+$, as $C({\cal H}) = {\cal H}\times_{r^2}{{\mathbb{R}}^+}$, endowed with the metric $\langle\cdots\rangle_{C({\cal H})} = dr^2 + r^2\langle\cdots\rangle_{\cal H}$. Here and below we use subscripts to metrics descriptively whenever convenient. The conical nature of $C({\cal H})$ is discerned by noting that there exists a group of diffeomorphisms of $C({\cal H})$, isomorphic to the multiplicative group of real numbers, ${\mathbb{R}}_+$, namely, $r\longmapsto\ t\; r$, for any positive real number $t$, which rescales the metric as: $\langle\cdots\rangle_{C({\cal H})} \longmapsto t^2\langle\cdots\rangle_{C({\cal H})}$. Moreover, if the space $\cal H$ admits a transitive $U(1)$-action, then one can define another space ${\cal H}_B$ by quotienting $\cal H$ by the $U(1)$. Thus, the space ${\cal H}_B$ can be written as $C({\cal H})/({\mathbb{R}}_+\times U(1))$, that, in view of the decomposition ${\mathbb{C}}^{\star} = {\mathbb{R}}_+\times U(1)$, can be expressed as $C({\cal H})/{\mathbb{C}}^{\star}$. Here ${\mathbb{C}}^{\star}$ denotes the complex-plane with the origin deleted, that is, ${\mathbb{C}}^{\star} = {\mathbb{C}}\setminus\{0\}$, or, in other words, the algebraic torus. The metric cone on $\cal H$ can therefore be viewed as a \emph{complex cone} over ${\cal H}_B$. In the sequel we shall refer to the metric cone simply as a \emph{cone}, while the complex cone will be explicitly mentioned. Also, we shall refer to the space $\cal H$ as the \emph{horizon} whilst ${\cal H}_B$ will be referred to as the \emph{base} (of the complex cone as well as of the horizon). Let us begin by briefly recalling some features of supergravity background solutions in the presence of D3--branes preserving some supersymmetry. The solution to the supergravity equations of motion for the ten-dimensional metric retaining some supersymmetry is sought in the following form \cite{keh} \begin{equation}\label{oops} ds^2 = e^{2\Upsilon(x^m)}\eta_{\mu\nu}dx^{\mu}dx^{\nu} + e^{-2\Upsilon(x^m)}h_{mn}dx^mdx^n\quad m,n = 1, {\ldots}, 6;\ \ \mu , \nu = 7, {\ldots}, 10 \end{equation} where the ten-dimensional space-time is taken to be split into two parts. Four of the ten directions are parallel to the world-volume of the D3--brane, coordinatized by $x^{\mu}$, ${\mu} = 7, {\ldots}, 10$, and $\eta_{\mu\nu}$ denotes the flat-metric. The remaining six coordinates are transverse to the D3--brane are denoted by $x^m$, $m=1, {\ldots}, 6$. Thus, the D3--brane is a point in the six-dimensional space coordinatized by $x^m, m = 1, {\ldots}, 6$. In \eq{oops}, $\Upsilon (x^m)$ is a harmonic function of the coordinates of the six-dimensional space transverse to the world-volume of the D3--brane. The six-dimensional metric is chosen to be of the form \begin{equation}\label{metricone} ds_{C({\cal H})}^2 = h_{mn}dx^mdx^n = dr^2 + r^2g_{ij}dx^idx^j, \quad i,j = 1, {\ldots}, 5 \end{equation} where the five-dimensional metric $g_{ij}$ is the metric on the horizon $\cal H$ that appears in the compactification of Type--IIB supergravity on $\ads{5}\times {\cal H}$. Finally, the horizon is envisaged as a $U(1)$-bundle over a four-dimensional space ${\cal H}_B$, with metric of the form \begin{equation} ds_{\cal H}^2= g_{ij}dx^idx^j = ds_{{\cal H}_B}^2 + 4(d\psi + A_adx^a)^2, \quad a,b = 1,2,3,4 \end{equation} where the metric on the four-dimensional base ${\cal H}_B$ is \begin{equation}\label{gab-ans} ds_{{\cal H}_B}^2 = g_{ab}dx^adx^b \quad a,b = 1,2,3,4. \end{equation} Now, if the four-dimensional space ${\cal H}_B$ is a complex K\"ahler surface, with metric $g_{ab}$, then the 1-form $A_adx^a$ is the $U(1)$ connection with field strength $F$ proportional to the K\"ahler two-form $\omega$ of ${\cal H}_B$, {\em viz.~} \begin{equation} F = i \omega . \end{equation} With the above ans\"atz, the condition for the existence of a covariant Killing spinor in the absence of D7--branes is found to be \cite{keh} \begin{equation}\label{normalzn} R_{ab} = 6g_{ab}, \end{equation} where $R_{ab}$ denotes the Ricci tensor of the metric $g_{ab}$ in \eq{gab-ans}. Thus, the question of finding out admissible background solutions reduces to the question of existence of K\"ahler--Einstein metric on the base ${\cal H}_B$. One can use the known four-dimensional spaces admitting K\"ahler--Einstein metric as candidates for ${\cal H}_B$ \cite{MP}. Some of the choices are related to different partial resolutions of the moduli space of a D3--brane on an orbifold ${\mathbb{C}}^3/({\mathbb{Z}}_2\times{\mathbb{Z}}_2)$, studied earlier \cite{subir,brg}. The connections are forged by considering the toric data of the D-brane orbifolds of ${\mathbb{C}}^3$ in the blown-down limit and then realizing the horizon $\cal H$ as a line-bundle over a certain base ${\cal H}_B$ inside the orbifold viewed as the level set of a combination of moment maps obtained by splitting the original one at blown-down \cite{MP}. Another interesting case is when the D-brane orbifold is ${\mathbb{C}}^3/{\mathbb{Z}}_3$. The corresponding base ${\cal H}_B$ is $\cp{2}$. In this case (and some others studied in \cite{MP}), the horizon $\cal H$ has the structure of a {Sasaki--Einstein}~ manifold, with the aforementioned $U(1)$ providing the corresponding contact structure. Moreover, the action of the $U(1)$ is regular, For such cases the existence of K\"ahler--Einstein metrics on the base ${\cal H}_B$ has been studied earlier \cite{FK,MP}. Let us point out that many of the mathematical results used in our considerations are strictly valid for compact spaces. However, we continue to use these results for the non-compact spaces at hand as the compact spaces approximate their non-compact counterparts in a sufficiently small neighborhood of the singular point. \section{K\"ahler--Einstein metrics on toric varieties}\label{guill} In this section we introduce the canonical toric metric \cite{gui} and its deformation \cite{abru,rk}. We also derive the equations governing the deformation by demanding the resulting metric to be Einstein. On the other hand, the K\"ahlerity of the deformed metric follows from its very construction. Let $(X,\omega)$ be a compact, connected $2d$-dimensional manifold. Let $\tau: {\mathbb{T}}^d \longrightarrow \qopname\relax o{Diff} (X,\omega)$ be an effective Hamiltonian action of the standard $d$-torus ${\mathbb{T}}^d$. Let ${\mu}: X\longrightarrow {\mathbb{R}}^d$ denote the associated moment-map and $\Delta$ the image of $X$ on ${\mathbb{R}}^d$ under the moment map, $\Delta = \mu(X)\subset{\mathbb{R}}^d$. The convex polytope $\Delta$ is referred to as the \emph{moment polytope}. The triple $(X,\omega ,\tau)$ is determined up to isomorphism by the moment polytope \cite{gui}. The polytope $\Delta$ in ${\mathbb{R}}^d$ is called \emph{Delzant} if there are $d$ edges meeting at each vertex $p$ of $\Delta$ and any edge meeting at $p$ can be given the form $p+sv_i$, for $0\leq s\leq\infty$, where $\{v_i\}$ is a basis of ${\mathbb{Z}}^d$ \cite{gui,abru}. Conversely, one can associate a toric variety $X_{\Delta}$ with the above properties to a Delzant polytope $\Delta$ in ${\mathbb{R}}^d$, such that $\Delta$ is the moment polytope of $X_{\Delta}$. Let $X_{\Delta} = X$ be the toric variety associated to $\Delta$ that is, $X_{\Delta} = {\mu}^{\star}(\Delta )$, where $\mu^{\star}$ denotes the pull-back of $\mu$. The moment polytope $\Delta$ can be described by a set of inequalities of the form $\langle y, u_i\rangle \geq \lambda_i$, $i=0,1,2,{\ldots}, d-1$. Here $u_i$ denotes the inward-pointing normal to the $i$-th $(d-1)$-dimensional face of $\Delta$ and is a primitive element of the lattice ${\mathbb{Z}}^d\subset {\mathbb{R}}^d$. The pairing $\langle~,~\rangle$ denotes the standard scalar product in ${\mathbb{R}}^d$ and $y$ represents a $d$-dimensional real vector. We can thus define a set of linear maps, $\ell_i : {\mathbb{R}}^n\longrightarrow {\mathbb{R}}$, \begin{eqnarray}\label{li} \ell_i(y) = \langle y, u_i\rangle - \lambda_i, \quad i=0,1,{\ldots}, d-1. \end{eqnarray} Denoting the interior of $\Delta$ by $\Delta^{\circ}$, $y\in\Delta^{\circ}$, if and only if $\ell_i(y) > 0$ for all $i$. On the open ${\mathbb{T}}^d_{{\mathbb{C}}}$-orbit in $X_{\Delta}$, associated to a Delzant polytope $\Delta$, the K\"ahler form $\omega$ can be written as \cite{gui,abru}: \begin{eqnarray}\label{om} \omega = {i}~\partial}\def\W{{\mathcal W}\bar{\partial}\def\W{{\mathcal W}}\mu^{\star} \left( \sum_{i=0}^{d-1}\lambda_i\ln\ell_i + \ell_{\infty} \right), \end{eqnarray} where we have defined $\ell_{\infty}$ as the sum, \begin{eqnarray} \ell_{\infty} = \sum_{i=0}^{d-1}\langle y, u_i\rangle . \end{eqnarray} The potential in the expression \eq{om} for the K\"ahler form $\omega$ is determined by the equations for faces of the Delzant polytope \eq{li}. If $\omega$ is a ${\mathbb{T}}^d$-invariant K\"ahler form on the complex torus ${\cal M} = {\mathbb{C}}^d/2{\pi}i{\mathbb{Z}}^d$, then there exists a function ${\cal F}(x)$, with $x = {{\goth R\goth e}} z$, $z\in{\mathbb{C}}$, on $\cal M$, such that $\omega = 2i\partial}\def\W{{\mathcal W}\bar{\partial}\def\W{{\mathcal W}}{\cal F}$. Moreover, the moment map $\mu:~{\cal M}\longrightarrow{\mathbb{R}}^d$ is given by \begin{equation}\label{mu-leg} \mu (z) = \frac{\partial}\def\W{{\mathcal W}{\mathcal{F}}}{\partial}\def\W{{\mathcal W} x} \end{equation} The K\"ahler form $\omega$ can be written in terms of this new function (K\"ahler potential) as: \begin{eqnarray} \omega = \frac{i}{2} \sum_{j,k=0}^{n-1} \frac{\partial}\def\W{{\mathcal W}^2{\mathcal{F}}}{\partial}\def\W{{\mathcal W} x_j\partial}\def\W{{\mathcal W} x_k} dz_j\wedge d\bar{z}_k. \end{eqnarray} The moment-map $\mu$ defines a Legendre transform through \eq{mu-leg} and thus we can define a new set of variables $y_i$, $i=0,{\ldots},d$, \begin{eqnarray}\label{legendre} y_i = \frac{\partial}\def\W{{\mathcal W}{\mathcal{F}}}{\partial}\def\W{{\mathcal W} x_i}, \end{eqnarray} conjugate to $x_i = {\goth R\goth e}~z_i$. Consequently, we can define a potential $\cal G$, conjugate to $\cal F$ under the Legendre transform \eq{legendre} on $\Delta^{\circ}$, such that \cite{gui,abru}, \begin{eqnarray}\label{defG} {\mathcal{G}} = \frac{1}{2}\sum_{k=0}^{d-1}\ell_k(y)\ln\ell_k(y). \end{eqnarray} The inverse of the Legendre transform \eq{legendre} can be shown to be of the form \cite{gui}: \begin{eqnarray}\label{invleg} x_i = \frac{\partial}\def\W{{\mathcal W}{\mathcal{G}}}{\partial}\def\W{{\mathcal W} y_i} + r_i, \quad i=0,\cdots , d-1, \end{eqnarray} where $r_i$ are constants. This means that up to a linear term in the coordinates $y_i$, ${\mathcal{G}}$ is the K\"ahler potential Legendre-dual to ${\mathcal{F}}$. Moreover, the $d{\times}d$ matrix \begin{eqnarray} {\mathcal{G}}_{ij} = \frac{\partial}\def\W{{\mathcal W}^2{\mathcal{G}}}{\partial}\def\W{{\mathcal W} y_i\partial}\def\W{{\mathcal W} y_j}, \end{eqnarray} evaluated at $y_i=\frac{\partial}\def\W{{\mathcal W}{\mathcal{F}}}{\partial}\def\W{{\mathcal W} x_i}$ \eq{legendre}, is the inverse of the matrix \begin{eqnarray}\label{metric-f} {\mathcal{F}}_{ij} = \frac{\partial}\def\W{{\mathcal W}^2{\mathcal{F}}}{\partial}\def\W{{\mathcal W} x_i\partial}\def\W{{\mathcal W} x_j}. \end{eqnarray} The Ricci-tensor for the metric \eq{metric-f} takes the following form: \begin{eqnarray}\label{Ric-1} R_{ij} &=& -\frac{1}{2}\frac{\partial}\def\W{{\mathcal W}^2\ln\det~{\mathcal{F}}}{\partial}\def\W{{\mathcal W} x_i\partial}\def\W{{\mathcal W} x_j}\\ \label{Ric-tensor} &=& -\frac{1}{2}\sum_{k,l=0}^{n-1} {\mathcal{G}}^{lj} \frac{\partial}\def\W{{\mathcal W}^2 {\mathcal{G}}^{ik}}{\partial}\def\W{{\mathcal W} y_k\partial}\def\W{{\mathcal W} y_l}, \end{eqnarray} where ${\mathcal{G}}^{ij}$ denotes the inverse of ${\mathcal{G}}_{ij}$. Note that, ${\mathcal{F}}$ in \eq{Ric-1} ( and in \eq{curv1} below) denotes the matrix ${\mathcal{F}}_{ij}$, and not the K\"ahler potential unlike elsewhere in this note. The Ricci-scalar for this metric is then derived by multiplying \eq{Ric-tensor} with ${\mathcal{G}}_{ij}$, which is the inverse of the metric ${\mathcal{F}}_{ij}$ in the $y$ coordinates, and is given by \cite{abru} \begin{eqnarray}\label{curv1} R &=& -\frac{1}{2}\sum_{i,j=0}^{n-1}{\mathcal{F}}^{ij} \frac{\partial}\def\W{{\mathcal W}^2 \ln\det{\mathcal{F}}}{\partial}\def\W{{\mathcal W} x_i\partial}\def\W{{\mathcal W} x_j} \\\label{curvature} &=& -\frac{1}{2}\sum_{i,j=0}^{n-1}\frac{\partial}\def\W{{\mathcal W}^2{\mathcal{G}}^{ij}} {\partial}\def\W{{\mathcal W} y_i\partial}\def\W{{\mathcal W} y_j}, \end{eqnarray} where ${\mathcal{F}}^{ij}$ denotes the inverse of ${\mathcal{F}}_{ij}$. For our purposes, it will be convenient to use the matrix ${\mathcal{G}}_{ij}$ in the coordinates $y$. One can, in principle, rewrite all the relevant expressions in terms of the coordinates $x_i$ and the matrix ${\mathcal{F}}_{ij}$. The expressions for the K\"ahler potential \eq{om}, the metric \eq{metric-f} and the curvature \eq{Ric-tensor} described above will be referred to as the \emph{canonical} ones. Moreover, if $\iota:X\longrightarrow\cp{n}$, denotes a projective embedding of $X$ in $\cp{n}$ for some $n$, then the canonical K\"ahler form $\omega$, given by \eq{om}, is ${\mathbb{T}}^d$-equivariantly symplectomorphic to the pull-back of the K\"ahler form $\omega_{\scriptscriptstyle\text{FS}}$ corresponding to the Fubini-Study metric on $\cp{n}$, namely $\iota^{\star}\omega_{\scriptscriptstyle \text{FS}}$. However, the curvature of the canonical toric metric is arbitrary in general. In order to obtain a metric with prescribed --- usually constant --- curvature, as in our cases, one may need to deform the canonical K\"ahler form to another one in the same K\"ahler class. This may be effected by adding a function, smooth on some open subset of ${\mathbb{R}}^n$ containing $\Delta$, to the potential ${\mathcal{G}}$, such that the Hessian of the new potential is positive definite on $\Delta^{\circ}$ in order for the new potential to be K\"ahler. This furnishes a deformed K\"ahler metric in the same K\"ahler class as the canonical one. The variety $X_{\Delta}$ is then endowed with two different K\"ahler forms related by a ${\mathbb{T}}^n$-equivariant symplectomorphism since the function $f$ is non-singular. One then finds out the form of this extra function by demanding the prescribed curvature. This method was used in deriving the extremal K\"ahler Metric on ${\mathbb{C}}{\mathbb{P}}^2\#{\mathbb{C}}{\mathbb{P}}^2$ in Calabi's form \cite{abru} as well as in finding out a Ricci-flat metric on the resolved D--brane orbifold ${\mathbb{C}}^3/{\mathbb{Z}}_3$ \cite{rk}. See \cite{cao} for another approach involving the Heat equation. Following the approach of \cite{abru,rk}, let us as add a function $f$ to ${\mathcal{G}}$, and define a new potential $\widetilde{{\mathcal{G}}} = {\mathcal{G}} + \frac{1}{2}f$. Now the matrix $\widetilde{{\mathcal{G}}}_{ij}$ corresponding to the potential $\widetilde{{\mathcal{G}}}$ assumes the form \begin{eqnarray} \widetilde{{\mathcal{G}}}_{ij} = {\mathcal{G}}_{ij} + \frac{1}{2}\frac{\partial}\def\W{{\mathcal W}^2f}{\partial}\def\W{{\mathcal W} y_i\partial}\def\W{{\mathcal W} y_j}. \end{eqnarray} One can then find out the K\"ahler metric by inverting $\widetilde{{\mathcal{G}}}_{ij}$ and hence the curvature for this new metric. This will also give rise to a new $\widetilde{{\mathcal{F}}}$ corresponding to ${\mathcal{F}}$ and also new coordinates $\widetilde{x}$. What we propose to do next is to write down the general form of the metric for a function $f$ and then determine $f$ by demanding that the Ricci-tensor given by the formula \eq{Ric-tensor} to be proportional to the new metric $\widetilde{{\mathcal{G}}}^{ij}$. This yields a differential equation for the function $f$. For the two cases considered here, the base ${\cal H}_B$ is two-(complex)-dimensional. Thus, we have $i=1,2$. Specialising to this case, the three components of $R_{ij}$ can be written as: \begin{eqnarray} \label{r11} -2R_{11} &=& \widetilde{{\mathcal{G}}}^{11}\left( \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{11}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_1} + \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} \right) + \widetilde{{\mathcal{G}}}^{12}\left( \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{11}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} + \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_2\partial}\def\W{{\mathcal W} y_2} \right), \\ \label{r12} -2R_{12} &=& \widetilde{{\mathcal{G}}}^{12}\left( \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{11}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_1} + \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} \right) + \widetilde{{\mathcal{G}}}^{22}\left( \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{11}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} + \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_2\partial}\def\W{{\mathcal W} y_2} \right), \\ -2R_{22} &=& \widetilde{{\mathcal{G}}}^{22}\left( \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{22}}{\partial}\def\W{{\mathcal W} y_2\partial}\def\W{{\mathcal W} y_2} + \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} \right) + \widetilde{{\mathcal{G}}}^{12}\left( \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{22}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} + \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_1} \right). \end{eqnarray} If there exists a K\"ahler--Einstein metric on ${\cal H}_B$, then for some choice of the deforming function the components of the curvature $R_{ij}$ must be equal to the components of the metric $\widetilde{{\mathcal{G}}}^{ij}$. We find out a solution for $f$ by imposing this condition on $R_{ij}$. Thus, the condition for the manifold ${\cal H}_B$ to be K\"ahler--Einstein, {\em i.e.~} \begin{equation}\label{reqg} R_{ij} = \Lambda\widetilde{{\mathcal{G}}}^{ij}, \end{equation} is satisfied if \begin{eqnarray} \label{reqn1} \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{11}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_1} &+& \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} = -2\Lambda, \\ \label{reqn2} \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{22}}{\partial}\def\W{{\mathcal W} y_2\partial}\def\W{{\mathcal W} y_2}&+& \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2} = -2\Lambda, \\ \label{reqn3} \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{22}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2}&+& \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_1} =0,\\ \label{reqn4} \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{11}}{\partial}\def\W{{\mathcal W} y_1\partial}\def\W{{\mathcal W} y_2}&+& \frac{\partial}\def\W{{\mathcal W}^2\widetilde{{\mathcal{G}}}^{12}}{\partial}\def\W{{\mathcal W} y_2\partial}\def\W{{\mathcal W} y_2} =0. \end{eqnarray} Let us note that the non-covariant notation for indices in \eq{reqg} originates from the fact that, abiding by common practice, we have denoted derivatives by subscripts and the metric under consideration has been written as $\widetilde{{\mathcal{G}}}^{ij} = \widetilde{{\mathcal{G}}}_{ij}^{-1}$ in $y$ variables. Here $\Lambda$ is a constant parameter that determines the Ricci scalar as $R= 2\Lambda$. However, in the following we keep this parameter arbitrary for book-keeping. In \S\ref{zed3} and \S\ref{zed5}, we use equations \eqs{reqn1}{reqn4} to prove the existence of a K\"ahler--Einstein metric on ${\cal H}_B$ for the two cases mentioned earlier. \section{K\"ahler--Einstein metric on the base ${\cal H}_B$ of ${\mathbb{C}}^3/{\mathbb{Z}}_3$}\label{zed3} In this section we consider the blown-down limit of the moduli space of a D3--brane on the orbifold ${\mathbb{C}}^3/{\mathbb{Z}}_3$. We first find out a toric description of the base ${\cal H}_B$ staring from the toric data of ${\mathbb{C}}^3/{\mathbb{Z}}_3$. Then, following the discussion in \S\ref{guill}, we find out the K\"ahler--Einstein metric on this base. In order to obtain the base ${\cal H}_B$ for blown-down ${\mathbb{C}}^3/{\mathbb{Z}}_3$, let us start from the charges of the fields. The toric data derived from F- and D-flatness conditions of the gauge theory is given by \cite{DGM} \begin{equation}\label{tordatz3} {\cal T} = \begin{pmatrix} -1&1&0&0\\ -1&0&1&0\\ 3&0&0&1 \end{pmatrix}. \end{equation} The blown-down moduli space is obtained in terms of charges specified by the kernel of $\cal T$ with the resolution (Fayet--Iliopoulos~) parameter vanishing. The charge matrix becomes \begin{equation} {\mathcal Q} = (\ker{{\cal T}})^{\text{T}} = \begin{pmatrix} 1 & 1 & 1 & -3 \end{pmatrix}, \end{equation} where a superscript $\scriptstyle{\text{T}}$ designates matrix transpose. The resulting moduli space is described by the following moment-map equation: \begin{equation}\label{z3-mom} |z_0|^2+|z_1|^2+|z_2|^2 - 3|z_3|^2 = 0, \end{equation} where $\{z_i|i=0,1,2,3\}$ are the homogeneous variables on the corresponding toric variety. The horizon $\cal H$ is obtained as a line-bundle from \eq{z3-mom}, following the construction of \cite{MP} \begin{equation}\label{momap1} |z_0|^2+|z_1|^2+|z_2|^2 =\zeta \qquad\text{and}\qquad 3|z_3|^2 = \zeta. \end{equation} The parameter $\zeta$ is left arbitrary in \eq{momap1} for book-keeping. It can be set to unity. Equations \eq{momap1} describe a space ${\mathbb{S}}^5\times{\mathbb{S}}^1$. The horizon $\cal H$ is obtained from \eq{momap1} after quotienting by a $U(1)$ as $({\mathbb{S}}^5\times{\mathbb{S}}^1)/U(1)$, with the $U(1)$-action on the homogeneous variables given by \begin{equation} U(1): (z_0, z_1, z_2, z_3)\longmapsto ( e^{i\theta}z_0, e^{i\theta}z_1, e^{i\theta}z_2, e^{-3i\theta}z_3). \end{equation} Now, to find the toric description for the base ${\cal H}_B$, we collect the charges from \eq{momap1} into another charge-matrix \begin{equation} {\mathcal Q}_B =\begin{pmatrix} 1 & 1 & 1 & 0\\ 0 & 0& 0 & 3 \end{pmatrix} \end{equation} and find out the toric data corresponding to ${\cal H}_B$ as the co-kernel of the transpose of ${\mathcal Q}_B$. The resulting toric data is given as \begin{equation}\label{tordat3} {\mathcal T}' = \coker{{\cal Q}_B^{\text{T}}} = \begin{pmatrix} -1&1&0&0\\ -1&0&1&0 \end{pmatrix} \end{equation} Leaving out the last column, this can be recognised as the toric data for $\cp{2}$, namely, \begin{equation}\label{tordatB3} {\mathcal T}_B = \begin{pmatrix} -1&1&0\\ -1&0&1 \end{pmatrix}. \end{equation} This description corresponds to the ``most efficient" description of the toric variety ${\cal H}_B$, as mentioned in \cite{MP}. The two-vectors described by the columns of ${\cal T}'$ are plotted in Figure~\ref{fig_z3}. As in \cite{MP}, the toric data ${\cal T}_B$ in \begin{figure}[h] \begin{center} \setlength{\unitlength}{0.00083333in} \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \begin{picture}(3109,3000)(0,-10) \put(1500,2625){\circle*{100}} \put(300,225){\circle*{100}} \put(2700,1425){\circle*{100}} \put(1500,1425){\circle{100}} \path(300,225)(1500,2625)(2700,1425)(300,225) \put(1300,2800){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(0,1)$}}}}} \put(-100,0){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(-1,-1)$}}}}} \put(2800,1375){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(1,0)$}}}}} \put(1300,1575){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(0,0)$}}}}} \end{picture} } \end{center} \caption{\small\sl Plot of the columns of the toric data ${\cal T}'$ for ${\mathbb{C}}^3/{\mathbb{Z}}_3$: the unfilled circle is omitted from the description in obtaining the Delzant polytope. The filled ones correspond to ${\cal T}_B$.} \label{fig_z3} \end{figure} \eq{tordatB3} for ${\cal H}_B$ is obtained after omitting the point in the interior of the triangle. Moreover, the toric data ${\cal T}'$ in \eq{tordat3} can be identified as the first two rows in the toric data $\cal T$ in \eq{tordatz3} for ${\mathbb{C}}^3/{\mathbb{Z}}_3$. We shall see shortly that the omission of the point in the interior of the polygon is related to the construction of the Delzant polytope corresponding to the variety. The corresponding canonical toric metric from ${\cal T}_B$ in \eq{tordatB3} is the Fubini-Study metric, as given in \cite{abru}. The Delzant polytope is given as \begin{equation}\label{poly-3} y_1 \geq 0,\qquad y_2\geq 0 \qquad\zeta-y_1-y_2\geq 0, \end{equation} where we have shifted the variables $y_i$, $i=1,2$ by the respective values of the support functions $a_i$, $i=1,2$, at each of the one-dimensional cone generators specified by the columns of ${\cal T}_B$, following \cite{rk}. The K\"ahler structure $\zeta$ is given as $\zeta=a_0+a_1+a_2$, and hence the first Chern-class of the variety, calculated as the sum of the coefficients of $a_i$ in $\zeta$ \cite{gui}, is constant, $\goth{c}_1(X_{\Delta}) = 1+1+1 =3$. Thus, the variety in question is Fano. This corroborates to the fact that the base ${\cal H}_B$ is $\cp{2}$, as pointed out in \cite{MP}. Let us point out that writing down the Delzant polytope provides a rationale for the omission of the interior points of the polygon shown in Figure~\ref{fig_z3} (this, however, is \emph{not} the Delzant polytope): they do not yield any new face for the Delzant polytope. The last column of \eq{tordat3} does not affect the polytope \eq{poly-3}. The canonical potential $\mathcal G$ corresponding to the Delzant polytope \eq{poly-3} is, by \eq{defG}, \begin{equation}\label{g3} {\mathcal G} = \frac{1}{2} \biggl[ y_1\ln y_1 + y_2\ln y_2 + (\zeta -y_1-y_2) \ln(\zeta - y_1-y_2 )\biggr]. \end{equation} The canonical metric ensuing from the potential \eq{g3} is given by \begin{equation}\label{metric3} {\mathcal G}^{ij} = \frac{2}{\zeta}\begin{pmatrix} y_1(\zeta - y_1)&-y_1y_2 \\-y_1y_2&y_2(\zeta-y_2). \end{pmatrix} \end{equation} The Ricci-tensor evaluated from \eq{metric3}, using \eq{Ric-tensor} can be checked to be proportional to the metric \eq{metric3}. That is, the equations \eqs{reqn1}{reqn4} are satisfied with $f=0$ and $\Lambda = 3/\zeta$. The Ricci-scalar for this metric is found from this as \begin{equation}\label{r-z3} R = -\frac{1}{2}\frac{2}{\zeta}(-2-1-1-2) = \frac{6}{\zeta} = 2\Lambda, \end{equation} as expected from \eq{curvature}. Thus, we have verified that there does exist a K\"ahler--Einstein metric on the base ${\cal H}_B$ of the blown-down D--brane orbifold ${\mathbb{C}}^3/{\mathbb{Z}}_3$. The metric is diffeomorphic to the Fubini-Study metric, with a constant scalar curvature as given in \eq{r-z3}. A comment is in order. A Ricci-flat metric on the resolved D3--brane orbifold ${\mathbb{C}}^3/{\mathbb{Z}}_3$ was found out earlier using similar techniques in \cite{rk}. This needed a deformation of the canonical toric metric by a function which was shown to be a solution of the differential equation \footnote{The expression \eq{fpp-sol} corrects a typographical error in equation (61) in \cite{rk}} \begin{eqnarray}\label{fpp-sol} f'' = \frac{9y_3(\xi+3y_3)^3 - (\xi+12y_3)[c + (\xi+3y_3)^3] }{y_3(\xi+3y_3)[c + (\xi + 3y_3)^3]}, \end{eqnarray} where $y_3$ denotes the third variable needed to form the Delzant polytope of the three-dimensional variety, $c$ is a constant and $\xi$ is the resolution parameter (set to zero in \eq{z3-mom}). The metric on the complex cone $C({\cal H})$ constructed from the metric \eq{metric3} using \eq{metricone} does not appear to be in the same form as the one derived in \cite{rk} for the resolved D--brane orbifold ${\mathbb{C}}^3/{\mathbb{Z}}_3$, with vanishing K\"ahler class. This discrepancy, however, may be attributed to the fact that constancy of the scalar curvature on the base was not imposed on the metric in \cite{rk}, as was pointed out there. However, since the restrictions of the metric \eq{metricone}, and the one derived in \cite{rk} on the base ${\cal H}_B$ are the same, with suitable normalisation of $y_3$, there exists some neighborhood of ${\cal H}_B$ in ${\mathbb{C}}^3/{\mathbb{Z}}_3$ on which the two metrics are diffeomorphic to one another, thanks to the Darboux-Weinstein theorem \cite{GS}. \section{K\"ahler--Einstein metric on the base ${\cal H}_B$ of ${\mathbb{C}}^3/{\mathbb{Z}}_5$}\label{zed5} In this section we consider the blown-down limit of the moduli space of a D3--brane on ${\mathbb{C}}^3/{\mathbb{Z}}_5$. First, let us find out the toric description of the base ${\cal H}_B$, from the toric data of the moduli space \cite{DGM}. We shall mimic the considerations of \S\ref{zed3}. Let us start from the following toric data for the orbifold ${\mathbb{C}}^3/{\mathbb{Z}}_5$ \cite{DGM}, \begin{equation} \label{t5} {\cal T} = \begin{pmatrix} -1 & 1 & 0 & 0& 0\\ -3 & 0 &1 &0 &-1\\ 5&0&0&1&2 \end{pmatrix}. \end{equation} The corresponding charge matrix, evaluated as the transpose of the kernel of $\cal T$ takes the form \begin{equation}\label{charge1} \begin{pmatrix} 1&1&0&1&-3\\ 0&0&1&-2&1 \end{pmatrix}, \end{equation} that leads to two moment map equations in the blown-down limit with vanishing Fayet--Iliopoulos~ parameters of the gauge theory: \begin{eqnarray} \label{mom5-1} |z_0|^2 + |z_1|^2 + |z_3|^2 - 3|z_4|^2 &=& 0\\ \label{mom5-2} |z_2|^2 - 2|z_3|^2 +|z_4|^2 &=& 0. \end{eqnarray} From the charge matrix \eq{charge1}, one obtains by row operations (changing basis) the charge matrix \begin{eqnarray} {\mathcal Q} = \begin{pmatrix} 1&1&3&-5&0\\ 0&0&1&-2&1 \end{pmatrix}. \end{eqnarray} $\cal Q$ corresponds to eliminating $z_4$ from \eq{mom5-1} by using \eq{mom5-2} and leads to the following equation for the global moment map: \begin{equation}\label{mom-5} |z_0|^2+|z_1|^2+3|z_2|^2 -5|z_3|^2 = 0. \end{equation} The horizon $\cal H$ is obtained from \eq{mom-5}, or from the first row of $\cal Q$ constructing the line-bundle \cite{MP}: \begin{equation}\label{split5} |z_0|^2+|z_1|^2+3|z_2|^2 = \zeta\qquad\text{and}\qquad 5|z_3|^2 = \zeta . \end{equation} Again, we have kept a free-parameter $\zeta$ for book-keeping. Equations \eq{split5} describe a direct product ${\mathbb{S}}^5\times{\mathbb{S}}^1$. The corresponding horizon is obtained from \eq{split5} quotienting by a $U(1)$ as ${\mathcal H} = ({\mathbb{S}}^5\times{\mathbb{S}}^1)/{U(1)}$, with the $U(1)$-action given by \begin{equation}\label{act-u1} U(1):\, (z_0, z_1, z_2, z_3)\longmapsto (e^{i\theta}z_0, e^{i\theta}z_1, e^{3i\theta}z_2, e^{-5i\theta}z_3). \end{equation} This $U(1)$-action \eq{act-u1} is not regular. For instance, the orbits of the points $(0,0,1,0)$ and $(1,0,0,0)$ are of different lengths. However, we proceed as before to derive the base ${\cal H}_B$ from this, and to this end write the corresponding charge-matrix as \begin{equation} {\mathcal Q}_B = \begin{pmatrix} 1&1&3&0&0\\ 0&0&0&5&0\\ 0&0&1&0&1\\ 0&0&0&2&0 \end{pmatrix}. \end{equation} This leads to the following toric data obtained as the co-kernel of the transpose of ${\mathcal Q}_B$, namely \begin{equation}\label{tordat5} {\mathcal T}' = \begin{pmatrix} 0&-3&1&0&-1\\ 1&-1&0&0&0 \end{pmatrix}. \end{equation} From ${\mathcal T}'$, ignoring the last two columns, corresponding to points in the interior of the triangle, as shown in Figure~\ref{figz5}, we obtain the toric data for ${\cal H}_B$ as the following rectangular matrix: \begin{equation}\label{tordatB5} {\mathcal T}_B = \begin{pmatrix} 0&-3&1\\1&-1&0 \end{pmatrix}. \end{equation} As before, one can identify the matrix ${\cal T}^{\prime}$ in \eq{tordat5} inside $\cal T$ in \eq{t5}: ${\cal T}'$ is $\cal T$ with the last row omitted. \begin{figure}[h] \begin{center} \setlength{\unitlength}{0.00083333in} \begingroup\makeatletter\ifx\SetFigFont\undefined% \gdef\SetFigFont#1#2#3#4#5{% \reset@font\fontsize{#1}{#2pt}% \fontfamily{#3}\fontseries{#4}\fontshape{#5}% \selectfont}% \fi\endgroup% \begin{picture}(5267,3030)(0,-10) \put(3750,1500){\circle{100}} \put(2550,1500){\circle{100}} \put(3750,2700){\circle*{100}} \put(150,300){\circle*{100}} \put(4950,1500){\circle*{100}} \path(150,300)(3750,2700)(4950,1500)(150,300) \put(-300,100){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(-3,-1)$}}}}} \put(2200,1300){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(-1,0)$}}}}} \put(3600,1650){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(0,0)$}}}}} \put(3600,2820){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(0,1)$}}}}} \put(5100,1425){\makebox(0,0)[lb]{\smash{{{\SetFigFont{12}{14.4} {\rmdefault}{\mddefault}{\updefault}$(1,0)$}}}}} \end{picture} } \end{center} \caption{\small\sl Plot of the columns of the toric data ${\cal T}'$ for ${\mathbb{C}}^3/{\mathbb{Z}}_5$: the unfilled circles are omitted from the description in obtaining the Delzant polytope. The filled ones correspond to ${\cal T}_B$.} \label{figz5} \end{figure} Having obtained the toric description for the base ${\cal H}_B$, we now proceed as in \S\ref{zed3}, to find out the toric metric. We first derive the canonical toric metric and the scalar curvature to learn that the latter is not constant, signalling that the canonical toric metric is not K\"ahler--Einstein. Following the considerations of \S\ref{guill}, we then deform the canonical toric metric and determine the deformation imposing the condition that the deformed metric is {K\"ahler--Einstein}, as discussed in \S\ref{guill}. Let us start by considering the Delzant polytope corresponding to the variety ${\cal H}_B$ as expressed in terms of the toric data ${\cal T}_B$ in \eq{tordatB5}. The Delzant polytope $\Delta$ is defined by the following inequalities dictated by the columns of ${\cal T}_B$: \begin{equation}\label{delz-z5} y_1\geq 0,\qquad y_2\geq 0, \qquad\zeta-3y_1-y_2\geq 0. \end{equation} Again the variables are shifted by the values of the respective support functions at the one-dimensional cone generators, given by the columns on ${\cal T}_B$. This leads to $\zeta= a_0+3a_1+a_2$, implying that the first Chern-class $\goth{c}_1(X_{\Delta})= 1+3+1=5$, whence the variety ${\mathcal H}_B$ is found to be Fano. Once again the omission of the last two columns of ${\cal T}'$ is justified by their being inconsequential for constructing the Delzant polytope. The canonical toric metric can be obtained from the potential \begin{equation}\label{g5} {\mathcal G} = \frac{1}{2} \biggl[y_1\ln y_1 + y_2\ln y_2 + (\zeta - 3y_1 -y_2)\ln (\zeta - 3y_1-y_2)\biggr], \end{equation} and the corresponding matrix ${\mathcal G}_{ij}$ is \begin{equation}\label{non-ek} {\mathcal G}_{ij}= \frac{1}{2A}\begin{pmatrix} 9+\frac{A}{y_1} & 3 \\ 3& 1 + \frac{A}{y_2} \end{pmatrix}, \end{equation} where we have defined $A = \zeta -3y_1-y_2$. The metric ${\mathcal G}^{ij}$ derived by inverting \eq{non-ek} is not Einstein. Indeed, the Ricci scalar for this canonical toric metric is: \begin{equation}\label{scalar-z5} R = \frac{26\zeta^2 + 60 \zeta y_1 + 72 y_1^2}{(\zeta + 6y_1)^3}. \end{equation} The Ricci scalar $R$ is, however, positive definite, since, by \eq{delz-z5}, $y_1 \geq 0$ on the Delzant polytope $\Delta$. Thus, as mentioned in \S\ref{guill}, let us now look for a deformed K\"ahler metric which is Einstein, by deforming the above potential $\mathcal G$ by a function $f$. As a simplifying ans\"atz, we choose $f$ to be a function of $y_1$ only, $f=f(y_1)$, since the Ricci-scalar \eq{scalar-z5} depends only on $y_1$. This leads to \begin{eqnarray}\label{g-ij} \widetilde{{\mathcal{G}}}_{ij} = \frac{1}{2A}\begin{pmatrix} 9 + \frac{A}{y_1} + Af'' & 3 \\ 3 & 1 + \frac{A}{y_2} \end{pmatrix}, \end{eqnarray} where a prime signifies differentiation with respect to $y_1$. Inverting ${\widetilde{{\mathcal{G}}}}_{ij}$, we obtain the following expression for $\widetilde{{\mathcal{G}}}^{ij}$: \begin{eqnarray}\label{gij} \widetilde{{\mathcal{G}}}^{ij} = \begin{pmatrix} \frac{2y_1}{F} (\zeta - 3y_1) & -\frac{6y_1y_2}{F} \\ -\frac{6y_1y_2}{F} & 2y_2 (1 - \frac{y_2\phi}{F}) \end{pmatrix}, \end{eqnarray} where $F = \zeta + 6y_1 + y_1f'' (\zeta - 3y_1)$ is a nowhere-vanishing function and we have defined $\phi = (F - 9y_1)/(\zeta - 3y_1)$. Now we solve \eqs{reqn1}{reqn4} using the expression \eq{gij} for $\widetilde{{\mathcal{G}}}^{ij}$. Equation \eq{reqn4} is identically satisfied as $f$ does not depend on $y_2$. Equations \eq{reqn2} and \eq{reqn3} lead to two equations respectively, \begin{eqnarray} \label{fineq2} -4\left(\frac{\phi}{F}\right) -6 \left(\frac{y_1}{F}\right)' = -2\Lambda,\\ \label{fineq3} -4\left(\frac{\phi}{F}\right)' -6 \left(\frac{y_1}{F}\right)'' = 0. \end{eqnarray} In deriving \eq{fineq3} from \eq{reqn3}, we have used the fact that $y_2\neq 0$, {\em i.e.~} $y\in\Delta^{\circ}$. Equation \eq{fineq3} can be obtained by differentiating \eq{fineq2} with respect to $y_1$. Finally, from \eq{reqn1} one derives, \begin{eqnarray} \label{fineq1} (\zeta - 3y_1) ( 2y_1{F'}^2 - y_1FF'') + (15y_1 - 2\zeta)FF' - 9F^2 +\Lambda F^3 = 0, \end{eqnarray} that, however, is solved identically on using \eq{fineq2} and \eq{fineq3}. Thus, one is left with one single equation for $f$, namely \eq{fineq2}, that can be rewritten as \begin{eqnarray}\label{fineq} 3y_1F' - (2\phi + 3) F + \Lambda F^2 = 0. \end{eqnarray} Using the expression for $\phi$, equation \eq{fineq} takes the following form: \begin{equation}\label{ffin} 3 y_1 (\zeta - 3 y_1) F' - 3 (\zeta - 9 y_1) F + \biggl[ (\zeta - 3 y_1) \Lambda - 2 \biggr] F^2 = 0. \end{equation} In order to solve this, we rewrite it further in terms of a new function $\chi = 1/F$ as: \begin{eqnarray}\label{chieqn} 3 y_1 (\zeta - 3 y_1) {\chi}' + 3 (\zeta - 9 y_1) \chi - \biggl[ (\zeta - 3 y_1) \Lambda - 2 \biggr] = 0. \end{eqnarray} Equation \eq{chieqn} is a linear first-order differential equation in $y_1$ and can be solved for $\chi$ and thence for $F$ resulting in \begin{eqnarray} F = \frac{9y_1}{\Theta}, \end{eqnarray} where $\Theta = 1 - \frac{\Lambda}{3}(\zeta - 3y_1) + 9c (\zeta - 3y_1)^{-2}$ and $c$ is a constant of integration. One is then lead to the following equation for $f$: \begin{equation}\label{fpp-z5} f'' = \frac{9y_1 - (\zeta + 6y_1)\Theta}{y_1(\zeta - 3y_1)\Theta} \end{equation} that can be explicitly integrated to determine the potential $\widetilde{{\mathcal{G}}}$ in $y$-variables. Using \eq{fpp-z5} in the expression \eq{gij} for $\widetilde{{\mathcal{G}}}^{ij}$ yields the desired K\"ahler--Einstein metric on the base ${\cal H}_B$ of the cone ${\mathbb{C}}^3/{\mathbb{Z}}_5$. \section{Conclusion}\label{conclusion} To conclude, in this note we have demonstrated the existence of K\"ahler--Einstein metrics on the bases of the complex cones obtained as moduli spaces of D3--branes on ${\mathbb{C}}^3/{\mathbb{Z}}_3$ and ${\mathbb{C}}^3/{\mathbb{Z}}_5$ orbifolds. These cones are, in turn, metric cones over horizons $\cal H$, and can be used as the internal space in Type--IIB string theory on $\ads{5}$, in the context of the conjectured AdS-CFT correspondence. The analysis validates the candidature of the horizon obtained from the blown-down orbifold ${\mathbb{C}}^3/{\mathbb{Z}}_5$ along with ${\mathbb{C}}^3/{\mathbb{Z}}_3$, the latter having already been pointed out in \cite{MP}. It seems possible to generalize the considerations in this note to establish similar results for the existence of K\"ahler--Einstein metrics on ${\cal H}_B$ for other D3--brane orbifolds already studied in literature \cite{muto1,brg,subir,MP}. However, it is not clear how to extend this analysis to cases that do not admit a toric description. The metrics presented here, like the one in \cite{rk}, are in rather special \emph{real} variables. It will be interesting to be able to write these metrics in terms of variables better-suited to studying their geometric properties. We hope to return to this issue in future. \section*{\small\slshape Acknowledgements} It is a pleasure to thank J~F Morales, S Mukhopadhyay, A Sagnotti and especially M Bianchi for illuminating discussions and useful comments during the course of this work. I also thank M Abreu for a useful communication.