__key__
stringlengths
9
15
__url__
stringclasses
18 values
json
dict
tiff
imagewidth (px)
13
2.68k
astro-ph0001011
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "Duration distribution of short GRBs. The solid line is the common BeppoSAX/Batse Sample distribution, the dashed one is the normalized 4B Batse distribution, for comparison." ], "images": [ null, "/dev/shm/tmpjlsu6_9l/IP-09-fig1.ps", null ], "texts": [ "What can BeppoSAX tell us about short GRBs: An update from the Subsecond GRB Project G. Gandolfi ^ (1) M.J.S. Smith ^ (2),(4) A. Coletta ^ (2) G. Celidonio ^ (2) L. Di Ciolo ^ (2) A. Paolino ^ (2) G. Tarei ^ (2) G. Tassone ^ (2) J.M. Muller ^ (2),(4) E. Costa ^ (1) M. Feroci ^ (1) F. Frontera ^ (3),(5) L. Piro ^ (1) ^ (1) IAS/CNR, Via del Fosso del Cavaliere, Roma, Italy ^ (2) BeppoSAX-SOC/TELESPAZIO, Via Corcolle 19, Roma, Italy ^ (3) TESRE/CNR, Via Gobetti 101, Bologna, Italy ^ (4) SRON, Sorbonnelaan 2, Utrecht, The Netherlands ^ (5) Dipartimento di Fisica, Universit a di Ferrara, Via del Paradiso 12, Ferrara, Italy We present some statistical considerations on the BeppoSAX hunt for subsecond GRBs at the Scientific Operation Center. Archive analysis of a BATSE/SAX sub-sample of bursts indicates that the GRB Monitor is sensitive to short ( 2 sec) events, that are in fact 22 of the total. The non-detection of corresponding prompt X-ray counterparts to short bursts in the Wide Field Cameras, in about 3 years of operations, is discussed: with present data no implications on the X-to- ray spectra of short vs long GRBs may be inferred. Finally, the status of searching procedures at SOC is reviewed. Introduction The nature and the origin of short bursts, namely the events with a duration 2 s, that clearly represent a separate class from longer ones (Kouvelioutou, 1993), is one of the most interesting and puzzling open problems in GRB studies. The celestial distribution of these events is clearly isotropic and their gamma spectrum predominantly harder than that of longer ones, while the two classes have the same peak flux range. A lot of questions naturally arise when considering the issue and comparing the phenomenology with the successful paradigm of Fireball models: are the short bursts caused by a different emission mechanism? Do they originate in extragalactic sources as well? Is their LogN-LogS euclidean (Tavani, 1998)? Are there other duration subclasses (Cline, 1999)? Do they possess an afterglow signature as the longer ones? At present no prompt counterpart detection at any wavelength is available and the recent afterglow discoveries refer only to the long bursts class, leaving a conspicuous gap in our knowledge of the phenomenon. Hence, the identification and localization of at least some event of this kind, that we call ``subsecond bursts'', remains a key objective of GRB research, an objective that can only be pursued at the moment by means of BeppoSAX satellite and, in the near future, by means of missions like Hete2, Integral and Swift. We present here some statistical consideration on BeppoSAX triggering sensitivity to short bursts with GRBM. Finally, the non detection of subsecond events in the sample of WFC selected GRBs is analysed and the new real-time procedure for on-ground triggering at Scientific Operation Centre is briefly resumed. Estimation of GRBM Efficiency in triggering Short Bursts The basic procedure for GRB counterpart detection at the BeppoSAX Scientific Operation Centre, performed since the very beginning of the mission, relies on the visual check of PDS Lateral Shield and WFC ratemeters around GRBM trigger times (Coletta, 1998). This procedure is being recently improved and extended, discriminating as possible the short, particle induced, GRBM fake events from real bursts (Feroci, 1999), monitoring the BATSE triggers as well and performing an additional on-ground search for excesses of counts on Lateral shields and WFCs below the trigger threshold (Smith, 1999; Gandolfi, 1999). In this work we address the general problem of the BeppoSAX sensitivity to GRBs with duration less than 2 s and we try to test the significance of experimental results in the field, i.e. the non detection of subsecond bursts at quick look level. The first point to evaluate is the efficiency of GRBM in triggering short events, since the fundamental method used to identify GRBs - the only one systematically applied to data for a long time, is the check of the counts at trigger times both in WFC and LS lightcurves with 1 s bins. If the sample of BeppoSAX triggers doesn't include the majority of short bursts we would expect a priori a very low probability of finding subsecond X-ray counterparts, because the general SOC monitoring of WFC transients at a quick look level has a poor sensitivity to excesses of counts that last less than 8 s (the standard bin adopted to scan lightcurves on an orbital basis). In order to estimate the GRBM sensitivity in the duration range of interest with respect to BATSE, we have selected the subsample of common triggers in the period 10/11/1996 - 15/9/1998 (BeppoSAX/SOC Catalog, 1999; BATSE Current Catalog, 1999). This guarantees the reality of SAX events (and automatically discriminates ``fake'' triggers), that are also analyzed in high time resolution mode (8 ms bin) in order to obtain the T90 (i.e. the time interval during which they emit the 90 of their fluence). We find 111 common triggers, with 24 events under the 2 s threshold that corresponds to a 22 of the total. Comparing this value with the 26 of BATSE duration distribution and assuming that the probability is described by a binomial, we conclude that the number of short bursts detected is still compatible with completeness with a probability of 15 , but is almost certainly affected by a bias for very brief events. In fact the bins with GRBs under 0.5 s exhibit a slight deficit of the order of 25 with respect to BATSE 4B duration distribution: coherently with expectations, the triggering efficiency is probably poor in that range due to the small fluence of shortest events (Fig. 1). An offline analysis of BATSE positions, WFCs field of view and ratemeters counts at the relevant trigger times confirms the non detection of counterpart in each one of the 24 cases.", null, "Increasing the Sensitivity with On-Ground Triggering: The BeppoSAX Subsecond Bursts Project The efficiency of the GRBM strongly depends from the spectral properties of the burst and from its maximum flux in the 1 s on-board trigger time bin. Other selection effects that could affect GRBs detection are mainly geometrical and depend on the effective area illuminated by the burst: in case of an event exactly on the axis of the LS (this is the most favorable condition to catch the counterpart in WFC field of view if involved lateral shield is unit 1 or unit 3) the area is minimum and the probability of triggering the GRBM is consequently minimized. For this reason an on-ground triggering procedure (part of the BeppoSAX subsecond GRB Project) has been implemented at SOC last summer (Gandolfi, 1999), with the aim of increasing the global detection efficiency of short and long-faint bursts. This semi-automatic procedure relies only on the comparison between lateral shields and WFC ratemeters in the whole orbit with a 1 s bin. The new trigger condition, that has been empirically chosen in order to achieve the best sensitivity to short events without overloading the amount of quick look work and is certainly more effective with respect to the on-board one, selects events which have a 3 excess of counts in the lateral shield and at the same time a 3.5 excess in the corresponding WFC ratemeter bin. The WFC threshold guarantees that all the bursts detectable in the celestial image, that is those with a global one bin fluence of at least 70 counts in the 2-10 keV range (Smith, 1999) are selected. Interesting examples of short events detected by the new semi-automatic routine, are GRB991014 and GRB991106: the first is slightly too long to be considered a subsecond event, the latter is probably a X-ray/GRB without a detected signal in the GRBM. Other candidate short bursts have been found, but the WFC excess of counts never satisfied the above condition and the transients were in fact not detectable in the corresponding WFC celestial image.The detection of a number of such events, even if it doesn't allow to localize the burst, could help to constrain the spectral properties of short GRBs with statistical significance. Experimental Results and Implications The present number of GRBs with X-ray prompt counterpart detected by BeppoSAX is 27 and no one is classified as short event. Is this result compatible with the expectations? If we assume the completeness in our GRBM detections in the subsecond duration range (and we have seen that we are at least almost complete) we would expect to detect the 26 of short bursts also in the sample of GRBs with revealed counterpart, i.e. 7. This is not the case, but we must remember that just 17 of these events have been triggered by GRBM, the only reliable detection method if the X-ray counterpart has a very low fluence (the GRB quick look procedure guarantees an optimal analysis of WFC, at 1 s or less resolution, only around trigger times). Hence any statistical consideration should be based on the GRBM trigger catalog, which has been - and is being - carefully inspected in real time by Duty Scientists at SOC. Furthermore, the most stringent limit to subsecond counterpart identification is the WFC ratemeter sensitivity in the 1 s bin inspected, that decreases with the decreasing duration of the transient. Considering the duration distribution of short GRBs and the X ray counterparts peak flux distribution for long bursts (Frontera, 1999; various IAUCs), we can estimate a rough global detection efficiency: with a minimum 3 signal in the 1 s bin of WFC ratemeter, in the hypotesis of an average offset of 10 degrees from the centre of the field of view and of a Crab-like spectrum for the source, we should miss about the 50 of expected events. The conversion from the peak flux to the fluence that is compared to the threshold in counts for the bin ( 36 in a standard WFC empty field), is done assuming a constant and maximum emission during T90. The expected number of subsecond counterparts in the GRBM-selected sample is therefore N_ xs = f_ s N_x , where f_ s is the percentage of short gamma bursts, the estimated global detection efficiency for short X-ray counterparts and N_x the global number of X ray counterparts detected by BeppoSAX GRBM. We find with all the above assumptions a value of 2 subsecond events. The non detection is not statistically significant, because considering again the binomial distribution (with p = f_ s the probability of detecting the counterpart of a short GRB instead of a long one, x the GRBM-selected global number of counterpart detections, in this case no one, and n the dimension of the sample, that is 17 events) its probability P(x,n,p) corresponds to P(0,17,0.13) 0.09. This result strongly encourages to increase the external trigger capability at quick look level in order to maximize the probability of finding a subsecond counterpart. In fact, extending to candidate events not selected by GRBM the real-time standard GRB procedure, with its robust and tested efficiency, will surely increase the chance of identifying low flux events. Conclusions At present, no sure prompt X-ray counterparts of subsecond GRBs have been noticed in BeppoSAX WFCs celestial images and we have shown that GRBM triggered detections in this range are compatible with completeness, with a probable slight deficit at shortest durations. This implies, taking into account WFC sensitivity limits and assuming a X to peak ratio similar to that of longer GRBs, that 2 events on a total of 17 triggered with revealed counterpart, should be subsecond bursts. The non detection is clearly not sufficient to test on a statistical basis any spectral hypotesis (i.e. the correctness of the assumption on X to peak ratio). Furthermore, we can sistematically rely on non triggered detections (that represents the 63 of the whole sample of discovered counterparts) just since about 6 months, thanks to the new on-ground triggering routine implemented at SOC. No inference can be made on the basis of the entire counterparts catalog (i.e. 3 / 4 events of 27), because untriggered events with a prompt counterpart could have been easily missed at a quick look level with the old standard GRB procedure. On the other side, archive analysis of the GRBM triggers catalog and of the WFC ratemeters of the whole mission are in progress, and the optimization of quick look procedures to identify and analyze at best the high time resolution data has been now obtained within the frame of the BeppoSAX Subsecond Bursts Project. Coletta A. et al. Proc. of the Active X-Ray Sky Workshop , Roma, Nuclear Physics B 69/1-3, 712 (1998) Cline D.B. et al. astro-ph/9905346 (1999) Feroci F. et. al. these proceedings Gandolfi G. et. al. http://www.ias.rm.cnr.it/ias-home/sax/subsec.html (1999) Kouveliotou C. et al. ApJ. , 413 , L101 (1993) Smith M.J.S. et. al. A A Suppl. Ser. , 138 , 561 (1999) Tavani M. ApJ. , 497 , L21 (1998) BeppoSAX/SOC GRBM Trigger Catalog, private communication (1999) Paciesas W.S. et al. 4B BATSE Catalog, http://cossc.gsfc.nasa.gov/cossc/batse" ] }
astro-ph0001016
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "Top panel: mass function of the Local Group (LG). The points show the space density of LG members containing , after correcting for incompleteness. The solid line shows the field from Zwaan et al. (1997), scaled vertically so as to fit the points. The dotted line is a with a steep upturn at the low mass end, recently proposed by Schneider et al. (1998) Second panel: mass functions for extragalactic HVCs. The dark histogram shows the space density of BSTHB HVCs if they are put at the critical radii for gravitational stability. The unshaded histogram shows the effect of a baryon to total mass ratio f varying from cloud to cloud. The lighter shaded histogram corresponds to BB HVCs, all at the same distance of 1 Mpc. Lines are as in top panel. Third panel: mass function for BSTHB HVCs for different values f (from right to left: f =0.2, 0.1, 0.05, 0.025, 0.0125). The HVC is consistent with the field if f 0.02 and the median distance 200 kpc. Bottom panel: Distribution of the expected detections in the for the BSTHB population. The light grey histogram corresponds to clouds surrounding galaxies, the dark grey histogram represents clouds in galaxy groups. See section 4 for explanation. fig1.fig", "Illustration of the two surveys strips and the distribution of spheres with radii of 1 Mpc around known galaxies (thin circles) and galaxy groups (thick circles). The grey areas indicate the Zone of Avoidance where |b|<10^ . The solid horizontal lines show the paths of the Arecibo beam. fig2.fig" ], "images": [ null, "/dev/shm/tmppegm5khq/fig1.eps", null, "/dev/shm/tmppegm5khq/fig2.eps", null ], "texts": [ "N_ HI H i AH i SS H i MF M_ HI M _ h_ 100 ^ 1 Mpc h_ 100 Mpc^ 1 km s^ 1 cm^ 2 r_ g The Space Density of Primordial Gas Clouds near Galaxies and Groups and their Relation to Galactic HVCs Martin A. Zwaan and Frank H. Briggs Kapteyn Astronomical Institute, Postbus 800, 9700 AV Groningen, Netherlands zwaan@astro.rug.nl, fbriggs@astro.rug.nl The Arecibo \\ Strip Survey probed the halos of 300 cataloged galaxies and the environments of 14 groups with sensitivity to neutral hydrogen masses 10^ 7 . The survey detected no objects with properties resembling the High Velocity Clouds (HVCs) associated with the Milky Way or Local Group. If the HVCs were typically =10^ 7.5 objects distributed throughout groups and galaxy halos at distances of 1 Mpc, the survey should have made 70 HVC detections in groups and 250 detections around galaxies. The null detection implies that HVCs are deployed at typical distances of 200 kpc from the galaxies or group barycenters. If the clouds are in virial equilibrium, their average dark matter fraction must be 98 or higher. Local Group --- intergalactic medium --- galaxies: ISM -- galaxies: luminosity function, mass function --- radio lines: ISM After an extensive review of the High Velocity Cloud (HVC) literature, wak97 concluded that no single origin can account for the properties of the HVC population of neutral hydrogen clouds that surround the Milky Way Galaxy. Instead, several mechanisms, including infalling extragalactic clouds, cloud circulation within the Galactic halo driven by a galactic fountain, and a warped outer arm extension to the Galaxy must be invoked. The Magellanic Stream and associated HVC complexes require an explanation due to tidal interactions within the Local Group (LG) cf. put98 . A defining property for the HVCs has been the lack of associated stellar emission. This also means that spectroscopic parallax methods cannot be used to measure distances to the clouds, and the lack of known distances, in turn, hinders the calculation of the clouds' physical properties. Recently, absorption lines cf. woe99 and Balmer recombination line emission driven by reprocessing of the ionizing radiation originating in the Galactic star forming regions bla98,bla99 have been used to specify distances to a few of the clouds, indicating a range of distances from within the Galactic halo to greater than 50 kpc. Recent interest has returned to the idea that HVCs could be primordial objects raining on the Galaxy, as either remnants from the formation of the LG or as representatives from an intergalactic population of dark matter dominated mini-halos in which hydrogen has collected and remained stable on cosmological time scales ( BSTHB bli99 ; BB bra99 ). The requirement of gravitational stability without star formation places lower limits on the distances of the clouds from the Sun -- a sort of independent distance indicator that can be applied to each cloud individually, depending on its \\ flux, angular extent and velocity profile width. Typical derived distances are of order 1 Mpc, implying that this segment of the HVC population (1) inhabits the LG rather than the Galactic halo, (2) has typical \\ mass per cloud greater than 10^7 , and (3) increases the LG \\ mass budget by contributing 4 10^ 10 in the case of the BSTHB sample. For the BB sample, which is restricted to 65 confirmed ``compact HVCs'', the integral \\ content adds 10^ 9 to the LG. Further impetus to search for a primordial population of low mass objects comes from simulations of galaxy and group formation cf. kly99 , which predict of order 10 times more 10^8 mini-halos in the LG than can be counted in the dwarf galaxy population. The association of HVCs with this missing population, as well as arguments based on the kinematics of the cloud population (BSTHB and BB), makes an appealing picture for the extragalactic/LG explanation. Similar concerns motivated the extragalactic line survey by wei91 , whose study of several representative environments (clusters and voids) found only gas-rich galaxies containing stars. Several extragalactic \\ surveys of substantially larger volumes to more sensitive \\ mass limits have also found no objects with HVC properties (i.e. \\ detections without associated starlight) zwa97,spi99,kil99 . Clearly, if the HVC phenomenon is a common feature of galaxy formation and evolution, then extragalactic surveys of the halos and group environments of nearby galaxies should show evidence for this population. We take two approaches to the problem of placing the local HVC population in an extragalactic context. The first (in sections 3 and 4) is to compute the \\ mass function for the local group, both for optically selected group members and for group members plus HVC populations as modeled by BSTHB and BB. The second, separate approach (section limits.sec ) is to calculate the probability that the narrow strip that the Arecibo The Arecibo Observatory is part of the National Astronomy and Ionosphere Center, which is operated by Cornell University under a cooperative agreement with the National Science Foundation \\ Strip Survey ( ) sor94,zwa97 makes through the halos of 200 galaxy halos and 14 groups would detect members of HVC populations in those systems. The \\ mass function ( ) is used in extragalactic astronomy to quantify the space density of gas-rich galaxies and possible intergalactic clouds as a function of \\ mass. For the field galaxy population, the \\ has been determined accurately for >10^ 7.5 h_ 65 ^ 2 and can be fit satisfactorily with a faint end slope 1.2 bri93,zwa97,kil99 . At lower <10^ 7.5 h_ 65 ^ 2 , where there is considerable uncertainty due to the small number of detections, sch98 have found evidence for a steep upturn in the tail of the . Although this steep tail has a tantalizing similarity to the signature of massive HVCs, it appears that at least one of the two \\ signals responsible for the rise comes from a normal galaxy, and the other is too close to a bright star to exclude faint optical emission spi99 . We construct the \\ for the LG from \\ measurements of all known LG members as compiled recently by mat98 . Within the LG volume, a total of 22 galaxies are known in which \\ has been detected. The statistics are therefore poor, and consequently the \\ is noisy. The LG \\ is in one sense the best measured , with data over six orders of magnitude, compared to three or four orders of magnitude for determination for the field galaxy population. On the other hand, the LG \\ may suffer from severe selection effects due to obscuration by dust in the disk of the Milky Way galaxy. In order to estimate how many galaxies might have escaped detection so far, mat98 plotted the cumulative number of galaxies as a function of 1- |b| , for the total sample of LG galaxies. If LG galaxies were distributed equally over the sky, the resulting histogram should be a straight line. We applied the same method to the sample of galaxies with \\ detections and found that 4 to 7 galaxies with \\ masses >10^4 are likely to be missing at low Galactic latitude. The missing galaxies would be predominantly the ones with low optical surface brightness, but since there is no clear correlation between surface brightness and \\ mass, it is not possible to make a more refined correction to the \\ than just adding one galaxy to each half decade mass bin below =10^8 . For larger galaxies with >10^8 , we assume that the census of the LG galaxy population is complete cf. hen98 .", null, "The result is presented in the top panel of Fig. fig1.fig , where the points represent the LG \\ and the line is the field \\ derived by zwa97 scaled vertically so as to fit the points in the region around =10^9 where the curve has been measured accurately. This scaling is justified given the fact that the \\ for optically selected and \\ selected galaxies is identical cf. bri93,zwa97 . The scaling accounts for the overdensity of the LG, which amounts to a factor of 25, assuming that the LG volume is 15 Mpc ^3 . Also shown is a \\ with a steep upturn proposed by sch98 . The large divergence between the extrapolated curves from zwa97 and sch98 illustrates the uncertainty in \\ below =10^ 7.5 . The LG \\ for optically selected galaxies is remarkably flat, with the faint-end slope of a Schechter function fit of 1.0 . Studies of the \\ content of galaxies in different environments (including voids [ szo96 ] , clusters [ mcm93 ] and groups [ kra99 ] ) have shown that the shape of the \\ for > 10^8 is independent of cosmic density. The fact that we find here that the \\ of optically selected LG members is flat down to \\ masses of a few 10^4 does not, however, insure that the field \\ is also flat to these low masses. Although the crossing time of the LG is approximately equal to a Hubble time, there are clear indications of interactions and references therein mat98 . The \\ distributions in the lowest luminosity LG dwarfs are often highly asymmetric, perhaps indicative of tidal distortions. It is quite possible that low mass systems are destroyed or merged, which could cause the LG \\ to be flatter than that of the field. We now derive H i MFs for the population of extragalactic HVCs as proposed by BSTHB and BB. For the compact, isolated HVCs identified by BB, we estimate \\ masses by using their measurements of the integrated fluxes and the assumption that all clouds are at the same 1 Mpc distance, suggested by BB. Typical \\ masses are then a few times 10^7 and the sizes are approximately 15 kpc. The resulting \\ is indicated by the light shaded histogram in the second panel of Fig. fig1.fig . For the BSTHB sample we use the compilation of HVCs by WW wak91 , which forms the main source for the BSTHB analysis. For each cloud the distance \\ at which the cloud is gravitationally stable is calculated. Adopting BHTHB's value of f=0.1 for the ratio of baryonic mass to total mass leads to typical distances of self-gravitating HVCs of 1 Mpc. The dark shaded histogram represents the \\ for the BSTHB clouds, assuming that all clouds are at their distance . The resulting \\ shows a clear peak at approximately 3 10^7 , equal to the typical \\ mass estimated by BSTHB. The largest uncertainty in the determination of \\ is f , which might vary from cloud to cloud. We tested the effect of a varying f on the \\ by applying to each individual cloud a random value of f in the range 0.03 to 0.3 . The resulting \\ (shown by the unshaded histogram) is not significantly different from the \\ based on a fixed f in the region of interest >10^7 . The space density of the BB HVC population is comparable to the \\ for optically selected LG galaxies, but it is obvious that the BSTHB HVCs outnumber normal galaxies by a factor 5 to 10 in the range 10^ 7.5 < < 10^9 . This implies that if the BSTHB cloud population is typical for galaxy groups, \\ surveys in groups should have encountered 5 to 10 dark \\ clouds for every detected galaxy. This is clearly at variance with the observations. At what distances from the Milky Way must the HVCs be located so that their \\ is not in conflict with the observed field ? Since the virial distance \\ is directly proportional to f , we can test this by varying f from 0.2 to 0.025 and calculate the resulting . The results are shown in the third panel of Fig. fig1.fig . The space density of HVCs in the LG can only be brought into agreement with the observed field \\ if the median value of f is lowered to 0.02 , a value much lower than what is normally observed in galaxies. The median distance of such clouds must be smaller than 200 kpc. The Arecibo \\ Strip Survey ( ), which is discussed in detail in sor94 and zwa97 , puts limits on the space density of primordial gas clouds in external galaxy groups and around galaxies. The survey was taken in drift-scan mode and consists of two strips at constant declinations, together covering 20^ h of RA over a redshift range from 0 to 7500 . The limiting column density was 10^ 18 , which is lower than that of most of the HVCs in the WW compilation and those presented in BB. The sky coverage is small (15 deg ^2 excluding the side lobes) but the survey strip passes through the halo regions of many groups and galaxies as shown in Fig. fig2.fig . The unique character of this survey makes it therefore more suitable to assess the HVC problem than other surveys of equal size. From the Lyon-Meudon Extragalactic Database (LEDA We have made use of the Lyon-Meudon Extragalactic Database (LEDA) supplied by the LEDA team at the CRAL-Observatoire de Lyon (France). ) we selected all known galaxies with projected distances <1 h^ 1 _ 65 Mpc from the two \\ strips. The circles in Fig. fig2.fig indicate shells with 1 Mpc radii around the galaxies. Since the discussion in BSTHB is primarily focussed on galaxy groups, we also selected all cataloged groups within 1 Mpc of the strips. Galaxy groups were drawn from wil97 , who used the the Mark III catalog, and gar93 who selected groups from the LEDA galaxy sample.", null, "We fill the volumes around the selected groups and galaxies with a synthetic population of HVCs similar to the one proposed in BSTHB. To construct such an ensemble we make use of the compilation of HVCs by WW as discussed in section hvchimfs.sec . Although BSTHB put particular emphasis on galaxy groups, we choose to consider clouds around individual galaxies as well. Hierarchical formation scenarios do not distinguish between galaxies and groups in the relative number of satellites kly99 . Further motivation comes from the fact that in the LG, subclustering of dwarfs is observed around the Milky Way and M31 mat98 . The 3' beam of the Arecibo telescope subtends d_ beam =0.87 D kpc at a distance D Mpc. The typical sizes of the HVCs discussed in BSTHB are 28 kpc and the lowest column density clouds could therefore be detected out to distances of 32 Mpc, beyond which the average HVC would no longer fill the beam. HVCs with column densities in excess of the limiting value of 10^ 18 could be detected to larger distances. The limiting column density for clouds at large distances where d_ cloud <d_ beam is N_ lim =10^ 18 (d_ beam /d_ cloud )^2 . For each group and galaxy, a fraction of the volume of the surrounding sphere is scanned by the Arecibo beam. The number of clouds within that volume is calculated, taking into account the column densities and sizes of the individual clouds. Table table1.tab lists the number of clouds that would have been detected in the Arecibo \\ strip survey if a population of extragalactic HVCs existed with the BSTHB properties. We calculate the numbers of clouds differently for groups and for galaxies. Since BSTHB do not specify the exact radial distribution of clouds, we tested three different radial distribution functions to fill the volumes with clouds: 1) a spherical volume of radius R , 2) a thin spherical shell of radius R , and 3) a thick spherical shell with clouds distributed according to a Gaussian about the radius R with dispersion =R/3 . The latter distribution most closely resembles the derived distribution of r_ g given in BSTHB. The numbers in the table are based on R=1 Mpc , the value preferred by both BSTHB and BB. The group halos are filled with 450 clouds, the number of HVCs identified by WW, excluding complexes A, C, M, the Outer Arm and the Magellanic Stream. To calculate the expected number of clouds around galaxies, the number of clouds associated with each galaxy is scaled in direct proportion to the ratio of the galaxy luminosity compared to the integral luminosity of the LG. This leads to a median number of clouds per galaxy of 40. Table table1.tab shows that the expected number of detections is essentially independent of how the clouds are distributed around the groups and galaxies. For these samples we should detect approximately 250 clouds around galaxies and 70 around groups if the HVC population of BSHTB is a common feature of nearby galaxies. Restricting our analysis to the compact clouds of BB, reduces these numbers to 39 and 9. For a uniformly filled spherical distribution of BSTHB clouds, the distribution of \\ masses of the expected detections is shown in the lowest panel in Fig. fig1.fig . This figure illustrates that our analysis is sensitive to typical \\ clouds (compare second panel of Fig. fig1.fig ), and not only to the most massive ones. The robustness of the result is demonstrated in Table table1.tab , where the numbers in parentheses indicate the expected number of detections if the detection threshold is increased from 5 to 7 . The average decrease is 25 . A 50 decrease occurs if the detection threshold is set at 10 . This extremely conservative threshold would still predict more than 100 detections. The hypothesis that HVCs are primordial gas clouds with typical \\ masses of a few 10^7 at distances of 1 Mpc from the Galaxy is not in agreement with observations of nearby galaxies and groups. Blind \\ surveys of the extragalactic sky would have detected these clouds if they exist around all galaxies or galaxy groups in numbers equal to those suggested for the Local Group. These results are highly significant: the Arecibo \\ strip survey would have detected approximately 250 clouds around individual galaxies and 70 in galaxy groups. We are grateful to B. Wakker for discussions and for providing the list of HVC parameters. J. Bland-Hawthorn, L. Blitz, R. Braun, J. van Gorkom, and H. van Woerden are thanked for useful comments. Bland-Hawthorn, J., Veilleux, S., Cecil, G. N., Putman, M. E., Gibson, B. K., Maloney, P. R. 1998, , 299, 611 Bland-Hawthorn, J., Maloney, P. R. 1999, , 510, L33 Blitz, L., Spergel, D. N., Teuben, P. J., Hartmann, D., Burton, W. B. 1999, , 514, 818 Braun, R., Burton, W. B 1999, , 341, 437 Briggs, F. H., Rao, S. 1993, , 417, 494 Garcia, A. M. 1993, A AS, 100, 47 Henning, P. A., Kraan-Korteweg, R. C., Rivers, A. J., Loan, A. J., Lahav, O., Burton, W. B 1998, AJ, 115, 584 Kilborn, V., Webster, R. L., Staveley-Smith 1999, PASA, 16, 8 Klypin, A. A., Kravtsov, A. V., Valenzuela, O., Prada, F. 1999, , submitted Kraan-Korteweg, R. C., van Driel, W., Briggs, F. H., Binggeli, B., Mostefaoui, T. I. 1999, A AS, 135, 255 Mateo, M. L. 1998, , 36, 435 McMahon, P. M. 1993, Ph.D Thesis, Columbia University Putman, M. E. et al. 1998, , 394, 752 Schneider, S. E., Spitzak, J. G., Rosenberg, J. L. 1998, , 507, L9 Sorar, E. 1994, Ph.D. Thesis, University of Pittsburgh Spitzak, J. G., Schneider, S. S. 1999, , 119, 159 Szomoru, A., van Gorkom, J. H., Gregg, M. D., Strauss, M. A. 1996, , 111, 2150 Szomoru, A., Guhathakurta, P., van Gorkom, J. H., Knapen, J. H., Weinberg, D. H., Fruchter, A. S. 1994, , 108, 491 Wakker, B. P., van Woerden, H. 1991, , 250, 509 Wakker, B. P., van Woerden, H. 1997, , 35, 217 van Woerden, H., Schwarz, U. J., Peletier, R. F., Wakker, B. P., Kalberla, P. M. W. 1999, , 400, L138 Weinberg, D. H., Szomoru, A., Guhathakurta, P., van Gorkom, J. H. 1991, , 372, L13 Willick J. A., Courteau S., Faber S. M., Burstein D., Dekel A., Strauss M. A. 1997, , 109, 333 Zwaan, M. A., Briggs, F. H., Sprayberry, D., Sorar, E. 1997, , 490, 173 lrrrrr 6 0pc Expected extragalactic HVCs detections & 3 BSTHB & & 2-4 & uniform & shell & Gaussian & BB & selected groups & 70(52) & 72(54) & 70(52) & 9(8) & 14 galaxies & 248(167) & 256(173) & 260(177) & 39(28) & 347 The number of expected HVC detections are calculated assuming a 5 detection threshold. The numbers in parentheses indicate the expected number if the threshold is raised to 7 . The last column shows the number of groups and galaxies that are included in the analysis. For the Gaussian distributions a selection radius of 2 Mpc has been used for inclusion in the sample." ] }
astro-ph0001018
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "bhpt Simple boxplots illustrating CN bandstrength ranges in RGB and AGB stars of NGC 6752, M13, and M4 are shown. For all of the individual abundance boxes, the horizontal line inside a box indicates the median value of S(3839). The vertical boundaries of a box show the interquartile range (the middle 50 of the data). The vertical tails extending from the boxes indicate the total range of S(3839), excluding outliers. Mild outliers (those between 1.5 and 3 times the interquartile range) are denoted by open circles. No severe outliers (those greater than 3 times the interquartile range) are present in these data. The number of stars included in each boxplot is noted in parentheses. The basic data (estimates of CN line blocking index S(3839)) were taken from the following sources: (a) NGC6752: Norris (1981), and we have used their suggested relationship of the lower limit of S(3839) to V magnitude to compute the relative CN bandstrengths S(3839); (b) M13: converted photometric CN index m (CN) values obtained by Suntzeff (1981). We first removed the variation of the index due to position in the -m diagram using Suntzeff's formula, and then converted the results to S(3839) values using the relationship we found for the NGC 6752 stars in common to the m (CN) values presented by Langer (1992) and the S(3839) values of Norris (1981). () M4: Norris (1981) or Suntzeff Smith (1991), or the mean of both, using Norris' suggested relationship of the lower limit of S(3839) to V magnitude.", "bhpt Boxplots of sodium abundances in M13 (Pilachowski 1996b) and field giant stars (Pilachowski 1996a). The statistical abundance distributions represented by each box's vertical boundaries, etc., are as described in Figure 1. The different boxes illustrate the sodium abundance distributions found for RGB tip stars, lower luminosity giants, and AGB stars. For contrast, a boxplot representing sodium abundances in field giants is also shown in this figure." ], "images": [ null, "/dev/shm/tmp5h77tp7r/deltacn2.eps", null, "/dev/shm/tmp5h77tp7r/nam132.eps", null ], "texts": [ "\\ The recent improvements in globular cluster colour-magnitude diagrams, coupled with an increase in large-sample spectroscopic abundance studies of cluster giants, finally allow some attempts at a statistically meaningful comparison of the chemical compositions of red giant branch and asymptotic branch cluster stars. We review some of the extant data here, concluding that in a few clusters the AGB stars show on average smaller amounts of high-temperature proton-capture synthesis products (low oxygen, high sodium and aluminum) at their surfaces than do the first-ascent RGB stars. This suggests that those RGB stars with envelopes that have been enriched with proton-capture material also have high helium contents. Such stars after the He-flash then take up residence on the bluest parts of the HB (as a consequence of their high envelope helium), probably never to return to the AGB during subsequent evolutionary stages. Asymptotic giant branch (AGB) stars of all populations have basically the same interior structures, with shell fusion zones of He and H surrounding C-O cores. Most AGB stars will undergo episodes of mass loss that eject their outer envelopes, leaving the exposed cores to fade away as white dwarfs. Thus low mass, metal-poor halo AGB stars and their higher mass, metal-rich disk counterparts exist in the same evolutionary domains and share the same eventual fates; all these stars appear to be theoretically quite similar. Observationally however, the properties of disk and halo population AGB stars are quite distinct. The high mass, high metallicity AGB stars are both extremely luminous and extremely cool. Sometimes they are surrounded by substantial gas/dust shells of their own making, and thus present unique photometric signatures (especially in the infrared). Often they exhibit spectroscopic peculiarities (strong carbon-containing molecular and neutron-capture element features) indicative of nuclear processing in their He fusion zones. These are the stars that are treated in most of the contributions to this workshop. In contrast, the low mass, low metallicity stars that can be positively associated with the AGB are photometrically and spectroscopically somewhat difficult to distinguish from first-ascent red giant branch (RGB) stars. In globular cluster colour-magnitude (-m) diagrams the AGB is a thinly-populated stream of stars connecting the red end of the horizontal branch (HB) and the end of the RGB. Over the brightest 1--2 magnitudes, the AGB and RGB are separated by less than 0.2 in V magnitude at a given B--V colour (a particularly clear example is the M3 -m diagram of Buonanno \\ 1986). Globular cluster AGB stars will not become extremely luminous because they are former HB stars, whose masses cannot in theory exceed M 0.6 M _ ; the second-ascent AGB tip effectively merges with the first-ascent RGB tip. Unfortunately, for many globular clusters, photometry precise enough to cleanly separate the AGB from the RGB for most candidate stars still does not exist. The spectra of globular cluster AGB stars also do not differ radically from those of RGB stars. CH stars, those possessing spectroscopic evidence of having possibly mixed He shell burning products (carbon, neutron-capture elements) to their surfaces, are apparently very rare in globular clusters; only a handful have been discovered ( , McClure Norris 1977; Cowley Crampton 1985; Vanture Wallerstein 1992; C ote \\ 1997). In recent years photometrists and spectroscopists have combined efforts to substantially increase the quantity and quality of data on AGB stars in globular clusters. In this paper, we look for chemical composition differences between AGB and RGB stars in three globular clusters, concluding that there is some evidence suggesting that AGB stars have less chemically evolved surface layers. This suggestion is then related to the ``second parameter problem'' of globular clusters. Several decades of spectroscopic investigations have established the reality of large-scale star-to-star abundance variations among light elements in globular cluster stars. The variations are not of the same magnitude in all clusters, and indeed each cluster seems to have a chemical composition signature that is not repeated exactly in other clusters. Most of the abundance inhomogeneities observed in globular clusters involve some aspects of so-called ``proton-capture'' nucleosynthesis. Extensive reviews of these abundance variations have been published by , Kraft (1979), Freeman Norris (1981), Smith (1987), Suntzeff (1993), Briley \\ (1994), Kraft (1994), and Sneden (1998,1999). Some general statements about cluster nucleosynthesis are summarized here without attribution to specific papers, and the reader is strongly encouraged to consult the reviews and the original papers quoted in them for details on these abundance trends. The CN cycle: The chief products of ordinary CN cycle fusion are observed at the surfaces of most RGB and AGB stars. That is, the carbon isotope ratios are uniformly low (4 10), carbon abundances are usually low (--0.3 [ C/Fe ] --1.3), and nitrogen abundances are correspondingly very high (+0.5 [ N/Fe ] +1.5). However, the N overabundances are sometimes far greater than the amounts that would be predicted from simple C N conversion. The ON cycle: Globular cluster giants, unlike almost all halo field giants, often exhibit very depleted oxygen abundances (--1.0 [ O/Fe ] +0.4). This suggests that the ON cycle, which requires higher temperatures ( T 40 10 ^6 K) in hydrogen fusion zones than does the CN cycle, has been active either in the giants that are being observed or in an earlier cluster generation. This cycle's major net effect is O N conversion, and can therefore account for the anomalously large N abundances mentioned above. Finally, in nearly all cluster giants with complete CNO abundance data, the C+N+O abundance sum appears to be conserved, adding further weight to the idea that the variations in these elements are simply due to the combined CN and ON element re-shufflings. The NeNa cycle: Sodium abundances also vary widely among globular cluster giants (--0.3 [ Na/Fe ] +0.4). The same globular cluster giants that have low O abundances almost invariably have high Na abundances; an anticorrelation between these abundances apparently occurs in all lower metallicity clusters ( [ Fe/H ] < --1) studied to date. This anticorrelation suggests that the NeNa proton-fusion cycle, which can work efficiently at the same temperatures as does the ON cycle, has at some time in globular cluster histories converted Ne (undetectable in cluster giant spectra) into Na. The MgAl cycle: Aluminum abundances also have large star-to-star variations that are anticorrelated with O abundances, and in some well-studied clusters the anticorrelation extends also to Mg abundances. Again, proton-capture fusion leading to Mg Al conversion is the probable culprit (Shetrone 1996), but the burning temperature requirements (T 70 10 ^6 K) are large enough that it is difficult to imagine low mass globular cluster giants performing the MgAl cycle, unless such such transmutations occur as the result of a thermal instability of the H or He shell source (Langer \\ 1997, Powell 1999). Alternatively, stars with abnormally large Al abundances might either have been born with them, created in previous higher mass stars, or have accreted them from the winds of higher mass AGB stars. Other Nucleosynthesis effects: Some significant cluster-to-cluster abundance differences are seen in heavier elements that cannot be altered in proton-capture synthesis reactions. For example, the very heavy elements Ba, La, and Eu can have very different abundance ratios in different clusters, indicating varying contributions of slow and rapid neutron-capture synthesis reactions to the creation of these elements. Among the elements that participate in the major nuclear fusion chains, silicon should only be altered during the last stages of very high mass stars. But its mean abundance varies from cluster to cluster; some globular clusters have Si abundances nearly a factor of two larger than those of typical halo field giants. And in addition to the star-to-star variations of Al abundances within individual clusters, the Al mean abundance level also differs substantially from cluster to cluster. All of these abundance anomalies point to nucleosynthesis contributions of multiple generations of stars in a given cluster, either from stars that died before the present stars were born or during their formation. Also, the relatively small numbers of high mass stars that must have existed in or preceded formation of each cluster probably produced supernovae of different masses in each cluster, creating distinct ``initial'' abundance distributions in each cluster. Perhaps the first suggestion that AGB and RGB stars in some clusters might on average have different compositions was made by Norris \\ (1981). In a large-sample study of CN bandstrengths among giants of NGC 6752, they found that there is a bi-modal distribution of CN bandstrengths that is nearly independent of RGB position. But they suggested that there is a nearly uni-modal set of CN bandstrengths among the AGB stars: their CN bands are almost all weak. Norris \\ presented this situation in their Figure 3, plotting the CN absorption index S(3839) as a function of V magnitude and B--V colour. We have used the formula developed by Norris \\ to convert S(3839) to a CN bandstrength indicator that is independent of stellar temperature/gravity effects, and in Figure 1 we show ``boxplots'' that illustrate the ranges in CN bandstrength found in RGB and AGB stars. The lower CN strengths of the AGB stars on average is obvious, but just as important is the near total lack of any CN strong AGB stars in this cluster. For comparison, we also show similar data for two other clusters, M4 and M13. The M4 CN bandstrength data are taken from either Norris (1981) or Suntzeff and Smith (1991), or the mean of both, where the variation of S(3839) with position in the -m diagram has been removed according to Norris' formula. The evolutionary status of the stars are those determined by Ivans in their H-R diagram of Figure 12, which illustrates the reddening-free positions of the stars. For the M13 data, we referred to Suntzeff (1981), where we converted the photometric m (CN) indices to relative photometric bandstrengths m (CN) using Suntzeff's suggested relationship of the lower limit of m (CN) to B--V colour index. We further transformed the m (CN) values to S(3839) relative bandstrengths employing the relationship we derived for m (CN) and S(3839) found using stars in common between the studies of NGC 6752 stars by Langer \\ (1992), who used m (CN), and Norris \\ (1981) who used S(3839). The evolutionary status of the M13 stars were those determined by Suntzeff (1981) and, in the cases where the photometry made the status ambiguous, we supplemented the information using the stars in common studied by Pilachowski \\ (1996b). Thus, the distributions shown in Figure 1 are all, in effect, on the S(3839) system of Norris \\ (1981) and only include the stars for which AGB vs RGB designations are unambiguous.", null, "Norris \\ (1981) offered two possible explanations for the relatively weak CN bandstrengths in NGC 6752 AGB stars; both explanations involve an inability of the strong-CN RGB stars to ascend the giant branch a second time after HB evolution. In one scenario, some cluster stars would have been born with abnormally large C and/or N abundances, accompanied by larger-than-average He/H ratios (presumably from the CN and/or ON cycles). Stellar evolution computations ( , Lee \\ 1994, and references therein) have shown that RGB stars with higher He contents will, after they undergo the He flash, take up residence in bluer parts of the HB than do otherwise identical stars with lower He contents. In fact, these stars may arrive at such a blue HB position that they may eventually evolve directly to the white dwarf track, entirely avoiding the AGB stage. In the other scenario, larger internal mixing in CN-strong stars during RGB evolution might drive large amounts of mass loss, leading to lower-than-average envelope masses after the He flash. Again, such stars would wind up on the bluer end of the HB, possibly never to return as AGB stars. Thus the stars observed on the AGB of NGC 6752 may have weak CN bands because they are the former RGB stars that had little mixing of CN cycle products (N and He) into their envelopes; they are the ones that survived the HB stage to rise again toward the giant branch tip. This latter hypothesis can be tested by comparing abundances of light proton-capture elements (C, N, O, Na, Mg, Al) in AGB and RGB stars of NGC 6752. Unfortunately, the extant high resolution spectroscopic studies of NGC 6752 giants (Gratton 1987, Norris Da Costa 1995, Minniti \\ 1996, Shetrone 1998) were only able to include the brightest stars near the RGB tip, where the distinction between RGB and AGB stars cannot be made. Pilachowski \\ (1996b) derived sodium abundances for 130 giants in M13; their program stars ranged from those at the RGB tip to ones about as faint as the HB. They showed that Na abundances of most M13 giants are greater than those of similar-metallicity halo field stars, but there are some significant differences between RGB and AGB stars in this cluster. We illustrate this situation in Figure 2 with another boxplot, in which we compare [ Na/Fe ] ratios for lower luminosity M13 RGB stars (those with log g > 1), RGB tip stars (log g < 1), and AGB stars, along with [ Na/Fe ] ratios for field stars in the metallicity range --1.2 > [ Fe/H ] > --1.9 (Pilachowski \\ 1996a). The higher Na abundances of M13 giants is obvious in this figure, but it is also clear that the AGB stars have lower mean Na abundances than do the RGB tip stars, and that they have a narrower range in Na. It is possible that the Na abundances for the very cool RGB tip stars must be corrected downward somewhat to correct for departures from LTE ( , Gratton \\ 1999). But oxygen abundances in cluster giants are always determined from the [ O I ] transitions, which do not suffer substantial departures from LTE. And in M13 not only do the RGB stars exhibit on average the largest Na abundances but they also have the lowest O abundances (Kraft \\ 1997). Therefore the difference between the the mean levels of Na in AGB and RB stars in M13 is probably real.", null, "Pilachowski \\ (1996) followed a line of reasoning similar to that of Norris \\ (1981) in supposing that the presently observed AGB stars in M13 are those whose envelope He contents remained relatively low when they were RGB stars; the RGB stars with elevated He took up residence on the blue part of the HB and never arrived on the AGB. However, the Pilachowski \\ scenario differed from that of Norris \\ in one important respect: the RGB stars with elevated He were those that had contaminated their atmospheres with material that had been processed through the CNO hydrogen-burning shell, in accordance with the deep mixing scenario and nuclear transmutation calculations of Langer \\ (1993), Langer Hoffman (1995) and Cavallo \\ (1998). Sweigart (1997a,b) showed that such ``deep mixed'' stars could indeed be moved sharply to the blue in their subsequent evolution onto the HB, largely as a result of increased mass loss prior to the helium core flash. Pilachowski \\ also noted that since M13 has the most extreme cases of Na and Al enhancements and O depletions among RGB stars, it also probably has a higher percentage of high-He stars than other globular clusters. If so, then the AGB of M13 ought to be relatively unpopulated. This view is supported by the statistics of Caputo \\ (1978) and Buzzoni \\ (1983), from which one finds that M13 has the lowest ratio (by a factor of 2) of AGB to RGB stars among the 16 clusters studied. M13 and NGC 6752 represent the clearest cases for chemical composition differences between AGB and RGB stars in globular clusters. But truth in advertising compels us to admit that the situation is probably far more complex than we have suggested so far. Smith Norris (1993) suggested that the AGB stars of M5 have a different CN bandstrength distribution: ``... the observations reported in this paper yield no consistent picture of the CN distributions among stars in more advanced stages of evolution. The asymptotic giant branch appears to be deficient in CN-weak stars for M5, but deficient in CN-strong stars for NGC 6752.'' Consideration of these differences has been made possible by the existence of very large bandstrength or abundance samples in these two clusters. Unfortunately, most other globulars have not been studied in sufficient detail to assess the chemical compositions of AGB stars. In their study of a large number of bright giants in M4, Ivans \\ (1999) found some of the same correlated variations in proton-capture elements that have been seen in other clusters. Their data were most extensive for the determination of oxygen abundances, and they concluded that the mean oxygen abundance of M4 AGB stars is slightly larger than that of the RGB stars. This provides mild further support for the suggestion that AGB stars in globular cluster are on average less chemical evolved in the proton-capture elements than are RGB stars. This problem cannot be effectively dealt with until stellar samples in many globular clusters include at least 10 AGB stars, as well as many more RGB stars over a large luminosity range. Enough detailed high resolution, large wavelength coverage spectroscopic studies of individual stars in selected globulars to make it clear that the proton-capture phenomenon is ``universal''. Thus in addition to the continued full-scale abundance analyses of the brightest cluster members, it will be especially fruitful to now survey cluster giant branches with multi-object spectrometers (in the manner of Pilachowski \\ 1996b) that concentrate on fairly complete descriptions of the abundance trends of just one or two elements that will stand as surrogates for the behaviour of the whole set of proton-capture elements. We thank Raffaele Gratton for helpful discussions on this work. This research was supported by NSF grants AST-9217970 to RPK AST-9618364 to CS. Travel support given by the Rome Observatory to CS is gratefully acknowledged. Briley, M. M., Bell, R. A., Hesser, J. E., Smith, G. H.: 1994, Can. J. Phys. 72 , 772. Buonanno, R., Buzzoni, A., Corsi, C. E., Fusi-Pecci, F., Sandage, A. R.: 1986, 57 391. Cavallo, R. M., Sweigart, A. V., Bell, R. A.: 1998, 492 575. C ote, P., Hanes, D. A., Mclaughlin, D. E., Bridges, T. J., Hesser, J. E., Harris, G. L. H.: 1997, 476 L15. Cowley, A. P., Crampton, D.: 1985, 97 835. Freeman, K. C., Norris, J.: 1981, 19 319. Gratton, R.: 1987, 177 177. Gratton, R. G., Carretta, E., Eriksson, K., Gustafsson, B.: 1999, 350 955. Ivans, I. I., Sneden, C., Kraft, R. P., Suntzeff, N. B., Smith, V. V., Langer, G. E., Fulbright, J.: 1999 118 1273. Kraft, R. P.: 1979, 17 309. Kraft, R. P.: 1994, 106 553. Kraft, R. P., Sneden, C., Smith, G. H., Shetrone, M. D., Langer, G. E., Pilachowski, C. A.: 1997, 113 279. Langer, G. E., Hoffman, R.: 1995, 107 1177. Langer, G. E., Hoffman, R., Sneden, C.: 1993, 105 301. Langer, G. E., Hoffman, R., Zaidins, C.: 1997, 109 244. Langer, G. E., Suntzeff, N. B., Kraft, R. P.: 1992 104 523. Lee, Y. -W., Demarque, P., Zinn, R.: 1994, 423 248. McClure, R. D., Norris, J.: 1977, 217 L101. Minniti, D., Peterson, R. C., Geisler, D., Claria, J.: 1996, 470 953 Norris, J.: 1981, 248 177. Norris, J., Cottrell, P. L., Freeman, K. C., da Costa, G. S.: 1981, 244 205. Norris, J. E., Da Costa, G. S.: 1995, 447 680. Pilachowski, C. A., Sneden, C., Kraft, R. P.: 1996a, 111 1689. Pilachowski, C. A., Sneden, C., Kraft, R. P., Langer, G. E.: 1996b, 112 545. Powell, D. C.: 1999, 111 1186. Shetrone, M. D. 1998, in Fundamental Stellar Properties: The Interaction between Observation and Theory , IAU Symp. 189, T. R. Bedding ed., School of Physics: Univ. of Sydney, p. 158 Shetrone, M. D. 1996, 112 2639. Smith, G. H.: 1987, 99 67. Smith, G. H., Norris, J. E.: 1993, 105 173. Sneden, C.: 1998, in Galaxy Evolution: Connecting the Distant Universe with the Local Fossil Record , M. Spite ed., Kluwer Academic Publ., in press. Sneden, C.: 1999, in 35th Liege International Astrophysics Colloquium: The Galactic Halo, from Globular Clusters to Field Stars , P. Magain, A. Noels eds., in press. Suntzeff, N. B.: 1981 47 1. Suntzeff, N. B. 1993, in The Globular Cluster-Galaxy Connection , G. H. Smith and J. B. Brodie eds., ASP Conf. Ser. 48, p. 167. Suntzeff, N. B., Smith, V. V.: 1991 381 160. Sweigart, A. V.: 1997a, 474 L23. Sweigart, A. V. 1997b, in Third Conference on Faint Blue Stars , A.G.D. Philip, J. W. Liebert and R.A. Safford eds., L. Davis Press, Schenectady, p. 3. Vanture, A. D., Wallerstein, G.: 1992, 104 888." ] }
astro-ph0001028
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "Fig1. Light curves of 2155-304; Fig 2. V_ m ^ 2 vs trial period of 2155-304; Fig 3. DCF analysis for 2155-304; Fig 4. The simulation light curve (solid curve) and the observation B light curve for 2155-304" ], "images": [ null, "/dev/shm/tmpftf06h3v/fan2155.eps" ], "texts": [ "The variability analysis of PKS 2155-304 J.H. Fan Center for Astrophysics, Guangzhou Normal University, Guangzhou 510400, China, e-mail: jhfan@guangztc.edu.cn Chinese Academy of Sciences-Peking University Joint Beijing Astrophysical Center (CAS-PKU.BAC), Beijing, China R.G. Lin Center for Astrophysics, Guangzhou Normal University, Guangzhou 510400, China Received <date>;accepted<date> Received<data>;accepted<Oct. 1999> In this paper, the post-1977 photometric observations of PKS 2155-304 are compiled and used to discuss the variation periodicity. Largest amplitude variations ( U = 1^ m .5 ; B = 1^ m .65 ; V = 1^ m .85 ; R = 1^ m .25 ; I = 1^ m .14 ) and color indices ( (B-V) = 0.30 0.06 ; (U-B) = -0.72 0.08 ; (B-R) = 0.62 0.07 ; (V-R) = 0.32 0.04 ) are found. The Jurkevich's method and DCF (Discrete Correlation Function) method indicate possible periods of 4.16-year and 7.0-year in the V light curve. BL Lacertae objects:individual: PKS 2155-304-galaxies:variation- periodicity BL Lac objects are a special subclass of active galactic nuclei (AGNs) showing some extreme properties: rapid and large variability, high and variable polarization, no or only weak emission lines in its classical definition. BL Lac objects are variable not only in the optical band, but also in radio, infrared, X-ray, and even ray bands. Some BL Lac objects show that the spectral index changes with the brightness of the source (Bertaud et al. 1973; Brown et al. 1989; Fan 1993), generally, the spectrum flattens when the source brightens, but different phenomenon has also been found (Fan et al. 1999). The nature of AGNs is still an open problem; the study of AGNs variability can yield valuable information about their nature, and the implications for quasars modeling are extremely important ( see Fan et al. 1998a). PKS 2155-304, the prototype of the X-ray selected BL Lac objects and TeV ray emitter (Chadwick et al. 1999), is one of the brightest and the best studied objects. Its spectrum from 3600 to 6800 appears blue (B-V < 0.1) and featureless (Wade et al. 1979). A 0.17 redshift was claimed from the probably detected weak [ O III ] emission feature ( Charles et al. 1979), which was not detected in Miller McAlister (1983) observation. Later, a redshift of 0.117 was obtained from several discrete absorption features (Bowyer et al. 1984). PKS 2155-304 varies at all observation frequencies and is one of the most extensively studied objects for both space-based observations in UV and X-ray bands (Treves et al. 1989; Urry et al. 1993; Pian et al. 1996; Giommi et al. 1998) and multiwavelength observations (Pesce et al. 1997 ). Variation over a time scale of one day was observed (Miller Carini 1991) and that over a time scale of as short as 15 minutes is also reported by Paltani et al. (1997) in the optical band. Differently brightness-dependent spectrum properties are found ( see Miller McAlister 1983; Smith Sitko 1991; Urry et al. 1993; Courvoisier et al. 1995; Xie et al. 1996; Paltani et al. 1997). In this paper, we will investigate the periodicity in the light curve and discuss the variation as well. The paper has been arranged as follows: In section 2, the variations are presented and the periodicities are searched, in section 3, some discussions and a brief conclusion are given. Light curves The optical data used here are from the literature: Brindle et al. (1986); Carini Miller (1992); Courvoisier et al. (1995); Griffiths et al. (1979); Hamuy Maza (1987); Jannuzi et al. (1993); Mead et al. (1990); Miller McAlister (1983); Pesce et al. (1997); Smith Sitko (1991); Treves et al. (1989); Urry et al. (1993); Xie et al. (1996) and shown in Fig. 1a-e. From the data, the largest amplitude variabilities in UBVRI bands are found: U = 1^ m .5 (11^ m .87- 13^ m .37) ; B = 1^ m .65 (12^ m .55 - 14^ m .20) ; V = 1^ m .85 (12^ m .27 - 14^ m .13) ; R = 1^ m .25 (11^ m .96 - 13^ m .21) ; I = 1^ m .14 (11^ m .55 - 12^ m .69) . and color indexes are found: (B-V) = 0.30 0.06 (N=140 pairs); (U-B) = -0.72 0.08 (N=105 pairs); (B-R) = 0.62 0.07 (N=90 pairs); (V-R) = 0.32 0.04 (N=98 pairs), the uncertainty is 1 dispersion. Periodicity The photometric observations of PKS 2155-304 indicate that it is variable on time scales ranging from days to years (Miller McAlister 1983). Is there any periodicity in the light curve? To answer this question, the Jurkevich (1971) method is used to search for the periodicity in the V light curve since there are more observations in this band. The Jurkevich method (Jurkevich 1971, also see Fan et al. 1998a) is based on the expected mean square deviation and it is less inclined to generate spurious periodicity than the Fourier analysis. It tests a run of trial periods around which the data are folded. All data are assigned to m groups according to their phases around each trial period. The variance V_i^2 for each group and the sum V_m^2 of all groups are computed. If a trial period equals the true one, then V_m^2 reaches its minimum. So, a ``good'' period will give a much reduced variance relative to those given by other false trial periods and with almost constant values. To show the significance of the trial periodicity, we adopted the F test (see Press et al. 1992). When the Jerkevich method is used to V measurements, some results are obtained and shown in Fig. 2 ( m = 10 ), which shows several minima corresponding to trial periods of less than 4.0-year and two broad minima corresponding to averaged periods of (4.16 0.2) and (7.0 0.16) years respectively. For the periods, which are smaller than 4.0-year, we found that the decrease of the V_m^2 is less than 3 times of the noise suggesting that it is difficult for one to take them as real signatures of periods, i.e., those periods should be discussed with more observations. For the two broad minima, the F test is used to check their reality. The significance level is 93.8 for the 4.16-year period and 96.2 for the 7.0-year period. PKS 2155-304 was observed more than 100 years ago. Griffiths et al. (1979) constructed the annually averaged B light curve up to the 1950's from Harvard photographic collection. But there are only a few observations during the period of 1950-1970. The periodicity obtained here (see Fig. 2) are based on the post-1977 data. For comparison, we adopted the DCF (Discrete Correlation Function) method to the V measurements. The DCF method, described in detail by Edelson Krolik (1988) (also see Fan et al. 1998b), is intended for analyses of the correlation of two data set. This method can indicate the correlation of two variable temporal series with a time lag, and can be applied to the periodicity analysis of a unique temporal data set. If there is a period, P , in the lightcurve, then the DCF should show clearly whether the data set is correlated with itself with time lags of = 0 and = P . It can be done as follows. Firstly, we have calculated the set of unbinned correlation (UDCF) between data points in the two data streams a and b , i.e. UDCF_ ij = (a_ i a ) (b_ j b ) _ a ^2 _ b ^2 , where a_ i and b_ j are points in the data sets, a and b are the average values of the data sets, and _ a and _ b are the corresponding standard deviations. Secondly, we have averaged the points sharing the same time lag by binning the UDCF_ ij in suitably sized time-bins in order to get the DCF for each time lag : DCF( ) = 1 M UDCF_ ij ( ), where M is the total number of pairs. The standard error for each bin is ( ) = 1 M-1 [ UDCF_ ij DCF( ) ] ^ 2 ^ 0.5 . The resulting DCF is shown in Fig. 3. Correlations are found with time lags of (4.20 0.2) and ( 7.31 0.16) years. In addition, there are signatures of correlation with time lags of less than 3.0 years. If we consider the two minima in both the right and left sides of the 7.0-year minimum, then we can say that the periods of 4.16 and 7.0-year found with Jerkevich method are consistent with the time lags of 4.2-year and 7.3-year found with the DCF method. These two periods are used to simulate the light curve (see the solid curve in Fig. 4). It is clear that the solid curve does not fit the observations so well. One of the reasons is that there are probable more than two periods ( 4.2 and 7.0 years) in the light curve as the results in Fig 2 and 3 indicate. Another reason is that the derived period is not so significant as Press (1978) mentioned. Press argued that periods of the order the third of the time span have a large probability to appear if longer-term variations exist. The data used here have a time coverage of about 16.0 years, i.e., about 3 times the derived periods. Therefore, these are only tentative and should be confirmed by independent work.\" From the data, the largest amplitude variations are found for UBVRI bands with I and R bands showing smaller amplitude variations. One of the reasons is from the fact that there are fewer observations for those two bands, another reason is perhaps from the effect of the host galaxy, which affects the two bands more seriously. In this paper, the post-1970 UBVRI data are compiled for 2155-304 to discuss the spectral index properties and to search for the periodicity. Possible periods of 4.16 and 7.0 years are found. We are grateful to the referee for his/her comments and suggestionsThis work is support by the National Pan Deng Project of China and the National Natural Scientific Foundation of China Bertaud CH., Wlerick G., V e ron P., et al. 1973, A AS 11, 77 Bowyer S., Brodie J., Clarke J. T., Henry J.P. 1984, ApJ 278, L103 Brindle C., Hough J.H., Bailey J.A., et al. 1986, MNRAS 221, 739 Brown L.M.J., Robson E.I., Gear W.K., Smith M.G.: 1989, ApJ 340, 150 Chadwick P.M. et al. 1999, ApJ 513, 161 Carini M.T., Miller H.R., 1992, ApJ 385, 146 Charles P., Thorstensen J., Bowyer S., 1979, Nat. 281, 285 Courvoisier T.J.L., Blecha A., Bouchet P., et al., 1995, ApJ 438, 108 Edelson, R.A. Krolik, J.H., 1988, ApJ, 333, 646 Fan J.H., Xie G.Z., Pecontal E., et al., 1998a, ApJ, 507, 173 Fan J.H., Adam G. Xie G.Z. et al., 1998b, A AS, 136, 217 Fan J.H.: 1993, Ph.D. Thesis, Yunnan Astronomic Observatory, Chinese Academy of Sciences. Fan J.H., Xie G.Z., Adam G. et al. 1999, ASP Conf. Ser. Vol. 159, Eds. L. Takalo A. Sillanpaa, p99 Giommi P., Barr P., Garilli B., et al. 1991, ApJ 356, 432 Giommi P., Fiore F., Guainazzi M., et al. 1998, A A 333, L5 Griffiths R. E., Briel U., Chaisson L., Tapia S., 1979, ApJ 234, 546 Hamuy M., Maza J., 1987, A AS 68, 383 Jurkevich I. 1971, Ap SS 13, 154 Jannuzi B.T., Smith P.S., Elston R., 1993, ApJS 85, 265 Mead A.R.G., Ballard K.R., Brand P.W.J.L., et al. 1990, A AS 83, 183 Miller H.R., Carini M.T., 1991, in Variability of Active Galactic Nuclei, eds. H.R. Miller and P.J. Wiita (Cambridge: Cambridge Univ. Press), p256 Miller H.R., McAlister H.A., 1983, ApJ 272, 26 Paltani S., Courvoisier T.J.L., Blecha A., Bratschi P., 1997, A A 327 539 Pesce J.E., Urry C.M., Maraschi L., et al. 1997, ApJ 486, 770 Pian E., Urry C.M., Pesce J.E. et al. 1996, in 'Blazar Continuum Variability' eds. H.R. Miller, J.R. Webb, and J.C. Noble, ASP Conf. Series VoL. 110, 417 Press W.H. 1978 Comm. on Astrophys. 7, 103 Press W.H., Teukolsky S.A., Vetterling W.T., Flannery B.P. 1992, in Numerical Recipes in C, The Art of Scientific Computing (2nd Edition), Cambridge Uni. Press. Smith P.S., Sitko M.L., 1991, ApJ 383, 580 Treves A., Morini M., Chiappetti L. et al. 1989, ApJ 341, 733 Urry C.M., Maraschi L., Edelson R. et al. 1993, ApJ 411, 614 Wade R.A., Szkody P., C o rdova F., 1979, IAU Circ., No. 3279. Zhang Y.H., Xie G.Z. 1996, A AS 119, 199 Xie G.Z. et al. 1996, AJ, 111, 1065 Figure Captions Fig. 1: a: The long-term U light curve of PKS 2155-304; b: The long-term B light curve of PKS 2155-304; : The long-term V light curve of PKS 2155-304; d: The long-term R light curve of PKS 2155-304; e: The long-term I light curve of PKS 2155-304. Fig. 2: Plot of V_ m ^ 2 vs. trial period, P , in years Fig. 3: DCF for the V band data. It shows that the V light curve is self-correlated with time lags of 4.2 and 7.31 years. In addition, there are also correlation with time lags of less than 4.0 years. Fig. 4: The observed V light curve (filled points) and the simulated V light curve (solid curve) with the periods of 4.16 and 7.0 considered.", null ] }
astro-ph0001042
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "Figure 1 As a demonstrating example, we have determined the potential for =2.8 , a=0.5 and K=0.2 . Figures (a) and (b) show the isocontours and a three dimensional view of , respectively.", "Figure 1 As a demonstrating example, we have determined the potential for =2.8 , a=0.5 and K=0.2 . Figures (a) and (b) show the isocontours and a three dimensional view of , respectively.", "Figure 2 The surface density isocontours corresponding to the potential function of Figure 1.", "Figure 3 The graphs of f(u) and g(v) for =2.8 , C=0.2 , E=1.2 and a=0.5 . The horizontal lines indicate the levels of I_2 and -I_2 in the graphs of f(u) and g(v) , respectively. (a) I_2=0.4 (b) -I_2=-0.4 () I_2=-0.25 (d) -I_2=0.25 (e) I_2=-0.1 (f) -I_2=0.1 .", "Figure 3 The graphs of f(u) and g(v) for =2.8 , C=0.2 , E=1.2 and a=0.5 . The horizontal lines indicate the levels of I_2 and -I_2 in the graphs of f(u) and g(v) , respectively. (a) I_2=0.4 (b) -I_2=-0.4 () I_2=-0.25 (d) -I_2=0.25 (e) I_2=-0.1 (f) -I_2=0.1 .", "Figure 3 The graphs of f(u) and g(v) for =2.8 , C=0.2 , E=1.2 and a=0.5 . The horizontal lines indicate the levels of I_2 and -I_2 in the graphs of f(u) and g(v) , respectively. (a) I_2=0.4 (b) -I_2=-0.4 () I_2=-0.25 (d) -I_2=0.25 (e) I_2=-0.1 (f) -I_2=0.1 .", "Figure 3 The graphs of f(u) and g(v) for =2.8 , C=0.2 , E=1.2 and a=0.5 . The horizontal lines indicate the levels of I_2 and -I_2 in the graphs of f(u) and g(v) , respectively. (a) I_2=0.4 (b) -I_2=-0.4 () I_2=-0.25 (d) -I_2=0.25 (e) I_2=-0.1 (f) -I_2=0.1 .", "Figure 3 The graphs of f(u) and g(v) for =2.8 , C=0.2 , E=1.2 and a=0.5 . The horizontal lines indicate the levels of I_2 and -I_2 in the graphs of f(u) and g(v) , respectively. (a) I_2=0.4 (b) -I_2=-0.4 () I_2=-0.25 (d) -I_2=0.25 (e) I_2=-0.1 (f) -I_2=0.1 .", "Figure 3 The graphs of f(u) and g(v) for =2.8 , C=0.2 , E=1.2 and a=0.5 . The horizontal lines indicate the levels of I_2 and -I_2 in the graphs of f(u) and g(v) , respectively. (a) I_2=0.4 (b) -I_2=-0.4 () I_2=-0.25 (d) -I_2=0.25 (e) I_2=-0.1 (f) -I_2=0.1 .", "Figure 4 The possible families of orbits: (a) an eccentric butterfly orbit (b) an aligned loop orbit () a horseshoe orbit (d) a lens orbit associated with I_2=C .", "Figure 4 The possible families of orbits: (a) an eccentric butterfly orbit (b) an aligned loop orbit () a horseshoe orbit (d) a lens orbit associated with I_2=C .", "Figure 4 The possible families of orbits: (a) an eccentric butterfly orbit (b) an aligned loop orbit () a horseshoe orbit (d) a lens orbit associated with I_2=C .", "Figure 4 The possible families of orbits: (a) an eccentric butterfly orbit (b) an aligned loop orbit () a horseshoe orbit (d) a lens orbit associated with I_2=C ." ], "images": [ null, "/dev/shm/tmpzy8uxmwx/fig1b.ps", "/dev/shm/tmpzy8uxmwx/fig1a.ps", null, "/dev/shm/tmpzy8uxmwx/fig2.ps", null, "/dev/shm/tmpzy8uxmwx/fig3d.ps", "/dev/shm/tmpzy8uxmwx/fig3e.ps", "/dev/shm/tmpzy8uxmwx/fig3c.ps", "/dev/shm/tmpzy8uxmwx/fig3a.ps", "/dev/shm/tmpzy8uxmwx/fig3b.ps", "/dev/shm/tmpzy8uxmwx/fig3f.ps", null, "/dev/shm/tmpzy8uxmwx/fig4b.ps", "/dev/shm/tmpzy8uxmwx/fig4d.ps", "/dev/shm/tmpzy8uxmwx/fig4a.ps", "/dev/shm/tmpzy8uxmwx/fig4c.ps", null ], "texts": [ "We introduce a class of eccentric discs with ``strong\" density cusps whose potentials are of St ackel form in elliptic coordinates. Our models exhibit some striking features: sufficiently close to the location of the cusp, the potential and surface density distribution diverge as ^ 1 and ^ 2 , respectively. As we move outward from the centre, the model takes a non-axisymmetric, lopsided structure. In the limit, when tends to infinity, the isocontours of and become spherically symmetric. It is shown that the configuration space is occupied by three families of regular orbits: eccentric butterfly , aligned loop and horseshoe orbits. These orbits are properly aligned with the surface density distribution and can be used to construct self-consistent equilibrium states. celestial mechanics, stellar dynamics -- galaxies: kinematics and dynamics High resolution observations based on Hubble Space Telescope photometry of nearby galaxies have increased our understanding of the central regions of elliptical and spiral galaxies. It was found that in most galaxies density diverges toward the centre in a power-law cusp. In the presence of a cusp, regular box orbits are destroyed and replaced by chaotic orbits (Gerhard Binney 1985). Through a fast mixing phenomenon, stochastic orbits cause the orbital structure to become axisymmetric at least near the centre (Merritt Valluri 1996). These results are confirmed by the findings of Zhao et al. (1999, hereafter Z99). Their study reveals that highly non-axisymmetric, scale-free mass models can not be constructed self-consistently. Among the models studied for self-consistency, one can refer to the integrable, cuspy models of Sridhar Touma (1997, hereafter ST97). Without a nuclear black hole (BH), centrophobic bananas are the only family of orbits presenting in ST97 discs. Although such orbits elongate in the same direction as density profile, the orbital angular momentum takes a local minimum somewhere rather than the major axis where the surface density has a maximum. This is the main obstacle for building self-consistent equilibria by regular bananas (Syer Zhao 1998; Z99). A similar situation occurs for anti-aligned tube and high resonance orbits for which one could not be able to fit the curvatures of orbits and surface density distribution near the major axis (Z99). According to the results of Miralda-Escud e Schwarzschild (1989), it is only possible to construct self-consistent models by certain families of fish orbits. The orbital structure of stellar systems is enhanced by central BHs in a different manner. Although nuclear BHs destroy box orbits, they enforce some degree of regularity in both centred and eccentric discs (Sridhar Touma 1999, hereafter ST99). In systems with analytical cores and central BHs, a family of long-axis tube orbits can help the host galaxy to maintain its non-axisymmetric structure within the BH sphere of influence (Jalali 1999). In this paper, we present a class of non-scale-free, lopsided discs, which display a collection of properties expected in self-consistent non-axisymmetric cuspy systems. Our models are of St ackel form in elliptic coordinates (e.g., Binney Tremaine 1987) for which the Hamilton-Jacobi equation separates and stellar orbits are regular. In central regions where the effect of the cusp dominates, the potential functions of our distributed mass models are proportional to ^ 1 as 0 . So, we attain an axisymmetric structure near the centre which is consistent with the predicted nature of density cusps. The slope of potential function changes sign as we depart from the centre and our model galaxies considerably become non-axisymmetric. Non-axisymmetric structure is supported by a family of eccentric loop orbits, which are aligned with the lopsidedness. Our potential functions have a local minimum around of which a family of eccentric butterfly orbits emerges. Close to the centre, loop orbits break down and give birth to a new family of orbits, horseshoe orbits. Stars moving in horseshoes lose their kinetic energy as they approach to the centre and contribute a large amount of mass to form a cusp. Our models can be applied to the study of dynamics in systems with double nucleus such as M31 (Tremaine 1995, hereafter T95) and NGC4486B (Lauer et al. 1996). Consider the Hamiltonian function H = 12 (p_x^2+p_y^2) + (x,y), which is described in cartesian coordinates, (x,y) . The variables p_x and p_y denote the momenta conjugate to x and y , respectively. is the potential due to the self-gravity of the disc. Let us express H in elliptic coordinates, (u,v) , through the following transformations x&=&a (1+ u v), y&=&a u v, u & & 0, 0 v 2 , where a is constant and 2a is the distance between the foci of confocal ellipses and hyperbolas defined by the curves of constant u and v , respectively. In the new coordinates, the Hamiltonian function becomes H = 1 2a^2( ^2 u + ^2 v) (p_u^2+p_v^2) + (u,v), with p_u and p_v being the new canonical momenta. We think of those potentials which take St ackel form in elliptic coordinates. The most general potential of St ackel form is (u,v)= F(u)+G(v) 2a^2( ^2 u + ^2 v) , where F and G are arbitrary functions of their arguments. By this assumption, the Hamilton-Jacobi equation separates and results in the second integral of motion, I_2 . We get I_2=p_u^2 - 2a^2 E ^2 u + F(u), or equivalently I_2=p_v^2 - 2a^2 E ^2 v + G(v), where E is the total energy of the system, E H . We now introduce a class of potentials with F(u) &=& C ( u)^ , G(v) &=& -C v | v|^ 1 , where C>0 and are constant parameters. One can readily verify that u = 1 2a (+s), v = 1 2a (-s), where ^2 = x^2+y^2, s^2 = (x-2a)^2+y^2. We substitute from ( 10 ) into ( 8 ) and express in the (x,y) coordinates: &=& K (+s)^ (-s)|-s|^ 1 2r s , K &=& C(2a)^ . The surface density distribution, associated with , is determined as (see Binney Tremaine 1987): (x',y')= 1 4 ^2 G ( ^2 ) dxdy (x'-x)^2+(y'-y)^2 . We examine the characteristics of the potential and surface density functions for small and large radii. Very close to the centre, we have s that simplifies ( 11 ) as follows = Ks^ 1 2 (1+ rs)^ + (1- rs)^ . One can expand (1+ rs)^ and (1- rs)^ in terms of /s to obtain = Ks^ 1 1+ _ n=1 ^ ( +1) (2n) ( 2n+1) ( rs )^ 2n , where is the well known Gamma function. As tends to zero, s is approximated by 2a and /s 0 . Therefore, Equation ( 14 ) reads K(2a)^ 1 . Dimensional considerations show that the surface density will approximately be proportional to ^ 2 . Thus, sufficiently close to the centre, we obtain a strong density cusp with spherical symmetry. When tends to infinity, the potential is approximated as K 2^ 1 ^ 2 . So, we find out ^ 3 . We have to select those values of for which the surface density distribution is plausible and orbits are bounded. According to ( 17 ), the surface density decays outward ( ) for <3 . Moreover, Equation ( 16 ) shows that orbits will be escaping if 2 . To verify this, consider the force exerted on a star, which is equal to . This force will always be directed outward for 2 and results in escaping motions. Therefore, we are confined to 2< <3 .", null, null, "We have used Equations ( 11 ) and ( 12 ) to compute (Figure 1) and (Figure 2) for =2.8 . Due to the complexity of ^2 , we have utilized a numerical scheme to evaluate the double integral of ( 12 ). The potential and surface density functions are symmetric with respect to the x axis and are cuspy at (x=0,y=0) . The potential has a local minimum at (x=a,y=0) that plays an important role in the evolution of orbits. This minimum point has no image in the plane of the surface density isocontours. The surface density monotonically decreases outward from the centre. As it is evident from Figure 2, a non-axisymmetric, lopsided structure is present at moderate distances from the centre.", null, "To this end, we classify orbit families. Having the two isolating integrals E and I_2 , one can find the possible regions of motion by employing the positiveness of p_u^2 and p_v^2 in ( 6 ) and ( 7 ). We define the following functions: f(u)&=& -2 a^2 E ^2 u + F(u), g(v)&=& -2 a^2 E ^2 v + G(v), where F(u) and G(v) are given as ( 8 ). By virtue of p_u^2 0 and p_v^2 0 one can write I_2-f(u) & & 0, I_2-g(v) & & 0. Due to the nature of , no motion exists for negative energies. Hence, E can only take positive values, E>0 . Our classification is based on the behavior of f(u) and g(v) . The most general form of f(u) is attained for C<4a^2E . In such a circumstance, f(u) takes a local maximum at u=0 , f_ M =f(0)=C , and a global minimum at u=u_ m , f_ m =f(u_ m ) , where u_ m = ( 4a^2E C )^ 1 2 , and f_ m =-2a^2E ^2 u_ m + C ( u_ m )^ . According to ( 20 ) we obtain I_2 f_ m . On the other hand, g(v) has a global maximum at v= , g_ M =g( )=C , and two global minima at v= /2 and v=3 /2 , g_ m = g( /2) = g(3 /2) = 2a^2E . Therefore, Inequality ( 21 ) implies I_2 2a^2E. By combining ( 24 ) and ( 25 ) one achieves f_ m I_2 2a^2E. It should be noted that 2a^2E>C . This is because of 2< < 3 . f_ m and in consequence I_2 , can take both positive and negative values. Depending on the value of I_2 , three general types of orbits are generated: (i) Eccentric Butterflies . For C<I_2<2a^2E , the allowed values for u and v are u u_0, v_ b,1 v v_ b,2 , v_ b,3 v v_ b,4 , where u_0 and v_ b,i ( i=1,2,3,4 ) are the roots of f(u)=I_2 and g(v)=-I_2 , respectively. As Figure 3a shows, the horizontal line that indicates the level of I_2 , intersects the graph of f(u) at one point, which specifies the value of u_0 . The line corresponding to the level of I_2 intersects g(v) at four points that give the values of v_ b,i s (Figure 3b). In this case the motion takes place in a region bounded by the coordinate curves u=u_0 and v=v_ b,i . The orbits fill the shaded region of Figure 4a. These are butterfly orbits (de Zeeuw 1985) displaced from the centre. We call them eccentric butterfly orbits.", null, null, null, null, null, null, "(ii) Aligned Loops . We now let I_2 be negative so that f_ m <I_2<-C . In this case the equation f(u)=I_2 has two roots, u_ l,1 and u_ l,2 , which can be identified by the intersections of f(u) and the level line of I_2 (see Figure 3c). The equation g(v)=-I_2 has no real roots and Inequality ( 21 ) is always satisfied (Figure 3d). The allowed ranges of u and v will be u_ l,1 u u_ l,2 , 0 v 2 . The orbits fill a tubular region as shown in Figure 4b. These orbits are bound to the curves of u=u_ l,1 and u=u_ l,2 and elongate in the same direction as lopsidedness. Following ST99, they are called aligned loops. (iii) Horseshoes . For C<I_2<C , we have a different story. In this case, both of the equations f(u)=I_2 and g(v)=-I_2 have two roots. We denote these roots by u=u_ h,i and v=v_ h,i ( i=1,2 ). In other words, the level lines of I_2 intersect the graphs of f(u) and g(v) at two points as shown in Figures 3e and 3f. The orbits fill the shaded region of Figure 4c, which looks like a horseshoe. We call these horseshoe orbits. The orbital angular momentum of stars moving in horseshoes ( G=xp_y-yp_x ) flips sign when stars arrive at one of the coordinate curves v=v_ h,1 or v=v_ h,2 .", null, null, null, null, "For C > 4a^2E , f(u) is a monotonically increasing function of u and eccentric butterflies are the only existing family of orbits. There are three transitional cases corresponding to I_2=C , I_2=2a^2E and I_2=f_ m . For I_2=C , eccentric butterflies extend to a lens orbit as shown in Figure 4d. For I_2=2a^2E , stars undergo a rectilinear motion on the line x=a with the amplitude of a u_0 in the y direction. For I_2=f_ m , loop orbits are squeezed to an elliptical orbit defined by u=u_ m . In this work we explore a credible model based on the self-gravity of stellar discs to explain how an eccentric disc, with strong density cusp, can be in equilibrium. Our mass models exhibit most features of eccentric stellar systems, especially, double nucleus ones such as M31 and NGC4486B. All of the orbits of our model discs are non-chaotic. Below, we clarify how the existing families of orbits help the eccentric disc to maintain the assumed structure. The force exerted on a star is equal to . The motion under the influence of this force can be tracked on the potential hill of Figure 1b. This helps us to better imagine the motion trajectories. As Figure 1b shows, the potential function is concave. A test particle released from distant regions with x>0 and ``small\" initial velocity, slides down on the potential hill and moves toward the local minimum at ( x=a,y=0 ). After passing through the neighborhood of this point (there are some trajectories that exactly visit the minimum point), the test particle climbs on the potential hill until its potential energy becomes maximum. Then, the particle begins to slip down again. This process is repeated and the trajectory of the particle fills an eccentric butterfly orbit. Stars moving in eccentric butterflies form a local group in the vicinity of ( x=a,y=0 ). The accumulation of stars around this local minimum of can create a second nucleus like P2 in M31 (see T95). The predicted second nucleus will approximately be located at the ``centre\" of loop orbits while the brighter nucleus (P1) is at the location of the cusp. Aligned loop orbits occur when the orbital angular momentum is high enough to prevent the test particle to slide down on the potential hill. The boundaries of loop orbits are defined by the ellipses u=u_ l,1 and u=u_ l,2 . The central cusp is located at one of the foci of these ellipses. Aligned loops have the same orientation as the surface density isocontours (compare Figures 2 and 4b). Thus, according to the results of Z99, it is possible to construct a self-consistent model using aligned loop orbits. Similarly, we can describe the behavior of horseshoe orbits. Stars that start their motion sufficiently close to the centre, are repelled from the centre because the force vector is not directed inward in this region. As they move outward, their orbits are bent and cross the x axis with non-zero angular momentum. These stars considerably lose their kinetic energy as they approach the centre (this is equivalent to their climb on the cuspy region of the potential hill). Meanwhile, the orbital angular momentum takes a minimum and switches sign somewhere on the boundary of horseshoe orbit. This boundary is defined by v=v_ h,1 (or v=v_ h,2 ) and can be chosen arbitrarily close to the centre. These stars spend much time near the centre and deposit a large amount of mass, which generates a cusp. Therefore, horseshoe orbits can be used to construct a self-consistent strong cusp. The method of Z99 is no longer applicable to horseshoes because such orbits don't cross the long axis (here the x axis) near the centre. In fact, horseshoe orbits are an especial class of boxlets that appropriately bend toward the centre. The lack of such a property in banana orbits causes the ST97 discs to be non-self-consistent. In the case of M31 and NGC4486B, if we suppose that loop and high-energy butterfly orbits control the overall shape of outer regions, horseshoe orbits together with low-energy butterflies (small-amplitude liberations around the local minimum of ) can support the existence and stability of a double nucleus. The parameter a will indicate the distance between P1 and P2. There remains an important question: what does happen to a star just at the centre? The centre of the model, where the cusp has been located, is inherently unstable. With a small disturbance, stars located at ( x=0,y=0 ) are repelled from the centre. But, the time that they spend near the centre will be much longer than that of distant regions when they move in horseshoes. We remark that the stars of central regions live in horseshoe orbits. Although one can place a point mass (black hole) at the centre without altering the St ackel nature of the potential, such a point mass will not remain in equilibrium and leaves the centre. Based on the results of this paper, we conjecture that there may not be any mass concentration just at the centre of cuspy galaxies. However, a very dense region exists arbitrarily close to the centreThis may be an explanation of dark objects at the centre of cuspy galaxies. The centre of our model galaxies is unreachable. Our next goal is to apply the method of Schwarzschild (1979,1993) for the investigation of self-consistency. Binney J., Tremaine S., 1987, Galactic Dynamics, Princeton University Press, Princeton de Zeeuw P.T., 1985, MNRAS, 216, 273 Gerhard O.E., Binney J., 1985, MNRAS, 216, 467 Jalali M.A., 1999, MNRAS, 310, 97 Lauer T.R., et al., 1996, ApJ, 471, L79 Merritt D., Quinlan G.D., 1998, ApJ, 498, 625 Merritt D., Valluri M., 1996, ApJ, 471, 82 Miralda-Escud e J., Schwarzschild M., 1989, ApJ, 339, 752 Schwarzschild M., 1979, ApJ, 232, 236 Schwarzschild M., 1993, ApJ, 409, 563 Sridhar S., Touma J., 1997, MNRAS, 287, L1-L4 (ST97) Sridhar S., Touma J., 1999, MNRAS, 303, 483 (ST99) Syer D., Zhao H.S., 1998, MNRAS, 296, 407 Tremaine S., 1995, AJ, 110, 628 (T95) Zhao H.S., Carollo C.M. de Zeeuw P.T., 1999, MNRAS, 304, 457 (Z99)" ] }
astro-ph0001046
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "The time evolution of non-thermal emission for various energy band: from top to bottom, inverse Compton scattering of EUVE band (65-245 eV), soft X-ray band (4-10 keV), and hard X-ray band (10-100 keV), and synchrotron radio emission (10 MHz - 10 GHz). The times are relative to the most contracting epoch.", "The inverse Compton scattering spectra at t=0.0 (solid lines) and 0.25 (dotted lines).", "Snapshots of synchrotron radio (10MHz-10GHz) surface brightness distribution (solid contours) and X-ray one of thermal ICM (dashed contours) seen from the direction perpendicular to the collision axis. Contours are equally spaced on a logarithmic scale and separated by a factor of 7.4 and 20.1 for radio and X-ray maps, respectively.", "Same as figure fig:syncmap , but for seen from the direction tilted at an angle of 30^ with respect to the collision axis.", "The synchrotron radiation spectra at t=0.0 (solid lines) and 0.25 (dotted lines)." ], "images": [ null, "/dev/shm/tmpdjaygpbm/f1.epsi", "/dev/shm/tmpdjaygpbm/f2.epsi", "/dev/shm/tmpdjaygpbm/f3.epsi", "/dev/shm/tmpdjaygpbm/f4.epsi", "/dev/shm/tmpdjaygpbm/f5.epsi" ], "texts": [ "NON-THERMAL EMISSION FROM RELATIVISTIC ELECTRONS IN CLUSTERS OF GALAXIES: A MERGER SHOCK ACCELERATION MODEL Motokazu Takizawa Department of Astronomy, Faculty of Science, Kyoto University, Sakyo-ku, Kyoto 606-8502, Japan Research Center for the Early Universe, Graduate School of Science, University of Tokyo, Bunkyo-ku, Tokyo, 113-0033, Japan takizawa@astron.s.u-tokyo.ac.jp Tsuguya Naito Yamanashi Gakuin University, Department of Management Information, Sakaori 2-4-5, Koufu-si, Yamanashi-ken, 400-8575, Japan tsuguya@ygu.ac.jp We have investigated evolution of non-thermal emission from relativistic electrons accelerated at around the shock fronts during merger of clusters of galaxies. We estimate synchrotron radio emission and inverse Compton scattering of cosmic microwave background photons from extreme ultraviolet (EUV) to hard X-ray range. The hard X-ray emission is most luminous in the later stage of merger. Both hard X-ray and radio emissions are luminous only while signatures of merging events are clearly seen in thermal intracluster medium (ICM). On the other hand, EUV radiation is still luminous after the system has relaxed. Propagation of shock waves and bulk-flow motion of ICM play crucial roles to extend radio halos. In the contracting phase, radio halos are located at the hot region of ICM, or between two substructures. In the expanding phase, on the other hand, radio halos are located between two ICM hot regions and shows rather diffuse distribution. galaxies: clusters: general --- hydrodynamics--- intergalactic medium --- particle acceleration --- radiation mechanisms: nonthermal Some clusters of galaxies have diffuse non-thermal synchrotron radio halos, which extend in a Mpc scale (e.g., Giovannini et al. 1993; R o ttgering et al. 1997; Deiss et al. 1997). This indicates that there exists a relativistic electron population with energy of a few GeVs (if we assume the magnetic field strength is an order of G) in intracluster space in addition to the thermal intracluster medium (ICM). Furthermore, it is well known that such clusters of galaxies have evidences of recent major merger in X-ray observations (e.g., Henriksen Markevitch 1996; Honda et al. 1996; Markevitch, Sarazin, Vikhlinin 1999; Watanabe et al. 1999). In such clusters of galaxies with radio halos, non-thermal X-ray radiation due to inverse Compton (IC) scattering of cosmic microwave background (CMB) photons by the same electron population is expected (Rephaeli 1979). Indeed, non-thermal X-ray radiation was recently detected in a few rich clusters (e.g., Fusco-Femiano et al. 1999; Rephaeli, Gruber, Blanco 1999; Kaastra et al. 1999) and several galaxy groups (Fukazawa 1999) although their origins are still controversial. In addition to such relatively high energy non-thermal emission, diffuse extreme ultraviolet (EUV) emission is detected from a number of clusters of galaxies (Lieu et al. 1996; Mittaz, Lieu, Lockman 1998; Lieu, Bonamente, Mittaz 1999). Although their origins are also unclear, one hypothesis is that EUV emission is due to IC emission of CMB. If the hypothesis is right, this indicates existence of relativistic electrons with energy of several handled MeVs in intracluster space. The origin of such relativistic electrons is still unclear. Certainly, there are point sources like radio galaxies in clusters of galaxies, which produces such a electron population. However, the electrons can spread only in a few kpc scale by diffusion during the IC cooling time. Clearly, this cannot explain typical spatial size of radio halos. One possible solution of this problem is a secondary electron model, where the electrons are produced through decay of charged pions induced by the interaction between relativistic protons and thermal protons in ICM (Dennison 1980). In the secondary electron model, however, too much gamma-ray emission is produced through neutral pion decay to fit the Coma cluster results (Blasi Colafrancesco 1999). Moreover, the secondary electron model cannot explicitly explain the association between merger and radio halos. From N-body + hydrodynamical simulations, it is expected that there exist shock waves and strong bulk-flow motion in ICM during merger (e.g, Schindler M u ller 1993; Ishizaka Mineshige 1996; Takizawa 1999, 2000). This suggests that relativistic electrons are produced at around the shock fronts through 1st order Fermi acceleration and that propagation of the shock waves and bulk-flow of ICM are responsible for extension of radio halos. Obviously, the merger shock acceleration model can explicitly explain the association between merger and radio halos. However, such hydrodynamical effects on time evolution and spatial distribution of relativistic electrons during merger are not properly considered in previous studies. In this paper, we investigate the evolution of a relativistic electron population and non-thermal emission in the framework of the merger shock acceleration model. We perform N-body + hydrodynamical simulations, explicitly considering the evolution of a relativistic electron population produced at around the shock fronts. The rest of this paper is organized as follows. In s:order we show order estimation about the spatial size of radio halos. In s:models we describe the adopted numerical methods and initial conditions for our simulations. In s:results we present the results. In s:conclusions we summarize the results and discuss their implications. In this section, we estimate several kinds of spatial length scale relevant to the extent of cluster radio halos. Diffusion Length According to Bohm diffusion approximation, a diffusion coefficient is, = E_ e 3 e B , where, is an enhanced factor from Bohm diffusion limit, E_ e is the total energy of an electron, is the velocity of light, e is the electron charge, and B is the magnetic field strength of intracluster space. Since IC scattering of CMB photons is the dominant cooling process for electrons with energy of GeVs in typical intracluster conditions (Sarazin 1999), electron cooling time in this energy range is, t_ IC = 1.1 10^9 yr ( E_ e GeV )^ 1 , where we assume that the cluster redsift is much less than unity. Thus, the diffusion length within the cooling time is L_ diff t_ IC 1.1 10^ 4 Mpc ( 10^2 )^ 1/2 ( B G )^ 1/2 . This is much less than the spatial size of radio halos. From this reason, electrons which are leaking from point sources such like AGNs cannot be responsible for the radio halos. Shock Wave Propagation Length From N-body + hydrodynamical simulations of cluster mergers, the propagation speed of shock wave is an order of 1000 km s ^ 1 (Takizawa 1999). Thus, the length scale where the shock front propagates during the cooling time of equation ( eq:tic ) is, L_ shock.prop 1.1 Mpc ( v_ shock 1000 km /s ) ( E_ e GeV )^ 1 , where v_ shock is the propagation speed of the shock front. Roughly speaking, L_ shock.prop is related to the extent of radio halos along to the collision axis since the accelerated electrons can emit synchrotron radio radiation only during the cooling time behind the shock front. Spatial Size of Shock Surfaces In cluster merger, the shock front spread over in a cluster scale ( Mpc). Thus, even if the shock is nearly standing, radio halos spread in a Mpc scale. Roughly speaking, this is related to the extent of radio halos perpendicular to the collision axis. In the merger shock acceleration model, therefore, propagation of shock fronts and spatial size of shock surfaces play more crucial role to the extent of radio halos than diffusion. Furthermore, the model naturally produce 3-dimensionally extended radio halos in a Mpc scale. We consider the merger of two equal mass ( 0.5 10^ 15 M_ ) subclusters. In order to calculate the evolution of ICM, we use the smoothed-particle hydrodynamics (SPH) method. Each subcluster is represented by 5000 N-body particles and 5000 SPH particles. The initial conditions for ICM and N-body components are the same as those of Run A in Takizawa (1999). The numerical methods and initial conditions for N-body and hydrodynamical parts are fully described in 3 of Takizawa (1999). Our code is fully 3-dimensional. To follow the evolution of a relativistic electron population, we should solve the diffusion-loss equation (see Longair 1994) for each SPH particle. Since the diffusion term is negligible as mentioned in s:order , the equation is, d N(E_ e , t) dt = E_ e [ b(E_ e , t)N(E_ e , t) ] + Q(E_ e , t), where N(E_ e , t)dE_ e is the total number of relativistic electrons per a SPH particle with kinetic energies in the range E_ e to E_ e +dE_ e (hereafter, we denote kinetic energy of an electron to E_ e ), b(E_ e , t) is the rate of energy loss for a single electron with an energy of E_ e , and Q(E_ e , t)dE_ e gives the rate of production of new relativistic electrons per a SPH particle. According to the standard theory of 1st order Fermi acceleration, we assume that Q(E_ e , t) E_ e ^ , where is described as (+2)/(-1) using the compression ratio of the shock front, . For the shocks appeared in this simulation, the ratio is roughly 10 (Takizawa 1999), which provides =2.4 . Since it is very difficult to monitor the compression ratio at the shock front for each SPH particle per each time step, we neglect the time dependence of . The influence of the changes in on the results is discussed in s:conclusions . The normalization of Q(E_ e , t) is proportional to the artificial viscous heating, which is nearly equal to the shock heating. We generate the relativistic electrons everywhere even if explicit shock structures do not appear in the simulation. We assume that sub-shock exists where a fluid element has enough viscous heating. Such sub-shocks are recognized in higher resolution simulations (e.g. Roettiger, Burns, Stone 1999). We assume that total kinetic energy of accelerated electrons from E_ e = 0 to + is 5 of the viscous energy, which is consistent with the recent TeV gamma ray observational results for the galactic supernova remnant SN 1006 (Tanimori et al. 1998; Naito et al. 1999). Note that equation eq:diffloss for the evolution of a relativistic electron population is linear in N( E_ e , t) . Thus, it is easy to rescale our results of N(E_ e , t) if we choose other parameters for the acceleration efficiency. We neglect energy loss of thermal ICM due to the acceleration. For b(E_ e , t) , we consider IC scattering of the CMB photons, synchrotron losses, and Coulomb losses. We neglect bremsstrahlung losses for simplicity, which is a good approximation in typical intracluster conditions (Sarazin 1999). Then, if we ignore weak energy dependence of coulomb losses, the loss function b(E_ e , t) becomes b(E_ e , t) = b_c(t) + b_1(t) E_ e ^2, where b_c(t) = 7.0 10^ 16 (n_ e (t) / cm ^ 3 ) and b_1 = 2.7 10^ 17 +2.6 10^ 18 (B(t)/ G )^2 (if b(E_ e , t) and E_ e are given in units of GeV s ^ 1 and GeV, respectively.). In the above expressions, n_ e is the number density of ICM electrons and B is strength of the magnetic field. To integrate equation ( eq:diffloss ) with the Courant and viscous timestep control (see Monaghan 1992), we use the analytic solution as follows. First, we integrate equation ( eq:diffloss ) from t to t+ t , regarding the second term on the right-hand side as being negligible small. Then, N(E_ e , t + t) = @ ll N( E_ e ,0 , t) b(E_ e ,0 ) b(E_ e ) , & ( E_ e <E_ e , max ), 0, & ( E_ e >E_ e , max ), where E_ e ,0 &=& b_c b_1 ( b_1 b_c E_ e + b_c b_1 t ), E_ e , max &=& b_c b_1 1 b_1 b_c t . Next, we add the contribution from the second term to the above N(E_ e , t + t) using the second-order Runge-Kutta method. In the present simulation, N(E_ e , t) is calculated on logarithmically equally spaced 300 points in the range E_ e = 0.05 to 50 GeVs for each SPH particle. Magnetic field evolution is included by means of the following method. We assume initial magnetic pressure is 0.1 of ICM thermal pressure. This corresponds to B = 0.1 G in volume-averaged magnetic field strength. For Lagrangean evolution of B , due to the frozen-in assumption we apply B(t)/B(t_0) = ( _ ICM (t)/ _ ICM (t_0) )^ 2/3 . Field changes due to the passage of the shock waves is not considered in this paper. The change may depend on field configuration at the shock front and have value of 1-4 . However, it is difficult to examine it in the present simulation even under high condition. We will try this problem in the future paper. Our model implies continuous production of power law distributed relativistic electrons at around the shock fronts. This is valid only when t_ acc is sufficiently shorter than the dynamical timescale of the system ( 10^9 yr), where t_ acc denotes acceleration time in which Q(E_ e ,t) becomes power law distribution. It is presented in the framework of the standard shock acceleration theory as t_ acc =3 u^ 2 (-1)^ 1 ( _ 1 + _ 2 ) , where u is the flow velocity of the upstream of the shock front, and _ 1, 2 are diffusion coefficients of the upstream and downstream, respectively (see e.g. Drury 1983). Assuming B_ 1 =B_ 2 and Bohm diffusion approximation of equation ( eq:kappa ), t_ acc =1.9 10^ 2 yr ( E_ e GeV ) ( 10^ 2 ) ( u 10^3 km \\ s ^ 1 )^ 2 ( B G )^ 1 . This value is certainly much shorter than the dynamical timescale. Figure fig:nthe shows the time evolution of non-thermal emission for various energy band: from top to bottom, IC emission of the Extreme Ultraviolet Explorer (EUVE) band (65-245 eV), soft X-ray band (4-10 keV), and hard X-ray band (10-100 keV), and synchrotron radio emission (10 MHz - 10 GHz). The times are relative to the most contracting epoch. The calculation of the luminosity for each band is performed in the simplified assumption that electrons radiate at a monochromatic energy given by E_ X = 2.5 keV (E_ e / GeV )^2 and = 3.7 MHz (B/ G ) (E_ e / GeV )^2 for IC scattering and synchrotron emission, respectively. Since cooling time is roughly proportional to E_ e ^ 1 in these energy range, the higher the radiation energy of IC emission is, the shorter duration of luminosity increase is. In other words, luminosity maximum comes later for lower energy band. Hard X-ray and radio emissions come from the electrons with almost the same energy range. The luminosity maximum in the hard X-ray band, however, comes slightly after the most contracting epoch. On the other hand, radio emission becomes maximum at most contracting epoch since the change of magnetic field due to the compression and expression plays an more crucial role than the increase of relativistic electrons. In any cases, radio halos and hard X-ray are well associated to merger phenomena. They are observable only when thermal ICM have definite signatures of mergers such as complex temperature structures, non-spherical and elongated morphology, or substructures. Soft X-ray emission, which is observable only in clusters (or groups) with relatively low temperature ( 1 keV) ICM, is still luminous in 1 Gyr after the merger. Thus, the association of mergers in this band is weaker than in the hard X-ray band. Moreover, EUV emission continues to be luminous after the signatures of the merger have been disappeared in the thermal ICM. Figure fig:icspec shows the IC spectra at t=0.0 (solid lines) and 0.25 (dotted lines). In lower energies, L_ ^ 0.7 , which is originated from the electron source spectrum Q(E_ e ) E_ e ^ 2.4 . On the other hand, in higher energies, the spectrums become close to steady solution, L_ ^ 1.2 , owing to the IC and synchrotron losses (Longair 1994). The break point of the spectrum moves toward lower energies as time proceeds. Figure fig:syncmap shows the snapshots of synchrotron radio (10MHz-10GHz) surface brightness distribution (solid contours) and X-ray one of thermal ICM (dashed contours) seen from the direction perpendicular to the collision axis. Contours are equally spaced on a logarithmic scale and separated by a factor of 7.4 and 20.1 for radio and X-ray maps, respectively. At t=-0.25 , the main shocks are located between the two X-ray peaks and relativistic electrons are abundant there. Thus, the radio emission peak is located between the two X-ray peaks although the magnetic field strength there is weaker. At t= 0.0 , relativistic electrons are concentrated around the central region since the main shocks are nearly standing and located near X 0.2 . Furthermore, gas infall compress ICM and the magnetic field. Thus, radio distribution shows rather strong concentration. In these phase (at t=-0.25 and 0.0 ), the radio halo is located at the high temperature region of ICM. On the other hand, at t=0.25 , relativistic electron distribution becomes rather diffuse since fresh relativistic electrons are producing in the outer regions as the shock waves propagate outwards. At t=0.25 the main shocks are located at X 1 . Between the shock fronts rather diffuse radio emission is seen. In this phase, the radio halo is located between two high temperature regions of ICM. As described above, the morphology of the radio halo is strongly depending on the phase of the merger when viewed from the direction perpendicular to the collision axis. When viewed nearly along the collision axis, however, this is not the case. Figure fig:syncmap2 shows the same as figure fig:syncmap , but for seen from the direction tilted at an angle of 30^ with respect to the collision axis. Radio and X-ray morphology are similar each other in all phases. When the cluster is viewed along the collision axis, the distribution of relativistic electrons roughly follows that of the thermal ICM since the shock fronts face to the observers and spread over the cluster. The distribution of magnetic field strength also roughly follows that of the thermal ICM. Therefor, the radio morphology follows X-ray one. Figure fig:syncspec shows the synchrotron radiation spectra at t=0.0 (solid lines) and 0.25 (dotted lines). At t=0.0 , the synchrotron spectrum follows that of IC emission. On the other hand, at t=0.25 there exits a bump at lower energies in the spectrum, which cannot be seen in that of IC emission. Emissivity of synchrotron radiation depends on not only relativistic electron density but also magnetic energy density, which is larger in the central region in this model. On the other hand, IC emissivity depends on the electron density and CMB energy density, which is homogeneous. Thus, the total synchrotron spectrum is more like that in the central region than the total IC spectrum. At t=0.0 , since relativistic electrons are centered, the emission from the outer region is negligible for both spectra. Thus, similar results are given. On the other hand, at t=0.25 , propagation of shocks makes a diffuse distribution of relativistic electrons. Therefor, the contribution from the outer region is not negligible for the total IC spectrum while the total synchrotron spectrum is still biased the central region as seen in figure fig:syncmap . In the central region, however, electrons produced by the main shocks at t 0 with energies more than several GeVs have already cooled down. Thus the spectrum in the central region have a bump in lower energies, which is present in the total spectrum. The emission above 30 MHz is mainly due to the electrons produced by the sub-shocks there. If such sub-shocks do not exist, the emission near the main shocks, where the magnetic field strength is rather weaker, can be seen in this energy range. Note that this feature of the synchrotron spectrum is sensitive to the spatial distribution of magnetic field. We have investigated evolution of non-thermal emission from relativistic electrons accelerated at around the shock fronts during merger of clusters of galaxies. Hard X-ray and radio radiations are luminous only while merger signatures are left in thermal ICM. Hard X-ray radiation becomes maximum in the later stage of merger. In our simulation, radio emission is the most luminous at the most contracting epoch. This is due to the magnetic field amplification by compression. According to the recent magnetohydrodynamical simulations (Roettiger, Stone, Burns 1999), however, it is possible that the field amplification occurs as the bulk flow is replaced by turbulent motion in the later stages of merger. If this is effective in real clusters, radio emission can increase by a factor of two or three than our results in the later stages of merger. EUV emission is still luminous after the merger signatures have been disappeared in thermal ICM. This is consistent with the EUVE results. Morphological relation between radio halos and ICM hot regions is described as follows. In the contracting phase, radio halos are located at the hot regions of thermal ICM, or between two substructures (see the left panel of figure fig:syncmap ). This may correspond to A2256 (R o ttgering et al. 1994). In the expanding phase, on the other hand, radio halos are located between the two hot regions of ICM and show rather diffuse distribution (see the right panel of figure fig:syncmap ). This may correspond to Coma (Giovannini et al. 1993; Deiss et al. 1997) and A2319 (Feretti, Giovannini, B o hringer 1997). In the further later phase, the shock fronts reach outer regions and the GeV electrons are already cooled in the central parts. Then, radio halos are located in the cluster outer regions near the shock fronts and we cannot detect radio emission in the central part of the cluster. However, observational correlation between ICM hot regions and radio halos is not clear since the electron temperature there is significantly lower than the plasma mean temperature due to the relatively long relaxation time between ions and electrons (Takizawa 1999, 2000). Note that until now we could only find the electron temperature through X-ray observations. This may correspond to A3667 (R o ttgering et al. 1997). It is possible that such radio halos located in the outer regions are classified into radio relics since their radio powers and spatial scales becomes weaker and smaller than those of typical radio halos, respectively. When the cluster is viewed nearly along the collision axis, however, such morphological relations between radio halo and ICM are unclear and radio and X-ray distributions become similar each other. We neglect the changes of the spectral index in the electron source term. Since the mach number is gradually increasing as merger proceeds (Takizawa 1999), the spectrum of relativistic electrons becomes flatter as time proceeds. We believe that such changes in the spectral index does not influence our results seriously because most of relativistic electrons are produced in the central high density region, where the mach number is almost constant. In the later stage of merger, however, contribution of relativistic electrons produced in the outer region cannot be negligible in higher energy range ( 10 GeV) since cooling time is relatively short. Thus, it is probable that the inverse Compton spectrum in the hard X-ray ( 10-100 keV) becomes flatter in the later stages. The lower energy part of the electron spectrum can emit hard X-ray through bremsstrahlung. Whether IC scattering or bremsstrahlung is dominant in the hard X-ray range is depending on the shape of the electron spectrum. Roughly speaking, when the spectrum of relativistic electrons is flatter than E_ e ^ 2.5 , IC scattering dominates the other and vice versa (see Appendix). Furthermore, the shape of the electron spectrum in the lower energy part is flatter than the originally injected form since the cooling time due to the coulomb loss is proportional to E_ e and very short (Sarazin 1999). In the present simulation, therefor, it is most likely that the contribution of the bremsstrahlung components in the hard X-ray range is negligible. More detailed calculations, including nonlinear effects for the shock acceleration (Jones Ellison 1991), by Sarazin Kempner (1999) show that IC scattering is dominant in the hard X-ray when the accelerated electron momentum spectrum is flatter than p_ e ^ 2.7 , which corresponds to the electron energy spectrum of E_ e ^ 2.7 in the fully relativistic range. Thus, the bremsstrahlung contribution in the hard X-ray should be considered in mergers with low mach numbers. A merger shock acceleration model also predicts some gamma-ray emission. Electrons which radiate EUV due to IC scattering also emit 100 MeV gamma-ray through bremsstrahlung. Furthermore, it is most likely that protons as well as electrons are accelerated at around the shock fronts. Such high energy protons also produce gamma-rays peaked at 100 MeV through decay of neutral pions. Although the energy density ratio between electrons and protons in acceleration site is uncertain, the contributions of protons and the bremsstrahlung to the emission become important in the higher energies. We think that it is interesting to investigate from hundreds MeV to multi TeV emissions, which are observable with operating instruments such as EGRET , ground-based air C erenkov telescopes, and planning projects like GLAST satellites. However, since the diffusion length of protons within the cooling time is much longer than that of electrons, treatment of the diffusion-loss equation for protons are more complex than the model in this paper. We would like to thank Drs. Y. Fukazawa and S. Shibata for helpful comments. MT thanks Drs. S. Mineshige and T. Shigeyama for continuous encouragement. MT is also grateful to S. Tsubaki for fruitful discussion. We estimate the contribution of the bremsstrahlung form the lower energy part of the electron spectrum to the hard X-ray emission in our model, which is neglected in this paper. Although the crude estimations discussed here are order-of magnitudes, it is helpful to explain which mechanism, IC scattering and bremsstrahlung radiation, is dominant. The standard theory of 1st order Fermi acceleration provides power law spectrum in momentum distribution. We, therefore, assume the momentum spectrum of accelerated electrons has a form of dN_ e dP_ e = N_ 0 P_ e m_ e ^ \\ \\ cm^ 3 \\ (eV \\ ^ 1 )^ 1 \\ , where P_ e is an electron momentum, is the speed of light, and m_ e is the electron rest mass. Using the relation between the momentum P_ e and the kinetic energy E_ e , E_ e = (P_ e ^2 ^2 + m_ e ^2 ^4)^ 1/2 m_ e ^2 , the spectrum for kinetic energy is given by dN_ e dE_ e = N_ 0 E_ e + m_ e ^2 m_ e ^2 P_ e m_ e ^ 1 \\ \\ cm^ 3 \\ (eV)^ 1 \\ . For relativistic electrons, E_ e m_ e ^2 , the spectrum becomes dN_ e dE_ e = N_ 0 E_ e m_ e ^2 ^ \\ , which is consistent with an assumption of Q(E_ e ,t) in section 3. Since the hard X-ray photons are produced from CMB photon field via IC scattering of relativistic electrons in our model, we use this form for the estimation of IC emissivity. For non-relativistic electrons, E_ e m_ e ^2 , the spectrum becomes dN_ e dE_ e = 2^ + 1 2 N_ 0 E_ e + m_ e ^2 m_ e ^2 E_ e m_ e ^2 ^ + 1 2 \\ , where we use the relation E_ e =P_ e ^2/(2 m_ e ) . Since the bremsstrahlung radiation at the hard X-ray range is emanated from electrons with almost the same energy range, we use this form for the estimation of bremsstrahlung emissivity. For simplicity, we approximate the emissivity, , for both IC and bremsstrahlung processes to be, = dN_ e dE_ e dE_ e d _ dE_ e dt \\ \\ erg \\ s^ 1 \\ cm^ 3 \\ eV^ 1 \\ , where the emission rate is assumed to be equal to the electron energy loss rate and _ is the photon energy. To obtain the bremsstrahlung emissivity, we assume that an electron with energy E_ e loses its energy to emit photon of energy _ =E_ e after it has traversed one mean free path X_ 0 . Hence, dE_ e d _ =1, and dE_ e dt _ v_ e X_ 0 \\ \\ erg \\ s^ 1 \\ , where v_ e is electron velocity. From _ T n_ 0 X_ 0 1 , we approximate dE_ e dt _ v_ e _ T n_ 0 \\ , where _ T denotes the cross section of Thomson scattering and n_ 0 denotes the density of ambient matter. Using equations ( eq:nespecnrel ), ( eq:emiss ), ( eq:dedphotonbrem ), and ( eq:coolbrem ), we estimate the bremsstrahlung emissivity in the hard X-ray energy range, _ = _ HXR , as _ brem 2^ 2 N_ 0 _ T n_ 0 m_ e ^2 1+ _ HXR m_ e ^2 _ HXR m_ e ^2 ^ 2 + 1 \\ , For IC scattering of CMB photons, the photon energy after the scattering by an electron with energy E_ e =m_ e _ e ^2 is approximated by single energy of _ = _ e ^2 _ CMB , where _ CMB is peak energy of CMB spectrum. Thus, dE_ e d _ = 1 2 _ CMB m_ e ^2 ^ 1 2 _ m_ e ^2 ^ 1 2 \\ . Since the scattering is in Thomson energy range ( _ CMB _ e m_ e ^2 ), we set dE_ e dt = 4 3 _ T _ e ^2 _ CMB n_ CMB \\ \\ erg \\ s^ 1 \\ , where n_ CMB is photon number density of CMB fields. Using equations ( eq:nespecrel ), ( eq:emiss ), ( eq:dedphotoic ), and ( eq:coolic ), we estimate the IC emissivity in the hard X-ray energy range, _ = _ HXR , as _ IC = 2 3 N_ 0 _ T _ CMB n_ CMB _ CMB m_ e ^2 ^ 3 2 _ HXR m_ e ^2 ^ 1 2 \\ . From equations ( eq:emissbrem ) and ( eq:emissic ), we derive the emissivity ratio as _ brem _ IC = 3 2 n_ 0 n_ CMB 1+ _ HXR m_ e ^2 _ CMB m_ e ^2 ^ 1 2 _ HXR m_ e ^2 ^ 1 2 2 _ CMB m_ e ^2 ^ 2 \\ . For _ brem < _ IC , we get the relation < 2 3 2 n_ 0 n_ CMB 1+ _ HXR m_ e ^2 ( _ CMB _ HXR )^ 1 2 m_ e ^2 2 _ CMB m_ e ^2 \\ . Substituting typical values for ICM at z 0 , n_ 0 =1 10^ 3 \\ \\ cm^ 3 , n_ CMB =400 \\ \\ cm^ 3 , _ CMB =6.57 10^ 4 \\ \\ eV , and _ HXR =10 \\ \\ keV , we obtain < 2.5 . At _ HXR =100 \\ \\ keV , the spectrum of equation ( eq:emissbrem ) becomes steeper as electrons enter the trans-relativistic energy range. As a result, the bremsstrahlung emissivity is reduced so that the larger index of electron spectrum is accepted for _ brem < _ IC . Blasi, P. Colafrancesco, S. 1999, Astrop. Phys., 12, 169 Deiss, B. M., Reich, W., Lesch, H., Wielebinski, R. 1997, A A, 321, 55 Dennison, B. 1980, ApJ, 239, L93 Drury, L. O'C. 1983, Rep. Prob. Phys., 46, 973 Feretti, L., Giovannini, G. B o hringer, H. 1997, New Astronomy, 2, 501 Fukazawa, Y. 1999, in Proc. ASCA Symposium on Heating and Acceleration in the Universe, ed. H. Inoue, T. Ohashi, T. Takahashi, in press Fusco-Femiano, R., Fiume, D. D., Feretti, L., Giovannini, G., Grandi, P., Matt, G., Molendi, S., and Santangelo, A. 1999, ApJ, 513, L21 Giovannini, G., Feretti, L., Venturi, T., Kim, K. T., and Kronberg, P. P. 1993, ApJ, 406, 399 Henriksen, M. J., Markevitch, M. 1996, ApJ, 466, L79 Honda, H., Hirayama, M., Watanabe, M., Kunieda, H., Tawara, Y., Yamashita, K., Ohashi, T., Hughs, J., P., Henry, J. P. 1996, ApJ, 473, L71 Ishizaka, C., Mineshige, S. 1996, PASJ, 48, L37 Jones, F. C., Ellison, D. C. 1991, Space Sci. Rev., 58, 259 Kaastra, J. S., Lieu, R., Mittaz, J. P. D., Bleeker, J. A. M., Mewe, R., Colafrancesco, S., Lockman, F. J. 1999, ApJ, 519, L119 Lieu, R., Mittaz, J. P. D., Bowyer, S., Lockman, F. J., Hwang, C.-Y., Schmitt, J. H. M. M. 1996 ApJ, 458, L5. Lieu, R., Bonamente, M., Mittaz, J. P. D. 1999, ApJ, 517, L91 Longair, M. S. 1994, High Energy Astrophysics (Cambridge: Cambridge University Press) Mittaz, J. P. D., Lieu, R., Lockman, F. J. 1998, ApJ, 498, L17 Markevitch, M., Sarazin, C. L., Vikhlinin, A. 1999, ApJ, 521, 526 Monaghan, J. J. 1992, ARA A, 30, 543 Naito, T., Yoshida, T., Mori, M. and Tanimori, T., in Proc. ASCA Symposium on Heating and Acceleration in the Universe, ed. H. Inoue, T. Ohashi, T. Takahashi, in press. Rephaeli, Y. 1979, ApJ, 227, 364 Rephaeli, Y., Gruber, D., Blanco, P. 1999, ApJ, 511, L21 Roettiger, K., Stone, J. M., Burns, J. O. 1999, ApJ, 518, 594 Roettiger, K., Burns, J. O., Stone, J. M. 1999, ApJ, 518, 603 R o ttgering, H., Snellen, I., Miley, G., de Jong, P., Hanisch, R. J., Perley, R. 1994, ApJ, 436, 654 R o ttgering, H. J. A., Wieringa, M. H., Hunstead, R. W., and Ekers, R. D. 1997, MNRAS, 290, 577 Sarazin, C. L. 1999, ApJ, 520, 529 Sarazin, C. L., Kempner, J. C. 1999, ApJ, in press, astro-ph/9911335 Schindler, S., M u ller, E. 1993, A A, 272, 137 Takizawa, M. 1999, ApJ, 520, 514 Takizawa, M. 2000, ApJ, in press, astro-ph/9910441 Tanimori, T. et al. 1998, ApJ, 497, L25 Watanabe, M., Yamashita, K., Furuzawa, A., Kunieda, H., Tawara, Y., Honda, H. 1999, ApJ, in press", null, null, null, null, null ] }
astro-ph0001047
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "(a) Distributions of the orientation angle `` alpha '' for gamma-ray-like events with respect to the center of the field of view, which for on-source data corresponds to the NW rim of RX J1713.7 - 3946. The solid line and dashed line are for on-source and off-source data respectively. (b) Distribution of the excess events of the on-source over the off-source level shown in Figure alpha a, shown as the shaded bins. The vertical bars for several bins indicate plus and minus one standard deviation which is approximately the same for all bins. The expected alpha distribution for a point source (dotted line), and disk-like sources with a radius of 0 ^ .2 (dashed line) and 0 ^ .4 (thick solid line) centered at the FOV by the Monte Carlo method are also shown. Here the curves are normalized to the actual excess number of gamma-ray-like events with alpha 30^ .", "Contour map of significance around the NW rim of RX J1713.7 - 3946 centered at the region brightest in hard X-ray emission (right ascension 17^ h 11^ m 56^ s .7 , declination -39^ 31^ 52^ .4 (J2000)). White lines indicate the significance level. The contours of the 0.5--10 keV band of the X-ray flux (from Tomida 1999) also are overlaid as solid lines. The solid circle indicates the size of our PSF." ], "images": [ null, "/dev/shm/tmpu3njhbb1/./bm012-f1.eps", null, "/dev/shm/tmpu3njhbb1/bm012-f2.eps", null ], "texts": [ "02 (08.19.5 RX J1713.7-3946; 13.07.2) Evidence for TeV gamma-ray emission from the shell type SNR RX J1713.7 3946 H. Muraishi 1 , T. Tanimori 2 , S. Yanagita 1 , T. Yoshida 1 , M. Moriya 2 , T. Kifune 3 , S. A. Dazeley 4 , P. G. Edwards 5 , S. Gunji 6 , S. Hara 2 , T. Hara 7 , A. Kawachi 3 , H. Kubo 2 , Y. Matsubara 8 , Y. Mizumoto 9 , M. Mori 3 , Y. Muraki 8 , T. Naito 7 , K. Nishijima 10 , J. R. Patterson 4 , G. P. Rowell 3,4 , T. Sako 8,11 , K. Sakurazawa 2 , R. Susukita 12 , T. Tamura 13 , and T. Yoshikoshi 3 H. Muraishi et al. TeV gamma-ray emission from the SNR RX J1713.7 3946 yanagita@mito.ipc.ibaraki.ac.jp Faculty of Science, Ibaraki University, Mito, Ibaraki 310-8521, Japan Department of Physics, Tokyo Institute of Technology, Meguro, Tokyo 152-8551, Japan Institute for Cosmic Ray Research, University of Tokyo, Tanashi, Tokyo 188-8502, Japan Department of Physics and Mathematical Physics, University of Adelaide, South Australia 5005, Australia Institute of Space and Astronautical Science, Sagamihara, Kanagawa 229-8510, Japan Department of Physics, Yamagata University, Yamagata 990-8560, Japan Faculty of Management Information, Yamanashi Gakuin University, Kofu, Yamanashi 400-8575, Japan STE Laboratory, Nagoya University, Nagoya, Aichi 464-8601, Japan National Astronomical Observatory, Tokyo 181-8588, Japan Department of Physics, Tokai University, Hiratsuka, Kanagawa 259-1292, Japan LPNHE, Ecole Polytechnique. Palaiseau CEDEX 91128, France Computational Science Laboratory, Institute of Physical and Chemical Research, Wako, Saitama 351-0198, Japan Faculty of Engineering, Kanagawa University, Yokohama, Kanagawa 221-8686, Japan Received 1 December 1999 / Accepted 23 December 1999 We report the results of TeV gamma-ray observations of the shell type SNR RX J1713.7 3946 (G347.3 0.5). The discovery of strong non-thermal X-ray emission from the northwest part of the remnant strongly suggests the existence of electrons with energies up to 100 TeV in the remnant, making the SNR a good candidate TeV gamma-ray source. We observed RX J1713.7 3946 from May to August 1998 with the CANGAROO 3.8m atmospheric imaging Cerenkov telescope and obtained evidence for TeV gamma-ray emission from the NW rim of the remnant with the significance of 5.6 . The observed TeV gamma-ray flux from the NW rim region was estimated to be (5.3 0.9 [ statistical ] 1.6 [ systematic ] ) 10 ^ 12 photons cm ^ 2 s ^ 1 at energies 1.8 0.9 TeV. The data indicate that the emitting region is much broader than the point spread function of our telescope. The extent of the emission is consistent with that of hard X-rays observed by ASCA. This TeV gamma-ray emission can be attributed to the Inverse Compton scattering of the Cosmic Microwave Background Radiation by shock accelerated ultra-relativistic electrons. Under this assumption, a rather low magnetic field of 11 G is deduced for the remnant from our observation. gamma-rays: observations -- supernova remnant Supernova remnants (SNRs) are currently believed to be a major source of galactic cosmic rays (GCRs) from the arguments of energetics, shock acceleration mechanisms (Blandford Eichler , Jones Ellison ), and the elemental abundances in the source of GCR (Yanagita et al. , Yanagita Nomoto ). EGRET observations suggest that the acceleration sites of GCRs at GeV energies are SNRs (Esposito et al. ). However, direct evidence for the SN origin of GCRs at TeV energies is scarce (e.g. \\ Koyama et al. \\ , Allen et al. \\ , Buckley et al. \\ ). Arguably the best evidence for the existence of relativistic electrons with energies around 100 TeV is the CANGAROO observation of TeV gamma-rays from the northeast rim of SN1006, which coincides with the region of maximum flux in the 2--10 keV band of the ASCA data (Tanimori et al. ). This TeV gamma-ray emission was explained as arising from 2.7 K Cosmic Microwave Background Radiation (CMBR) photons being Inverse Compton (IC) up-scattered by electrons with energies up to 100 TeV and allowed, together with the observation of non-thermal radio and X-ray emission, the estimation of the physical parameters of the remnant, such as the magnetic field strength (Pohl , Mastichiadis , Mastichiadis de Jager , Yoshida Yanagita , Naito et al. ). The shell type SNR RX J1713.7 3946 was discovered in the ROSAT All-Sky Survey (Pfeffermann Aschenbach ). The remnant has a slightly elliptical shape with a maximum extent of 70 ^ . The 0.1--2.4 keV X-ray flux from the whole remnant is 4.4 10 ^ 10 erg cm ^ 2 s ^ 1 ranking it among the brightest galactic supernova remnants. Subsequent observations of this remnant by the ASCA Galactic Plane Survey revealed strong non-thermal hard X-ray emission from the northwest (NW) rim of the remnant that is three times brighter than that from SN1006 (Koyama et al. ). The non-thermal emission from the NW rim dominates the X-ray emission from RX J1713.7 3946, and the SNR X-ray emission as a whole is dominated by non-thermal emission (Slane et al. , Tomida ). It is notable that the observed emission region of hard X-rays extends over an area 0^ .4 in diameter. Slane et al. ( ) carried out 843 MHz radio observations using the Molonglo Observatory Synthesis Telescope, and discovered faint emission which extends along most of the SNR perimeter, with the most distinct emission from the region bright in X-rays. Slane et al. ( ) suggest the distance to RX J1713.7 3946 is about 6 kpc based upon the observation of CO emission from molecular clouds which are likely to be associated with the remnant. The dominance of non-thermal emission from the shell is reminiscent of SN1006. Koyama et al. ( ) proposed from the global similarity of the new remnant to SN1006 in its shell type morphology, the non-thermal nature of the X-ray emission, and apparent lack of central engine like a pulsar, that RX J1713.7 3946 is the second example, after SN1006, of synchrotron X-ray radiation from a shell type SNR. These findings from X-ray observations would suggest that TeV gamma-ray emission could be expected, as observed in SN1006, from regions in the remnant extended over an area larger than the point spread function of a typical imaging telescope ( 0^ .2). Both SN1006 and RX J1713.7 3946 show notably lower radio flux densities and relatively lower matter densities in their ambient space when compared with those for the other shell type SNRs (Green ) for which the Whipple group (Buckley et al. ) and CANGAROO group (Rowell et al. ) have reported upper limits to the TeV gamma-ray emission. These characteristics might be related to the reason why TeV gamma-rays have been detected only for SN1006 and not from other shell type SNRs: the lower radio flux may indicate a weaker magnetic field which may result in a higher electron energies due to reduced synchrotron losses. In addition, the lower matter density would suppress the production of ^ 0 decay gamma-rays. An observation of TeV gamma-rays from RX J1713.7 3946 would provide not only further direct evidence for the existence of very high energy electrons accelerated in the remnant but also other important information on some physical parameters such as the strength of the magnetic field which are relevant to the particle acceleration phenomena occurring in the remnant, and would also help clarify the reason why TeV gamma-rays have until now been detected only from SN1006. With the above motivation, we have observed RX J1713.7 3946 with the CANGAROO imaging TeV gamma-ray telescope in 1998. Here we report the result of these observations. The CANGAROO 3.8m imaging TeV gamma-ray telescope is located near Woomera, South Australia (136 ^ 47'E, 31 ^ 06'S) (Hara et al. ). A high resolution camera of 256 photomultiplier tubes (Hamamatsu R2248) is installed in the focal plane. The field of view of each tube is about 0 ^ .12 0 ^ .12, and the total field of view (FOV) of the camera is about 3 ^ . The pointing accuracy of the telescope is 0^ .02 , determined from a study of the trajectories of stars of magnitude 5 to 6 in the FOV. RX J1713.7 3946 was observed in May, June and August in 1998. During on-source observations, the center of the FOV tracked the NW rim (right ascension 17^ h 11^ m 56^ s .7 , declination 39^ 31^ 52^ .4 (J2000)), which is the brightest point in the remnant in hard X-rays (Koyama et al. ). An off-source region having the same declination as the on-source but a different right ascension was observed before or after the on-source observation for equal amounts of time each night under moonless and usually clear sky conditions. The total observation time was 66 hours for on-source data and 64 hours for off-source data. After rejecting data affected by clouds, a total of 47.1305 hours for on-source data and 45.8778 hours for off-source data remained for this analysis. The standard method of image analysis was applied for these data which is based on the well-known parameterization of the elongated shape of the Cerenkov light images using `` width ,''`` length ,''`` concentration '' (shape), `` distance '' (location), and the image orientation angle `` alpha '' (Hillas , Weekes et al. , Reynolds et al. ). However, the emitting region of TeV gamma-rays in this target may be extended, as in the case of SN1006. For extended sources, use of the same criteria as for a point source in the shower image analysis is not necessarily optimal. We made a careful Monte Carlo simulation for extended sources of various extents and found the distribution of the shower image parameter of width , length , and concentration for gamma-ray events is essentially the same within a statistical fluctuation as in the case of a point source. However, the simulation suggests that we should allow a wider range dependent on the extent of the source for the parameter of distance and alpha to avoid overcutting gamma-ray events. In this analysis, gamma-ray--like events were selected with the criteria of 0. ^ 01 width 0. ^ 11, 0. ^ 1 length 0. ^ 45, 0.3 concentration 1.0 and 0. ^ 5 distance 1. ^ 2. Figure alpha a shows the resultant alpha distribution when we analyzed the distribution centered at the tracking point (right ascension 17^ h 11^ m 56^ s .7 , declination 39^ 31^ 52^ .4 (J2000)), which is the brightest point in the remnant in hard X-rays (Koyama et al. ). The solid line and the dashed line indicate the on-source and off-source data respectively. Here we have normalized the off-source data to the on-source data to take into account the difference in observation time and the variation of trigger rates due to the difference in zenith angle between on- and off-source data and due to subtle changes in weather conditions. The value of the normalization factor is estimated to be 1.03 from the difference in total obsevation time for on- and off-source measurements. On the other hand, the actual value of the normalization factor is estimated to be 0.99 from the ratio of N_ on / N_ off , where N_ on and N_ off indicate the total number of gamma-ray-like events with alpha between 40 ^ and 90 ^ for on- and off-source data respectively. We selected the region with alpha > 40 ^ to avoid any ``contamination'' by gamma-rays from the source, in the knowledge that the source may be extended. The small discrepancy in the two estimates of the value might come from a slight change in the mirror reflectivity during the observation due to dew. Here we adopt the value 0.99 for in the following analysis by taking the small discrepancy into the systematic errors due to the uncertainty in the mirror reflectivity as shown below. Figure alpha b shows the alpha distribution of the excess events for the on-source over the off-source distribution shown in Figure alpha a. A rather broad but significant peak can be seen at low alpha , extending to 30^ . The alpha distributions expected for a point source and several disk-like extended sources of uniform surface brightness with various radii centered at our FOV were calculated using the Monte Carlo method. These distributions are shown in the same figure. The alpha distribution of the observed excess events appears to favour a source radius of 0^ .4 , which suggests the emitting region of TeV gamma-rays is extended around the NW rim of RX J1713.7 3946. The statistical significance of the excess is calculated by ( N_ on ( ) - N_ off ( ) ) / N_ on ( ) + ^ 2 N_ off ( ) , where N_ on ( ) and N_ off ( ) are the numbers of gamma-ray--like events with alpha less than in the on- and off-source data respectively. The significance at the peak of the X-ray maximum was 5.6 when we chose a value of alpha 30^ considering the result of the Monte Carlo simulation shown in Figure alpha b. In order to verify this extended nature, we examined the effects of the cut in shape parameters on the alpha distribution by varying each cut parameter over wide ranges. We also produced alpha distributions for different energy ranges and data sub-sets. Similar broad peaks in the alpha distribution persisted through these examinations. Also we examined more recent data from PSR1706 44 from July and August 1998 and obtained a narrow peak at alpha < 15^ , as expected for a point source. This confirms that the extended nature of the TeV gamma-ray emitting region does not come from some malfunction of our telescope system and/or systematic errors in our data analysis. A similar, but not as broad, alpha peak was seen for SN1006 (Tanimori et al. ). In order to see the extent of the emitting region, we made a significance map of the excess events around the NW rim of RX J1713.7 3946. Significances for alpha 30^ were calculated at all grid points in 0^ .1 steps in the FOV. Figure map-add shows the resultant significance map of the excess events around the NW rim of RX J1713.7 3946 plotted as a function of right ascension and declination, in which the contours of the hard X-ray flux (Tomida ) are overlaid as solid lines. The solid circle indicates the size of the point spread function (PSF) of our telescope which is estimated to have a standard deviation of 0^ .25 for alpha 30^ based upon Monte Carlo simulations for a point source with a Gaussian function. The area which shows the highest significance in our TeV gamma-ray observation coincides almost exactly with the brightest area in hard X-rays. The region which shows the emission of TeV gamma-rays with high significance ( 3 level) extends wider than our PSF and appears to coincide with the ridge of the NW rim that is bright in hard X-rays. It extends over a region with a radius of 0^ .4 . This region persisted in similar maps calculated for several values of alpha narrower than 30^ .", null, "The integral flux of TeV gamma-rays was calculated, assuming emission from a point source, to be (5.3 0.9 [ statistical ] 1.6 [ systematic ] ) 10 ^ 12 photons cm ^ 2 s ^ 1 ( 1.8 0.9 TeV). The flux value and the statistical error were estimated from the excess number of N_ on (30^ )- N_ off (30^ ) , where the value of 30^ for alpha is chosen by the argument mentioned before. The causes of the systematic errors are categorized by uncertainties in (a) assumed differential spectral index, (b) the loss of gamma-ray events due to the parameter cuts, () the estimate of core distance of showers by the Monte Carlo method, (d) the trigger condition, (e) the conversion factor of the ADC counts to the number of photo-electrons, and (f) the reflectivity of the reflector. These errors from (a) to (f) are estimated as 15 , 22 , 3 , 12 , 10 , and 8 for the integral flux and 24 , 2 , 8 , 20 , 29 , and 17 for the threshold energy, respectively. The total systematic errors shown above are obtained by adding those errors quadratically. To summarise, all our observed data support the hypothesis that the emitting region of the NW rim is extended. In general, the value of the effective detection area of the telescope system for extended sources would be reduced by some factor from that for a point source, because the gamma-ray detection efficiency decreases with the distance of emitting points from the center of the FOV when we observe with a single dish. We calculated the efficiency as a function of the distance by the Monte Carlo method by analyzing the data with the same criteria as applied to the actual data. We estimated the value of the correction factor to the effective area to be 1.2 for our target by integrating the efficiency over the distance for an extended disk-like source of uniform surface brightness with a radius of 0^ .4 . The factor of 1.2 is less significant than the systematic errors estimated above.", null, "The SNR RX J1713.7 3946 is reminiscent of SN1006 both in the synchrotron X-ray emission from the shell far from the centre of the remnant and also in the TeV gamma-ray emission from an extended region coincident with that of the non-thermal X-rays. This suggests that the particles responsible for the emission of the high energy photons are accelerated in shocks. There are several possible emission processes of TeV gamma-rays: the emission induced by accelerated protons (by the ^ 0 decay process) and by electrons -- through bremsstrahlung and/or the Inverse Compton (IC) process. The expected integral flux of gamma-rays above our threshold energy of 1.8 TeV by the ^ 0 decay process is estimated to be < 4 10^ 14 photons cm ^ 2 s ^ 1 (Drury et al. , Naito Takahara ), where we assume the distance and the upper limit for the number density in the ambient space of the remnant as 6 kpc and 0.28 atoms/cm ^ 3 , respectively (Slane et al. ). This flux value is too low to explain our observed flux, even taking into account the large uncertainties in the estimates of the distance and the ambient matter density of the remnant (Slane et al. , Tomida ). However, there remains the possibility of some contribution of the ^ 0 decay process if the remnant is interacting with a molecular cloud located near the NW rim (Slane et al. ). The relative contribution in emissivity of the bremsstrahlung process compared to ^ 0 decay process is estimated as 10 , assuming the flux ratio of electrons to protons is 1/100 and that both have power law spectra with the index of 2.4 (Gaisser ), indicating this process is also unlikely to dominate. Therefore, the most likely process for TeV gamma-ray emission seems to be the IC process. Under this assumption, the magnetic field strength in the supernova remnant can be deduced from the relation L_ syn /L_ IC = U_ B /U_ ph between the IC luminosity L_ IC and synchrotron luminosity L_ syn , where U_ B = B^ 2 /8 and U_ ph are the energy densities of the magnetic field and the target photon field, respectively. L_ syn and L_ IC in the above formula must be due to electrons in the same energy range. The value of L_ syn which should be compared with our TeV gamma-ray data is estimated from the ASCA result to be L_ syn =L_ ASCA _ E^ min _ syn ^ E^ 1.44 dE / _ 0.5keV ^ 10keV E^ 1.44 dE , extrapolating the synchrotron spectrum with the same power law out of the energy range of 0.5 10 keV covered by ASCA (Tomida ). Here L_ ASCA =2.0 10^ 10 erg cm ^ 2 s ^ 1 is the X-ray luminosity in the 0.5 10 keV energy band observed by ASCA from the NW rim of the remnant and the power law index of 1.44 is the mean value for index of X-rays in the same energy range (Tomida ). E^ min _ syn =0.14(B/10 G ) keV is a typical synchrotron photon energy emitted by electrons which emit 1.8 TeV photons (the threshold energy of our observation) by the IC process when we assume the target photons to be from the CMBR. The value of L_ IC is calculated to be 4.2 10^ 11 erg cm ^ 2 s ^ 1 from our result for the number of photons of TeV gamma-rays, and using the fact that the spectra of synchrotron photons and IC photons follow the same power law when the electrons have a power law spectrum. Thus inserting L_ syn , L_ IC , and U_ ph =4.2 10^ 13 erg cm ^ 3 of the energy density for the CMBR into the above relation, we can solve for the magnetic field strength B . Finally, the magnetic field at the NW rim is estimated to be 10.8 G. The extrapolation used to estimate L_ syn is reasonable, because E^ min _ syn is estimated to be 0.15 keV; this is not so different from the minimum energy of the ASCA band (0.5 keV). The electrons responsible for the synchrotron and IC photon emissions are likely to have been accelerated by the shocks in the remnant as discussed above. If the maximum electron energy is limited by synchrotron losses, this maximum energy can be estimated by equating the cooling time due to synchrotron losses with the time scale of acceleration by the first Fermi process in a strong shock as 50 (V_ s /2000 km s ^ 1 ) (B/10 G )^ 0.5 TeV, where V _ s is the shock velocity (Yoshida Yanagita ). On the other hand, equating the acceleration time with the age of the remnant, the maximum energy can be expressed 180 (V_ s /2000 km s ^ 1 )^ 2 (B/10 G ) (t_ age /10000 year ) TeV. In either case, whether it is synchrotron losses or the age of the remnant that limits the maximum electron energy (Reynolds Keohane ), electrons should exist with energies high enough to emit the observed synchrotron X-rays and TeV gamma-rays by the IC process. It is notable that both RX J1713.7 3946 and SN1006 have relatively low magnetic field strengths and low matter densities in their ambient space. These common features may have arisen if the magnetic field was `frozen in' to the matter without amplification other than by compression by shocks and may be the reason why electrons are accelerated to such high energies. These facts may also explain the radio quietness (Green ) and the weak emissivity of ^ 0 decay gamma-rays of the remnants. For SN1006, the low matter density in the ambient space might result from the remnant being located far off the galactic plane and the supernova being of type Ia. For RX J1713.7 3946, the low matter density may be caused by material having been swept out by the stellar wind of the supernova progenitor (Slane et al. ). The low magnetic field and the low matter density in the ambient space of SN1006 and RX J1713.7 3946 may explain why TeV gamma-rays have been detected so far only for these two remnants. In conclusion, we have found evidence for TeV gamma-ray emission from RX J1713.7 3946 at the level of 5.6 sigma. If confirmed ( a la Weekes ), this would be the second case after SN1006 to show directly that particles are accelerated up to energies of 100 TeV in the shell type SNR. We sincerely thank H. \\ Tomida and K. \\ Koyama for providing us the ASCA data. We thank the referee very much for his helpful comments on the paper. This work is supported by a Grant-in-Aid for Scientific Research from Japan's Ministry of Education, Science, and Culture, a grant from the Australian Research Council and the (Australian) National Committee for Astronomy (Major National Research Facilities Program), and the Sumitomo Foundation. The receipt of JSPS Research Fellowships (SH, AK, GPR, KS, and TY) is also acknowledged. Allen, G. E., et al., 1997, ApJ 487, L97 Blandford, R. D. Eichler, D., 1987, Phys. Rep. 154, 1 Buckley, J. H., et al., 1998, A A 329, 639 Drury, L. O'C., et al., 1994, A A 287, 959 Esposito, J. A., et al., 1996, ApJ 461, L820 Gaisser, T. K., 1990, Cosmic Rays and Particle Physics, Cambridge University Press Green, D. A., 1998, A Catalogue of Galactic Supernova Remnants (1998 September version), Mullard Radio Astronomy Observatory, Cambridge, UK (http://www.mrao.cam.ac.uk/surveys/snrs/) Hara, T., et al., 1993, Nucl. \\ Instr. \\ Meth. \\ Phys. \\ Res. \\ A 332, 300 Hillas, A. M., 1985, Proc. 19th Int. Cosmic Ray Conf. (La Jolla) 3, 445 Jones, F. C. Ellison, D. C., 1991, Space Sci. \\ Rev. \\ 58, 259 Koyama, K., et al., 1995, Nature 378, 255 Koyama, K., et al., 1997, PASJ 49, L7 Mastichiadis, A., 1996, A A 305, L53 Mastichiadis, A. de Jager, .O.C., 1996, A A 311, L5 Naito, T. Takahara, F., 1994, J. Phys. G 20, 477 Naito, T., et al., 1999, submitted to Astron. Nachr. Pfeffermann, E. Aschenbach, B., 1996, in `Roentgenstrahlung from the Universe', MPE Report 263, P267 Pohl, M., 1996, A A 307, 57 Reynolds, P. T., et al., 1993, ApJ 404, 206 Reynolds, S. P. Keohane, J. W., 1999, ApJ 525, 368 Rowell, G. P., et al., 1999, submitted to A A Slane, P., et al., 1999, ApJ 525, 357 Tanimori, T, et al., 1994, ApJ 429, L61 Tanimori, T, et al., 1998, ApJ 497, L25 Tomida, H., 1999, Ph. D. thesis, Kyoto Univ. (unpublished) Yanagita, S., Nomoto, K., and Hayakawa, S., 1990, Proc. 21st Int. Cosmic Ray Conf. (Adelaide) 4, 44 Yanagita, S. Nomoto, K., 1999, Proc. 3rd INTEGRAL Workshop `The Extreme Universe' in press, Astrophys. Letters Communications Yoshida, T. Yanagita, S., 1997, Proc. 2nd INTEGRAL Workshop `The Transparent Universe' ESA SP-382, 85 Weekes, T.C., et al., 1989, ApJ 342, 379 Weekes, T.C., 1999, astro-ph/9910394, to appear in `GeV-TeV Astrophysics: Toward a Major Atmospheric Cherenkov Telescope VI,' Snowbird, Utah" ] }
astro-ph0001050
hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_001_wds.tar
{ "captions": [ "Distribution of the identified objects (thick circle) using the redshift catalog of Fern andez-Soto et al. (1999) (thin triangle). The filled circles denote our five targets in figure 1. The size of the symbols corresponds to the K' magnitudes.", "N212 -K' aperture color versus N212 magnitude diagram for all the object detected in either frame. The circles are the identified objects shown in figure 2, while the triangles are unidentified objects. The filled circles denote our five targets in figure 1. The size of the symbols corresponds to the K' magnitudes. The simulated data points are represented by thin dots, and the 98 confidence level is indicated by the broken lines. The solid lines show the flux of emission line within 1 . - '' 16 aperture with the unit of W m ^ -2 .", "Redshift distribution of the identified objects. The circles are plotted based on the spectroscopic redshifts, while the triangles are the photometric redshifts calculated by Fern andez-Soto et al. (1999) for the data whose spectroscopic redshifts are not available. The filled circles denote our five targets in figure 1. The size of the symbols corresponds to the K' magnitudes.", "(a) SEDs of the objects A--C whose H lines expected to be redshifted into the bandpass of the N212 filter. The triangles are the photometric data by Fern andez-Soto et al. (1999), and the filled circle and filled square represent of our photometric results for N212 - and K' -band respectively. (b) The SEDs of the objects D and E with the [ O iii ] emission lines. Symbols are the same in these figures.", "(a) SEDs of the objects A--C whose H lines expected to be redshifted into the bandpass of the N212 filter. The triangles are the photometric data by Fern andez-Soto et al. (1999), and the filled circle and filled square represent of our photometric results for N212 - and K' -band respectively. (b) The SEDs of the objects D and E with the [ O iii ] emission lines. Symbols are the same in these figures.", "Line flux normalized by UV flux density compared to UV spectral index . The asterisks show a local sample of UV selected starbursts and the thick cross represent seven Lyman-break galaxies (Meurer et al. 1999 and references therein). The filled symbols labeled A--E are the five targets in figure 1, while open symbols are other candidates whose H or [ O iii ] lines expected to be redshifted into the bandpass of the N212 filter from their photometric redshift data calculated by Fern andez-Soto et al. (1999). The circles indicate the emission-line data detected with more than a 1 confidence level, while the inverted triangle correspond to the 1 upper limit for the data in which the targeted emission line was not detected. The small symbols are the two redshifted H data points (see figure 4) converted to the H flux by assuming the case-B condition.", "Volume-averaged SFR derived by various observing techniques. The points by the rest-UV continuum are Treyer et al. (1998; open triangle); Lilly et al. (1996; open pentagons); Connolly et al. (1997; open stars); Madau et al. (1998; open squares); Steidel et al. (1999; inverted open triangles). The points by the rest-optical emission lines are Gallego et al. (1995; filled triangle); Tresse, Maddox (1998; filled diamond); Glazebrook et al. (1999; filled square); Yan et al. (1999; inverted filled triangle); this work (filled circle); after a reddening correction using -parameter (open circle)." ], "images": [ null, "/dev/shm/tmpdocqdi5z/fig2.ps", "/dev/shm/tmpdocqdi5z/fig3.ps", "/dev/shm/tmpdocqdi5z/fig4.ps", "/dev/shm/tmpdocqdi5z/fig5a.ps", "/dev/shm/tmpdocqdi5z/fig5b.ps", "/dev/shm/tmpdocqdi5z/fig6.ps", "/dev/shm/tmpdocqdi5z/fig7.ps" ], "texts": [ "Near-infrared emission-line galaxies in the Hubble Deep Field North Fumihide Iwamuro , Kentaro Motohara , Toshinori Maihara , Jun'ichi Iwai , Hirohisa Tanabe , Tomoyuki Taguchi , Ryuji Hata , Hiroshi Terada , and Miwa Goto Department of Physics, Kyoto University, Kitashirakawa, Kyoto 606-8502 E-mail(FI): iwamuro@cr.scphys.kyoto-u.ac.jp [ ] Shin Oya Communications Research Laboratory, Koganei, Tokyo 184-8975 [ ] Masanori Iye , Michitoshi Yoshida , and Hiroshi Karoji Optical and Infrared Astronomy Division, National Astronomical Observatory, Mitaka, Tokyo 181-8588 and Ryusuke Ogasawara , and Kazuhiro Sekiguchi Subaru Telescope, National Astronomical Observatory, 650 North Aohoku Place, Hilo, HI 96720, USA We present the 2.12 m narrow-band image of the Hubble Deep Field North taken with the near-infrared camera (CISCO) on the Subaru telescope. Among five targets whose H or [ O iii ] emission lines are redshifted into our narrow-band range expected from their spectroscopic redshift, four of them have strong emission lines, especially for the two [ O iii ] emission-line objects. The remaining one target shows no H emission in spite of its bright rest-UV luminosity, indicating that this object is already under the post-starburst phase. The volume-averaged SFR derived from the detected two H emission is roughly consistent with that evaluated from the rest-UV continuum. cosmology: observations --- galaxies: formation --- infrared: galaxies headings The Hubble Deep Field North (HDF-N) has been the deepest and the most scrutinizingly observed area since the first images taken by WFPC2 (Williams et al. 1996). The acquired data dramatically pushed the study of the global star-formation history of the early Universe with the establishment of the technique to determine the photometric redshift. The volume-averaged star-formation rate ( SFR ) derived from UV-to-optical data in this field shows a declination from the peak at z 1.5 toward the higher- z (Madau et al. 1996), while there are some suggestions that the SFR remains almost at a constant value when the reddening correction (Calzetti 1998; Steidel et al. 1999) or the selection effect of the rest-UV surface brightness (Pascarelle et al. 1998) is taken into account. On the other hand, the volume-averaged SFR derived from the H luminosity density tends to be larger than that evaluated from the rest-UV continuum (Pettini et al. 1998; Glazebrook et al. 1999; Yan et al. 1999), possibly due to a difference in the reddening effect between the UV continuum and the H emission line. Now, the HDF-N has become a standard area for studying distant galaxies, and various catalogs are available (Fern andez-Soto et al. 1999; Thompson et al. 1999; etc.). In this paper, we report on the results of the observation of the HDF-N by the near-infrared camera CISCO mounted on the Subaru telescope. This observation was planned not only for verifying the system performance in the engineering phase just after the first light observing run, but also for studying the H or [ O iii ] luminosity density in this field. The details of the observation and the data reduction are described in section 2, the detection of the near-infrared emission-line objects and the numerical results are reported in section 3, and the properties of these objects and the volume-averaged star-formation rate are discussed in section 4. Throughout this paper, we assume the Hubble constant of H_0=50 km s ^ 1 Mpc ^ 1 and the deceleration parameter of q_0=0.5 . Observations of the HDF-N were made at the Subaru telescope on 1999 February 23,25, and 27 with the Cooled Infrared Spectrograph and Camera for OHS (CISCO; Motohara et al. 1998). The field was imaged through a standard K' (1.96--2.30 m) and H _2 1--0 S(1) (2.110--2.130 m, hereafter N212 ) filters. Within the 2' 2' field-of-view of the CISCO, there are five candidates whose rest-frame optical emission lines are redshifted into the bandpass of the N212 filter on the basis of their spectroscopic redshifts. and several more candidates expected similarly from their photometric redshifts (Fern andez-Soto et al. 1999 and references therein). The images were taken using an octagonal-star-like dither pattern with a diameter of 12 '' . In the K' band case, 20 s 12 sequential exposures were made at each position (60 s 6 in the case of N212 ), yielding a total exposure time of 1920 s for the standard round of the observation. Unfortunately, the seeing size was unstable during this observing run, and varied from 0 . '' 4 to 0 . '' 9 in each set of exposures. The observations are summarized in table 1. All image reductions were carried out using IRAF. The initial reduction to make a ``quick look image'' followed that of standard infrared-image processing: a flat fielding using the ``standard K' sky flat'', median-sky subtraction, and shift-and-add to combine the image. Then, the positions of the bright sources were listed in a ``mask table'' and the median-sky image was reconstructed by masking out the listed bright objects. In the K' band case, the median smoothed sky image was also used as the self-sky-flat image, and all images were re-flattened by this frame. After applying secondary self-sky flat fielding and sky subtraction, the following reduction processes were carried out: a bad-pixel correction, self-fitting to remove any minor residual of the bias pattern with masking the listed bright object, and shift-and-add of all images with a 3 clipping algorithm. The reduced K' and N212 images are shown in figure 1. The effective exposure times are equal to the total exposure time of 5280 s in the K' and 9360 s in the N212 band, respectively, because there was no discarded frame. The FWHMs for the point source in the final images are 0 . '' 58 ( K' ) and 0 . '' 67 ( N212 ). After a Gaussian smoothing was applied to the K' image to make the image size identical to that of the N212 image, object detection was performed on each frame using SExtractor (Bertin, Arnouts 1996). We employed the ``BEST'' (Kron-like) magnitude output as the total magnitude, while the N212 K' color within 1 . '' 16 (10 pix) aperture was measured by running SExtractor in the double-image mode : K' magnitudes for the N212 selected sample were derived using the N212 frame as a reference image for detection and the K' image for measurements only and vice versa for the K' selected sample. Next, each catalog was compared with the redshift catalog of Fern andez-Soto et al. (1999) under the condition that the nearest corresponding pair having the position difference smaller than 0 . '' 9 is regarded as the same object. The size of the effective area is 3 arcmin ^2 and the distribution of the identified objects is shown in figure 2. The identified and unidentified objects are plotted on the N212 K' color versus N212 magnitude diagram (figure 3) together with the simulated data of the artificial point sources whose intrinsic N212 K' colors are 0 mag. Here, the horizontal N212 mag is the aperture magnitude, because the scatter of the simulated data becomes smaller than in the case when the total magnitude is applied. The 98 confidence level estimated from the simulated data points is expressed as N212 K'=2.5 log (1 10^ N212 21.45 2.5 ), corresponding to the 2.3 detection limit of the emission line of 2.6 10^ 20 W m^ 2 . Figure 4 shows the redshift distribution of the identified objects. Among the objects with available spectroscopic redshift, seven are located where the emission lines are redshifted into the N212 filter bandpass. The SEDs of three objects (A,B,C) located in the redshift range for H are shown in figure 5a. The brightness and the shape of these SEDs are quite similar, while the H strength of object C is far different from that of A or B. The SEDs of other two objects (D,E) for [ O iii ] are also shown in figure 5b. We could not detect any emission from the remaining two objects for redshifted H and the other emission-line candidates expected from the photometric redshifts. The intensity of the emission lines and the SFR s for objects A--E are listed in table 2. Here, we assumed an emission line ratio of H / [ N ii ] 6583,6548=2.3 (Kennicutt 1992; Gallego et al. 1997; Yan et al. 1999), and employed the conversion factor from Kennicutt (1998): SFR_ H ( M_ yr ^ 1 )=7.9 10^ 42 L (H ) (erg s ^ 1 ), SFR_ UV ( M_ yr ^ 1 )=1.4 10^ 28 L_ (1600 ) (erg s ^ 1 Hz ^ 1 ). Recently, Meurer et al. (1999) reported that the UV spectral index ( f_ ^ ) is a good indicator of dust extinction. They found a simple relationship between the dust absorption at 1600 ( A_ 1600 ) and the parameter: A_ 1600 =4.43+1.99 . We assume that this correlation is also applicable to our results. First, we checked our results by using the same diagram of Meurer et al. (1999) (figure 6). The vertical axis is the ``normalized line flux'' defined as the ratio of line flux to flux density f_ 1600 . In this diagram, the ratio of H /f_ 1600 indicates the rough duration time of post-starburst phase, because the H decays with a shorter timescale ( < 20 Myr) than that for the UV continuum ( < 100 Myr) (Kennicutt 1998). The condition of SFR_ H = SFR_ UV without dust extinction implies a constant SFR of the galaxy, displayed as the solid line in figure 6a. We employed the 'net' obscuration expression of the stellar continuum reported by Calzetti (1998) to correct for the reddening factor at the wavelength of H . As can be seen, objects A and B are probably undergoing a starburst phase with a constant SFR in contrast to object C, which has already finished starburst activity. In the case of the [ O iii ] emission line (figure 6b), objects D and E are located in a higher region compared with the distribution of the local starburst galaxies, and reside in a region similar to other Lyman-break galaxies. Although the [ O iii ] emission line is not a good quantitative SFR indicator, due to its large diversity (Kennicutt 1992), it is still notable that the star-forming galaxies at high- z tend to have strong [ O iii ] emission lines. The remaining data points correspond to the other candidates from the photometric redshifts (open symbols) and the two redshifted H (small filled symbols) converted to the H flux assuming the case B condition. Although one marginal value with 1 <S/N< 2.3 is included (open circle in figure 6a), the accuracy of the photometric redshift ( 10 ) does not seem to be sufficient to select the emission-line candidates for a narrow-band survey like this. On the other hand, one of two data points corresponding to the redshifted H emission is located at a higher position than the constant SFR line in figure 6a. This is also a marginal data point, however, indicating that this object is a possible young starburst galaxy. Next, the volume-averaged SFR s were calculated using the H luminosity function at z 0.7--1.9 obtained by Yan et al. (1999). The 2.3 detection limit of 2.6 10^ 20 W m^ 2 corresponds to the H luminosity of L_ lim = 6.4 10^ 41 erg s^ 1 . The survey volume ( V ) is 585 \\ Mpc^3 for the redshifted H emission line. The sum of the SFR_ H for the two H detected objects ( SFR_ sum ) is 22.5 M_ yr ^ 1 (see table 2). We then calculated the volume-averaged SFR ( ^ ) using the following formulae: &=&-1.35, L^ &=&7 10^ 42 \\ erg \\ s^ 1 , ^ &=&1.7 10^ 3 \\ Mpc^ 3 , L_ sum (x)&=& _x^ L ^ L L^ ^ exp L L^ d L L^ , ^ &=& SFR _ sum V L_ sum (0) L_ sum (L_ lim ) &=&0.05 M_ yr^ 1 Mpc^ 3 . Our value is slightly larger than the previous results derived from the rest-UV continuum at z 2.2 (see figure 7), showing the same tendency as the previous H survey reported by Glazebrook et al. (1999) or Yan et al. (1999). If we take the reddening effect expected from the parameter into account, the resultant value of 0.09 M_ yr ^ 1 Mpc ^3 is almost consistent with the reddening-corrected value calculated from the rest-UV continuum (Calzetti 1998; Steidel et al. 1999). Although the error is relatively large, the reddening correction by the parameter works well as far as the present results are concerned. Finally, we calculated the [ O iii ] luminosity density in the same manner as mentioned above to compare with the H luminosity density expected from the previous results for the UV continuum. Because the resultant ratio of the [ O iii ] /H luminosity density is about 5, it cannot be explained as being the same star-forming activity in the local universe. The most conceivable cause is that the high- z star-forming galaxies emit a larger amount of ionizing photons than local starburst galaxies due to some different type of activity, making the [ O iii ] lines brighter. There is also another possibility, that oxygen becomes the major cooling species, owing to the deficiency of metal elements, similar to the nearby blue-compact-dwarf galaxies with metallicity of 0.1 solar (Kunth, Ostlin 1999). Infrared spectroscopy is necessary to investigate such activity in detail, and further observations by the OH-airglow Suppressor spectrograph on the Subaru telescope (Maihara et al. 1993; Iwamuro et al. 1994) are expected to provide the infrared spectra of high- z star forming galaxies. The 2.12 m narrow-band image of the HDF-N was taken with the near-infrared camera on the Subaru telescope. The main conclusions of the present paper are the following: We examined three objects whose H lines should be redshifted into our N212 narrow-band filter as expected from their known spectroscopic redshifts of z 2.2. We found that two of the three targets are undergoing a starburst phase similar to the local starburst galaxies, while the last one target has already finished starburst activity, since no H emission was detected from this object. We also detected strong [ O iii ] emission lines from both targets with the known spectroscopic redshift of z 3.2. We infer that the high- z star-forming galaxy tends to show strong [ O iii ] emission, just like the Lyman-break galaxies reported by Pettini et. al (1998). The volume-averaged SFR derived from the detected H lines is slightly larger than the values previously reported from the rest-UV continuum. They become almost consistent when the reddening correction is applied using the UV spectral index . The present result was accomplished during an engineering observing run of the Subaru Telescope, and is indebted to all members of Subaru Observatory, NAOJ, Japan. We would like to express thanks to the engineering staff of Mitsubishi Electric Co. for their fine operation of the telescope, and the staff of Fujitsu Co. for the timely provision of control software. References Bertin E., Arnouts S. 1996, A AS 117, 393 Calzetti D. 1998, in Proc. the Ringberg Workshop on Ultraluminous Galaxies (astro-ph/9902107) Connolly A.J., Szalay A.S., Dickinson M., Subbarao M.U., Brunner R.J. 1997, ApJ 486, L11 Fern andez-Soto A., Lanzetta K.M., Yahil A. 1999, ApJ 513, 34 Gallego J., Zamorano J., Arag on-Salamanca A., Rego M. 1995, ApJ 455, L1 Gallego J., Zamorano J., Rego M., Vitores A.G. 1997, ApJ 475, 502 Glazebrook K., Blake C., Economou F., Lilly S., Colless M. 1999, MNRAS 306, 843 Iwamuro F., Maihara T., Oya S., Tsukamoto H., Hall D.N.B., Cowie L.L., Tokunaga A.T., Pickles A.J. 1994, PASJ 46, 515 Kennicutt R.C.Jr. 1992, ApJ 388, 310 Kennicutt R.C.Jr. 1998, ARA A 36, 189 Kunth D., Ostlin G. 2000, A AR in press (astro-ph/9911094) Lilly S.J., Le Fevre O., Hammer F., Crampton D. 1996, ApJ 460, L1 Madau P., Pozzetti L., Dickinson M.E. 1998, ApJ 498, 106 Madau P., Ferguson H.C., Dickinson M.E., Giavalisco M., Steidel C.C., Fruchter A. 1996, MNRAS 283, 1388 Maihara T., Iwamuro F., Hall D.N.B., Cowie L.L., Tokunaga A.T., Pickles A.J. 1993, Proc. SPIE 1946, 581 Meurer G.R., Heckman T.M., Calzetti D. 1999, ApJ 521, 64 Motohara K., Maihara T., Iwamuro F., Oya S., Imanishi M., Terada H., Goto M., Iwai J. et al. 1998, Proc. SPIE 3354, 659 Pascarelle S.M., Lanzetta,K.M., Fern andez-Soto A. 1998, ApJ 508, L1 Pettini M., Kellogg M., Steidel C.C., Dickinson M., Adelberger K.L., Giavalisco M. 1998, ApJ 508, 539 Steidel C.C., Adelberger K.L., Giavalisco M., Dickinson M., Pettini M. 1999, ApJ 519, 1 Teplitz H.I., Malkan M.A., McLean I.S. 1999, ApJ 514, 33 Thompson R.I., Storrie-Lombardi L.J., Weymann R.J., Rieke M.J., Schneider G., Stobie E., Lytle D. 1999, AJ 117, 17 Tresse L., Maddox S.J. 1998, ApJ 495, 691 Treyer M.A., Ellis R.S., Milliard B., Donas J., Bridges T.J. 1998, MNRAS 300, 303 Williams R.E., Blacker B., Dickinson M., Dixon W.V.D., Ferguson H.C., Fruchter A.S., Giavalisco M., Gilliland R.L., et al. 1996, AJ 112, 1335 Yan L., McCarthy P.J., Freudling W., Teplitz H.I., Malumuth E.M., Weymann R.J., Malkan M.A. 1999, ApJ 519, L47 t Table 1. @ p 8pc ccc [ ] Date&Filter&Exposure&Seeing [ ] [ ] 1999 February 23& K' &1680 s(20 s 12 7)&0 . '' 5--0 . '' 8 & N212 &4320 s(60 s 6 12)&0 . '' 5--0 . '' 9 [ ] [ ] 1999 February 25& K' &1920 s(20 s 12 8)&0 . '' 5--0 . '' 9 & N212 &5040 s(60 s 6 14)&0 . '' 6--1 . '' 0 [ ] [ ] 1999 February 27& K' &1680 s(20 s 12 7)&0 . '' 4--0 . '' 8 [ ] [ ] t Table 2. @ p 6pc ccccc [ ] Object&Spectroscopic&Line& F (line)& SFR_ H & SFR_ UV &redshift&&10 ^ 20 W m ^ 2 & 2 M_ yr ^ 1 [ ] [ ] A&2.237&H + [ N ii ] &5.0 1.1 &9.8 2.2 &5.2 B&2.233&H + [ N ii ] &6.5 1.1 &12.7 2.2 &7.3 C&2.232&H + [ N ii ] & < 1.1 & < 2.2 &0.7 D&3.233& [ O iii ] &5.7 1.1 &(69.3 13.7) ^ &5.9 E&3.216& [ O iii ] &10.0 1.1&(121.7 13.7) ^ &18.5 [ ] [ ] * SFR_ O III ( M_ yr ^ 1 )=1.5 10^ 41 L ( [ O iii ] ) (erg s ^ 1 ) is assumed (Teplitz et al. 1999).", null, null, null, null, null, null, null ] }
astro-ph0001051
"hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_0(...TRUNCATED)
{"captions":["Distribution of the FWHM of star A during the observation. Only images whose FWHM was (...TRUNCATED)
astro-ph0001060
"hf://datasets/mlfoundations/MINT-1T-ArXiv@70eeb4039972880f5fded4a28b00de52a2428bbf/arXiv_src_0001_0(...TRUNCATED)
{"captions":[null,"fig2 Temperature structures for the model atmospheres of T_ eff = 1400 K and log (...TRUNCATED)

πŸƒ MINT-1T:
Scaling Open-Source Multimodal Data by 10x:
A Multimodal Dataset with One Trillion Tokens

πŸƒ MINT-1T is an open-source Multimodal INTerleaved dataset with 1 trillion text tokens and 3.4 billion images, a 10x scale-up from existing open-source datasets. Additionally, we include previously untapped sources such as PDFs and ArXiv papers. πŸƒ MINT-1T is designed to facilitate research in multimodal pretraining. πŸƒ MINT-1T is created by a team from the University of Washington in collaboration with Salesforce Research, other academic institutions including Stanford University, University of Texas at Austin, and University of California Berkeley.

You are currently viewing the ArXiv subset of πŸƒ MINT-1T. For HTML and PDF subsets, please refer to the πŸƒ MINT-1T collection.

Examples

Dataset Details

Dataset Sources

Uses

Direct Use

πŸƒ MINT-1T is designed to facilitate research in multimodal pretraining. The dataset can be used for training multimodal models that can reson about interleaved text and images sequences such as Idefics2, XGen-MM, and Chameleon.

Out-of-Scope Use

πŸƒ MINT-1T was built to make research into large multimodal models more accessible. Using the dataset to train models that ingest or generate personally identifying information (such as images of people’s faces and other sensitive content) as well as military applications are all inappropriate use cases of πŸƒ MINT-1T.

Dataset Creation

Curation Rationale

πŸƒ MINT-1T was created to address a significant gap in the open-source domain by providing a large-scale multimodal interleaved dataset for pre-training large multimodal models. This dataset aims to be a valuable resource for the research community, facilitating open science in multimodal pretraining.

Source Data

The dataset is a comprehensive collection of multimodal documents from various sources:

  • HTML documents: Filtered from CommonCrawl WARC dumps spanning from 2017 to 2024
  • PDF documents: Extracted from CommonCrawl WAT dumps covering 2023 to 2024
  • ArXiv documents: A subset of papers from the ArXiv repository

In total, πŸƒ MINT-1T contains 1056.8 million documents, broken down as follows:

  • 1029.4 million HTML documents
  • 26.8 million PDF documents
  • 0.6 million ArXiv documents

Data Collection and Processing

The data collection and processing involved several steps:

  1. Document Extraction:

    • HTML documents were parsed from CommonCrawl WARC files
    • PDF documents were extracted from CommonCrawl WAT files
    • ArXiv papers were directly sourced from ArXiv S3 buckets
  2. Filtering Process:

    • Applied text quality filters to ensure content relevance and readability
    • Removed duplicate content at both paragraph and document levels
    • Filtered out undesirable content based on predefined criteria
    • Verified image availability and quality for HTML documents
    • Limited PDF size to 50MB and 50 pages to manage dataset size and quality
  3. Image Processing:

    • Used NSFW image detection to remove pornographic or otherwise undesirable images
    • Removed images smaller than 150 pixels or larger than 20,000 pixels
    • Adjusted aspect ratio thresholds for HTML (2:1) and PDF (3:1) to preserve scientific figures
  4. Text Processing:

    • Used fasttext for language identification, focusing on English content
    • Masked personally identifiable information such as email addresses and IP addresses
    • Applied paragraph and document-level deduplication using Bloom filters
  5. PDF Specific Processing:

    • Used PyMuPDF for parsing PDFs and extracting reading order
    • Clustered text blocks based on columns and ordered from top left to bottom right
  6. ArXiv Specific Processing:

    • Used TexSoup to parse LaTeX source code and interleave images with text
    • Cleaned up LaTeX code by removing imports, bibliography, tables, and citation tags

Various open-source tools were utilized in this process, including fasttext, PyMuPDF, and DCLM and bff for deduplication and content filtering.

Personal and Sensitive Information

Despite sourcing from public web data, significant efforts were made to minimize the inclusion of personal and sensitive information:

  • Email addresses and IP addresses were masked to protect privacy
  • An NSFW image classifierto remove inappropriate visual content
  • URLs containing substrings associated with undesirable or sensitive content were filtered out

However, users should be aware that as the data originates from the public web, it may still contain some sensitive or personal information. The dataset creators acknowledge this limitation and advise users to exercise caution and potentially apply additional filtering based on their specific use cases.

Bias, Risks, and Limitations

Several potential biases, risks, and limitations have been identified:

  1. Data Bias: As the dataset is sourced from web crawls, it may inherit biases present in online content.

  2. Content Risks: Despite extensive filtering, there's a possibility that some offensive, insensitive, or inappropriate content may remain in the dataset.

  3. Image Availability: The dataset relies on external image URLs, which may become unavailable over time due to link rot, potentially affecting the dataset's long-term usability.

  4. PDF Parsing Limitations: The current method for extracting reading order from PDFs may not always accurately capture the intended flow, especially for documents with complex layouts.

  5. Potential Legal and Ethical Concerns: While efforts were made to respect robots.txt files and remove sensitive information, there may still be content that individuals did not explicitly consent to include.

Recommendations

Given these considerations, the following recommendations are provided:

  1. Additional Filtering: Users are strongly encouraged to apply additional filtering based on their specific use case and ethical considerations.

  2. Inappropriate Use Cases: The dataset is not recommended for applications involving the processing or generation of personally identifying information, nor for military applications.

  3. Legal Compliance: Users should independently verify compliance with applicable laws before employing MINT-1T for commercial purposes.

  4. Bias Awareness: Researchers and developers should be cognizant of potential biases in the dataset and consider their impact on model training and outputs.

License

We release πŸƒ MINT-1T under a CC-BY-4.0 license, designating it primarily as a research artifact. While the dataset is freely available, users are responsible for ensuring its legal use in commercial settings. Users must independently verify compliance with applicable laws before employing MINT-1T for commercial purposes.

Citation

@article{awadalla2024mint1t,
      title={MINT-1T: Scaling Open-Source Multimodal Data by 10x: A Multimodal Dataset with One Trillion Tokens}, 
      author={Anas Awadalla and Le Xue and Oscar Lo and Manli Shu and Hannah Lee and Etash Kumar Guha and Matt Jordan and Sheng Shen and Mohamed Awadalla and Silvio Savarese and Caiming Xiong and Ran Xu and Yejin Choi and Ludwig Schmidt},
      year={2024}
}
Downloads last month
2
Edit dataset card

Collection including mlfoundations/MINT-1T-ArXiv