Filename
stringlengths 22
64
| Paragraph
stringlengths 8
5.57k
|
---|---|
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | The 11 ASKAP detections with optical counterparts in this galaxy group span a range of H i masses from 5.4×109 M⊙ down to 0.5×109 M⊙. The two H i clouds we follow up with the Gemini telescope stand out as they are more massive in H i than several of the other detections in IC 1459 which have bright optical counterparts. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | The H i emission associated with IC 5270 has a helio- centric velocity of vhel ∼ 1790−2110 km/s in the ASKAP spectra obtained by Serra et al. (2015), in agreement with the optical redshift of 1858 km/s (Da Costa et al. 1998). The two H i clouds have emission at radial veloc- ities of vhel ∼ 1900 − 2110 km/s. The HIPASS data of this system shows that IC 5270 and the two neighbour- ing hydrogen clouds are linked by a diffuse and extended H i envelope. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | The largest of the two clouds (North cloud) has a slight velocity gradient in H i – Serra et al. (2015), their figure 6. The second, lighter, North-East cloud has a far more homogeneous velocity field (moment-one) in the ASKAP data. The HIPASS name for the IC 5270 system (galaxy + clouds) is HIPASS J2258–35. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 1. 3-cm ATCA continuum image of OH 354.88-0.54. The large cir- cle indicates the approximate optical extent of the PN. The ring is 39′′ in diameter, centred on the stellar position which is marked by the small cross. Radio contours plotted correspond to -4, -2, 2, 4, 8, 12, 16, 20 σ. 1σ is 0.04 mJy beam−1. The beam size is shown in the bottom left corner. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 2. 6-cm ATCA continuum image of OH 354.88-0.54 as for Fig. 1. Radio contours plotted correspond to -4, -2, 2, 4, 8, 12, 16, 20, 30, 40, 50, 60 σ. 1σ is 0.03 mJy beam−1. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 3. 13-cm ATCA continuum image of OH 354.88-0.54 as for Fig. 1. Radio contours have been chosen to emphasize a possible weak structure linking the two continuum components and correspond to -1.5, 1.5, 2, 3, 6, 8, 12, 14, 21, 30, 35 and 40 mJy beam−1. 1σ is 0.14 mJy beam−1. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 4. 20-cm ATCA continuum image of OH 354.88-0.54 as for Fig. 1. Radio contours plotted correspond to -2, 2, 4, 8, 12, 16, 20σ. 1σ is 0.18 mJy beam−1. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | seen on the front of the nebula while source B is seen at the back. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Table 2. Summary of best-fit continuum results to single Gaussians for sources A and B. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 5. Spectral energy distributions of sources A and B fitted by power laws. Filled triangles indicate upper limits at 2 cm from the VLA and 1.3 cm from our data for A, and at 20 cm, from our data, for source B. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 7. Hα/SR quotient image of the PN with variable PSF-matching before dividing the Hα and SR data. The linear feature east of the PN is the track of a satellite. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 8. An enlarged version of Fig. 7 overlaid by white contours of 6-cm radio continuum emission corresponding to 0.3, 0.5, 0.8, 1.1, 1.4, 1.6, 1.9, 2.2 mJy beam−1. White circle marks the approximate outer boundary of the PN; cross marks the position of the star and OH masers. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | tion, corresponding to AHα ≈2.8. Thus the intrinsic Hα flux is 3.3×10−13 erg cm−2 s−1, and the corresponding Hβ flux would be 1.1×10−13 erg cm−2 s−1. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | Figure 9. 6-cm continuum emission of source A shown both as a greyscale image and overlaid by its own contours in white, for levels of 0.2, 0.4, 0.7, 0.9, 1.2, 1.4, 1.6, 1.85, and 2.0 mJy beam−1. The black dashes mark the locus of the ends of the major axes of the white contours. Note the distinct curvature of this locus. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | kinetic energy (KE) flux calculated from the known mass loss rates ( ˙M) of WRs (a few ×10−5 M⊙ yr−1) and wind velocities (Vwind∼2000 km s−1), 1 2, is typically 6×1037 ergs s−1. |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | the interaction region, so the KE flux at the interaction is about 2×1035 ergs s−1 (see Chapman et al. (1999) for details). |
Processed_Radio_observations_of_the_planetary_nebula_around_.txt | KE flux reaches the detected shocked regions giving a total KE flux of 4.4×1036 ergs s−1. For Vf ast, the velocity of the fast wind, we adopt 1000 km s−1. From these values we can crudely estimate the mass-loss rate of the fast wind to be ˙Mf ast≈1.2×10−5 M⊙ yr−1. |
Processed_Sound_and_Image_Processing_with_Optical_Biocompute.txt | This paper was presented at ICIAM – 5th International Congress on Industrial and Applied Mathematics, 7-11th July, Sydney 2003. For the 6th Engineering Mathematics and Applications Conference. EML150N-007. 11th July, 2pm, 2003. |
Processed_Sound_and_Image_Processing_with_Optical_Biocompute.txt | This paper introduces a new method to perform signal processing using cells or atoms to synthesize variation in digital data stored in optical format. Interfacing small-scale biological or chemical systems with information stored on CD, DVD and related disc media has been termed molecular media. This is defined as micron or nanoscale interference and/or augmentation of optical data translation with a specific focus on sound and image processing. The principle of this approach uses cells with different to morphology and size or atom clusters geometrically alter the way that binary information is translated into an output signal. These experiments demonstrate that living and non-living, micron-sized cellular systems like bacteria, fungi or yeast [1] and chemical particles, clusters, emulsions or atom arrays [2] having micron or nanosize dimensions can be successfully interfaced with traditional computing devices. This results in multiple hybrid, yet functional information processing devices that can be understood as a molecular c o m p u t e r that demonstrates a plurality of useful functions. Comparative studies may be found in [3]. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | We have mapped the radio emission in the error ellipse of GeV J1417-6100 (2EGS J1418-6049) at 13cm and 20cm using the Australia Telescope Compact Array. We find a large shell with extended wings , at the edge of which is a non-thermal, polarized structure with a center filled morphology (the ‘Rabbit’), coincident with an extended, hard X-ray source. We discuss the various sources seen within the ellipse as potential counterparts of the γ−ray source. We conclude that the most likely scenario is that the Rabbit is a wind nebula surrounding a radio-quiet γ−ray pulsar. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The nature of the unidentified galactic plane γ−ray sources is one of the outstanding mysteries of high energy astrophysics. There are ∼40 sources that have been detected by EGRET at b < 10◦, many with significant flux above 1 GeV. The only known persistent, discrete sources of GeV emission which have firm identifications are blazars and young pulsars. In addition, several EGRET sources are coincident with young supernova remnants. These source types are moderately bright sources of hard X-rays, and in most cases have associated radio emission as well. We argue that any source drawn from the identified classes can be detected with X-ray/radio observations. Any error box which does not contain a source consistent with the known GeV source types would then be a strong candidate for a new source class. Finding low energy counterparts has been complicated by the complex γ−ray background and the high absorption in the regions of star formation where these objects are found. To minimize the effects of the gas and dust in these regions, it is best to focus on energies above 1 GeV in γ− rays, above 2 keV in X-rays, and in the radio regime. We have begun a campaign to characterize these sources by taking the best positions determined from the > 1 GeV EGRET photons (eg. Lamb & Macomb 1997), imaging them in hard X-rays with ASCA, and following up with radio observations. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The only GeV survey objects in the galactic plane with firm identifications are 5 young pulsars, one of which, Geminga, is radio quiet. Models of the beaming geometries required to reproduce the observed multi- wavelength pulse profiles suggest that most of the GeV sources in the plane are radio-quiet young pulsars as well (Yadigaroglu & Romani 1997). A firm identification as a pulsar requires the detection of a pulse, which is not possible without a-priori knowledge of the pulse period (and even then is difficult), given the limited γ−ray data available. However, the known pulsars have been well studied at many wavelengths, showing properties which allow the identification of promising candidates. In particular, X-ray images of young pulsars show regions of extended hard (power law photon indices Γ ∼ 1 − 2) emission that may be synchrotron nebulae excited by the pulsar wind (Kawai et al. 1997). At radio wavelengths, 3 of the 5 GeV pulsars (Crab, Vela, and PSR B1706-44), have associated Pulsar Wind Nebulae (PWN), which are characterized by radio spectral indices −0.3 ∼< αR ∼< −0.1 (Fν ∝ ναR), amorphous morphologies, and high fractional polarizations (∼ 10 − 30%) (Frail & Scharringhausen 1997). Sources with these X-ray and radio properties within GeV error ellipses are prime candidates to be γ−ray pulsars. The combination of X-ray and radio images of GeV error ellipses can also be used to screen out the only other confirmed class of persistent GeV point source emission, blazars. Since these active galactic nuclei both emit hard x-rays (Γ ∼ 1.7, Fx ∼> 10−13ergs s−1cm−2; Sambruna, Maraschi, & Urry 1996) and are radio-loud (Fr ∼> 0.5 Jy; Mattox et al. 1997), it is relatively simple to check the possibility of a particular low latitude source being a background blazar. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | GeV J1417-6100 (Lamb & Macomb 1997, also 2EGS J1418-6049) was first recognized as a separate source in the supplement to the second EGRET catalog (Thompson et al. 1995), having previously been confused with the nearby, softer source 3EG J1410-6147 (Hartmann et al. 1998). X-ray observations of this region (Roberts and Romani 1998, hereafter RR98) with the ASCA telescope revealed three hard sources X1, X2, and X3 (ps1, ps2, and ps3 in RR98) as well as diffuse emission above 2 keV (Figure 1). One of these sources (X1) is extended and well-centered within the GeV error ellipse; RR98 proposed it to be a plerion candidate, and the likely γ−ray counterpart. The other two appear point-like, with one (X2) possibly having short-term variability. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | In the radio, the low resolution 5GHz Parkes survey image shows a ring structure with a bright source (G313.5+0.2) at the northern end which has been identified as an H ii region at a probable distance of 13.4 kpc (Caswell & Haynes 1987). The IRAS 60µm image shows both the ring and the H ii region. A higher resolution image at 843 MHz, made with the Molonglo Observatory Synthesis Telescope (MOST) during a survey of the galactic plane (Green 1994) resolves the structure of the ring and also shows wings that are coincident with the diffuse X-ray emission. There is a bright radio enhancement at the location of the plerion candidate, X1. There is no source at the position of the X-ray source X2, while there is a faint (∼ 7 mJy) point source consistent with the position of X3, leading RR98 to conclude that they both have properties consistent with background Seyfert galaxies. In this paper we report on synthesis images at 20 and 13 cm of the ring and plerion candidate made with the Australia Telescope Compact Array (ATCA). |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | of each was 128 MHz. Full Stokes parameters are recorded at each wavelength. In order to image the entire complex and to ensure that the strong H ii region at the edge of the ring was well within the main beam, a mosaic with 2 pointings was done, with pointing centers near the plerion candidate (14h 18m 40s, −61◦01′00′′[J2000]) and the H ii region (14h 19m 10s, −61◦51′00′′[J2000]). PKS B1934-638 was used for absolute flux calibration, while PKS B1329-665 was used as a phase calibrator and to calibrate the antenna gains as a function of time. Data reduction and analysis of the data were done using the MIRIAD package (Sault & Killeen 1998). |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The field has also been observed a second time by MOST, at 843 MHz with a 3 kHz bandwidth, as part of a new galactic plane survey using a wider field of view (2.7◦, Green 1997). This has resulted in an improved image without the artifacts of the original survey image caused by proximity to the field edge (figure 1). |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | nearby at a Σ-D distance of 1.9kpc; several young stellar associates are also at similar distances (Yadigaroglu & Romani 1997). |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Total intensity images at 20cm (figure 2) and 13cm were made without using the 6th antenna’s baselines, for a longest baseline of 1.5 km, in order to have reasonably complete UV coverage. The inversion and cleaning of the data was done using the mosaicing routines in MIRIAD, which use a maximum entropy method to remove artifacts (Sault, Stavely-Smith, & Brouw 1996). The baseline weightings were chosen to minimize sidelobes while maintaining a high dynamic range. In addition, high resolution maps (figure 3) were generated by including the 6th antenna for a maximum baseline of 6km, with uniform weighting chosen to maximize point source sensitivity. All the images show a large (peak contour ∼ 12′ in diameter), circular shell with a broad wing (K2) extending to the northeast and a narrower wing extending to the southwest (K4). An additional spur of emission stretches to the northwest, connecting to a bright, unresolved source (G313.2+0.3). The entire complex reminds us of an overweight bird; since it was observed in Australia, we refer to it as the Kookaburra. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Fig. 2.— 20cm ATCA image showing the ‘Kookaburra’, the ‘Rabbit’ and a number of apparent background sources. Only baselines ≤ 1.5km have been included. Restored beam 49.1′′ × 39.8′′, rms noise level of 0.5 mJy. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | At the north end of the shell is the H ii region (G313.5+0.2), which is seen to be compact (∼ 25′′). |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The high resolution images resolve this source for the first time. Near the center of the shell is another source (G313.4+0.1) which the high resolution images show to be a typical extra-galactic core-jet. The shell itself is brightest in the northern half, except for a long, ∼ 8′, bright arc (K1) near the southern edge. Our plerion candidate is clearly seen as a diffuse, rabbit-shaped structure at the west edge (figure 2). In the MOST image and the 20cm map there can be seen an additional diffuse shell to the southwest of the Kookaburra (G313.0-0.1). Due to the smaller primary beam, this is out of the field of view of the 13cm images. The MOST image also includes the nearby supernova remnant G312.4-0.4 (figure 1). |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Fig. 3.— High resolution 13cm image of part of the shell, with 2.9′′ × 2.7′′ beam, rms noise = 0.2 mJy/beam made with uniform weighting of the longest baselines. Contour levels of 8, 12, and 20 mJy/Beam are from the 13cm, baseline ≤1.5 km, 31.8′′ × 25.9′′ beam maps which better preserve flux. Positions of the 3 X-ray sources are indicated. The H ii region and central source are clearly resolved. No unresolved point sources are seen in the Rabbit. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | measurements with ATCA and MOST observations. This method consists of spatially filtering observations at different frequencies so that their effective UV coverage is identical. Images are then produced with identical beam sizes and resolutions of ∼ 1 pixel/beam. Several adjacent pixels are used for flux comparison. If the region has a constant spectral index, pixel values from images of different wavebands plotted on a temperature-temperature (T-T) plot (Turtle et al. 1962) should well fit a line F1 = aF2 + b. The slope of this line is used to calculate the spectral index (αR = loga/log[ν1/ν2]) in order to account for any systematic offset in the flux levels. Such a plot for the Rabbit is shown in figure 4. Using this method, we have measured spectral indices between the 843 MHz, 1384 MHz, and 2496MHz spectral bands for various parts of the field. In order to better constrain the fits, only pixels with 20cm flux densities above 20mJy/beam for region K2 and 12 mJy/beam for K4 were included. Spectral indices are summarized in table 1. Due to the smaller primary beam size, the 13cm maps had problems with artifacts at the edges of regions K2 and K4, so these areas were only fit between the 20cm and 36cm bands. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The bright point source G313.2+0.3, the double source near the center of the shell, and the point source associated with the X-ray source X3 from RR98, all are too small to have a significant number of pixels to fit in the above manner. We therefore measured their fluxes directly from the high resolution maps. Since these sources are mostly unresolved, their flux levels should not be affected by differing UV coverage. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Fig. 4.— Sample T-T plot, showing the spectral fit of the Rabbit to the 20cm and 13cm maps. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | full Q and U maps. These were then combined to make total polarization maps and polarization angle maps of each 8 MHz channel. All 13 of the 13cm channels were averaged together to make the total linear polarization maps. Due to the different central frequency of the 1.5 km configuration, there were only 8 overlapping channels at 20cm to average. The polarization angle maps were fit pixel by pixel in order to generate rotation measure maps. Pixels that could not be well fit were blanked. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | In order to determine fractional polarization, total intensity maps were generated using the same inversion, CLEANing, and linear mosaicing parameters as the polarization maps, but using the full 13 channel bandwidth. The final polarization maps were then divided by these total intensity maps. The noise level was relatively high, so pixels which were not ∼ 3σ above background in both polarized intensity and total intensity were blanked to avoid spurious measurements of high polarization. This means there may be regions of high fractional polarization but low total intensity that were not included in the maps. It should be noted that, since the ATCA simultaneously observes at 2 different frequencies using a combination feed, only one of the frequencies is at the optimum focus, in this case the 20cm band. This means that the 13cm polarization maps are more susceptible to instrumental polarization leakages for bright sources. This is evidenced by significant polarized flux coming from the H ii region in the 13cm maps. Therefore, we use the 20cm maps for quantitative analysis, although we note that morphological features seen in the 20cm total polarization map are reproduced in the 13cm map, and that the apparent fractional polarization is higher at 13cm. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | and 14 mJy/beam superimposed. Fractional polarization reaches ∼ 15% in the ‘paw’ regions P1 and P2. A spur stretching to the NW also has high polarization; residual polarization in the rest of the image is quite low. The circle marks the core position of the ASCA x-ray emission. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The H ii region (G313.5+0.2) is seen to be small, with a linear size of ∼ 1.5 pc, and a total 20cm flux of ∼ 650 mJy. It has two central peaks, separated by ∼ 0.4pc, and no evidence of a shell component (figure 3). The spectrum does not seem to have turned over to the optically thin regime at our observed frequencies. This source is quite bright in the IRAS image, with the ratio of 60 µm to 843 MHz flux density RIR/R ∼ 2000. The small size, high turnover frequency, high RIR/R, plus the observation of associated H2O and OH masers (Caswell et al. 1989), suggest that this is a compact H ii regions (Gordon 1988, and references therein); such sources may be involved with the early stages of star formation. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | spectrum (α ∼ −1). There is no evidence of associated X-rays in the ASCA image, and no infrared excess in the 60µm IRAS image. We therefore identify it with the class of extra-galactic compact steep spectrum radio sources which make up a large fraction of radio point sources (O’Dea 1998). The double source near the center of the Kookaburra (G313.4+0.1) has the typical morphology of a radio galaxy, a steep spectrum, 20cm fluxes of ∼ 25 mJy and ∼ 29 mJy for the two components, and no evidence of X-ray or 60 µm emission. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The radio source coincident with the hard X-ray source X3 is unresolved (figure 3). We measure an improved (J2000) position of (14h 18m 37.7s, −60◦45′00′′). Its spectrum appears flat between 20cm and 13cm, with a flux of ∼ 1.8 mJy. The MOST 843 MHz flux varied between the first and second survey images, changing from ∼ 7 mJy to ∼ 0.3 mJy. The variability and overall flux are consistent with its identification as a radio-quiet AGN. Upper limits of ∼ 0.5 mJy at 20cm and ∼ 0.2 mJy at 13cm can be placed on any point source emission associated with X2 of RR98. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | just SW of the Rabbit up into the NE wing (region K2). Unfortunately, the southern arc (region K1) and much of region K4 are either outside or right at the edge of the ASCA GIS field of view. However, the observed X-rays do not appear to be tracing out the radio shell which makes up the body of the Kookaburra, suggesting that the wings and the shell may be unrelated. If this is so, K1 would be the best region to examine to determine the character of the shell. The fit spectra are consistent with thermal emission, although the scatter in the T-T plot may indicate a non-thermal component to the shell. However, taking into consideration the associated 60µm flux seen by IRAS (RIR/R ∼ 500), it is likely that the shell’s emission is predominantly thermal in nature. Assuming the H ii region, the Rabbit, the central source, and the wings are not part of the shell, we estimate the total 20cm flux to be ∼ 6.5 Jy. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | having a large spread in its T-T plot. The analysis is complicated as well by the smaller beam size at 13cm, limiting the northern regions to comparisons only between the 20cm map and the MOST image. Spectral measurements of K3, while having large uncertainties, are steeper than expected from a thermal source. The non-thermal nature of this area is supported by the hard X-ray spectrum and a lack of correlated 60µm emission. Region K4 also shows filamentary structure. It has an overall lower intensity than K2, with a total flux of ∼ 400 mJy, making it more difficult to measure its spectrum. Again we find no correlation with the IRAS flux in this region. In the 20cm polarization map and rotation measure map, an excess of pixels in the wings (K2 and K4) survive our goodness of fit cuts, suggesting they may be significantly polarized, although the overall low signal to noise prevents accurate estimates of the polarized flux. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The low resolution images of the Rabbit show it to have a centrally peaked morphology, with the peak of the X-ray emission coincident with the front “paw” region P1, and a total 20cm radio flux of ∼ 400 mJy. A direct comparison of the X-ray morphology to the radio is difficult given the broad ASCA PSF and the extreme distortion when it is near the edge of the ASCA GIS field of view. However, it should be noted that the fit 2D Gaussian profile of the X-ray source, when compared to that of a point source is extended in a north-south direction, as is the Rabbit. The high resolution radio images (figure 3) fail to detect any unresolved sources (upper limits of ∼ 0.8 mJy at 20cm, ∼ 0.3 mJy at 13cm), although there is a faint, ∼ 6 mJy, point source just south of the Rabbit. The radio spectrum of the Rabbit is clearly non-thermal, with a spectral index of −0.25 ± 0.1. The IRAS image shows slight evidence of increased IR emission in this portion of the shell, with the ratio of the excess IR emission to excess radio RIR/R ∼< 35. In the 20 cm polarization map, the Rabbit’s ‘paws’, P1 and P2, are seen to be the only sources of polarized flux in the entire field with intensities greater than 2 mJy/beam (figure 5). This implies linear polarization fractions of ∼ 10%. The rotation measure across the 20cm band varies with position, with the polarization peak at P1 having RM= −150 ± 20 rad m−2, and the peak P2 having RM=−240 ± 20 rad m−2. There is an additional spur of polarized emission extending to the northwest. The polarization fraction at 13cm appears even higher, suggesting significant Faraday depolarization at 20cm. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | of a low energy counterpart difficult. However, several of the sources (G313.4+0.1, G313.2+0.3, X3) are readily identified with common extra-galactic sources which are not known to be γ−ray emitters. This is true of the H ii region as well. Both known γ−ray source classes (pulsars and blazars) have hard X-ray counterparts. This leaves the Rabbit and portions of the wings as viable radio counterparts to the GeV source. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The nature of the body of the Kookaburra is uncertain. The common wisdom is that if a shell is thermal, which the spectrum of region K1 and the infrared flux seem to suggest, it is not a supernova remnant. Whiteoak and Green (1996) list typical values of RIR/R for various objects, and the observed value for the shell is consistent with that of a planetary nebulae. The lack of a bright central infrared or optical source argues against a massive stellar wind bubble interpretation, although this remains a possibility. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | wings are part of a large (∼ 2◦) old SNR shell; in figure 1 there is faint radio emission continuing in an arc to the west. The presence of X-rays and non-thermal radio suggest that region K2 has been re-energized with non-thermal electrons. Region K3, having both excess radio and hard X-rays, might be a site of nonthermal e± acceleration (eg. a pulsar), but with the current data the evidence for this is far from compelling. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The nature of the emission from the Rabbit is certainly non-thermal; its significant polarized flux and low IR/R emission imply a supernova remnant identification. SNR with αR ∼> −0.3, polarization fractions ∼> 10%, and a filled center morphology are usually classified as plerionic, or a PWN. Comparing the polarization fraction of P1 and P2 and the total spectral index with PWN around known γ−ray pulsars (Frail & Scharringhausen 1997), we see that the Rabbit is within the observed range of values. Although there are no direct measurements of the distance, if we assume that it is related to the young objects in the region (Clust 3, Cl Lynga 2, or SNR G312.4-0.4, Yadigaroglu & Romani 1997 and references therein) we infer a distance of ∼< 2kpc. For αR = −0.25, F20cm = 400 mJy, and a distance of 2 kpc, the luminosity of the Rabbit integrated from 107 − 1011 Hz is ∼ 1031.9 ergs/s, which is near the lower end of the range of other known PWN (Frail & Scharringhausen 1997). However, the wind nebula associated with PSR B0906-49 has an Lr ∼ 1030 ergs/s (Gaensler et al. 1998b), and the γ-ray pulsar PSR B1055-52 shows no evidence of extended radio emission (Gaensler, private communication), indicating PWN luminosities can be even lower than the Rabbit’s. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Fig. 6.— Broad-band spectrum of the Rabbit. A break νB ∼< 0.3THz is required between the radio and X-ray regimes, presumably associated with electron cooling. The ASCA determination of αX is at present quite poor. Dashed lines show bounds of the multi-parameter 90% errors (not one-parameter as stated in RR98), with the long dashed line extended to indicate the upper limit to the spectral break. The well-constrained index and flux of the GeV emission clearly requires an extra spectral component, plausibly emission from a young pulsar. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Kookaburra. The bright point source G313.2+0.3 is of course likely in the background. If the Rabbit is as distant as the H ii region at 13.4kpc, it’s radio luminosity would be comparable to the brightest PWN. On the other hand, if the GeV source is associated with the Rabbit it would be some 20 times more luminous than the Crab – we therefore feel that the 2kpc distance is more plausible and the H ii region is also an unassociated background source. The non-thermal wings and Rabbit might well be at the same distance, but it is unclear if these are associated with the shell. HI measurements of the shell and Rabbit would be challenging, but very illuminating in sorting out this region’s associations. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | somewhere between the radio and X-ray bands. If we assume the observed radio and X-ray fluxes are synchrotron emission coming from a single population of electrons N (γ) ∼ γp, we expect the electron power law to steepen from p ≈ −1.5 to p ≈ −2.5 above a break energy corresponding to the synchrotron lifetime. The implied αX ≈ −0.75 is consistent with the present large error bars; the resulting break frequency is νB ∼< 3 × 1011Hz. If this is the case, the EGRET photons are clearly an additional component, lying well above the extrapolated radio to X-ray spectrum with a steeper spectral index. If the X-rays are dominated by a different electron population which is also the source of the γ−ray emission, a spectral break is required between the X-rays and γ−rays. This would require the cooling break in the radio to be lower than implied above (of course if the break is intrinsic to the e± population it is at present poorly constrained). |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | At present, the radio/X-ray data suggest a plerion, but do not prove a pulsar origin of the GeV emission. However, if we adopt the pulsar plus plerion interpretation, we can use the various spectral components to constrain the putative pulsar’s properties and to check the global consistency of the model. For example, our estimated νB implies B ∼ 370τ −2/3 µG for a nebula age of 104τ4y. If we assume the post-shock nebular field is contributed by a pulsar of surface dipole 1012B12G and age 104τ4y with a toroidal field out to a LC rs) ∼ 230(B12τ4dkpcθ′)−1µG with wind shock at angle θs = θ′ arcmin, then we expect Bs ∼ 3B∗r3 rs = 0.3dkpcθ′pc. Then, as RR98 noted, if we adopt the empirical PWN LX( ˙ESD) law of Kawai, et al. ˙ESD ≈ 3 × 1035d1.6 kpc erg/s which translates to the constraint 1998, we infer a spindown luminosity of B12 ≈ 18τ −1 kpc . The fiducial pulsar values B12 ≈ 3, τ4 ≈ 3 satisfy these constraints and also predict (Romani 1996) a GeV pulsar luminosity comparable to the observed flux of GeV J1417-6100. The wind shock field estimated from these parameters agrees with the spectral estimate from νB above if θs is rather small, ∼ 5′′. Such a shock standoff would be reasonable if the pulsar has a modest space velocity θs ≈ ( ˙ESD/4πnmHc)1/2d−1v−1 field amplifies beyond equipartition. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | The rotation measure gives us an independent estimate of Bneb, if we assume a local electron density. The Galactic RM in this direction we estimate to be ∼ −100 rad/m2 at ∼ 2kpc (Rand & Lyne 1994), which is consistent with the values measured for the surviving pixels in the wings. This gives an intrinsic RM of ∼ −50 rad/m2 for P1, and ∼ −140 rad/m2 for P2. So since RM= 0.8 R neBkµdlpc we infer an average value hneBki ∼ 200(θ′dkpc)−1µG cm−3 in the nebula. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | electrons were injected into the remnant. The west half is also coincident with the softer γ−ray source 3EG J1410-6147 (figure 1). However, as noted in RR98, this site would require a high transverse velocity vp = 1.1 × 103(dkpc/2)(t4/3)−1km/s, if the Σ − D distance, which is often unreliable, of ∼ 1.9 kpc is correct. The small shell G313.0-0.1, which is only ∼ 25′ away, is also a potential SNR. It is interesting to note that the morphology of the ‘paws’, plus the position of the X-ray source near this region (figure 5) suggests a bow shock, with the body of the Rabbit a wake, implying an origin to the west. On the other hand, a line drawn from the polarization peak P1 through P2 passes close to the centers of both G313.0-0.1 and SNR G312.4-0.4. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Finally, the “wings” may themselves represent a large, old shell birthsite. With an age of 105y, and distance of 1-2kpc, the implied transverse pulsar velocity of ∼ 170dkpc km/s is reasonable. Then, as in CTB80/PSR1951+32, the old shell is being re-energized by PWN electrons (cf. Shull, Fesen, & Saken 1989). High resolution X-ray imaging, coupled with our radio maps will be very helpful in checking the scale and orientation of the implied bow shock morphology. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | are consistent with it being a pulsar powered plerion, with the most likely location of the pulsar being near P1. The Kookaburra is a thermal shell of unclear origin, with non-thermal wings that may not be physically associated. High resolution X-ray images would be useful in determining which features in the radio are not associated with the shell, searching for the magnetospheric flux of any neutron star within the Rabbit and checking if the ‘paw’ region has a bow shock morphology. Stronger limits on point sources within the Rabbit are needed to rule out a pulsar visible in the radio, but statistically we should not be surprised if this is a GeV pulsar with a radio beam missing Earth. If true, this scenario will only receive detailed confirmation with a pulse period detection at high energies. |
Processed_The_`Rabbit':_A_Potential_Radio_Counterpart_of_GeV.txt | Telescope is funded by the commonwealth of Australia for operation as a National Facility managed by the CSIRO. The MOST is owned and operated by the University of Sydney, with support from the Australian Research Council and the Science Foundation within the School of Physics. This work was supported in part by NASA grants NAG5-3333 and NAGW-4562 and a grant from the Cottrell Foundation. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | An exactly soluble one-dimensional model of electrons interacting with order parameter flucta- tions associated with short-range order is considered. The energy and momentum dependence of the electronic self energy and spectral function are calculated and found to exhibit non-Fermi-liquid features similar to that seen for the two-dimensional Hubbard model: a pseudogap, shadow bands, anomalies in the self energy, and breakdown of the quasiparticle picture. Deviations from Fermi liquid behavior are largest close to the Fermi surface and as the correlation length increases. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | The question as to whether the metallic properties of high-Tc superconductors1 and quasi-one-dimensional materials2 can be described by the Fermi liquid picture that works so well in conventional metals has recently re- ceived considerable attention. Many theoretical studies have been made of strongly correlated electron models such as the Hubbard and t − J models3. The availability of high quality photoemission data4,5,6,7 has recently fo- cussed attention on the electron spectral weight function A(k, E) which is related to the probability of observing an electron with momentum k and energy E. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | A consistent picture is gradually emerging from studies of the two-dimensional Hubbard model using quantum Monte Carlo simulations8,9,10,11 and studies using the fluctuation exchange (FLEX) approximation12,13,14,15. Some common features are observed as the temperature is lowered, the hole doping is decreased, or the Coulomb repulsion U is increased. It has been suggested that these changes have a common origin11: they correspond to an increase in ξ, the correlation length associated with short range antiferromagnetic order16. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | that such shadow bands are observable when the corre- lation length is larger than a couple of lattice spacings10. The corresponding photoemission peaks have been ob- 5,7 and insulating served in metallic Bi2Sr2CaCu2O8+x 6. (6) The self energy associated with the Sr2CuO2Cl2 shadow bands is singular18. As ξ increases the imagi- nary part of the self energy is much larger on the shadow Fermi surface than on the regular Fermi surface15. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | The purpose of this Rapid Communication is to point out that there is a simple one-dimensional model which also has many of the features listed above. It is easy to study because it has an exact analytic solution, found by Sadovskii19. This solution can be used to test differ- ent approximation schemes and methods of analytic con- tinuation used on two-dimensional models. Hopefully, the results presented here will also provide physical in- sight into how short range order produces non-Fermi- liquid behavior in two dimensions. The model is directly relevant to the related issues for quasi-one-dimensional materials20,21. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | The common non-Fermi-liquid features observed are: (1) As ξ increases a pseudogap develops, i.e., there is a suppression of the density of states near the Fermi energy8,9,10,11,13,14. (2) As ξ increases peaks in the elec- tron spectral function become smaller and broader14. Near the Fermi surface a breakdown of the quasiparticle picture may occur. (3) On the Fermi surface the real part of the self energy Σ(k, E) can develop a positive slope at the Fermi energy14. (4) The magnitude of the imaginary part of the self energy has a local maximum at the Fermi energy14. (5) “Shadow bands” exist due to the incipient antiferromagnetic order17, i.e., in addition to the peak in the spectral function at the energy E and momentum k there is a much smaller peak at k + Q where Q = (π, π) is the wave vector associated with antiferromagnetic or- der (and the nesting vector of the Fermi surface for half filling). This means the spectral function A(k, E) It is estimated has peaks for |k| > kF but E < EF . |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | scale and a length scale ξ0 ≡ vF /ψ. It will be seen below that the ratio of the correlation length to ξ0 determines to what extent the short-range order causes deviation from Fermi liquid behavior. For the commensurate case of a half-filled band ∆(z) is real. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | and ψ then defines the strength of the coupling of the electrons to the spin fluctuations. In this interpreta- tion we are assuming that the energy scale associated with the spin fluctuations is so much smaller than the electronic energy scale that the spin fluctations can be treated as static. Deisz, Hess, and Serene14 noted that for the two-dimensional Hubbard model at half filling anomalies in the self energy were always accompanied by a spin fluctuation T-matrix Tσσ(q, ωn) that is strongly peaked near the wave vector q = Q and Matsubura fre- quency ωn = 0. In this regime there is a natural map- ping onto the model considered here. At half-filling the Fermi surface is square and has perfect nesting so is some- what “one-dimensional”. The one-dimensional momen- tum then corresponds to that along the (π, π) direction in the two dimensional model. ψ2 corresponds to the weight of the peak in Tσσ(q, 0) within ξ−1 of Q (compare equation (5) in Ref.14). |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | here are for the incommensurate case. Qualitatively sim- ilar results are obtained for the commensurate case. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | Generally we find that as the correlation length in- creases and the Fermi surface is approached the prop- erties of the self energy and spectral function deviate significantly from the quasiparticle picture of Fermi liq- uid theory. Many of the anomalous features we see are similar to those found for the two-dimensional Hubbard model at half- filling treated in the fluctuation-exchange approximation14. Figure 1 shows how the spectral func- tion A(k, E) broadens significantly as the electron mo- mentum approaches the Fermi momentum. This happens even for correlation lengths of the order of ξ0 = vF /ψ. This means that for strong coupling, i.e., ψ of the order of the band width, the correlation length ξ can be of the order of a lattice constant. Shadow bands are present and become larger as ξ increases and |k| decreases. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | Figure 2 shows how the quasiparticle picture breaks down as the correlation length ξ increases. As ξ increases from 0.2ξ0 to 100ξ0 the spectral function on the Fermi surface evolves from a single narrow peak to two broad peaks which are the precursors of conduction and va- lence bands associated with long range spin-density-wave order. For similar reasons as the correlation length in- creases a pseudogap develops in the total density of states (Figure 2 inset). It should be pointed out that the two peak structure is not seen in FLEX results14 or quan- tum Monte carlo for weak coupling on large lattices8. Whether this absence of the two peaks is a real prop- erty of the 2D Hubbard model or a result of the FLEX approximation or not being able to go to low enough temperatures is not clear. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | The self-energy has anomalous features similar to those found for the 2D Hubbard model at half-filling14. At the Fermi energy the real part of the self energy has a positive slope (Figure 3). The magnitude of the imaginary part of the self energy, which in a quasiparticle picture is related to the quasiparticle lifetime and develops a maximum at the Fermi energy (Figure 3). In contrast, in a Fermi liquid ReΣ has a negative slope and |ImΣ| has a minimum (which goes to zero as the temperature goes to zero) at the Fermi energy. The opposite behavior observed here means that although for our model there is a pole in the spectral function at E = 0 it is not possible to define a quasiparticle solution there. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | Ek which satisfies Ek − kvF − ReΣ(k, Ek) = 0. Con- sequently, the shadow band peak is not a replica of the regular quasiparticle peak. These results can be com- pared to those of Chubukov18 for a two-dimensional spin- fluctuation model. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | Figure 5 shows the spectral function calculated for this model using second order perturbation theory. This corresponds to termination of the continued fraction at l = 1. Such a perturbative form for the Greens func- tion has been used in calculations concerning the role of order parameter fluctuations in quasi-one-dimensional materials24. However, the discrepancy with the exact re- sults shown in Figure 1 shows this is quite unreliable for |E| < ψ when |k−kF | < ψ/vF and ξ > vF /ψ. In particu- lar perturbation theory greatly underestimates the width of the spectral function. The inset of Figure 5 shows how perturbation theory gives unreliable results for the total density of states if ξ > ξ0. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | the simple model considered here has many features similar to those seen for the two- dimensional Hubbard model and provides insight into how short range order can produce non-Fermi-liquid be- havior. Of particular interest is the following. (a) The ratio of the correlation length to ξ0 = vF /ψ determines deviations from Fermi liquid behavior. (b) Non-Fermi- liquid behavior occurs even when the correlation length is sufficiently short that there is only a weak pseudogap. (c) The “shadow band” is associated with singularities in the self energy and only corresponds to poles in the spectral function for very large correlation lengths. (d) Perturbation theory gives unreliable results in the non- Fermi-liquid regime. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | This work was stimulated by discussions with J. R. Schrieffer. We thank H. Castella, J. Deisz, D. Hess, M. Steiner, and J. Voit for very helpful discussions. This work was supported by the Australian Research Council. Note added in proof. Recent angle-resolved photoe- mission measurements [H. Ding et al., Nature 382, 51 (1996); A. G. Loeser et al., Science 273, 325 (1996)] have measured a pseudogap in the normal state of Bi2Sr2CaCu2O8+x. The observed opening up of the pseudogap with decreasing doping and temperature are qualitatively similar to the how the pseudogap considered here opens up with increasing correlation length (Fig. 2 inset). |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | FIG. 1. Shadow bands and the breakdown of the quasi- particle picture near the Fermi surface. The energy depen- dence of the electron spectral function A(k, E) is shown for three different momenta k. The correlation length associ- ated with the short range order is equal to 5ξ0 = 5vF /ψ where vF is the Fermi velocity and ψ is a measure of the strength of the coupling of the electrons to the order param- eter fluctuations. Note the presence of “shadow bands”, i.e., small peaks in A(k, E) with E < EF = 0 and k > kF = 0. The spectral function becomes significantly broader as the Fermi surface is approached. Opposite behavior occurs in a Fermi liquid: the quasiparticles are better defined close to the Fermi surface. These results can be compared to those for the two-dimensional Hubbard model at half filling (see Figure 1 in Ref.14 and Figure 2 in Ref.10). |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | FIG. 2. Breakdown of the quasiparticle picture with in- creasing correlation length. The energy dependence of the spectral function at the Fermi momentum is shown for sev- eral correlation lengths. The spectral function broadens con- siderably and evolves into two bands as the correlation length increases. These results can be compared to those for the 2D Hubbard model at half-filling as the temperature is lowered (see Figure 1 in Ref.14). The inset shows how the pseudo- gap in the total density of states opens up as the correlation length increases. Note that although for ξ ∼ ξ0 the pseudo- gap is rather weak deviations from Fermi liquid behavior still occur. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | FIG. 3. Anomalous behaviour of the self energy. The real part of the self energy at the Fermi momentum, ReΣ(0, E), has a positive slope at the Fermi energy (E = 0). This means that a quasiparticle weight cannot be defined. As the cor- relation length increases a maximum at E = 0 develops in the magnitude of the imaginary part of the self energy at the Fermi momentum |ImΣ(0, E)|. (Compare Figure 2 in Ref.14). |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | FIG. 4. Singular behavior of the shadow band self energy. Away from the Fermi surface the self energy near E ∼ kvF has Fermi liquid properties. In contrast, near E ∼ −kvF which is associated with the shadow band the self energy becomes increasingly singular as the correlation length increases. In- tersections of the ReΣ curves with the dot-dashed straight line correspond to poles of the spectral function. Hence, the shadow band is only associated with a pole for very large correlation lengths. |
Processed_Non-Fermi-liquid_behavior_due_to_short_range_order.txt | FIG. 5. Second order perturbation theory is unreliable near the Fermi surface. The spectral function is calculated for the same parameter values as in Figure 1. (Note the vertical scale is three times larger here). The discrepancy with the exact results is large for |k| < 1/ξ0 and becomes larger as ξ increases. The inset shows the total density of states for the same parameter values as in Figure 2. (Note the vertical scale is two times larger here). |
Processed_Neutrino_Mixing_and_CP_Violation_in_Matter.txt | Abstract. Within the framework of three lepton families I present a transparent analytical relationship between the neutrino mixing and CP-violating parameters in vacuum and thos in matter. Such a model- and parametrization-independent result will be particularly useful to recast the fundamental lepton flavor mixing matrix from the future long-baseline neutrino experiments. |
Processed_Neutrino_Mixing_and_CP_Violation_in_Matter.txt | Today strong evidence, that neutrinos are massive and lepton flavors are mixed, has been accumulated from a variety of neutrino experiments. The mixing matrix of three different lepton families may in general consist of non-removable complex phases, leading to CP or T violation. Leptonic CP violation can manifest itself in neutrino oscillations. The best way to observe CP- and T-violating effects is to carry out the long-baseline neutrino experiments. In such experiments the earth- induced matter effects, which are likely to deform the neutrino oscillation behaviors in vacuum and fake the genuine CP-violating signals, must be taken into account. To single out the “true” theory of lepton mass generation and CP violation depends crucially upon how accurately the fundamental parameters of lepton flavor mixing can be measured and disentangled from the matter effects. It is therefore desirable to explore the most transparent analytical relationship between the genuine flavor mixing matrix and the matter-corrected one. |
Processed_Neutrino_Mixing_and_CP_Violation_in_Matter.txt | where (α, β, γ) and (i, j, k) run over (e, µ, τ ) and (1, 2, 3), respectively. In the specific models of fermion mass generation V can be derived from the mass matrices of charged leptons and neutrinos [3]. To test such theoretical models one has to compare their predictions for V with the experimental data of neutrino oscillations. The latter may in most cases be involved in the potential matter effects and must be carefully handled. |
Processed_Neutrino_Mixing_and_CP_Violation_in_Matter.txt | Such an elegant relation has already been observed in Ref. [6]. It indicates that the matter contamination to CP- and T-violating observables, which must be depen- m and λi are complicated dent upon functions of the matter parameter A. |
Processed_Neutrino_Mixing_and_CP_Violation_in_Matter.txt | We observe that matter effects can be significant for the elements in the first 10−5 eV2. In comparison, the magnitudes and the second columns of V , if A may be drastically enhanced or suppressed only for A > of 10−3 eV2. The neutrinos are relatively more sensitive to the matter effects than the antineutrinos. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | ABSTRACT Details are presented of the HI Jodrell All Sky Survey (HIJASS). HIJASS is a blind neutral hydrogen (HI) survey of the northern sky (δ>22◦), being conducted using the multibeam receiver on the Lovell Telescope (FWHM beamwidth 12 arcmin) at Jodrell Bank. HIJASS covers the velocity range –3500 km s−1 to 10000 km s−1, with a velocity resolution of 18.1 km s−1 and spatial positional accuracy of ∼2.5 arcmin. Thus far about 1115 deg2 of sky have been surveyed. The average rms noise during the early part of the survey was around 16 mJy beam−1. Following the first phase of the Lovell telescope upgrade (in 2001), the rms noise is now around 13 mJy beam−1. We describe the methods of detecting galaxies within the HIJASS data and of measuring their HI parameters. The properties of the resulting HI-selected sample of galaxies are described. Of the 222 sources so far confirmed, 170 (77 per cent) are clearly associ- ated with a previously catalogued galaxy. A further 23 sources (10 per cent) lie close (within 6 arcmin) to a previously catalogued galaxy for which no previous redshift exists. A further 29 sources (13 per cent) do not appear to be associated with any previously catalogued galaxy. The distributions of peak flux, integrated flux, HI mass and cz are discussed. We show, using the HIJASS data, that HI self-absorption is a significant, but often overlooked, effect in galaxies with large inclination angles to the line of sight. Properly accounting for it could increase the derived HI mass density of the local Universe by at least 25 per cent. The effect this will have on the shape of the HI Mass Function (HIMF) will depend on how self-absorption affects galaxies of different morphological types and HI masses. We also show that galaxies with small inclinations to the line of sight may also be excluded from HI-selected samples, since many such galaxies will have observed velocity-widths which are too narrow for them to be distinguished from narrow-band radio frequency interference. This effect will become progressively more serious for galaxies with smaller intrinsic velocity-widths. If, as we might expect, galaxies with smaller intrinsic velocity-widths have smaller HI masses, then compensating for this effect could significantly steepen the faint-end slope of the derived HIMF. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | Key words: surveys – galaxies: evolution – galaxies: luminosity function, mass func- tion – galaxies: distances and redshifts – large-scale structure of Universe. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | surveys. There is an inevitable bias in such surveys against low luminosity objects (dwarfs), but also against low surface brightness (LSB) objects (see, e.g. Disney 1976; Disney & Phillipps 1987; Impey & Bothun 1997; Disney 1999). |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | However, it has become clear that low luminosity and low surface brightness galaxies play a key role in many cosmological and cosmographical problems. For example, dwarf/LSB galaxies can clearly play a major role in help- ing us to understand large-scale structure and its influence on galaxy formation and evolution. Their numbers and dis- tribution place constraints on the increasingly sophisticated numerical and semi-analytic models of galaxy formation (e.g., Baugh, Cole & Frenk 1996; Kauffmann et al. 1997), while their morphologies and stellar contents may reflect the local physics which define the star formation process in galaxies (e.g., Bell & Bower 2000; Bell & de Jong 2000). |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | Recent observational studies have fully supported the view (expounded by, e.g., Phillipps et al. 1987 and Impey, Bothun & Malin 1988) that low luminosity and low surface brightness galaxies numerically dominate the galaxy popu- lation in the local Universe (see e.g. McGaugh 1996; Cross et al. 2001). However, optical surveys of the local Universe for faint/LSB objects are problematic due to the very long ex- posure times required, the large areas which need to be sur- veyed and the need to measure a redshift for each faint/LSB object found. Consequently, our knowledge of the local pop- ulation of galaxies at low luminosity and low surface bright- ness is still relatively limited. This inhibits our knowledge of many broader cosmological/cosmographical issues. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | Given the limitations of optical surveys for detecting low luminosity / LSB objects, an alternative method to sam- ple the extragalactic population is to use the 21-cm neutral hydrogen (HI) line. This provides a way of potentially cir- cumventing optical selection effects operating against low luminosity and/or LSB objects, since a galaxy’s HI content may be relatively uncorrelated with its optical emission. For example, it is well known that elliptical galaxies contain lit- tle HI whereas we might expect to find large amounts of HI in galaxies where star formation has been inefficient, e.g. in low luminosity and LSB galaxies. However, until compar- atively recently, most HI surveys were limited to HI mea- surements of galaxies previously detected in optical or IR surveys. The advent of the 21-cm multibeam receiving sys- tems at Parkes and Jodrell Bank has made possible, for the first time, blind HI surveys of large areas of sky to reasonable sensitivity over comparatively large volumes. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | The HI Jodrell All Sky Survey (HIJASS) is the northern counterpart to HIPASS. HIJASS will survey the northern sky above δ=22◦ to similar sensitivity to HIPASS, using the Multibeam 4-beam cryogenic receiver mounted on the 76-m Lovell Telescope. HIJASS was begun in 2000. So far ≃1115 deg2 have been surveyed. We recently presented re- sults from the HIJASS data covering the M81 group (Boyce et al. 2001). The survey reveals several new aspects to the complex morphology of the HI distribution in the group and illustrates that a blind HI survey of even such a nearby, well studied group of galaxies can add much new information. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | This paper presents a detailed description of the HI Jo- drell All Sky Survey and of the properties of the HI-selected sample of galaxies which has been compiled so far from the HIJASS data. We use the sample of confirmed HIJASS sources to study the effect that a galaxy’s inclination to the line of sight has on its inclusion within an HI-selected sam- ple. We show that both highly inclined galaxies and galaxies close to face-on are subject to selection effects which could have led to their being under-represented in previous deter- minations of the HI Mass Function (HIMF) and HI mass density, ΩHI, from HI-selected samples of galaxies. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | Section 2 describes the hardware, the observing strategy and survey parameters and also describes the data reduction methods. Section 3 describes the methods by which galaxies have been detected within HIJASS data and their param- eters measured. Section 4 is a discussion of the properties of the sample of galaxies found in HIJASS data thus far. In Section 5 we use the sample of confirmed HIJASS sources to study the inclination-dependent selection effects on the inclusion of a galaxy in an HI-selected sample and discuss the implications of this. Section 6 presents some concluding remarks. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | HIJASS uses a cryogenic Multibeam receiver (Bird 1997) of similar design to the Multibeam receiver used at Parkes for HIPASS (Staveley-Smith et al. 1996). The Multibeam system installed at Jodrell Bank has four dual linearly po- larised receivers covering a frequency range of 1200 MHz to 1550 MHz. The feed horn array consists of 4 stepped cir- cular horns (which were designed at the CSIRO) arranged in a rhombic pattern, the apertures of which are located at the telescope prime focus. Stepped circular horns were cho- sen because of their good pattern symmetry, low spillover and good cross polarization properties (Bird 1994). The horns couple directly into a low temperature, high vacuum, cryogenic dewar. The output from the feed horn arrays are then fed to a set of 3-stage high electron mobility transistor (HEMT) pre-amplifiers which are cooled to a temperature of around ∼25 K. Following amplification, each receiver RF band is then down converted to an IF bandpass which can be set anywhere between 30 MHz and 245 MHz. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | Figure 1. A section of a HIJASS cube at roughly constant R.A.. Visible are the residue of Galactic HI at 0 km˙s−1, the galaxy UGC06534 at V⊙ ≃1300 km s−1; and the radio frequency interference between 4500→7500 km s−1. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | the focus cabin to the observing room via about 300 m of low loss coaxial cable which is terminated into N-socket connec- tions in the Lovell Observing Room. These outputs are next patched into a set of digitally programmable attenuators and are then fed into a set of equaliser and splitter units. Each IF is equalised for frequency dependent cable losses and then split into two outputs. The first output connects to a filter bank which sets the bandpass which is presented to the correlator. A second set of outputs are used for pulsar survey measurements. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | multibeam correlator, one of these chips, operating in au- tocorrelation mode, is used on each of the 8 sampled data streams, thereby providing, after Fourier transformation, a measurement of the input spectrum at 1024 contiguous fre- quencies spaced at 62.5 kHz (equivalent to 13.2 km s−1 at the rest frequency of HI). |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | tem temperature is ∼30 K. Bandpass correction and cali- bration are applied using the software package LIVEDATA (see Barnes et al. 2001). The spectra are gridded into three- dimensional 8◦×8◦ datacubes (α,δ,V⊙) using the software package GRIDZILLA (Barnes et al. 2001). The observed spectra are smoothed online by applying a 25 per cent Tukey filter to reduce ‘ringing’ caused by strong Galactic signals entering through the side-lobes. This reduces the actual ve- locity resolution in the gridded datacubes to 18.1 km s−1. The spatial pixel size of the datacubes is 4 arcmin×4 ar- cmin. The effects of continuum emission on the baselines of the spectra in each cube are then removed by the pro- gram POLYCON written by Daniel Zambonini and Robert Minchin. This program fits and then subtracts a 5th order polynomial baseline to each individual spectrum in a dat- acube: fitting is only performed on parts of the spectrum free of interference or line emission. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | HIJASS has been conducted during three observing runs: in April-June 2000; in Jan-Feb 2001; and in Jan 2002. Ta- ble 1 notes those areas of the northern sky so far surveyed by HIJASS and during which run the data were taken. The whole of a strip in R.A. between Decl.=70◦→78◦ has been surveyed (795 deg2). Two areas of the R.A. strip between Decl.=62◦→70◦have also been surveyed (192 deg2), along with smaller areas at Decls.=58◦, 34◦, and 26◦. In total 1115 deg2 has been surveyed thus far. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | Fig. 1 shows an example of the data. This is a plot of Decl. against Velocity at roughly constant R.A.. The HIJASS source HIJASS J1133+63 can be seen at Decl.≃63◦, V⊙=1300 km s−1. This has been identified as UGC 06534. Prominent in the data is a broad band of radio fre- quency interference (RFI) which affects all velocities from ≃4500 km s−1 to 7500 km s−1. This is mainly due to off-site radio and data communication emissions generated in the locality. |
Processed_First_Results_from_the_HI_Jodrell_All_Sky_Survey:_.txt | Between the observing runs in 2001 and 2002, the Lovell dish underwent the first stage of a major upgrade. A new telescope drive system was installed and around half of the dish surface was replaced. |
Subsets and Splits