Fig 11 - uploaded by Lorenzo Moncelsi
Content may be subject to copyright.
-Plot of the 'dust-mass function', the space density of galaxies as a function of dust mass. We have estimated this for five redshift slices, and the key is the same as for Fig. 8. The thick dashed line shows the Schechter function that is the best fit to the dust-mass function in the lowest redshift slice.

-Plot of the 'dust-mass function', the space density of galaxies as a function of dust mass. We have estimated this for five redshift slices, and the key is the same as for Fig. 8. The thick dashed line shows the Schechter function that is the best fit to the dust-mass function in the lowest redshift slice.

Source publication
Article
Full-text available
The Balloon-borne Large Aperture Submillimeter Telescope (BLAST) has recently surveyed ~8.7 deg^2 centered on GOODS-South at 250, 350, and 500 microns. In Dye et al. (2009) we presented the catalogue of sources detected at 5-sigma in at least one band in this field and the probable counterparts to these sources in other wavebands. In this paper, we...

Contexts in source publication

Context 1
... which n is the number of galaxies with dust masses and redshifts that fall within this bin and V is calculated using equation (8). Figure 11 shows the results for the five redshift slices without making any correction for missing counterparts. There is clearly strong evolution, in the sense that the space-density of the galaxies with the highest dust masses increases steadily with redshift. ...
Context 2
... et al. (2009) concluded that there was no evolution in the comoving density of dust in the universe. However, Pascale et al. effectively measured φ < M d > in each redshift slice, and Figure 11 shows that this does not change very much: the average dust mass of the galaxies detected at low redshift is lower than at high redshift but their space-density is higher. It is only by comparing the space-density at different redshifts but at the same dust mass that it is possible to see the evolution. ...
Context 3
... that the result in Fig. 11 is very insensitive to our assumptions about temperature, because on the Rayleigh-Jeans tail the Planck function in equation 10 only depends on the first power of dust temperature. The strength of the evolution would be less if the temperature of the bulk of the dust at high redshift were higher than that at low redshift. But even if ...
Context 4
... strength of the evolution would be less if the temperature of the bulk of the dust at high redshift were higher than that at low redshift. But even if the temperature were a factor of two greater at high redshift, the effect on the high-redshift points in Figure 11 would be to move them a factor of two to the left, which is not enough to remove the result. It is possible to think of scenarios in which the evolution was caused by temperature. ...
Context 5
... the evolution was caused by temperature. Suppose that as one moves to higher redshift, the fraction of BLAST galaxies that contain a luminous but obscured quasar gradually increases, and by a redshift of ∼ 1 the temperature of the dust heated by the hidden quasar is a factor of 10 greater than at z = 0. This would explain the evolution seen in Fig. 11. However, because of the strong dependence of bolometric luminosity on temperature, this increase in temperature would correspond to a increase in bolometric luminosity of at least a factor of 10 5 . Therefore, it is much harder to explain the evolution visible in Fig. 11 as a temperature effect than as an increase in the number of ...
Context 6
... of 10 greater than at z = 0. This would explain the evolution seen in Fig. 11. However, because of the strong dependence of bolometric luminosity on temperature, this increase in temperature would correspond to a increase in bolometric luminosity of at least a factor of 10 5 . Therefore, it is much harder to explain the evolution visible in Fig. 11 as a temperature effect than as an increase in the number of galaxies with high dust ...
Context 7
... the increased luminosity-density is caused by an increase in the global star-formation rate, it is possible, for example, that this is caused by a larger number of galaxy interactions at high redshift, which trigger starbursts, and not necessarily by the larger amount of interstellar material in galaxies. However, Figure 11 shows that the space-density of galaxies with high dust masses, and thus presumably large reservoirs of interstellar material, is also evolving strongly with redshift. ...

Similar publications

Article
Full-text available
We show that populations of high redshift galaxies grow more luminous as they grow in linear size. This is because the brightness per unit area has a distinct upper limit due to the self-regulating nature of star-formation. As a corollary, we show that the observed increase in characteristic luminosity of Lyman Break Galaxies (L*) with time can be...
Article
Full-text available
In this study we predict the total distributions of powerful (FR II) active double-lobed radio galaxies and ghost sources, and their observable distribution in the X-ray sky. We develop an analytic model for the evolution of the lobe emission at radio and X-ray energies. During jet activity, a double radio source emits synchrotron radiation in the...
Article
Full-text available
The surface brightness profile of H-alpha emission in galaxies is generally thought to be confined by a sharp truncation, sometimes speculated to coincide with a star formation threshold. Over the past years, observational evidence for both old and young stellar populations, as well as individual H II regions, has demonstrated that the outer disk i...
Article
Full-text available
We investigate and quantify the observed scatter in the empirical relationship between the broad line region size R and the luminosity of the active galactic nucleus (AGN), in order to better understand its origin. This study is motivated by the indispensable role of this relationship in the mass estimation of cosmologically distant black holes, bu...
Article
Full-text available
We have analyzed the radial distribution of old stars in a sample of 218 nearby face-on disks, using deep 3.6 μm images from the Spitzer Survey of Stellar Structure in Galaxies. In particular, we have studied the structural properties of those disks with a broken or down-bending profile. We find that, on average, disks with a genuine single-exponen...

Citations

... This process provides a total spectroscopic sample of 14811 cross-matched sources in XMM-LSS and 18811 sources in COSMOS. Within CDFS, a spectroscopic catalogue of 23947 cross-matched sources was obtained from the Spitzer Data Fusion database (Vaccari 2015), featuring results from surveys including VVDS (LeFèvre et al. 2013), BLAST (Eales et al. 2009), ACES (Cooper et al. 2012), GOODS-CANDELS (Hsu et al. 2014), OzDES (Yuan et al. 2015;Childress et al. 2017) and VUDS (Tasca et al. 2017). We also include new spectroscopic redshifts obtained from the LADUMA collaboration using the Anglo-Australian Telescope (Wu et al. in prep). ...
Article
Full-text available
We present measurements of the rest-frame ultraviolet luminosity function at redshifts z = 3, z = 4 and z = 5, using 96894, 38655 and 7571 sources respectively to map the transition between AGN and galaxy-dominated ultraviolet emission shortly after the epoch of reionization. Sources are selected using a comprehensive photometric redshift approach, using 10 deg2 of deep extragalactic legacy fields covered by both HSC and VISTA. The use of template fitting spanning a wavelength range of 0.3–2.4μm achieves 80–90 per cent completeness, much higher than classical colour-colour cut methodology. The measured LF encompasses −26 < MUV < −19.25. This is further extended to −28.5 < MUV < −16 using complementary results from other studies, allowing for the simultaneous fitting of the combined AGN and galaxy LF. We find that there are fewer UV luminous galaxies (MUV < −22) at z ∼ 3 than z ∼ 4, indicative of an onset of widespread quenching alongside dust obscuration, and that the evolution of the AGN LF is very rapid, with their number density rising by around 2 orders of magnitude from 3 < z < 6. It remains difficult to determine if a double power law (DPL) functional form is preferred over the Schechter function to describe the galaxy UV LF. Estimating the Hydrogen ionizing photon budget from our UV LFs, we find that AGN can contribute to, but cannot solely maintain, the reionization of the Universe at z = 3 − 5. However, the rapidly evolving AGN LF strongly disfavours a significant contribution within the EoR.
... Table 1 lists selected photometry for the source in the infrared and radio. Optical spectroscopy has yielded two independent redshift measurements of z = 0.5245 (Eales et al. 2009) and z = 0.5247 (Mao et al. 2012), neither with a quoted uncertainty; in this paper, we average these two measurements and adopt the ∼100 km s −1 uncertainty that is typical for AAT redshifts of similar vintage (see, e.g., Section 1 of Baldry et al. 2014), yielding z opt = 0.5246 ± 0.0005. ...
Article
Full-text available
In the local universe, OH megamasers (OHMs) are detected almost exclusively in infrared-luminous galaxies, with a prevalence that increases with IR luminosity, suggesting that they trace gas-rich galaxy mergers. Given the proximity of the rest frequencies of OH and the hyperfine transition of neutral atomic hydrogen (H i ), radio surveys to probe the cosmic evolution of H i in galaxies also offer exciting prospects for exploiting OHMs to probe the cosmic history of gas-rich mergers. Using observations for the Looking At the Distant Universe with the MeerKAT Array (LADUMA) deep H i survey, we report the first untargeted detection of an OHM at z > 0.5, LADUMA J033046.20−275518.1 (nicknamed “Nkalakatha”). The host system, WISEA J033046.26−275518.3, is an infrared-luminous radio galaxy whose optical redshift z ≈ 0.52 confirms the MeerKAT emission-line detection as OH at a redshift z OH = 0.5225 ± 0.0001 rather than H i at lower redshift. The detected spectral line has 18.4 σ peak significance, a width of 459 ± 59 km s ⁻¹ , and an integrated luminosity of (6.31 ± 0.18 [statistical] ± 0.31 [systematic]) × 10 ³ L ⊙ , placing it among the most luminous OHMs known. The galaxy’s far-infrared luminosity L FIR = (1.576 ±0.013) × 10 ¹² L ⊙ marks it as an ultraluminous infrared galaxy; its ratio of OH and infrared luminosities is similar to those for lower-redshift OHMs. A comparison between optical and OH redshifts offers a slight indication of an OH outflow. This detection represents the first step toward a systematic exploitation of OHMs as a tracer of galaxy growth at high redshifts.
... Table 1 lists selected photometry for the source in the infrared and radio. Optical spectroscopy has yielded two independent redshift measurements of z = 0.5245 (Eales et al. 2009) and z = 0.5247 (Mao et al. 2012), neither with a quoted uncertainty; in this paper, we average these two measurements and adopt the ∼ 100 km s −1 uncertainty that is typical for AAT redshifts of similar vintage (see, e.g., §1 of Baldry et al. 2014), yielding z opt = 0.5246±0.0005. We can determine the host galaxy's infrared luminosity using the source redshift and far-infrared photom-2 https://docs.ilifu.ac.za/#/about/what is etry. ...
Preprint
Full-text available
In the local Universe, OH megamasers (OHMs) are detected almost exclusively in infrared-luminous galaxies, with a prevalence that increases with IR luminosity, suggesting that they trace gas-rich galaxy mergers. Given the proximity of the rest frequencies of OH and the hyperfine transition of neutral atomic hydrogen (HI), radio surveys to probe the cosmic evolution of HI in galaxies also offer exciting prospects for exploiting OHMs to probe the cosmic history of gas-rich mergers. Using observations for the Looking At the Distant Universe with the MeerKAT Array (LADUMA) deep HI survey, we report the first untargeted detection of an OHM at $z > 0.5$, LADUMA J033046.20$-$275518.1 (nicknamed "Nkalakatha"). The host system, WISEA J033046.26$-$275518.3, is an infrared-luminous radio galaxy whose optical redshift $z \approx 0.52$ confirms the MeerKAT emission line detection as OH at a redshift $z_{\rm OH} = 0.5225 \pm 0.0001$ rather than HI at lower redshift. The detected spectral line has 18.4$\sigma$ peak significance, a width of $459 \pm 59\,{\rm km\,s^{-1}}$, and an integrated luminosity of $(6.31 \pm 0.18\,{\rm [statistical]}\,\pm 0.31\,{\rm [systematic]}) \times 10^3\,L_\odot$, placing it among the most luminous OHMs known. The galaxy's far-infrared luminosity $L_{\rm FIR} = (1.576 \pm 0.013) \times 10^{12}\,L_\odot$ marks it as an ultra-luminous infrared galaxy; its ratio of OH and infrared luminosities is similar to those for lower-redshift OHMs. A comparison between optical and OH redshifts offers a slight indication of an OH outflow. This detection represents the first step towards a systematic exploitation of OHMs as a tracer of galaxy growth at high redshifts.
... All stamps have a 20 arcsec side and are produced using the H V library. Further details in Section 5.2 archive 7(Ochsenbein et al. 2000) querying the catalogues from previous spectroscopic surveys of the CDFS(Cooper et al. 2012;Eales et al. 2009; Cowie et al. 2011). 7 https://vizier.u-strasbg.fr/ ...
Preprint
We present a sample of 16 likely strong gravitational lenses identified in the VST Optical Imaging of the CDFS and ES1 fields (VOICE survey) using Convolutional Neural Networks (CNNs). We train two different CNNs on composite images produced by superimposing simulated gravitational arcs on real Luminous Red Galaxies observed in VOICE. Specifically, the first CNN is trained on single-band images and more easily identifies systems with large Einstein radii, while the second one, trained on composite RGB images, is more accurate in retrieving systems with smaller Einstein radii. We apply both networks to real data from the VOICE survey, taking advantage of the high limiting magnitude (26.1 in the r-band) and low PSF FWHM (0.8" in the r-band) of this deep survey. We analyse $\sim21,200$ images with $mag_r<21.5$, identifying 257 lens candidates. To retrieve a high-confidence sample and to assess the accuracy of our technique, nine of the authors perform a visual inspection. Roughly 75% of the systems are classified as likely lenses by at least one of the authors. Finally, we assemble the LIVE sample (Lenses In VoicE) composed by the 16 systems passing the chosen grading threshold. Three of these candidates show likely lensing features when observed by the Hubble Space Telescope. This work represents a further confirmation of the ability of CNNs to inspect large samples of galaxies searching for gravitational lenses. These algorithms will be crucial to exploit the full scientific potential of forthcoming surveys with the Euclid satellite and the Vera Rubin Observatory
... Recent galaxy surveys have measured the dust mass function (Dunne, Eales & Edmunds 2003;Vlahakis, Dunne & Eales 2005;Dunne et al. 2011;Eales et al. 2009;Clemens et al. 2013) and many scaling relations between dust and the fundamental properties of galaxies. These include the relationship between dust mass and stellar mass (Santini et al. 2014) and the relationship between dust mass and star formation rate (SFR; da Cunha et al. 2010;Santini et al. 2014). ...
... When considering dust, observers typically report the properties of the ISM (e.g. Eales et al. 2009;Rémy-Ruyer et al. 2014;Santini et al. 2014;da Cunha et al. 2015;Mancini et al. 2015;Nersesian et al. 2019). Although some works have extended dust measurements to include the CGM (Ménard et al. 2010;Peek et al. 2015), such observations are rare. ...
... number of these galaxies, especially at low redshift (e.g. Dunne et al. 2003;Eales et al. 2009;Clemens et al. 2013;Rémy-Ruyer et al. 2014;da Cunha et al. 2015). ...
Article
We explore the relation between dust and several fundamental properties of simulated galaxies using the Dusty SAGE semi-analytic model. In addition to tracing the standard galaxy properties, Dusty SAGE also tracks cold dust mass in the interstellar medium (ISM), hot dust mass in the halo and dust mass ejected by feedback activity. Based on their ISM dust content, we divide our galaxies into two categories: ISM dust-poor and ISM dust-rich. We split the ISM dust-poor group into two subgroups: halo dust-rich and dust-poor (the latter contains galaxies that lack dust in both the ISM and halo). Halo dust-rich galaxies have high outflow rates of heated gas and dust and are more massive. We divide ISM dust-rich galaxies based on their specific star formation rate (sSFR) into star-forming and quenched subgroups. At redshift z = 0, we find that ISM dust-rich galaxies have a relatively high sSFR, low bulge-to-total (BTT) mass ratio, and high gas metallicity. The high sSFR of ISM dust-rich galaxies allows them to produce dust in the stellar ejecta. Their metal-rich ISM enables dust growth via grain accretion. The opposite is seen in the ISM dust-poor group. Furthermore, ISM dust-rich galaxies are typically late-types, while ISM dust-poor galaxies resemble the early-type population, and we show how their ISM content evolves from being dust-rich to dust-poor. Finally, we investigate dust production from z = 3 to z = 0 and find that all groups evolve similarly, except for the quenched ISM dust-rich group.
... Recent galaxy surveys have measured the dust mass function (Dunne et al. 2003;Vlahakis et al. 2005;Dunne et al. 2011;Eales et al. 2009;Clemens et al. 2013) and many scaling relations between dust and the fundamental properties of galaxies. These include the relationship between dust mass and stellar mass (Santini et al. 2014) and the relationship between dust mass and star formation rate (SFR) (da Cunha et al. 2010;Santini et al. 2014). ...
... When considering dust, observers typically report the properties of the ISM (e.g., Eales et al. 2009;Rémy-Ruyer et al. 2014;Santini et al. 2014;Mancini et al. 2015;da Cunha et al. 2015;Nersesian et al. 2019). Although some works have extended dust measurements to include the CGM (Ménard et al. 2010;Peek et al. 2015), such observations are rare. ...
... Dust obscures the intrinsic stellar spectra of such galaxies and re-emits the radiation in the infrared (Witt et al. 1992;Witt & Gordon 2000). Infrared and sub-millimeter surveys have found a substantial number of these galaxies, especially at low redshift (e.g., Eales et al. 2009;Dunne et al. 2003;da Cunha et al. 2015;Clemens et al. 2013;Rémy-Ruyer et al. 2014). ...
Preprint
Full-text available
We explore the relation between dust and several fundamental properties of simulated galaxies using the Dusty SAGE semi-analytic model. In addition to tracing the standard galaxy properties, Dusty SAGE also tracks cold dust mass in the interstellar medium (ISM), hot dust mass in the halo and dust mass ejected by feedback activity. Based on their ISM dust content, we divide our galaxies into two categories: ISM dust-poor and ISM dust-rich. We split the ISM dust-poor group into two subgroups: halo dust-rich and dust-poor (the latter contains galaxies that lack dust in both the ISM and halo). Halo dust-rich galaxies have high outflow rates of heated gas and dust and are more massive. We divide ISM dust-rich galaxies based on their specific star formation rate (sSFR) into star-forming and quenched subgroups. At redshift z=0, we find that ISM dust-rich galaxies have a relatively high sSFR, low bulge-to-total (BTT) mass ratio, and high gas metallicity. The high sSFR of ISM dust-rich galaxies allows them to produce dust in the stellar ejecta. Their metal-rich ISM enables dust growth via grain accretion. The opposite is seen in the ISM dust-poor group. Furthermore, ISM dust-rich galaxies are typically late-types, while ISM dust-poor galaxies resemble the early-type population, and we show how their ISM content evolves from being dust-rich to dust-poor. Finally, we investigate dust production from z=3 to z=0 and find that all groups evolve similarly, except for the quenched ISM dust-rich group.
... The first measurements of the local DMF were made using sub-mm observations at 450 m and 850 m, using an indirect technique and were hampered by small number statistics (Dunne et al. 2000;Dunne & Eales 2001;Vlahakis et al. 2005). Early efforts to probe beyond the local universe ( = 1, Eales et al. 2009;= 2.5, Dunne et al. 2003a) also suffered from poor statistics and a variety of other experimental limitations. ...
Preprint
We investigate the evolution in galactic dust mass over cosmic time through i) empirically derived dust masses using stacked submillimetre fluxes at 850um in the COSMOS field, and ii) dust masses derived using a robust post-processing method on the results from the cosmological hydrodynamical simulation IllustrisTNG. We effectively perform a self-calibration of the dust mass absorption coefficient by forcing the model and observations to agree at low redshift and then compare the evolution shown by the observations with that predicted by the model. We create dust mass functions (DMFs) based on the IllustrisTNG simulations from 0 < z < 0.5 and compare these with previously observed DMFs. We find a lack of evolution in the DMFs derived from the simulations, in conflict with the rapid evolution seen in empirically derived estimates of the low redshift DMF. Furthermore, we observe a strong evolution in the observed mean ratio of dust mass to stellar mass of galaxies over the redshift range 0 < z < 5, whereas the corresponding dust masses from IllustrisTNG show relatively little evolution, even after splitting the sample into satellites and centrals. The large discrepancy between the strong observed evolution and the weak evolution predicted by IllustrisTNG plus post-processing may be explained by either strong cosmic evolution in the properties of the dust grains or limitations in the model. In the latter case, the limitation may be connected to previous claims that the neutral gas content of galaxies does not evolve fast enough in IllustrisTNG.
... At submm wavelengths, tentative evidence for similar evolution was provided by Eales et al. (2009), based on BLAST balloon observations of the GOODS-South field: the data suggest strong evolution out to z = 1 in both monochromatic luminosity function and dust mass function, particularly among the higher luminosity/mass systems. One of the first results of the Herschel mission was a confirmation of this tentative evidence. ...
Preprint
We present infrared luminosity functions and dust mass functions for the EAGLE cosmological simulation, based on synthetic multi-wavelength observations generated with the SKIRT radiative transfer code. In the local Universe, we reproduce the observed infrared luminosity and dust mass functions very well. Some minor discrepancies are encountered, mainly in the high luminosity regime, where the EAGLE-SKIRT luminosity functions mildly but systematically underestimate the observed ones. The agreement between the EAGLE-SKIRT infrared luminosity functions and the observed ones gradually worsens with increasing lookback time. Fitting modified Schechter functions to the EAGLE-SKIRT luminosity and dust mass functions at different redshifts up to $z=1$, we find that the evolution is compatible with pure luminosity/mass evolution. The evolution is relatively mild: within this redshift range, we find an evolution of $L_{\star,250}\propto(1+z)^{1.68}$, $L_{\star,\text{TIR}}\propto(1+z)^{2.51}$ and $M_{\star,\text{dust}}\propto(1+z)^{0.83}$ for the characteristic luminosity/mass. For the luminosity/mass density we find $\varepsilon_{250}\propto(1+z)^{1.62}$, $\varepsilon_{\text{TIR}}\propto(1+z)^{2.35}$ and $\rho_{\text{dust}}\propto(1+z)^{0.80}$, respectively. The mild evolution of the dust mass density is in relatively good agreement with observations, but the slow evolution of the infrared luminosity underestimates the observed luminosity evolution significantly. We argue that these differences can be attributed to increasing limitations in the radiative transfer treatment due to increasingly poorer resolution, combined with a slower than observed evolution of the SFR density in the EAGLE simulation and the lack of AGN emission in our EAGLE-SKIRT post-processing recipe.
... Dust properties in galaxies have been intensively studied through statistics and scaling relations, of which three particularly interesting are dust mass functions (DMFs; Dunne, Eales & Edmunds 2003;Vlahakis, Dunne & Eales 2005;Eales et al. 2009;Dunne et al. 2011;Clemens et al. 2013;Beeston et al. 2018), dust-to-gas ratios (DGRs) and dust-to-metal ratios (DTMs) as a function of galaxy metallicity or stellar mass (Issa, MacLaren & Wolfendale 1990;Lisenfeld & Ferrara 1998;Hirashita, Tajiri & Kamaya 2002;Draine et al. 2007;Galametz et al. 2011;De Cia et al. 2013, 2016Zafar & Watson 2013;Rémy-Ruyer et al. 2014;Sparre et al. 2014;Giannetti et al. 2017;Wiseman et al. 2017;Chiang et al. 2018; Kahre et al. 2018;De Vis et al. 2019). These relationships provide a convenient method for determining gas masses in galaxies, as well as providing constraints on the baryon cycle that governs galaxy evolution at low and high redshifts (e.g. ...
... When comparing to observational data sets, we select galaxies within particular redshift bins as follows. ForEales et al. (2009), we plot data from 0.6 < z < 1.0. ForDunne et al. (2011) andBeeston et al. ( ...
Article
We present predictions for the evolution of the galaxy dust-to-gas ratio (DGR) and dust-to-metal ratio (DTM) from z = 0 → 6, using a model for the production, growth, and destruction of dust grains implemented into the simba cosmological hydrodynamic galaxy formation simulation. In our model, dust forms in stellar ejecta, grows by the accretion of metals, and is destroyed by thermal sputtering and supernovae. Our simulation reproduces the observed dust mass function at z = 0, but modestly underpredicts the mass function by ∼×3 at z ∼ 1–2. The z = 0 DGR versus metallicity relationship shows a tight positive correlation for star-forming galaxies, while it is uncorrelated for quenched systems. There is little evolution in the DGR–metallicity relationship between z = 0 and 6. We use machine learning techniques to search for the galaxy physical properties that best correlate with the DGR and DTM. We find that the DGR is primarily correlated with the gas-phase metallicity, though correlations with the depletion time-scale, stellar mass, and gas fraction are non-negligible. We provide a crude fitting relationship for DGR and DTM versus the gas-phase metallicity, along with a public code package that estimates the DGR and DTM given a set of galaxy physical properties.
... The 'A' sample determines dust masses using a dust temperature obtained from isothermal SED fitting, and the 'B' DMF has been calculated using a dust temperature of 20 K. Clemens et al. (2013) combined Herschel data with Wide-field Infrared Survey Explorer (WISE), Spitzer and Infrared Astronomical Satellite (IRAS) observations to investigate the properties of a flux-limited sample of local star-forming galaxies. They fit their SEDs with modified blackbody spectra using β 2 and dust temperatures in the range 10-25 K. Eales et al. (2009) use data obtained from the Balloon-borne Large Aperture Submillimeter Telescope (BLAST), using the grey body relation assuming a dust temperature of 20 K. ...
... This result is consistent with that of the previous section, that we overpredict the dust content of many massive galaxies at z = 0 in our model. On comparing our model predictions to the Eales et al. (2009) data for z = 1, we instead appear to slightly underpredict the DMF at high masses. It is worthwhile to note that this is a general feature seen in other models of galaxy formation tracking dust growth (e.g. ...
... The black line shows the prediction of our model using the underlying dark matter Millennium simulation, and the red line for Millennium-II. Observations are taken fromVlahakis et al. (2005) andClemens et al. (2013) at z = 0, and fromEales et al. (2009) for z = 1. ...
Article
We implement a detailed dust model into the L-Galaxies semi-analytical model which includes: injection of dust by type II and type Ia supernovae (SNe) and AGB stars; grain growth in molecular clouds; and destruction due to supernova-induced shocks, star formation, and reheating. Our grain growth model follows the dust content in molecular clouds and the inter-cloud medium separately, and allows growth only on pre-existing dust grains. At early times, this can make a significant difference to the dust growth rate. Above z ∼ 8, type II SNe are the primary source of dust, whereas below z ∼ 8, grain growth in molecular clouds dominates, with the total dust content being dominated by the latter below z ∼ 6. However, the detailed history of galaxy formation is important for determining the dust content of any individual galaxy. We introduce a fit to the dust-to-metal (DTM) ratio as a function of metallicity and age, which can be used to deduce the DTM ratio of galaxies at any redshift. At z ≲ 3, we find a fairly flat mean relation between metallicity and the DTM, and a positive correlation between metallicity and the dust-to-gas (DTG) ratio, in good agreement with the shape and normalization of the observed relations. We also match the normalization of the observed stellar mass–dust mass relation over the redshift range of 0–4, and to the dust mass function at z = 0. Our results are important in interpreting observations on the dust content of galaxies across cosmic time, particularly so at high redshift.