text
stringlengths
0
643k
meta
stringlengths
137
151
# Introduction Let \(G\) be a simple undirected graph with the *vertex set* \(V(G)\) and the *edge set* \(E(G)\). A vertex with degree one is called a *pendant vertex*. The distance between the vertices \(u\) and \(v\) in graph \(G\) is denoted by \(d_G(u,v)\). A cycle \(C\) is called *chordless* if \(C\) has no *cycle chord* (that is an edge not in the edge set of \(C\) whose endpoints lie on the vertices of \(C\)). The *Induced subgraph* on vertex set \(S\) is denoted by \(\langle S\rangle\). A path that starts in \(v\) and ends in \(u\) is denoted by \(\stackrel\frown{v u}\). A *traceable* graph is a graph that possesses a Hamiltonian path. In a graph \(G\), we say that a cycle \(C\) is *formed by the path* \(Q\) if \(| E(C) \setminus E(Q) | = 1\). So every vertex of \(C\) belongs to \(V(Q)\). In 2011 the following conjecture was proposed: Conjecture \(\,\) also appears in Problem 516. There are a few partial results known for Conjecture . Kostochka noticed that the Petersen graph, the prisms over cycles, and many other graphs have a decomposition desired in Conjecture . Ozeki and Ye proved that the conjecture holds for 3-connected cubic plane graphs. Furthermore, it was proved by Bachstein that Conjecture \(\,\) is true for every 3-connected cubic graph embedded in torus or Klein-bottle. Akbari, Jensen and Siggers showed that Conjecture \(\,\) is true for Hamiltonian cubic graphs. In this paper, we show that Conjecture \(\,\) holds for traceable cubic graphs. # Results Before proving the main result, we need the following lemma. **Remark 1.** [\[remark:1\]]{#remark:1 label="remark:1"} Let \(C\) be a cycle formed by the path \(Q\). Then clearly there exists a chordless cycle formed by \(Q\). Now, we are in a position to prove the main result. **Remark 2.** [\[remark:2\]]{#remark:2 label="remark:2"} Indeed, in the proof of the previous theorem we showed a stronger result, that is, for every traceable cubic graph there is a decomposition with at most two cycles.
{'timestamp': '2016-07-19T02:04:55', 'yymm': '1607', 'arxiv_id': '1607.04768', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04768'}
# Principle of nano strain-amplifier Figure [\[fig:fig1\]](#fig:fig1){reference-type="ref" reference="fig:fig1"}(a) and 1(b) show the concept of the conventional structures of piezoresistive sensors. The piezoresistive elements are either released from, or kept on, the substrate. The sensitivity (\(S\)) of the sensors is defined based on the ratio of the relative resistance change (\(\Delta R/R\)) of the sensing element and the strain applied to the substrate (\(\varepsilon_{sub}\)): \[S = (\Delta R/R)/\varepsilon_{sub} \label{eq:sensitivity}\] In addition, the relative resistance change \(\Delta R/R\) can be calculated from the gauge factor (\(GF\)) of the material used to make the piezoresistive elements: \(\Delta R/R = GF \varepsilon_{ind}\), where \(\varepsilon_{ind}\) is the strain induced into the piezoresistor. In most of the conventional strain gauges as shown in Fig. [\[fig:fig1\]](#fig:fig1){reference-type="ref" reference="fig:fig1"} (a,b), the thickness of the sensing layer is typically below a few hundred nanometers, which is much smaller than that of the substrate. Therefore, the strain induced into the piezoresistive elements is approximately the same as that of the substrate (\(\varepsilon_{ind} \approx \varepsilon_{sub}\)). Consequently, to improve the sensitivity of strain sensors (e.g. enlarging \(\Delta R/R\)), electrical approaches which can enlarge the gauge factor (\(GF\)) are required. Nevertheless, as aforementioned, the existence of the large gauge factor in nanowires due to quantum confinement or surface state, is still considered as controversial. It is also evident from Eq. [\[eq:sensitivity\]](#eq:sensitivity){reference-type="ref" reference="eq:sensitivity"} that the sensitivity of strain sensors can also be improved using a mechanical approach, which enlarges the strain induced into the piezoresistive element. Figure [\[fig:fig1\]](#fig:fig1){reference-type="ref" reference="fig:fig1"}(c) shows our proposed nano strain-amplifier structure, in which the piezoresistive nanowires are locally fabricated at the centre of a released bridge. The key idea of this structure is that, under a certain strain applied to the substrate, a large strain will be concentrated at the locally fabricated SiC nanowires. The working principle of the nano strain-amplifier is similar to that of the well-known dogbone structure, which is widely used to characterize the tensile strength of materials. That is, when a stress is applied to the dogbone-shape of a certain material, a crack, if generated, will occur at the middle part of the dogbone. The large strain concentrated at the narrow area located at the centre part with respect to the wider areas located at outer region, causes the crack. Qualitative and quantitative explanations of the nano strain-amplifier are presented as follows. For the sake of simplicity, the released micro frame and nanowire (single wire or array) of the nano strain-amplifier can be considered as solid springs, Fig. [\[fig:fig1\]](#fig:fig1){reference-type="ref" reference="fig:fig1"}(d). The stiffness of these springs are proportional to their width (\(w\)) and inversely proportional to their length (l): \(K \propto w/l\). Consequently, the model of the released nanowire and micro frames can be simplified as a series of springs, where the springs with higher stiffness correspond to the micro frame, and the single spring with lower stiffness corresponds to the nanowire. It is well-known in classical physics that, for serially connected springs, a larger strain will be concentrated in the low--stiffness string, while a smaller strain will be induced in the high--stiffness string. The following analysis quantitatively explained the amplification of the strain. When a tensile mechanical strain (\(\varepsilon_{sub}\)) is applied to the substrate, the released structure will also be elongated. Since the stiffness of the released frame is much smaller than that of the substrate, it is safe to assume that the released structure will follows the elongation of the substrate. The displacement of the released structure \(\Delta L\) is: \[\Delta L = \Delta L_m + \Delta L_n = L_m \varepsilon_m + L_n \varepsilon_n \label{eq:displacement}\] where \(L_m\), \(L_n\) are the length; \(\Delta L_m\), \(\Delta L_n\) are the displacement; and \(\varepsilon_m\), \(\varepsilon_n\) are the strains induced into the micro spring and nano spring, respectively. The subscripts m and n stand for the micro frames and nanowires, respectively. Furthermore, due to the equilibrium of the stressing force (\(F\)) along the series of springs, the following relationship is established: \(F= K_m\Delta L_m = K_n \Delta L_n\), where \(K_m\), \(K_n\) are the stiffness of the released micro frames and nanowires, respectively. Consequently the relationship between the displacement of the micro frame (higher stiffness) and nanowires (lower stiffness) is: \[\frac{\Delta L_m}{\Delta L_n}=\frac{K_n}{K_m}=\frac{L_mw_n}{L_nw_m} \label{eq:euili}\] Substituting Eqn. [\[eq:euili\]](#eq:euili){reference-type="ref" reference="eq:euili"} into Eqn. [\[eq:displacement\]](#eq:displacement){reference-type="ref" reference="eq:displacement"}, the strain induced into the locally fabricated nanowires is: \[\varepsilon_n = \frac{\Delta L_n}{L_n} = \frac{1}{1-\frac{w_m-w_n}{w_m}\frac{L_m}{L}}\varepsilon_{sub} \label{eq:strainamp}\] Equation [\[eq:strainamp\]](#eq:strainamp){reference-type="ref" reference="eq:strainamp"} indicates that increasing the ratio of \(w_m/w_n\) and \(L_m/L_n\) significantly amplifies the strain induced into the nanowire from the strain applied to the substrate. This model is also applicable to the case of nanowire arrays, in which \(w_n\) is the total width of all nanowires in the array. The theoretical model is then verified using the finite element analysis (FEA). In the FEA simulation, we compare the strain induced into (i) non released nanowires, (ii) the conventionally released nanowires, and (iii) our nano strain-amplifier structure, using COMSOL Multiphysics . In our nano strain amplifying structure, the width of the released frame was set to be 8 \(\mu\)m, while the width of each nanowire in the array (3 wires) was set to be 370 nm. The nanowires array structure was selected as it can enhance the electrical conductance of the SiC nanowires resistor which makes the subsequent experimental demonstration easier. The ratio between the length of nanowires and micro bridge was set to be 1: 20. With this geometrical dimensions, strain induced into nanowires array \(\varepsilon_n\) was numerically calculated to be approximately 6 times larger than \(\varepsilon_{sub}\), Eqn. [\[eq:strainamp\]](#eq:strainamp){reference-type="ref" reference="eq:strainamp"}. The simulation results show that for all structure, the elongation of non-released and released nanowires follow that of the substrate. In addition, strain was almost completely transferred into conventional released and non-released structures. Furthermore, the ratio of the strain induced in to the locally fabricated nanowires was estimated to be 5.9 times larger than that of the substrate, Fig. [\[fig:fig2\]](#fig:fig2){reference-type="ref" reference="fig:fig2"}. These results are in solid agreement with the theoretical analysis presented above. For a nanowire array with an average width of 470 nm, the amplified gain of strain was found to be 4.5. Based on the theoretical analysis, we conducted the following experiments to demonstrate the high sensitivity of SiC nanowire strain sensors using the nano strain-amplifier. A thin 3C-SiC film with its thickness of 300 nm was epitaxially grown on a 150 mm diameter Si wafer using low pressure chemical vapour deposition. The film was *in situ* doped using Al dopants. The carrier concentration of the p-type 3C-SiC was found to be \(5 \times 10^{18}\) cm\(^{-3}\), using a hot probe technique. The details of the characteristics of the grown film can be found elsewhere. Subsequently, I-shape p-type SiC resistors with aluminum electrodes deposited on the surface were patterned using inductive coupled plasma (ICP) etching. As the piezoresistance of p-type 3C-SiC depends on crystallographic orientation, all SiC resistors of the present work were aligned along \[110\] direction to maximize the piezoresistive effect. Next, the micro scale SiC resistors were then released from the Si substrate using dry etching (XeF\(_2\)). Finally, SiC nanowire arrays were formed at the centre of the released bridge using focused ion beam (FIB). Two types of nanowire array were fabricated with three nanowires for each array. The average width of each nanowire in each type were 380 nm and 470 nm, respectively. Figure [\[fig:fig3\]](#fig:fig3){reference-type="ref" reference="fig:fig3"} shows the SEM images of the fabricated samples, including the conventional released structure, non-released nanowires, and the nano strain-amplifier. The current voltage (I-V) curves of all fabricated samples were characterized using a HP 4145  parameter analyzer. The linear relationship between the applied voltage and measured current, indicated that Al made a good Ohmic contact with the highly doped SiC resistance, Fig. [\[fig:IV\]](#fig:IV){reference-type="ref" reference="fig:IV"}. Additionally, the electrical conductivity of both nanowires and micro frame estimated from the I-V curve and the dimensions of the resistors shows almost the same value. This indicated that the FIB process did not cause a significant surface damage to the fabricated nanowires. The bending experiment was used to characterize the piezoresistive effect in micro size SiC resistors and locally fabricated SiC nanowire array. In this experiment one end of the Si cantilever (with a thickness of 625 \(\mu\)m, and a width of 7 mm) was fixed while the other end was deflected by applying different forces. The distance from the fabricated nanowires to the free end of the Si cantilever was approximately 45 mm. The strain induced into the Si substrate is \(\varepsilon_\text{sub} = Mt/2EI\), where \(M\) is the applied bending moment; and \(t\), \(E\) and \(I\) are the thickness, Young's modulus and the moment of inertia of the Si cantilever, respectively. The response of the SiC resistance to applied strain was then measured using a multimeter (Agilent A). The relative resistance change (\(\Delta R/R\)) of the micro and nano SiC resistors was plotted against the strain induced into the Si substrate \(\varepsilon_{sub}\), Fig. [\[fig:DRR\]](#fig:DRR){reference-type="ref" reference="fig:DRR"}(a). For all fabricated samples, the relative resistance change shows a good linear relationship with the applied strain (\(\varepsilon_{sub}\)). In addition, with the same applied strain to the Si substrate, the resistance change of the SiC nanowires using the nano strain-amplifier was much larger than that of the the SiC micro resistor and the conventional non-released SiC nanowires. In addition, reducing the width of the SiC nanowires also resulted in the increase of the sensitivity. The magnitude of the piezoresistive effect in the nano strain-amplifier as well as conventional structures were then quantitatively evaluated based on the effective gauge factor (\(GF_{eff}\)), which is defined as the ratio of the relative resistance change to the applied strain to the substrate: \(GF_{eff} = (\Delta R/R)/\varepsilon_{sub}\). Accordingly, the effective gauge factor of the released micro SiC was found to be 28, while that of the non-released SiC nanowires was 35. From the data shown in Fig. [\[fig:DRR\]](#fig:DRR){reference-type="ref" reference="fig:DRR"}, the effective gauge factor of the 380 nm and 470 nm SiC nanowires in the nano strain-amplifier were calculated as 150 and 124, respectively. Thus for nanowire arrays with average widths of 380 nm and 470 nm, the sensitivity of the nano strain-amplifier was 5.4 times and 4.6 times larger than the bulk SiC, respectively. These results were consistent with analytical and numerical models presented above. The relative resistance change of the nano strain-amplifier also showed excellent linearity with the applied strain, with a linear regression of above 99%. The resistance change of the nano strain-amplifier can also be converted into voltage signals using a Wheatstone bridge, Fig. [\[fig:DRR\]](#fig:DRR){reference-type="ref" reference="fig:DRR"}(b). The output voltage of the nano strain-amplifier increases with increasing tensile strains from 0 ppm to 180 ppm, and returned to the initial value when the strain was completely removed, confirming a good repeatability after several strain induced cycles. The linearity of the relative resistance change, and the repeatability indicate that the proposed structure is promising for strain sensing applications. In conclusion, this work presents a novel mechanical approach to obtain highly sensitive piezoresistance in nanowires based on a nano strain-amplifier. The key factor of the nano strain-amplifier lies on nanowires locally fabricated on a released micro structure. Experimental studies were conducted on SiC nanowires, confirming that by utilizing our nano strain-amplifier, the sensitivity of SiC nanowires was 5.4 times larger than that of conventional structures. This result indicated that the nano strain-amplifier is an excellent platform for ultra sensitive strain sensing applications.
{'timestamp': '2016-07-18T02:07:38', 'yymm': '1607', 'arxiv_id': '1607.04531', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04531'}
# Introduction {#intro} Gas has a fundamental role in shaping the evolution of galaxies, through its accretion on to massive haloes, cooling and subsequent fuelling of star formation, to the triggering of extreme luminous activity around super massive black holes. Determining how the physical state of gas in galaxies changes as a function of redshift is therefore crucial to understanding how these processes evolve over cosmological time. The standard model of the gaseous interstellar medium (ISM) in galaxies comprises a thermally bistable medium (@Field:1969) of dense (\(n \sim 100\) cm\(^{-3}\)) cold neutral medium (CNM) structures, with kinetic temperatures of \(T_{\rm k} \sim 100\) K, embedded within a lower-density (\(n \sim 1\) cm\(^{-3}\)) warm neutral medium (WNM) with \(T_{\rm k} \sim 10^{4}\) K. The WNM shields the cold gas and is in turn ionized by background cosmic rays and soft X-rays (e.g. @Wolfire:1995 [@Wolfire:2003]). A further hot (\(T_{\rm k} \sim 10^{6}\) K) ionized component was introduced into the model by, to account for heating by supernova-driven shocks within the inter-cloud medium. In the local Universe, this paradigm has successfully withstood decades of observational scrutiny, although there is some evidence (e.g. @Heiles:2003b; @Roy:2013b; @Murray:2015) that a significant fraction of the WNM may exist at temperatures lower than expected for global conditions of stability, requiring additional dynamical processes to maintain local thermodynamic equilibrium. Since atomic hydrogen (H i) is one of the most abundant components of the neutral ISM and readily detectable through either the 21 cm or Lyman \(\alpha\) lines, it is often used as a tracer of the large-scale distribution and physical state of neutral gas in galaxies. The 21 cm line has successfully been employed in surveying the neutral ISM in the Milky Way (e.g. @McClure-Griffiths:2009 [@Murray:2015]), the Local Group (e.g. @Kim:2003 [@Bruns:2005; @Braun:2009; @Gratier:2010]) and low-redshift Universe (see @Giovanelli:2016 for a review). However, beyond \(z \sim 0.4\) (@Fernandez:2016) H i emission from individual galaxies becomes too faint to be detectable by current 21 cm surveys and so we must rely on absorption against suitably bright background radio (21 cm) or UV (Lyman-\(\alpha\)) continuum sources to probe the cosmological evolution of H i. The bulk of neutral gas is contained in high-column-density damped Lyman-\(\alpha\) absorbers (DLAs, \(N_{\rm HI} \geq 2 \times 10^{20}\) cm\(^{-2}\); see @Wolfe:2005 for a review), which at \(z \gtrsim 1.7\) are detectable in the optical spectra of quasars. Studies of DLAs provide evidence that the atomic gas in the distant Universe appears to be consistent with a multi-phase neutral ISM similar to that seen in the Local Group (e.g. @Lane:2000; @Kanekar:2001c; @Wolfe:2003b). However, there is some variation in the cold and warm fractions measured throughout the DLA population (e.g. @Howk:2005; @Srianand:2005 [@Lehner:2008]; @Jorgenson:2010; @Carswell:2011 [@Carswell:2012; @Kanekar:2014a]; @Cooke:2015; @Neeleman:2015). The 21-cm spin temperature affords us an important line-of-enquiry in unraveling the physical state of high-redshift atomic gas. This quantity is sensitive to the processes that excite the ground-state of H i in the ISM (@Purcell:1956 [@Field:1958; @Field:1959b; @Bahcall:1969]) and therefore dictates the detectability of the 21 cm line in absorption. In the CNM the spin temperature is governed by collisional excitation and so is driven to the kinetic temperature, while the lower densities in the WNM mean that the 21 cm transition is not thermalized by collisions between the hydrogen atoms, and so photo-excitation by the background Ly \(\alpha\) radiation field becomes important. Consequently the spin temperature in the WNM is lower than the kinetic temperature, in the range \(\sim\)`<!-- -->`{=html}1000--5000 K depending on the column density and number of multi-phase components (@Liszt:2001). Importantly, the spin temperature measured from a single detection of extragalactic absorption is equal to the harmonic mean of the spin temperature in individual gas components, weighted by their column densities, thereby providing a method of inferring the CNM fraction in high-redshift systems. Surveys for 21 cm absorption in known redshifted DLAs have been used to simultaneously measure the column density and spin temperature of H i (see @Kanekar:2014a and references therein). There is some evidence for an increase (at \(4\,\sigma\) significance) in the spin temperature of DLAs at redshifts above \(z = 2.4\), and a difference (at \(6\,\sigma\) significance) between the distribution of spin temperatures in DLAs and the Milky Way (@Kanekar:2014a). The implication that at least 90 per cent of high-redshift DLAs may have CNM fractions significantly less than that measured for the Milky Way has important consequences for the heating and cooling of neutral gas in the early Universe and star formation (e.g. @Wolfe:2003a). However, these targeted observations rely on the limited availability of simultaneous 21 cm and optical/UV data for the DLAs and assumes commonality between the column density probed by the optical and radio sight-lines. The first issue can be overcome by improving the sample statistics through larger 21 cm line surveys of high-redshift DLAs, but the latter requires improvements to our methodology and understanding of the gas distribution in these systems. There are also concerns about the accuracy to which the fraction of the source structure subtended by the absorber can be measured in each system, which can only be resolved through spectroscopic very long baseline interferometry (VLBI). It has been suggested that the observed evolution in spin temperature could be biased by assumptions about the radio-source covering factor (@Curran:2005) and its behaviour as a function of redshift (@Curran:2006b [@Curran:2012b]). In this paper we consider an approach using the statistical constraint on the average spin temperature achievable with future large 21 cm surveys using precursor telescopes to the Square Kilometre Array (SKA). This will enable independent verification of the evolution in spin temperature at high redshift and provide a method of studying the global properties of neutral gas below \(z \approx 1.7\), where the Lyman \(\alpha\) line is inaccessible using ground-based observatories. In an early attempt at a genuinely blind 21 cm absorption survey, used pilot data from the Arecibo Legacy Fast Arecibo L-band Feed Array (ALFALFA) survey to obtain upper limits on the column density frequency distribution from 21 cm absorption at low redshift (\(z \lesssim 0.06\)). However, they also noted that the number of detections could be used to make inferences about the ratio of the spin temperature to covering factor. Building upon this work, found that their upper limits on the frequency distribution function measured from the 40 per cent ALFALFA survey (\(\alpha\).40; @Haynes:2011) could only be reconciled with measurements from other low-redshift 21 cm surveys if the typical spin temperature to covering factor ratio was greater than 500 K. At higher redshifts, found that the number density of 21 cm absorbers in known Mg ii absorbers appeared to decrease with redshift above \(z \sim 1\), consistent with a reduction in the CNM fraction. We pursue this idea further by investigating whether future wide-field 21 cm surveys can be used to measure the average spin temperature in distant galaxies that are rich in atomic gas. # The expected number of intervening H i absorbers {#section:expected_number} We estimate the expected number of intervening H i systems towards a sample of background radio sources by evaluating the following integral over all sight-lines \[\label{equation:expected_number} \mu = \iint{f(N_{\rm HI},X)\,\mathrm{d}X\,\mathrm{d}N_{\rm HI}},\] where \(f(N_{\rm HI}, X)\) is the frequency distribution as a function of column density (\(N_{\rm HI}\)) and comoving path length (\(X\)). We use the results of recent surveys for 21 cm emission in nearby galaxies (e.g. @Zwaan:2005) and high-redshift Lyman-\(\alpha\) absorption in the Sloan Digitial Sky Survey (SDSS; e.g. @Prochaska:2005; @Noterdaeme:2009), which show that \(f(N_{\rm HI}, X)\) can be parametrized by a gamma function of the form \[f(N_{\rm HI}, X) = \left({f_{\ast} \over N_{\ast}}\right)\left({N_{\rm HI} \over N_{\ast}}\right)^{-\beta}\exp{\left(-{N_{\rm HI} \over N_{\ast}}\right)}\,\mathrm{cm}^{2},\] where \(f_{\ast} = 0.0193\), \(\log_{10}(N_{\ast}) = 21.2\) and \(\beta = 1.24\) at \(z = 0\) (@Zwaan:2005), and \(f_{\ast} = 0.0324\), \(\log_{10}(N_{\ast}) = 21.26\) and \(\beta = 1.27\) at \(z \approx 3\) (@Noterdaeme:2009). While the observational data do not yet constrain models for evolution of the H i distribution at intermediate redshifts between \(z \sim 0.1\) and \(3\)[^1], it is known to be much weaker than the significant decline seen in the global star-formation rate and molecular gas over the same epoch (e.g. @Lagos:2014). We therefore carry out a simple linear interpolation between the low and high redshift epochs to estimate \(f(N_{\rm HI},X)\) as a function of redshift. The probability of detecting an absorbing system of given column density depends on the sensitivity of the survey, the flux density and structure of the background source and the fraction of H i in the lower spin state, given by the spin temperature. We express the column density (\(N_{\rm HI}\); in atoms cm\(^{-2}\)) in terms of the optical depth (\(\tau\)) and spin temperature (\(T_{\rm spin}\); in K) by \[\label{equation:column_density} N_{\rm HI} = 1.823\times10^{18}\,T_{\rm spin} \int{\tau(v)\mathrm{d}v},\] where the integral is performed across the spectral line in the system rest-frame velocity \(v\) (in km s\(^{-1}\)). We then express the optical depth in terms of the observables as \[\tau =-\ln\left[1 + {\Delta{S}\over c_{\rm f}S_{\rm cont}}\right],\] where \(\Delta{S}\) is the observed change in flux density due to absorption, \(S_{\rm cont}\) is the background continuum flux density and \(c_{\rm f}\) is the (often unknown) fraction of background flux density subtended by the intervening gas. We assume that a single intervening system can be described by a Gaussian velocity distribution of full width at half maximum (FWHM) dispersion (\(\Delta{v_{\rm 50}}\)) and peak optical depth (\(\tau_{\rm peak}\)), so that can be re-written as \[\label{equation:column_density_gaussian} N_{\rm HI} = 1.941\times10^{18}\,T_{\rm spin}\,\tau_{\rm peak}\,\Delta{v_{\rm 50}}.\] If we further assume that the rms spectral noise is Gaussian, with a standard deviation \(\sigma_{\rm chan}\) per independent channel \(\Delta{v_{\rm chan}}\), then the 5\(\sigma\) column density detection limit is given by \[N_{5\sigma} \approx 1.941\times10^{18}\,T_{\rm spin}\,\tau_{\rm 5\sigma}\,\Delta{v_{\rm conv}},\] where \[\label{equation:optical_depth_limit} \tau_{5\sigma} \approx-\ln\left[1-{5\,\sigma_{\rm chan}\over c_{\rm f}\,S_{\rm cont}}\sqrt{\Delta{v}_{\rm chan}\over \Delta{v_{\rm conv}}}\right],\] and \(\Delta{v_{\rm conv}} \approx \sqrt{\Delta{v}_{\rm chan}^{2} + \Delta{v}_{50}^{2}}\), which is the observed width of the line, given by the convolution of the physical velocity distribution and the spectral resolution of the telescope. We now redefine \(\mu\) as the expected number of intervening H i detections in our survey as a function of the column density sensitivity along each sight-line where each comoving path element \(\delta{X}(z)\)[^2] in the integral defined by is given by \[\delta{X}(z)= \begin{cases} {\delta{z}\,(1+z)^{2}\over \sqrt{(1+z)^{2}(1+z\Omega_{\rm M})-z(z+2)\Omega_{\rm \Lambda}}}, & \text{if}\ N_{\rm HI} \geq N_{5\sigma}, \\ 0, & \text{otherwise}. \end{cases}\] To calculate the column density sensitivity for each comoving element we draw random samples for \(\Delta{v}_{50}\) and \(c_{\rm f}\) from continuous prior distributions based on existing evidence. In the case of \(\Delta{v}_{50}\) we use a log-normal distribution obtained from a simple least-squares fit to the sample distribution from previous 21-cm absorption surveys reported in the literature (see )[^3], assuming that this correctly describes the true distribution for the population of DLAs. However, direct measurement of the H i covering factor is significantly more difficult and so for the purposes of this work we draw random samples assuming a uniform distribution between 0 and 1. In , we show a comparison between this assumption and the sample distribution estimated by from their main sample of 37 quasars. Kanekar et al. used VLBI synthesis imaging to measure the fraction of total quasar flux density contained within the core, which was then used as a proxy for the covering factor. By carrying out a two-tailed Kolmogorov-Smirnov (KS) test of the hypothesis that the Kanekar et al. data are consistent with our assumed uniform distribution, we find that this hypothesis is rejected at the 0.05 level, but not at the 0.01 level (this outcome is dominated by the paucity of quasars in the sample with \(c_{\rm f} \lesssim 0.2\)). It is therefore possible that the population distribution of H i covering factors may deviate somewhat from the uniform distribution assumed in this work. We discuss the implications of this further in . # A 21 cm absorption survey with ASKAP {#section:all_sky_survey} We use the Australian Square Kilometre Array Pathfinder (ASKAP; @Johnston:2007) as a case study to demonstrate the expected results from planned wide-field surveys for 21 cm absorption (e.g. the ASKAP First Large Absorption Survey in H i--Sadler et al., the MeerKAT Absorption Line Survey--Gupta et al., and the Search for HI absorption with AperTIF--Morganti et al.). ASKAP is currently undergoing commissioning. Proof-of-concept observations with the Boolardy Engineering Test Array (@Hotan:2014) have already been used to successfully detect a new H i absorber associated with a probable young radio galaxy at \(z = 0.44\) (@Allison:2015a). Here we predict the outcome of a future 2 h-per-pointing survey of the entire southern sky (\(\delta \leq +10\degr\)) using the full 36-antenna ASKAP in a single 304 MHz band between 711.5 and 1015.5 MHz, equivalent to H i redshifts between \(z = 0.4\) and 1.0. Our expectations of the ASKAP performance are based on preliminary measurements by using the prototype Mark II phase array feed. We estimate the noise per spectral channel using the radiometer equation \[\sigma_{\rm chan} = {S_{\rm system} \over \sqrt{n_{\rm pol}\,n_{\rm ant}\,(n_{\rm ant}-1)\,\Delta{t}_{\rm in}\,\Delta{\nu}_{\rm chan}}},\] where \(S_{\rm system}\) is the system equivalent flux density, \(n_{\rm pol}\) is the number of polarizations, \(n_{\rm ant}\) is the number of antennas, \(\Delta{t}_{\rm in}\) is the on-source integration time and \(\Delta{\nu}_{\rm chan}\) is the spectral resolution in frequency. The sensitivity of the telescope in the 711.5-1015.5 MHz band is expected to vary between \(S_{\rm system} \approx 3200\) and \(2000\) Jy, with the largest change in sensitivity between 700 and 800 MHz. ASKAP has dual linear polarization feeds, 36 antennas and a fine filter bank that produces 16 416 independent channels across the full 304 MHz bandwidth, so the expected noise per 18.5 kHz channel in a 2 h observation is approximately 5.5-3.5 mJy beam\(^{-1}\) across the band. In the case of an actual survey, the true sensitivity will of course be recorded in the spectral data as a function of redshift (see e.g. @Allison:2015a), but for the purposes of the simulated survey presented in this work we split the band into several frequency bins to capture the variation in sensitivity and velocity resolution (which is in the range 7.8 km s\(^{-1}\) at 711.5 MHz to 5.5 km s\(^{-1}\) at 1015.5 MHz). In order to simulate a realistic survey of the southern sky we select all radio sources south of \(\delta = +10\degr\) from catalogues of the National Radio Astronomy Observatory Very Large Array Sky Survey (NVSS, \(\nu = 1.4\) GHz, \(S_{\rm src} \gtrsim 2.5\) mJy; @Condon:1998), the Sydney University Molonglo SkySurvey (\(\nu = 843\) MHz, \(S_{\rm src} \gtrsim 10\) mJy; @Mauch:2003) and the second epoch Molonglo Galactic Plane Survey (\(\nu = 843\) MHz, \(S_{\rm src} \gtrsim 10\) mJy; @Murphy:2007). The source flux densities, used to calculate the optical depth limit in , are estimated at the centre of each frequency bin by extrapolating from the catalogue values and assuming a canonical spectral index of \(\alpha =-0.7\). In , we show the resulting cumulative distribution of radio sources in our sample as a function of flux density across the band. For any given sight-line, the redshift interval over which absorption may be detected is dependent upon the distance to the continuum source. The lack of accurate spectroscopic redshift measurements for most radio sources over the sky necessitates the use of a statistical approach based on a model for the source redshift distribution. We therefore apply a statistical weighting to each comoving path element \(\delta{X}(z)\) such that the expected number of absorber detections is now given by \[\label{equation:weighted_sum_number} \mu = \iint{f(N_{\rm HI},X)\,\mathcal{F}_{\rm src}(z^{\prime} \geq z)\,\mathrm{d}X\,\mathrm{d}N_{\rm HI}},\] where \[\mathcal{F}_{\rm src}(z^{\prime} \geq z) = {\int_{z}^{\infty} \mathcal{N}_{\rm src}(z^{\prime})\mathrm{d}z^{\prime}\over\int_{0}^{\infty}\mathcal{N}_{\rm src} (z^{\prime})\mathrm{d}z^{\prime}},\] and \(\mathcal{N}_{\rm src}(z)\) is the number of radio sources as a function of redshift. To estimate \(\mathcal{N}_{\rm src}(z)\) we use the Combined EIS-NVSS Survey Of Radio Sources (CENSORS; @Brookes:2008), which forms a complete sample of radio sources brighter than 7.2 mJy at 1.4 GHz with spectroscopic redshifts out to cosmological distances. In we show the distribution of CENSORS sources brighter than 10 mJy beyond a given redshift \(z\), and the corresponding analytical function derived from the model fit of, given by \[\mathcal{N}_{\rm src}(z) \approx 1.29 + 32.37z-32.89z^{2} + 11.13z^{3}-1.25z^{4},\] which we use in our analysis. For the redshifts spanned by our simulated ASKAP survey, the fraction of background sources evolves from 87 per cent at \(z = 0.4\) to 53 per cent at \(z = 1.0\). We assume that this redshift distribution applies to any sight-line irrespective of the continuum flux density. However, this assumption is only true if the source population in the target sample evolves such that the effect of distance is nullified by an increase in luminosity. Given this criterion, and the sensitivity of our simulated survey, we limit our sample to sources with flux densities between 10 and 1000 mJy, which are dominated by the rapidly evolving population of high-excitation radio galaxies and quasars (e.g. @Jackson:1999 [@Best:2012; @Best:2014; @Pracy:2016]) and for which the redshift distribution is known to be almost independent of flux density (e.g. @Condon:1984 [@Condon:1998]). In , we show the number of sources from this sub-sample as a function of opacity sensitivity \[as defined by \], drawing random samples of the line FWHM and covering factor from the distributions shown in and . There are approximately 190 000 sightlines with sufficient sensitivity to detect absorption of optical depth greater than \(\tau_{5\sigma} \approx 1.0\) and 25 000 sensitive to optical depths greater than \(\tau_{5\sigma} \approx 0.1\). Since this distribution converges at optical depth sensitivities greater than \(\tau_{5\sigma} \approx 5\), the population of sources fainter than 10 mJy, which are excluded from our simulated ASKAP survey, would not significantly contribute to further detections of absorption. Similarly, while sources brighter than 1 Jy are good probes of low-column-density H i gas, they do not constitute a sufficiently large enough population to significantly affect the total number of absorber detections expected in the survey and can also be safely excluded. Based on these assumptions, we can estimate the number of absorbers we would expect to detect in our survey with ASKAP as a function of spin temperature. In , we show the expected detection yield as a cumulative function of column density. We show results for two scenarios where the spin temperature is fixed at a single value of either 100 or 1000 K, and the line FWHM and covering factors are drawn from the random distributions shown in and . We find that for both these cases the expected number of detections is not sensitive to column densities below the DLA definition of \(N_{\rm HI} = 2\times 10^{20}\) cm\(^{-2}\). We also show in the expected total detection yield (integrated over all H i column densities) as a function of a single spin temperature \(T_\mathrm{spin}\) and line FWHM \(\Delta{v}_\mathrm{50}\). We find that for typical spin temperatures of a few hundred kelvin (consistent with the typical fraction of CNM observed in the local Universe) and a line FWHM of approximately \(20\) km s\(^{-1}\), a wide-field 21 cm survey with ASKAP is expected to yield \(\sim 1000\) detections. However, even moderate evolution to a higher spin temperature in the DLA population should see significant reduction in the detection yield from this survey. # Inferring the average spin temperature {#section:spin_temp} We cannot directly measure the spin temperatures of individual systems without additional data from either 21 cm emission or Lyman-\(\alpha\) absorption. However, from it is evident that the total number of absorbing systems expected to be detected with a reasonably large 21 cm survey is strongly dependent on the assumed value for the spin temperature. Therefore, by comparing the actual survey yield with that expected from the known H i distribution, we can infer the average spin temperature of the atomic gas within the DLA population for a given redshift interval. Assuming that the total number of detections follows a Poisson distribution, the probability of detecting \(\mathcal{N}\) intervening absorbing systems is given by \[p(\mathcal{N}|\overline{\mu}) = {\overline{\mu}^{\mathcal{N}} \over \mathcal{N}!} \mathrm{e}^{-\overline{\mu}},\] where \(\overline{\mu}\) is the expected total number of detections given by the integral \[\overline{\mu} = \iiint{\mu(T_{\rm spin},\Delta{v}_{50},c_{\rm f})\rho(T_{\rm spin},\Delta{v}_{50},c_{\rm f})\mathrm{d}T_{\rm spin}\mathrm{d}\Delta{v}_{50}\mathrm{d}c_{\rm f}},\] and \(\rho\) is the distribution of systems as a function of spin temperature, line FWHM and covering factor. We assume that all three of these variables are independent[^4] so that \(\rho\) factorizes into functions of each. We then marginalise over the covering factor and line width distributions shown in and so that the expression for \(\overline{\mu}\) reduces to \[\label{equation:harmonic_spin_temp} \overline{\mu} = \int{\mu(T_{\rm spin})\rho(T_{\rm spin})\mathrm{d}T_{\rm spin}} = \mu(\overline{T}_{\rm spin}),\] where \(\overline{T}_{\rm spin}\) is the harmonic mean of the unknown spin temperature distribution, weighted by column density. This is analogous to the spin temperature inferred from the detection of absorption in a single intervening galaxy averaged over several gaseous components at different temperatures (e.g. @Carilli:1996). In the event of the survey yielding \(\mathcal{N}\) detections, we can calculate the posterior probability density of \(\overline{T}_{\rm spin}\) using the following relationship between conditional probabilities \[p(\overline{T}_{\rm spin}|\mathcal{N}) = {p(\mathcal{N}|\overline{T}_{\rm spin})p(\overline{T}_{\rm spin})\over p(\mathcal{N})},\] where \(p(\overline{T}_{\rm spin})\) is our prior probability density for \(\overline{T}_{\rm spin}\) and \(p(\mathcal{N})\) is the marginal probability of the number of detections, which can be treated as a normalizing constant. The minimally informative Jeffreys prior for the mean value \(\mu\) of a Poisson distribution is \(1/\sqrt{\mu}\) (@Jeffreys:1946)[^5]. From it therefore follows that a suitable form for the non-informative spin temperature prior is \(p(\overline{T}_{\rm spin}) = 1/\sqrt{\overline{\mu}}\), so that \[\label{equation:tspin_prob} p(\overline{T}_{\rm spin}|\mathcal{N}) = C^{-1}\,{\overline{\mu}^{(\mathcal{N}-1/2)} \over \mathcal{N}!} \mathrm{e}^{-\overline{\mu}},\] where the distribution is normalised to unit total probability by evaluating the integral \[C = \int{{\overline{\mu}^{(\mathcal{N}-1/2)} \over \mathcal{N}!} \mathrm{e}^{-\overline{\mu}}}\,\mathrm{d}\overline{T}_{\rm spin}.\] The probabilistic relationship given by and the expected detection yield derived in can be used as a frame-work for inferring the harmonic-mean spin temperature using the results of any homogeneous 21-cm survey. We have assumed that we can accurately distinguish intervening absorbing systems from those associated with the host galaxy of the radio source. However, any 21-cm survey will be accompanied by follow-up observations, at optical and sub-mm wavelengths, which will aid identification. Furthermore, future implementation of probabilistic techniques to either use photometric redshift information or distinguish between line profiles should provide further disambiguation. Of course we have not yet accounted for any error in our estimate of \(\overline{\mu}\), which will increase our uncertainty in \(\overline{T}_\mathrm{spin}\). In the following section we discuss these possible sources of error and their effect on the result. # Sources of error {#section:errors} Our estimate of the expected number of 21 cm absorbers is dependent upon several distributions describing the properties of the foreground absorbing gas and the background source distribution. For future large-scale 21 cm surveys, the accuracy to which we can infer the harmonic mean of the spin temperature distribution will eventually be limited by the accuracy to which we can measure these other distributions. In this section, we describe these errors and their propagation through to the estimate of \(\overline{T}_\mathrm{spin}\), summarizing our results in . ## The covering factor {#section:covering_factor} ### Deviation from a uniform distribution between 0 and 1 The fraction \(c_{\rm f}\) by which the foreground gas subtends the background radiation source is difficult to measure directly and is thereby a significant source of error for 21 cm absorption surveys. In this work, we have assumed a uniform distribution for \(c_{\rm f}\), taking random values between 0 and 1. In , we tested this assumption by comparing it with the distribution of flux density core fractions in a sample of 37 quasars, used by as a proxy for the covering factor. By carrying out a two-tailed KS test, we found some evidence (at the 0.05 level) that this quasar sample was inconsistent with our assumption of a uniform distribution between 0 and 1. Noticeably there seems to be an under-representation of quasars in the Kanekar et al. sample with estimated \(c_{\rm f} \lesssim 0.2\). In the low optical depth limit, the detection rate is dependent on the ratio of spin temperature to covering factor, in which case a fractional deviation in \(c_{\rm f}\) will propagate as an equal fractional deviation in \(\overline{T}_\mathrm{spin}\). Based on the difference seen in the covering factor distribution of the Kanekar et al. sample and the uniform distribution, we assume that the spin temperature can deviate by as much as \(\pm\)`<!-- -->`{=html}10 per cent. ### Evolution with redshift We also consider that the covering factor distribution may evolve with redshift, which would mimic a perceived evolution in the average spin temperature. Such an effect was proposed by and, who claimed that the relative change in angular-scale behaviour of absorbers and radio sources between low-and high-redshift samples could explain the apparent evolution of the spin temperature found by. To test for this effect in their larger DLA sample, considered a sub-sample at redshifts greater than \(z = 1\), for which the relative evolution of the absorber and source angular sizes should be minimal. While the significance of their result was reduced by removing the lower redshift absorbers from their sample, they still found a difference at \(3.5\,\sigma\) significance between spin temperature distributions in the two DLA sub-samples separated by a median redshift of \(z = 2.683\). Future surveys with ASKAP and the other SKA pathfinders will search for H i absorption at intermediate redshifts (\(z \sim 1\)), where the relative evolution of the absorber and source angular sizes is expected to be more significant than for the higher redshift DLA sample considered by. We therefore consider the potential effect of this cosmological evolution on the inferred value of \(\overline{T}_\mathrm{spin}\). We approximate the covering factor using the following model of \[\label{equation:covering_factor} c_{f} \approx \begin{cases} \left({\theta_{\rm abs}\over \theta_{\rm src}}\right)^{2}, & \text{if}\ \theta_{\rm abs} < \theta_{\rm src}, \\ 1, & \text{otherwise}, \end{cases}\] where \(\theta_{\rm abs}\) and \(\theta_{\rm src}\) are the angular sizes of the absorber and background source, respectively. Under the small-angle approximation \(\theta_{\rm abs} \approx {d_{\rm abs}/D_{\rm abs}}\) and \(\theta_{\rm abs} \approx {d_{\rm src}/D_{\rm src}}\), where \(d_{\rm abs}\) and \(D_{\rm abs}\) are the linear size and angular diameter distance of the absorber, and likewise \(d_{\rm src}\) and \(D_{\rm src}\) are the linear size and angular diameter distance of the background source. Assuming that the ratio \(d_{\rm abs}/d_{\rm src}\) is randomly distributed and independent of redshift, any evolution in the covering factor is therefore dominated by relative changes in the angular diameter distances. We calculate the expected angular diameter distance ratio at a redshift \(z\) by \[\left\langle{D_{\rm abs}\over D_{\rm src}}\right\rangle_{z} = D_{\rm abs}(z){\int_{z}^{\infty} \mathcal{N}_{\rm src}(z^{\prime})D_{\rm src}(z^{\prime})^{-1}\mathrm{d}z^{\prime}\over\int_{z}^{\infty}\mathcal{N}_{\rm src} (z^{\prime})\mathrm{d}z^{\prime}},\] which, for the source redshift distribution model given by, evolves from 0.7 at \(z = 0.4\) to 1.0 at \(z = 1.0\) (see ). We note that this is consistent with the behaviour measured by for the total sample of DLAs observed at 21 cm wavelengths. By applying this as a correction to the otherwise uniformly distributed covering factor (using ), we find that the inferred value of \(\overline{T}_{\rm spin}\) systematically increases by approximately 30 per cent. ## The \(\bmath{N_{\rm HI}}\) frequency distribution ### Uncertainty in the measurement of \(f(N_{\rm HI},X)\) We assume that \(f(N_{\rm HI}, X)\) is relatively well understood as a function of redshift by interpolating between model gamma functions fitted to the distributions at \(z = 0\) and \(3\). However, these distributions were measured from finite samples of galaxies, which of course have associated uncertainties that need to be considered. In the case of the data presented by and, both have typical measurement uncertainties in \(f(N_{\rm HI}, X)\) of approximately 10 per cent over the range of column densities for which our simulated ASKAP survey is sensitive (see ). This will propagate as a 10 per cent fractional error in the expected number of absorber detections, and contribute a similar percentage uncertainty in the inferred average spin temperature. ### Correcting for 21 cm self-absorption In the local Universe, showed that self-absorption from opaque H i clouds identified in high-resolution images of the Local Group galaxies M31, M33 and the Large Magellanic Cloud may necessitate a correction to the local atomic mass density of up to 30 per cent. Although it is not yet clear whether this small sample of Local Group galaxies is representative of the low-redshift population, it is useful to understand how this effect might propagate through to our average spin temperature measurement. We therefore replace the gamma-function parametrization of the local \(f(N_{\rm HI})\) given by with the non-parametric values given in table 2 of, and recalculate \(\overline{T}_{\rm spin}\). For an all-sky survey with the full 36-antenna ASKAP we find that \(\overline{T}_{\rm spin}\) increases by \(\sim\)`<!-- -->`{=html}30 for 100 detections and \(\sim\)`<!-- -->`{=html}10 per cent for 1000 detections. Note that the correction increases for low numbers of detections, which are dominated by the highest column density systems. ### Dust obscuration bias in optically-selected DLAs At higher redshifts, it is possible that the number density of optically-selected DLAs could be significantly underestimated as a result of dust obscuration of the background quasar (@Ostriker:1984). This would cause a reduction in the \(f(\mbox{H\,{\sc i}}, X)\) measured from optical surveys, thereby significantly underestimating the expected number of intervening 21 cm absorbers at high redshifts. The issue is further compounded by the expectation that the highest column density DLAs (\(N_{\rm HI} \gtrsim 10^{21}\) cm\(^{-2}\)), for which future wide-field 21 cm surveys are most sensitive (see ), may contain more dust than their less-dense counterparts. This conclusion was supported by early analyses of the existing quasar surveys at that time (e.g. @Fall:1993), which indicated that up to 70 per cent of quasars could be missing from optical surveys through the effect of dust obscuration, albeit with large uncertainties. However, subsequent optical and infrared observations of radio-selected quasars (e.g. @Ellison:2001; @Ellison:2005; @Jorgenson:2006), which are free of the potential selection biases associated with these optical surveys, found that the severity of this issue was substantially over-estimated and that there was minimal evidence in support of a correlation between the presence of DLAs and dust reddening. Furthermore, the H i column density frequency distribution measured by was found to be consistent with the optically-determined gamma-function parametrization of, with no evidence of DLA systems missing from the SDSS sample at a sensitivity of \(N_{\rm HI} \lesssim 5 \times 10^{21}\) cm\(^{-2}\). Although radio-selected surveys of quasars are free of the selection biases associated with optical surveys, they do typically suffer from smaller sample sizes and are therefore less sensitive to the rarer DLAs with the highest column densities. Another approach is to directly test whether optically-selected quasars with intervening DLAs, selected from the SDSS sample, are systematically more dust reddened than a control sample of non-DLA quasars. Comparisons in the literature are based on several different colour indicators, which include the spectral index (e.g. @Murphy:2004 [@Murphy:2016]), spectral stacking (e.g. @Frank:2010 [@Khare:2012]) and direct photometry (e.g. @Vladilo:2008; @Fukugita:2015). The current status of these efforts is summarized by, showing broad support for a missing DLA population at the level of \(\sim\)`<!-- -->`{=html}5 per cent but highlighting that tension still exists between different dust measurements. No substantial evidence has yet been found to support a correlation between the dust reddening and H i column density in these optically selected DLA surveys (e.g. @Vladilo:2008 [@Khare:2012; @Murphy:2016]). In an attempt to reconcile the differences and myriad biases associated with these techniques, carried out a statistically-robust meta-analysis of the available optical and radio data, using a Bayesian parameter estimation approach to model the dust as a function of column density and metallicity. They found that the expected fraction of DLAs missing from optical surveys is 7 per cent, with fewer than 28 per cent missing at 3 \(\sigma\) confidence. Based on this body of work we therefore assume that approximately 10 per cent of DLAs are missing from the SDSS sample of and consider the affect on our estimate of \(\overline{T}_{\rm spin}\). We further assume that there is no dependance on column density, an assumption which is supported by the aforementioned observational data for the range of column densities to which our 21 cm survey is sensitive. We find that increasing the high-redshift column density frequency distribution by 10 per cent introduces a systematic increase of approximately 3 per cent in the expected number of detections for the redshifts covered by our ASKAP surveys. We note that this error will increase significantly for 21 cm surveys at higher redshifts where the optically derived \(f(\mbox{H\,{\sc i}}, X)\) dominates the calculation of the expected detection rate. ## The radio source background As described in , we weight the comoving path-length for each sight-line by a statistical redshift distribution in order to account for evolution in the radio source background. We use the parametric model of, which is derived from fitting the measured redshifts of for CENSORS sources brighter than 10 mJy, and assume that this applies to all sources in the range 10-1000 mJy. In , we show the cumulative distribution of sources located behind a given redshift and the associated measurement uncertainty given by the errorbars. For the intermediate redshifts covered by the ASKAP survey, the fractional uncertainty in this distribution increases from \(\sigma_{\mathcal{F}_{\rm src}}/\mathcal{F}_{\rm src} \approx 3.5\) to 8 per cent between \(z = 0.4\) and 1.0, which propagates through to a similar fractional uncertainty in \(\overline{T}_{\rm spin}\). However, for higher redshifts this fractional uncertainty increases rapidly at \(z > 2\), to more than 50 per cent at \(z = 3\), reflecting the paucity of optical spectroscopic data for the high-redshift radio source population. Understanding how the radio source population is distributed at lower flux densities and at higher redshifts is therefore a concern for the future 21 cm absorption surveys undertaken with the SKA mid-and low-frequency telescopes (see @Kanekar:2004 and @Morganti:2015 for reviews). ::: # Expected results for future 21-cm absorption surveys {#section:tspin_results} In the top panel of we show the results of applying our method for inferring \(\overline{T}_{\rm spin}\) to the simulated all-southern-sky H i absorption survey with ASKAP described in . We account for the uncertainties in the expected detection rate \(\overline{\mu}\), discussed in , by using a Monte Carlo approach and marginalizing over many realizations. A yield of 1000 absorbers from such a survey would imply an average spin temperature of \(\overline{T}_\mathrm{spin} = 127^{+14}_{-14}\,(193^{+23}_{-23})\) K[^6], where values in parentheses denote the alternative posterior probability resulting from the systematic errors discussed in . This scenario would indicate that a large fraction of the atomic gas in DLAs at these intermediate redshifts is in the classical stable CNM phase. Conversely, a yield of only 100 detections would imply that \(\overline{T}_\mathrm{spin} = 679^{+64}_{-65}\,(1184^{+116}_{-120})\) K, indicating that less than 10 per cent of the atomic gas is in the CNM and that the bulk of the neutral gas in galaxies is significantly different at intermediate redshifts compared with the local Universe. We also consider the effect of reducing the sky area and array size, which is relevant for planned early science surveys with ASKAP and other SKA pathfinder telescopes. In the bottom panel of , we show the spin temperatures inferred when observing a random 1000 deg\(^{2}\) field for 12 h per pointing, between \(z_{\rm HI} = 0.4\) and \(1.0\), using a 12-antenna version of ASKAP. We find that detection yields of 30 and 3 from such a survey would give inferred spin temperatures of \(\overline{T}_{\rm spin} =134^{+23}_{-27}\,(209^{+40}_{-47})\) and \(848^{+270}_{-430}\,(1535^{+513}_{-837})\) K, respectively. The significant reduction in telescope sensitivity and sky-area, compensated by the increase in integration time per pointing planned for early-science, results in a factor of 30 decrease in the expected number of detections and therefore an increase in the sample variance and uncertainty in \(\overline{T}_{\rm spin}\). However, this result demonstrates that we expect to be able to distinguish between the limiting cases of CNM-rich or deficient DLA populations even during the early-science phases of the SKA pathfinders. For example 30 detections with the early ASKAP survey rules out an average spin temperature of 1000 K at high probability. # Conclusions We have demonstrated a statistical method for measuring the average spin temperature of the neutral ISM in distant galaxies, using the expected detection yields from future wide-field 21 cm absorption surveys. The spin temperature is a crucial property of the ISM that can be used to determine the fraction of the cold (\(T_{\rm k} \sim 100\) K) and dense (\(n \sim 100\) cm\(^{-2}\)) atomic gas that provides sites for the future formation of cold molecular gas clouds and star formation. Recent 21 cm surveys for H i absorption in Mg ii absorbers and DLAs towards distant quasars have yielded some evidence of an evolution in the average spin temperature that might reveal a decrease in the fraction of cold dense atomic gas at high redshift (e.g. @Gupta:2009 [@Kanekar:2014a]). By combining recent specifications for ASKAP, with available information for the population of background radio sources, we show that strong statistical constraints (approximately \(\pm10\) per cent) in the average spin temperature can be achieved by carrying out a shallow 2-h per pointing survey of the southern sky between redshifts of \(z = 0.4\) and \(1.0\). However, we find that the accuracy to which we can measure the average spin temperature is ultimately limited by the accuracy to which we can measure the distribution of the covering factor, the \(N_{\rm HI}\) frequency distribution function and the evolution of the radio source population as a function of redshift. By improving our understanding of these distributions we will be able to leverage the order-of-magnitude increases in sensitivity and redshift coverage of the future SKA telescope, allowing us to measure the evolution of the average spin temperature to much higher redshifts.
{'timestamp': '2016-08-09T02:04:29', 'yymm': '1607', 'arxiv_id': '1607.04828', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04828'}
# Introduction Given \(\rho>0\), we consider the problem \[\label{eq:main_prob_U} \begin{cases}-\Delta U + \lambda U = |U|^{p-1}U & \text{in }\Omega,\smallskip\\ \int_\Omega U^2\,dx = \rho, \quad U=0 & \text{on }\partial\Omega, \end{cases}\] where \(\Omega\subset{\mathbb{R}}^N\) is a Lipschitz, bounded domain, \(1<p<2^*-1\), \(\rho>0\) is a fixed parameter, and both \(U\in H^1_0(\Omega)\) and \(\lambda\in{\mathbb{R}}\) are unknown. More precisely, we investigate conditions on \(p\) and \(\rho\) (and also \(\Omega\)) for the solvability of the problem. The main interest in [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} relies on the investigation of standing wave solutions for the nonlinear Schrödinger equation \[i\frac{\partial \Phi}{\partial t}+\Delta \Phi+ |\Phi|^{p-1}\Phi=0,\qquad (t,x)\in {\mathbb{R}}\times \Omega\] with Dirichlet boundary conditions on \(\partial\Omega\). This equation appears in several different physical models, both in the case \(\Omega={\mathbb{R}}^N\), and on bounded domains. In particular, the latter case appears in nonlinear optics and in the theory of Bose-Einstein condensation, also as a limiting case of the equation on \({\mathbb{R}}^N\) with confining potential. When searching for solutions having the wave function \(\Phi\) factorized as \(\Phi(x,t)=e^{i\lambda t} U(x)\), one obtains that the real valued function \(U\) must solve \[\label{eq:NLS}-\Delta U + \lambda U = |U|^{p-1}U ,\qquad U\in H^1_0(\Omega),\] and two points of view are available. The first possibility is to assign the chemical potential \(\lambda\in{\mathbb{R}}\), and search for solutions of [\[eq:NLS\]](#eq:NLS){reference-type="eqref" reference="eq:NLS"} as critical points of the related action functional. The literature concerning this approach is huge and we do not even make an attempt to summarize it here. On the contrary, we focus on the second possibility, which consists in considering \(\lambda\) as part of the unknown and prescribing the mass (or charge) \(\|U\|_{L^2(\Omega)}^2\) as a natural additional condition. Up to our knowledge, the only previous paper dealing with this case, in bounded domains, is, which we describe below. The problem of searching for normalized solutions in \({\mathbb{R}}^N\), with non-homogeneous nonlinearities, is more investigated, even though the methods used there can not be easily extended to bounded domains, where dilations are not allowed. Very recently, also the case of partial confinement has been considered. Solutions of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} can be identified with critical points of the associated energy functional \[\mathcal{E}(U) = \frac12\int_\Omega|\nabla U|^2\,dx-\frac{1}{p+1} \int_\Omega|U|^{p+1}\,dx\] restricted to the mass constraint \[{\mathcal{M}}_\rho=\{U\in H_0^1(\Omega): \|U\|_{L^2(\Omega)}=\rho\},\] with \(\lambda\) playing the role of a Lagrange multiplier. A cricial role in the discussion of the above problem is played by the Gagliardo-Nirenberg inequality: for any \(\Omega\) and for any \(v\in H^1_0(\Omega)\), \[\label{sobest} \|v\|^{p+1}_{L^{p+1}(\Omega)} \leq C_{N,p} \| \nabla v \|_{L^2(\Omega)}^{N(p-1)/2} \| v \|_{L^2(\Omega)} ^{(p+1)-N(p-1)/2},\] the equality holding only when \(\Omega={\mathbb{R}}^N\) and \(v=Z_{N,p}\), the positive solution of \(-\Delta Z + Z = Z^{p}\) (which is unique up to translations ). Accordingly, the exponent \(p\) can be classified in relation with the so called *\(L^2\)-critical exponent* \(1+4/N\) (throughout all the paper, \(p\) will be always Sobolev-subcritical and its criticality will be understood in the \(L^2\) sense). Indeed we have that \({\mathcal{E}}\) is bounded below and coercive on \({\mathcal{M}}_\rho\) if and only if either \(p\) is subcritical, or it is critical and \(\rho\) is sufficiently small. The recent paper deals with problem [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} in the case of the spherical domain \(\Omega = B_1\), when searching for positive solutions \(U\). In particular, it is shown that the solvability of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} is strongly influenced by the exponent \(p\), indeed: - in the subcritical case \(1<p<1+4/N\), [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} admits a unique positive solution for every \(\rho>0\); - if \(p=1+4/N\) then [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} admits a unique positive solution for \[0<\rho<\rho^*=\left(\frac{p+1}{2C_{N,p}}\right)^{N/2}=\|Z_{N,p}\|^2_{L^2({\mathbb{R}}^N)},\] and no positive solutions for \(\rho\geq\rho^*\); - finally, in the supercritical regime \(1+4/N<p<2^*-1\), [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} admits positive solutions if and only if \(0<\rho\leq\rho^*\) (the threshold \(\rho^*\) depending on \(p\)), and such solutions are at least two for \(\rho<\rho^*\). In this paper we carry on such analysis, dealing with a general domain \(\Omega\) and with solutions which are not necessarily positive. More precisely, let us recall that for any \(U\) solving [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} for some \(\lambda\), it is well-defined the Morse index \[m(U) = \max\left\{k: \begin{array}{l} \exists V\subset H^1_0(\Omega),\,\dim(V)= k:\forall v\in V\setminus\{0\}\smallskip\\ \displaystyle\int_\Omega |\nabla v|^2 + \lambda v^2-p|U|^{p-1}v^2\,dx<0 \end{array} \right\}\in{\mathbb{N}}.\] Then, if \(\Omega=B_1\), it is well known that a solution \(U\) of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} is positive if and only if \(m(U)=1\). Under this perspective, the results in can be read in terms of Morse index one--solutions, rather than positive ones: introducing the sets of admissible masses \[{\mathfrak{A}}_k ={\mathfrak{A}}_k(p,\Omega) := \left\{\rho>0: \begin{array}{l} \eqref{eq:main_prob_U} \text{ admits a solution \(U\) (for some \(\lambda\))}\\ \text{having Morse index }m(U)\leq k \end{array} \right\},\] then implies that \({\mathfrak{A}}_1(p,B_1)\) is a bounded interval if and only if \(p\) is critical or supercritical, while \({\mathfrak{A}}_1(p,B_1)={\mathbb{R}}^+\) in the subcritical case. On the contrary, when considering general domains and higher Morse index, the situation may become much more complicated. We collect some examples in the following remark. Motivated by the previous remark, the first question we address in this paper concerns the boundedness of \({\mathfrak{A}}_k\). We provide the following complete classification. The proof of such result, which is outlined in Section [2](#sec:blow-up){reference-type="ref" reference="sec:blow-up"}, is obtained by a detailed blow-up analysis of sequences of solutions with bounded Morse index, via suitable a priori pointwise estimates (see ). In this respect, the regularity assumption on \(\partial\Omega\) simplifies the treatment of possible concentration phenomena towards the boundary. The argument, which holds for solutions which possibly change sign, is inspired by, where the case of positive solutions is treated. Once Theorem [\[thm:bbd_index\]](#thm:bbd_index){reference-type="ref" reference="thm:bbd_index"} is established, in case \(p\geq 1 + 4/N\) two questions arise, namely: 1. is it possible to provide lower bounds for \(\sup{\mathfrak{A}}_k\)? Is it true that \(\sup{\mathfrak{A}}_k\) is strictly increasing in \(k\), or, at least, that \(\sup{\mathfrak{A}}_k > \sup{\mathfrak{A}}_1\) for some \(k\)? 2. is [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} solvable for every \(\rho\in(0,\sup{\mathfrak{A}}_k)\), or at least can we characterize some subinterval of solvability? It is clear that both issues can be addressed by characterizing values of \(\rho\) for which existence (and multiplicity) of solutions with bounded Morse index can be guaranteed. To this aim, it can be useful to restate problem [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} as \[\label{eq:main_prob_u} \begin{cases}-\Delta u + \lambda u = \mu|u|^{p-1}u & \text{in }\Omega,\\ \int_\Omega u^2\,dx = 1, \quad u=0 & \text{on }\partial\Omega, \end{cases} \qquad\text{where}\quad \begin{cases} U=\sqrt{\rho} u\\ \mu = \rho^{(p-1)/2}, \end{cases}\] where now \(\mu>0\) is prescribed. Since \[\label{Emu} \text{both } \mathcal{E}_{\mu}(u) := \frac{1}{2}\int_{\Omega}|\nabla u|^2-\frac{\mu}{p+1}\int_{\Omega}| u|^{p+1} \qquad \text{and }{\mathcal{M}}={\mathcal{M}}_1=\{u: \|u\|_{L^2(\Omega)}=1\}\] are invariant under the \({\mathbb{Z}}_2\)-action of the involution \(u\mapsto-u\), solutions of [\[eq:main_prob_u\]](#eq:main_prob_u){reference-type="eqref" reference="eq:main_prob_u"} can be found via min-max principles in the framework of index theories (see e.g. ). Notice that in the supercritical case \({\mathcal{E}}_\mu\) is not bounded from below on \({\mathcal{M}}\). Following, it can be convenient to parameterize solutions to [\[eq:main_prob_u\]](#eq:main_prob_u){reference-type="eqref" reference="eq:main_prob_u"} with respect to the \(H^1_0\)-norm, therefore we introduce the sets \[\label{eq:defBU} \mathcal{B}_\alpha:=\left\{u\in {\mathcal{M}}:\,\int_\Omega |\nabla u|^2\,dx<\alpha\right\},\quad\quad \mathcal{U}_\alpha:=\left\{u\in {\mathcal{M}}:\,\int_\Omega |\nabla u|^2\,dx=\alpha\right\}.\] Introducing the first Dirichlet eigenvalue of \(-\Delta\) in \(H^1_0(\Omega)\), \(\lambda_1(\Omega)\), we have that the sets above are non-empty whenever \(\alpha> \lambda_1(\Omega)\). Since we are interested in critical points having Morse index bounded from above, following we introduce the following notion of genus. We remark that this notion of genus is different from the classical one of *Krasnoselskii genus*, which is well suited for estimates of the Morse index from below, rather than above. Nonetheless, \(\gamma\) shares with the Krasnoselskii genus most of the main properties of an index. In particular, by the Borsuk-Ulam Theorem, any set \(A\) homeomorphic to the sphere \({\mathbb{S}}^{m-1} := \partial B_1 \subset {\mathbb{R}}^m\) has genus \(\gamma(A) = m\). Furthermore, we show in Section [3](#sec:2const){reference-type="ref" reference="sec:2const"} that \(\Sigma^{(k)}_{\alpha}\) is not empty, provided \(\alpha>\lambda_k(\Omega)\) (the \(k\)-th Dirichlet eigenvalue of \(-\Delta\) in \(H^1_0(\Omega)\)). Equipped with this notion of genus we provide two different variational principles for solutions of [\[eq:main_prob_u\]](#eq:main_prob_u){reference-type="eqref" reference="eq:main_prob_u"} (and thus of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"}). The first one is based on a variational problem with *two constraints*, which was exploited as the main tool in proving the results in. As a matter of fact, the results in were obtained by a detailed analysis of the map \(\alpha \mapsto \mu_\alpha\) in the case \(k=1\), i.e. when dealing with \[M_{\alpha,1} = \max\left\{\|u\|_{L^{p+1}}^{p+1}: \|u\|_{L^2}^2=1,\,\|\nabla u\|_{L^2}^2=\alpha \right\}.\] In the present paper we do not investigate the properties of the map \(\alpha \mapsto \mu_\alpha\) for general \(k\), but we rather prefer to exploit the characterization of \(M_{\alpha,k}\) in connection with a second variational principle, which deals with only *one constraint*. The link between Theorem [\[thm:genus_2constr\]](#thm:genus_2constr){reference-type="ref" reference="thm:genus_2constr"} and Theorem [\[thm:genus_1constr\]](#thm:genus_1constr){reference-type="ref" reference="thm:genus_1constr"} is that we can provide explicit estimates of \(\hat \mu_k\) (and hence of \(\hat\rho_k\)) in terms of the map \(\alpha\mapsto M_{\alpha,k}\) (see Section [4](#sec:1const){reference-type="ref" reference="sec:1const"}). We stress that the above results hold for any Lipschitz \(\Omega\). As a first consequence, this allows to extend the existence result in to non-radial domains. We observe that the exponent of \(\lambda_1(\Omega)\) in the supercritical threshold is negative, therefore such threshold decreases with the size of \(\Omega\). Once the first thresholds have been estimated, we turn to the higher ones: by exploiting the relations between \(M_{\alpha,k}\) and \(c_k\), we can show that the thresholds obtained for Morse index one--solutions in Theorem [\[thm:intro_GS\]](#thm:intro_GS){reference-type="ref" reference="thm:intro_GS"} can be increased, by considering higher Morse index--solutions, at least for some exponent. Beyond existence results for [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"}, also multiplicity results can be achieved. A first general consideration, with this respect, is that Theorem [\[thm:genus_1constr\]](#thm:genus_1constr){reference-type="ref" reference="thm:genus_1constr"} holds true also when using the standard Krasnoselskii genus instead of \(\gamma\); this allows to obtain critical points having Morse index bounded from below (see ), and therefore to obtain infinitely many solutions, at least when \(\rho\) is less than some threshold. More specifically, we can also prove the existence of a second solution in the supercritical case, thus extending to any \(\Omega\) the multiplicity result obtained in for the ball. Indeed, on the one hand, in the supercritical case \({\mathcal{E}}_\mu\) is unbounded from below; on the other hand the solution obtained in Theorem [\[thm:genus_1constr\]](#thm:genus_1constr){reference-type="ref" reference="thm:genus_1constr"}, for \(k=1\), is a local minimum. Thus the Mountain Pass Theorem applies on \(\mathcal{M}\), and a second solution can be found for \(\mu<\hat\mu_1\), see Proposition [\[mpcritlev\]](#mpcritlev){reference-type="ref" reference="mpcritlev"} for further details (and also Remark [\[rem:further_crit_lev\]](#rem:further_crit_lev){reference-type="ref" reference="rem:further_crit_lev"} for an analogous construction for \(k\ge2\)). To conclude this introduction, let us mention that the explicit lower bounds obtained in Theorem [\[thm:intro_GS\]](#thm:intro_GS){reference-type="ref" reference="thm:intro_GS"} can be easily applied in order to gain much more information also in the case of special domains, as those considered in Remark [\[rem:specialdomains\]](#rem:specialdomains){reference-type="ref" reference="rem:specialdomains"}. For instance, we can prove then following. Therefore our starting problem in \(\Omega=B\) can be solved for any mass value also in the critical and supercritical regime, at least for \(p\) smaller than this further critical exponent \(1+4/(N-1) > 1+ 4/N\). Of course, higher masses require higher Morse index--solutions. In particular, since by we know that \({\mathfrak{A}}_1(B,1+4/N) = (0,\|Z_{N,p}\|_{L^2})\), we have that for larger masses, even though no positive solution exists, nodal solutions with higher Morse index can be obtained: in such cases [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} admits *nodal ground states with higher Morse index*. The paper is structured as follows: in Section [2](#sec:blow-up){reference-type="ref" reference="sec:blow-up"} we perform a blow-up analysis of solutions with bounded Morse index, in order to prove Theorem [\[thm:bbd_index\]](#thm:bbd_index){reference-type="ref" reference="thm:bbd_index"}; Section [3](#sec:2const){reference-type="ref" reference="sec:2const"} is devoted to the analysis of the variational problem with two constraints [\[maxmin\]](#maxmin){reference-type="eqref" reference="maxmin"} and to the proof of Theorem [\[thm:genus_2constr\]](#thm:genus_2constr){reference-type="ref" reference="thm:genus_2constr"}; that of Theorems [\[thm:genus_1constr\]](#thm:genus_1constr){reference-type="ref" reference="thm:genus_1constr"}, [\[thm:intro_GS\]](#thm:intro_GS){reference-type="ref" reference="thm:intro_GS"} and Proposition [\[thm:intro_3\>1\]](#thm:intro_3>1){reference-type="ref" reference="thm:intro_3>1"} is developed in Section [4](#sec:1const){reference-type="ref" reference="sec:1const"}, by means of the variational problem with one constraint [\[infsuplev\]](#infsuplev){reference-type="eqref" reference="infsuplev"}; finally, Section [5](#sec:symm){reference-type="ref" reference="sec:symm"} contains the proof of Theorem [\[pro:symm\]](#pro:symm){reference-type="ref" reference="pro:symm"}. **Notation.** We use the standard notation \(\{\varphi_k\}_{k\geq1}\) for a basis of eigenfunctions of the Dirichlet laplacian in \(\Omega\), orthogonal in \(H^1_0(\Omega)\) and orthonormal in \(L^2(\Omega)\). Such functions are ordered in such a way that the corresponding eigenvalues \(\lambda_k(\Omega)\) satisfy \[0<\lambda_1(\Omega)<\lambda_2(\Omega)\leq\lambda_3(\Omega)\leq\dots,\] and \(\varphi_1\) is chosen to be positive on \(\Omega\). \(C_{N,p}\) denotes the universal constant in the Gagliardo-Nirenberg inequality [\[sobest\]](#sobest){reference-type="eqref" reference="sobest"}, which is achieved (uniquely, up to translations and dilations) by the positive, radially symmetric function \(Z_{N,p}\in H^1({\mathbb{R}}^N)\), with \[\|Z_{N,p}\|^2_{L^2({\mathbb{R}}^N)}=\left(\frac{p+1}{2C_{N,p}}\right)^{N/2}.\] Finally, \(C\) denotes every (positive) constant we need not to specify, whose value may change also within the same formula. # Blow-up analysis of solutions with bounded Morse index {#sec:blow-up} Throughout this section we will deal with a sequence \(\{(u_n,\mu_n,\lambda_n)\}_n \subset H^1_0(\Omega)\times{\mathbb{R}}^+\times{\mathbb{R}}\) satisfying \[\label{eq:auxiliary_n}-\Delta u_n+\lambda_n u_n=\mu_n |u_n|^{p-1}u_n,\qquad\int_\Omega u_n^2\, dx=1,\qquad \int_\Omega |\nabla u_n|^2\, dx=:\alpha_n.\] To start with, we recall the following result (actually, in, the result is stated for positive solution, but the proof does not require such assumption). Next we turn to the study of sequences having arbitrarily large \(H^1_0\)-norm. In particular, we will focus on sequences of solutions having a common upper bound on the Morse index \[m(u_n) = \max\left\{k: \begin{array}{l} \exists V\subset H^1_0(\Omega),\,\dim(V)= k:\forall v\in V\setminus\{0\}\smallskip\\ \displaystyle\int_\Omega |\nabla v|^2 + \lambda_n v^2-p\mu_n|u_n|^{p-1}v^2\,dx<0 \end{array} \right\}.\] Throughout this section we will assume that \[\label{eq:mainass_secMorse} \text{the sequence }\{(u_n,\mu_n,\lambda_n)\}_n\text{ satisfies \eqref{eq:auxiliary_n}, with }\alpha_n\to+\infty\text{ and }m(u_n)\leq \bar k,\] for some \(\bar k\in{\mathbb{N}}\) not depending on \(n\). The local description of the asymptotic behaviour of the solutions \(U_n\) to [\[equn\]](#equn){reference-type="eqref" reference="equn"} with bounded Morse index can be carried out more conveniently by defining the sequence (see ) \[\label{defVn} V_n(y)=\varepsilon_n^{\frac{2}{p-1}}\, U_n(\varepsilon_n\,y+P_n),\quad y\in \Omega_n :=\frac{\Omega-P_n}{\varepsilon_n},\] where \(P_n\) is defined before [\[tildepsn\]](#tildepsn){reference-type="eqref" reference="tildepsn"}, and \(\varepsilon_n=\frac{1}{\sqrt{\lambda_n}}\to 0\). Then, \(V_n\) satisfies \[\nonumber \left\{ \begin{array}{ll} -\Delta V_n+ V_n=| V_n|^{p-1} V_n, & \hbox{in}\,\,\Omega_n;\\ |V_n|\le | V_n(0)|=\big ({\varepsilon_n/\tilde\varepsilon_n}\big )^{\frac{2}{p-1}}\rightarrow \tilde\lambda^{-\frac{1}{p-1}}, & \hbox{in}\,\, \Omega_n;\\ V_n=0, & \hbox{on}\,\, \partial\Omega_n. \end{array} \right.\] As before, we have (up to a subsequence) \(V_n\rightarrow V\) in \(\mathcal{C}^1_{\mathrm{loc}}(\overline H)\) where \(H\) is either \({\mathbb{R}}^N\) or a half space and \(V\) solves \[\label{limprob2} \left\{ \begin{array}{ll} -\Delta V+ V=| V|^{p-1} V, & \hbox{in}\,\, H;\\ |V|\le | V(0)|=\tilde\lambda^{-\frac{1}{p-1}}, & \hbox{in}\,\, H;\\ V=0, & \hbox{on}\,\, \partial H. \end{array} \right.\] By recalling the discussion following [\[limprob1\]](#limprob1){reference-type="eqref" reference="limprob1"} we also have \(m(V)<+\infty\). We collect some well known property of such a \(V\) in the following result. Following the same pattern as in, we now analyze the global behaviour of a sequence \(\{U_n\}\) of solutions to [\[equn\]](#equn){reference-type="eqref" reference="equn"} for \(\lambda_n\to +\infty\), assuming that \(\lim_{n\to +\infty} m(U_n)\leq\bar k<\infty.\) By the previous discussion, if \(P^1_n\) is a sequence of points such that \(|U_n(P^1_n)|=\|U_n\|_{L^{\infty}(\Omega)}\), we have \(|U_n(P^1_n)|\rightarrow +\infty\) and \({\lambda_n}\,d(P^1_n,\partial\Omega)^2\rightarrow +\infty\). We now look for other possible sequences of (local) extremum points \(P^i_n\), \(i=2,3..\), along which \(|U_n|\) goes to infinity. For any \(R>0\), consider the quantity \[\nonumber h_1(R)=\limsup_{n\to +\infty} \Bigl (\lambda_n^{-\frac{1}{p-1}}\max_{|x-P^1_n|\ge R\,\lambda_n^{-1/2}} |U_n(x)| \Bigr ).\] We will prove that if \(h_1(R)\) is *not vanishing* for large \(R\), then there exists a 'blow-up' sequence \(P^2_n\) for \(u_n\), 'disjoint' from \(P^1_n\). Indeed, let us suppose that \[\nonumber \limsup_{R\to +\infty} h_1(R)=4\delta>0.\] Hence, up to a subsequence and for arbitrarily large \(R\), we have \[\label{ass1} \lambda_n^{-\frac{1}{p-1}}\max_{|x-P^1_n|\ge R\,\lambda_n^{-1/2}} |U_n(x)| \ge 2\delta.\] Since \(U_n\) vanishes on \(\partial\Omega\), there exists \(P^2_n\in\Omega\backslash B_{R\,\lambda_n^{-1/2}}(P_n^1)\) such that \[\label{pn2} |U_n(P_n^2)|=\max_{|x-P^1_n|\ge R\,\lambda_n^{-1/2}}|U_n(x)|.\] Clearly, assumption [\[ass1\]](#ass1){reference-type="eqref" reference="ass1"} implies that \(|U_n(P_n^2)|\rightarrow +\infty\). We first prove that the sequences \(P_n^1\) and \(P_n^2\) are far away each other. Furthermore, we also have that the blow-up points stay far away from the boundary. We can now iterate the previous arguments: let us define, for \(k\ge 1\), \[\label{defhn} h_k(R)=\limsup_{n\to +\infty} \Bigl (\lambda_n^{-\frac{1}{p-1}}\max_{d_{n,k}(x)\ge R\,\lambda_n^{-1/2}} |U_n(x)| \Bigr ),\] where \[d_{n,k}(x): =\min\{|x-P^i_n|\,:\, i=1...,k\}\] and the sequences \(P^i_n\) are such that \[\nonumber \sqrt{\lambda_n}\,d(P^i_n,\partial\Omega)\rightarrow +\infty;\quad \lambda_n^{1/2}|P_n^i-P^j_n|\rightarrow +\infty,\quad\quad i,j=1...,k,\quad i\neq j\] as \(n\to +\infty\). Assume that \[\limsup_{n\to +\infty} h_k(R)=4\delta>0.\] As before, up to a subsequence and for arbitrarily large \(R\), we have \[\label{assk} \lambda_n^{-\frac{1}{p-1}}\max_{d_{n,k}(x)\ge R\,\lambda_n^{-1/2}} |U_n(x)| \ge 2\delta\] and there exist \(P^{k+1}_n\) so that \[\nonumber |U_n(P_n^{k+1})|=\max_{d_{n,k}(x)\ge R\,\lambda_n^{-1/2}}|U_n(x)|\] with \(\lim_{n\to +\infty}|U_n(P_n^{k+1})|=+\infty\). Moreover, as in Lemma [\[disj\]](#disj){reference-type="ref" reference="disj"} we deduce that, for every \(i=1...,k\) \[\label{limblowseqi} \lambda_n^{-\frac{1}{p-1}}\,U_n(\lambda_n^{-1/2}\,y+P^i_n): = V^i_n(y)\rightarrow V^i(y)\quad \textrm{in} \,\, \mathcal{C}^1_{{\mathrm{loc}}}({\mathbb{R}}^N)\] as \(n\to +\infty\); hence, by [\[assk\]](#assk){reference-type="eqref" reference="assk"} and again from the vanishing of \(V\) at infinity, we conclude that \[\lambda_n^{1/2}|P_n^{k+1}-P^i_n|\rightarrow +\infty\] as \(n\to \infty\), for every \(i=1...,k\). Setting now \[\nonumber \tilde\varepsilon^{k+1}_n: =|U_n(P^{k+1}_n)|^{-\frac{p-1}{2}}\quad \mathrm{and} \quad R_n^{(k+1)}:=\frac{1}{2}\,\frac{d_{n,k}(P^{k+1}_n)}{\tilde\varepsilon^{k+1}_n}\] we still have \(\tilde\varepsilon^{k+1}_n\to 0\) and, by [\[assk\]](#assk){reference-type="eqref" reference="assk"}, \(R_n^{(k+1)} \to +\infty\) as \(n\to \infty\) (see Lemma [\[distbd\]](#distbd){reference-type="ref" reference="distbd"}). Then, by the same arguments as in Lemma [\[distbd\]](#distbd){reference-type="ref" reference="distbd"}, we get \[\label{maxpkn} |U_n(P_n^{k+1})|=\max_{\Omega\cap B_{R^{(k+1)}_n\tilde\varepsilon^{k+1}_n}(P^{k+1}_n)} |u_n|\,,\] and furthermore \[\lim_{n\to +\infty}\tilde\varepsilon_n^{k+1}\,\sqrt{\lambda_n}>0\,,\] so that by defining \(R_n=:R^{(k+1)}_n\tilde\varepsilon_n^{k+1}\,\sqrt{\lambda_n}\rightarrow +\infty\) we have \[\label{maxpknball} |U_n(P_n^{k+1})|=\max_{\Omega\cap B_{R_n\lambda^{-1/2}_n}(P^{k+1}_n)}| U_n|.\] Now, by the same arguments as in, it turns out that the iterative procedure must stop after *at most* \(\bar k-1\) steps, where \(\bar k =\lim_{n\to +\infty} m(u_n)\). Thus, we have proved: We now show that the sequence \(U_n\) decays exponentially away from the blow-up points. We now exploit the previous results to show that suitable rescalings of the solutions to [\[eq:auxiliary_n\]](#eq:auxiliary_n){reference-type="eqref" reference="eq:auxiliary_n"} converge (locally) to some bounded solution \(V\) of \[\label{eqV}-\Delta V+ V=| V|^{p-1} V\] in \({\mathbb{R}}^N\). The previous lemma allows us to gain some information on the asymptotic behavior of the sequences \(\lambda_n\), \(\mu_n\) and \(\|u_n\|_{L^{p+1}(\Omega)}\). We first provide some bounds for the solutions of the limit problem [\[eqV\]](#eqV){reference-type="eqref" reference="eqV"} which will be useful in the sequel. # Max-min principles with two constraints {#sec:2const} In this section we deal with the maximization problem with two constraints introduced in, aiming at considering more general max-min classes of critical points. Let \({\mathcal{M}}\) be defined in [\[Emu\]](#Emu){reference-type="eqref" reference="Emu"} and, for any fixed \(\alpha>\lambda_1(\Omega)\), let \(\mathcal{B}_\alpha\), \(\mathcal{U}_\alpha\) be defined as in [\[eq:defBU\]](#eq:defBU){reference-type="eqref" reference="eq:defBU"}. We will look for critical points of the \(\mathcal{C}^2\) functional \[f(u)=\int_{\Omega}|u|^{p+1},\quad\quad\quad u\in {\mathcal{M}},\] constrained to \(\mathcal{U}_\alpha\). To start with, we notice that the topological properties of such set depend on \(\alpha\). Of course \({\mathcal{U}}_\alpha\) is closed and odd, for any \(\alpha\). Recalling Definition [\[def:genus\]](#def:genus){reference-type="ref" reference="def:genus"} we deduce that its genus \(\gamma({\mathcal{U}}_\alpha)\) is well defined. Now we turn to the properties of the functional \(f\). To start with, it satisfies the Palais-Smale (P.S. for short) condition on \(\overline{\mathcal{B}}_{\alpha}\); more precisely, the following holds. We can combine the previous lemmas to prove one of the main results stated in the introduction. We conclude this section with the following estimate. # Min-max principles on the unit sphere in \(L^2\) {#sec:1const} According to equation [\[Emu\]](#Emu){reference-type="eqref" reference="Emu"}, let \({\mathcal{M}}\subset H^1_0(\Omega)\) denote the unit sphere with respect to the \(L^2\) norm and \(\mathcal{E}_{\mu}\) the energy functional associated to [\[eq:main_prob_u\]](#eq:main_prob_u){reference-type="eqref" reference="eq:main_prob_u"}. In this section we are concerned with critical points of \(\mathcal{E}_{\mu}\) on \({\mathcal{M}}\) (which, in turn, correspond to solutions of our starting problem [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"}). By the Gagliardo-Nirenberg inequality [\[sobest\]](#sobest){reference-type="eqref" reference="sobest"}, setting \(\|\nabla u\|^2_{L^2}=\alpha\), one obtains \[\label{eq:boundonboundEmu} \frac12\,\alpha-\mu\frac{C_{N,p}}{p+1}\,\alpha^{N(p-1)/4} \leq \mathcal{E}_{\mu}(u)\le \frac12\alpha.\] In particular, \(\mathcal{E}_{\mu}\) is bounded on any bounded subset of \({\mathcal{M}}\), and it is bounded from below (and coercive) on the entire \({\mathcal{M}}\) for subcritical \(p<1+4/N\) and for critical \(p=1+4/N\) whenever \(\mu< \frac{p+1}{2}C_{N,p}^{-1}\). In these cases, one can easily show that \(\mathcal{E}_{\mu}\) satifies the P.S. condition and apply the classical minimax principle for even functionals on a closed symmetric submanifold (see e.g. ). In the complementary case, when \(p\) is either supercritical, i.e. \(p>1+4/N\), or critical and \(\mu\) is large, then \(\mathcal{E}_{\mu}\) is not bounded from below (see e.g. [\[minusinfty\]](#minusinfty){reference-type="eqref" reference="minusinfty"} below). In order to provide a minimax principle suitable for this case, we recall the Definition [\[def:genus\]](#def:genus){reference-type="ref" reference="def:genus"} of genus and that of \(\mathcal{B}_\alpha\) (see equation [\[eq:defBU\]](#eq:defBU){reference-type="eqref" reference="eq:defBU"}). Furthermore, we denote with \(K_{c}\) the (closed and symmetric) set of critical points of \({\mathcal{E}}_\mu\) at level \(c\) contained in \(\mathcal{B}_\alpha\). The following theorem is an adaptation of well known arguments of previous critical point theorems relying on index theory. We now provide a sufficient condition to guarantee the validity of assumption [\[ass2\]](#ass2){reference-type="eqref" reference="ass2"}. Exploiting the results above, we are ready to prove our main existence results. To conclude this section we prove that in the supercritical case, if \(\mu\) is not too large, in addition to \((c_k)_k\) there is a further sequence of critical levels \((\bar c_k)\) of \(\mathcal{E}_{\mu}\) constrained to \(\mathcal{M}\). For concreteness, let us first consider the case \(k=1\): since in such case \(c_1\) is a local minimum of \({\mathcal{E}}_\mu\) in \(\mathcal{M}\), and \({\mathcal{E}}_\mu\) is unbounded from below in \(\mathcal{M}\), the critical level \(\bar c_1\) is of mountain pass type. # Results in symmetric domains {#sec:symm} This section is devoted to the proof of Theorem [\[pro:symm\]](#pro:symm){reference-type="ref" reference="pro:symm"}, therefore we assume \(1+4/N \leq p < 2^*-1\). We perform the proof in the case of \(\Omega=B\), but it will be clear that the main assumption on \(\Omega\) is the following: - there is a tiling of \(\Omega\), made by \(h\) copies of a subdomain \(D\), in such a way that from any solution \(U_D\) of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} on \(D\) one can construct, using reflections, a solution \(U_\Omega\) of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} on \(\Omega\). Then \(U_\Omega\) has \(h\) times the mass of \(U_D\), and recalling Theorem [\[thm:intro_GS\]](#thm:intro_GS){reference-type="ref" reference="thm:intro_GS"} we deduce that [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} on \(\Omega\) is solvable for any \(\rho< h \cdot D_{N,p} \lambda_1(D)^{\frac{2}{p-1}-\frac{N}{2}}\). At this point, for a sequence \((D_k,h_k)_k\) of tilings satisfying **(T)**, we obtain the solvability of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} on \(\Omega\) whenever \[\rho< h_k \cdot D_{N,p} \lambda_1(D_k)^{\frac{2}{p-1}-\frac{N}{2}},\] and if we can show that \[\label{eq:finaltarget} \frac{ h_k }{ \lambda_1(D_k)^{\frac{N}{2}-\frac{2}{p-1}}} \to +\infty\qquad\text{as }k\to+\infty,\] we deduce the solvability of [\[eq:main_prob_U\]](#eq:main_prob_U){reference-type="eqref" reference="eq:main_prob_U"} on \(\Omega\) for every mass. Having this scheme in mind, it is easy to prove analogous results on rectangles and also in other kind of domains. Then let \(B\subset{\mathbb{R}}^N\) be the ball (w.l.o.g. of radius one), and let \[D_k:=\left\{(r\cos\theta, r\sin\theta,x_3,\dots,x_N)\in B:-\frac{\pi}{k} < \theta < \frac{\pi}{k}\right\}\] Then \(D_k\) satisfies **(T)**, with \(h_k=k\). In order to estimate \(\lambda_1(D_k)\) we observe that, by elementary trigonometry, \[B'_k = B_{\frac{\sin(\pi/k)}{\sin(\pi/k)+1}}\left(\frac{1}{\sin(\pi/k)+1},0,0,\dots,0\right) \subset D_k,\] and therefore \[\lambda_1(D_k) \le \lambda_1(B'_k) \le C k^2,\] for some dimensional constant \(C=C(N)\) and \(k\) large. Then \[\frac{ h_k }{ \lambda_1(D_k)^{\frac{N}{2}-\frac{2}{p-1}}}\ge C \frac{k}{k^{{N}-\frac{4}{p-1}}} = C k^{1-{N}+\frac{4}{p-1}} = C k^{\frac{N-1}{p-1}\left[1+\frac{4}{N-1}-p\right]},\] and finally [\[eq:finaltarget\]](#eq:finaltarget){reference-type="eqref" reference="eq:finaltarget"} holds true whenever \(p< 1+\frac{4}{N-1}\), thus completing the proof of Theorem [\[pro:symm\]](#pro:symm){reference-type="ref" reference="pro:symm"}.
{'timestamp': '2016-07-18T02:07:27', 'yymm': '1607', 'arxiv_id': '1607.04520', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04520'}
# Introduction {#sec:intro} Despite the immense popularity and availability of online video content via outlets such as Youtube and Facebook, most work on object detection focuses on static images. Given the breakthroughs of deep convolutional neural networks for detecting objects in static images, the application of these methods to video might seem straightforward. However, motion blur and compression artifacts cause substantial frame-to-frame variability, even in videos that appear smooth to the eye. These attributes complicate prediction tasks like classification and localization. Object-detection models trained on images tend not to perform competitively on videos owing to domain shift factors. Moreover, object-level annotations in popular video data-sets can be extremely sparse, impeding the development of better video-based object detection models. Girshik *et al*demonstrate that even given scarce labeled training data, high-capacity convolutional neural networks can achieve state of the art detection performance if first pre-trained on a related task with abundant training data, such as 1000-way ImageNet classification. Followed the pretraining, the networks can be fine-tuned to a related but distinct domain. Also relevant to our work, the recently introduced models Faster R-CNN and You Look Only Once (YOLO) unify the tasks of classification and localization. These methods, which are accurate and efficient, propose to solve both tasks through a single model, bypassing the separate object proposal methods used by R-CNN. In this paper, we introduce a method to extend unified object recognition and localization to the video domain. Our approach applies transfer learning from the image domain to video frames. Additionally, we present a novel recurrent neural network (RNN) method that refines predictions by exploiting contextual information in neighboring frames. In summary, we contribute the following: - A new method for refining a video-based object detection consisting of two parts: (i) a *pseudo-labeler*, which assigns provisional labels to all available video frames. (ii) A recurrent neural network, which reads in a sequence of provisionally labeled frames, using the contextual information to output refined predictions. - An effective training strategy utilizing (i) category-level weak-supervision at every time-step, (ii) localization-level strong supervision at final time-step (iii) a penalty encouraging prediction smoothness at consecutive time-steps, and (iv) similarity constraints between *pseudo-labels* and prediction output at every time-step. - An extensive empirical investigation demonstrating that on the YouTube Objects dataset, our framework achieves mean average precision (mAP) of \(68.73\) on test data, compared to a best published result of \(37.41\) and \(61.66\) for a domain adapted YOLO network. # Methods {#sec:method} In this work, we aim to refine object detection in video by utilizing contextual information from neighboring video frames. We accomplish this through a two-stage process. First, we train a *pseudo-labeler*, that is, a domain-adapted convolutional neural network for object detection, trained individually on the labeled video frames. Specifically, we fine-tune the YOLO object detection network, which was originally trained for the 20-class PASCAL VOC dataset, to the Youtube-Video dataset. When fine-tuning to the 10 sub-categories present in the video dataset, our objective is to minimize the weighted squared detection loss (equation [\[eqn:obj_det_loss\]](#eqn:obj_det_loss){reference-type="ref" reference="eqn:obj_det_loss"}) as specified in YOLO. While fine-tuning, we learn only the parameters of the top-most fully-connected layers, keeping the \(24\) convolutional layers and \(4\) max-pooling layers unchanged. The training takes roughly 50 epochs to converge, using the RMSProp optimizer with momentum of \(0.9\) and a mini-batch size of \(128\). As with YOLO, our fine-tuned \(pseudo-labeler\) takes \(448 \times 448\) frames as input and regresses on category types and locations of possible objects at each one of \(S \times S\) non-overlapping grid cells. For each grid cell, the model outputs class conditional probabilities as well as \(B\) bounding boxes and their associated confidence scores. As in YOLO, we consider a *responsible* bounding box for a grid cell to be the one among the \(B\) boxes for which the predicted area and the ground truth area shares the maximum Intersection Over Union. During training, we simultaneously optimize classification and localization error (equation [\[eqn:obj_det_loss\]](#eqn:obj_det_loss){reference-type="ref" reference="eqn:obj_det_loss"}). For each grid cell, we minimize the localization error for the *responsible* bounding box with respect to the ground truth only when an object appears in that cell. Next, we train a Recurrent Neural Network (RNN), with Gated Recurrent Units (GRUs). This net takes as input sequences of *pseudo-labels*, optimizing an objective that encourages both accuracy on the target frame and consistency across consecutive frames. Given a series of *pseudo-labels* \(\mathbf{x}^{(1)},..., \mathbf{x}^{(T)}\), we train the RNN to generate improved predictions \(\hat{\mathbf{y}}^{(1)},..., \hat{\mathbf{y}}^{(T)}\) with respect to the ground truth \(\mathbf{y}^{(T)}\) available only at the final step in each sequence. Here, \(t\) indexes sequence steps and \(T\) denotes the length of the sequence. As output, we use a fully-connected layer with a linear activation function, as our problem is regression. In our final experiments, we use a \(2\)-layer GRU with \(150\) nodes per layer, hyper-parameters determined on validation data. The following equations define the forward pass through a GRU layer, where \(\mathbf{h}^{(t)}_l\) denotes the layer's output at the current time step, and \(\mathbf{h}^{(t)}_{l-1}\) denotes the previous layer's output at the same sequence step: \[\label{eqn:GRU} \begin{aligned} \mathbf{r}^{(t)}_l &= \sigma(\mathbf{h}^{(t)}_{l-1}W^{xr}_l + \mathbf{h}^{(t-1)}_lW^{hr}_l + \mathbf{b}^r_l)\\ \mathbf{u}^{(t)}_l &= \sigma(\mathbf{h}^{(t)}_{l-1}W^{xu}_l + \mathbf{h}^{(t-1)}_lW^{hu}_l + \mathbf{b}^u_l)\\ \mathbf{c}^{(t)}_l &= \sigma(\mathbf{h}^{(t)}_{l-1}W^{xc}_l + r_t \odot(\mathbf{h}^{(t-1)}_lW^{hc}_l) + \mathbf{b}^c_l)\\ \mathbf{h}^{(t)}_l &= (1-\mathbf{u}^{(t)}_l)\odot \mathbf{h}^{(t-1)}_l + \mathbf{u}^{(t)}_l\odot \mathbf{c}^{(t)}_l \end{aligned}\] Here, \(\sigma\) denotes an element-wise logistic function and \(\odot\) is the (element-wise) Hadamard product. The reset gate, update gate, and candidate hidden state are denoted by \(\textbf{r}\), \(\textbf{u}\), and \(\textbf{c}\) respectively. For \(S = 7\) and \(B=2\), the pseudo-labels \(\mathbf{x}^{(t)}\) and prediction \(\hat{\mathbf{y}}^{(t)}\) both lie in \(\mathbb{R}^{1470}\). ## Training We design an objective function (Equation [\[eqn:objective\]](#eqn:objective){reference-type="ref" reference="eqn:objective"}) that accounts for both accuracy at the target frame and consistency of predictions across adjacent time steps in the following ways: \[\label{eqn:objective} \mbox{loss} = \mbox{d\_loss} + \alpha \cdot \mbox{s\_loss} + \beta \cdot \mbox{c\_loss} + \gamma \cdot \mbox{pc\_loss}\] Here, d_loss, s_loss, c_loss and pc_loss stand for detection_loss, similarity_loss, category_loss and prediction_consistency_loss described in the following sections. The values of the hyper-parameters \(\alpha=0.2\), \(\beta=0.2\) and \(\gamma=0.1\) are chosen based on the detection performance on the validation set. The training converges in 80 epochs for parameter updates using RMSProp and momentum \(0.9\). During training we use a mini-batch size of \(128\) and sequences of length \(30\). ### Strong Supervision at Target Frame On the final output, for which the ground truth classification and localization is available, we apply a multi-part object detection loss as described in YOLO. \[\label{eqn:obj_det_loss} \begin{aligned} \mbox{detection\_loss} &= \lambda_{coord}\sum^{S^2}_{i=0}\sum^{B}_{j=0}\mathbbm{1}^{obj}_{ij}\big(\mathit{x}^{(T)}_i-\hat{\mathit{x}}^{(T)}_i\big)^2 + \big(\mathit{y}^{(T)}_i-\hat{\mathit{y}}^{(T)}_i\big)^2 \\ & + \lambda_{coord}\sum^{S^2}_{i=0}\sum^{B}_{j=0}\mathbbm{1}^{obj}_{ij}\big(\sqrt{w_i}^{(T)}-\sqrt{\hat{w}^{(T)}_i}\big)^2 + \big (\sqrt{h_i}^{(T)}-\sqrt{\hat{h}^{(T)}_i} \big)^2 \\ & + \sum^{S^2}_{i=0}\sum^{B}_{j=0}\mathbbm{1}^{obj}_{ij}(\mathit{C}_i-\hat{\mathit{C}_i})^2 \\ & + \lambda_{noobj}\sum^{S^2}_{i=0}\sum^{B}_{j=0}\mathbbm{1}^{noobj}_{ij}\big(\mathit{C}^{(T)}_i-\hat{\mathit{C}}^{(T)}_i\big)^2 \\ & + \sum^{S^2}_{i=0}\mathbbm{1}^{obj}_{i}\sum_{c \in classes}\big(p_i^{(T)}(c)-\hat{p_i}^{(T)}(c)\big)^2 \end{aligned}\] where \(\mathbbm{1}^{obj}_{i}\) denotes if the object appears in cell \(i\) and \(\mathbbm{1}^{obj}_{ij}\) denotes that \(j\)th bounding box predictor in cell \(i\) is *responsible* for that prediction. The loss function penalizes classification and localization error differently based on presence or absence of an object in that grid cell. \(x_i, y_i, w_i, h_i\) corresponds to the ground truth bounding box center coordinates, width and height for objects in grid cell (if it exists) and \(\hat{x_i}, \hat{y_i}, \hat{w_i}, \hat{h_i}\) stand for the corresponding predictions. \(C_i\) and \(\hat{C_i}\) denote confidence score of *objectness* at grid cell \(i\) for ground truth and prediction. \(p_i(c)\) and \(\hat{p_i}(c)\) stand for conditional probability for object class \(c\) at cell index \(i\) for ground truth and prediction respectively. We use similar settings for YOLO's object detection loss minimization and use values of \(\lambda_{coord} = 5\) and \(\lambda_{noobj} = 0.5\). ### Similarity between *Pseudo-labels* and Predictions Our objective function also includes a regularizer that penalizes the dissimilarity between *pseudo-labels* and the prediction at each time frame \(t\). \[\label{auto_enc_loss} \mbox{similarity\_loss} = \sum^T_{t=0}\sum^{S^2}_{i=0}\hat{C}^{(t)}_i\Big(\mathbf{x}^{(t)}_i-\hat{\mathbf{y}_i}^{(t)} \Big)^2\] Here, \(\mathbf{x}^{(t)}_i\) and \(\hat{\mathbf{y}_i}^{(t)}\) denote the *pseudo-labels* and predictions corresponding to the \(i\)-th grid cell at \(t\)-th time step respectively. We perform minimization of the square loss weighted by the predicted confidence score at the corresponding cell. ### Object Category-level Weak-Supervision Replication of the static target at each sequential step has been shown to be effective in. Of course, with video data, different objects may move in different directions and speeds. Yet, within a short time duration, we could expect all objects to be present. Thus we employ target replication for classification but not localization objectives. We minimize the square loss between the categories aggregated over all grid cells in the ground truth \(\mathbf{y}^{(T)}\) at final time step \(T\) and predictions \(\hat{\mathbf{y}}^{(t)}\) at all time steps \(t\). Aggregated category from the ground truth considers only the cell indices where an object is present. For predictions, contribution of cell \(i\) is weighted by its predicted confidence score \(\hat{C}^{(t)}_i\). Note that cell indices with positive detection are sparse. Thus, we consider the confidence score of each cell while minimizing the aggregated category loss. \[\label{category_supervision} \mbox{category\_loss} = \sum^T_{t=0}\bigg(\sum_{c \in classes} \Big(\sum^{S^2}_{i=0} \hat{C}^{(t)}_i\big(\hat{p}^{(t)}_i(c)\big)-\sum^{S^2}_{i=0}\mathbbm{1}^{obj^{(T)}}_i \big(p_i^{(T)}(c)\big)\Big) \bigg)^2\] ### Consecutive Prediction Smoothness Additionally, we regularize the model by encouraging smoothness of predictions across consecutive time-steps. This makes sense intuitively because we assume that objects rarely move rapidly from one frame to another. \[\label{prediction_smoothness} \mbox{prediction\_consistency\_loss} = \sum^{T-1}_{t=0}\Big(\hat{\mathbf{y}_i}^{(t)}-\hat{\mathbf{y}_i}^{(t+1)} \Big)^2\] ## Inference The recurrent neural network predicts output at every time-step. The network predicts \(98\) bounding boxes per video frame and class probabilities for each of the \(49\) grid cells. We note that for every cell, the net predicts class conditional probabilities for each one of the \(C\) categories and \(B\) bounding boxes. Each one of the \(B\) predicted bounding boxes per cell has an associated *objectness* confidence score. The predicted confidence score at that grid is the maximum among the boxes. The bounding box with the highest score becomes the *responsible* prediction for that grid cell \(i\). The product of class conditional probability \(\hat{p}^{(t)}_i(c)\) for category type \(c\) and *objectness* confidence score \(\hat{C}^{(t)}_i\) at grid cell \(i\), if above a threshold, infers a detection. In order for an object of category type \(c\) to be detected for \(i\)-th cell at time-step \(t\), both the class conditional probability \(\hat{p}^{(t)}_i(c)\) and *objectness score* \(\hat{C}^{(t)}_i\) must be reasonably high. Additionally, we employ Non-Maximum Suppression (NMS) to winnow multiple high scoring bounding boxes around an object instance and produce a single detection for an instance. By virtue of YOLO-style prediction, NMS is not critical. # Experimental Results {#sec:results} In this section, we empirically evaluate our model on the popular **Youtube-Objects** dataset, providing both quantitative results (as measured by mean Average Precision) and subjective evaluations of the model's performance, considering both successful predictions and failure cases. The **Youtube-Objects** dataset is composed of videos collected from Youtube by querying for the names of 10 object classes of the PASCAL VOC Challenge. It contains 155 videos in total and between 9 and 24 videos for each class. The duration of each video varies between 30 seconds and 3 minutes. However, only \(6087\) frames are annotated with \(6975\) bounding-box instances. The training and test split is provided. ## Experimental Setup We implement the domain-adaption of YOLO and the proposed RNN model using Theano. Our best performing RNN model uses two GRU layers of \(150\) hidden units each and dropout of probability \(0.5\) between layers, significantly outperforming domain-adapted YOLO alone. While we can only objectively evaluate prediction quality on the labeled frames, we present subjective evaluations on sequences. ## Objective Evaluation We compare our approach with other methods evaluated on the Youtube-Objects dataset. As shown in Table [1](#table:per_category_results){reference-type="ref" reference="table:per_category_results"} and Table [2](#table:final_mAP){reference-type="ref" reference="table:final_mAP"}, Deformable Parts Model (DPM) )-based detector reports mean average precision below \(30\), with especially poor performance in some categories such as *cat*. The method of Tripathi *et al*(VPO) uses consistent video object proposals followed by a domain-adapted AlexNet classifier (5 convolutional layer, 3 fully connected) in an R-CNN-like framework, achieving mAP of \(37.41\). We also compare against YOLO (\(24\) convolutional layers, 2 fully connected layers), which unifies the classification and localization tasks, and achieves mean Average Precision over \(55\). In our method, we adapt YOLO to generate *pseudo-labels* for all video frames, feeding them as inputs to the refinement RNN. We choose YOLO as the *pseudo-labeler* because it is the most accurate among feasibly fast image-level detectors. The domain-adaptation improves YOLO's performance, achieving mAP of \(61.66\). Our model with RNN-based prediction refinement, achieves superior aggregate mAP to all baselines. The RNN refinement model using both input-output similarity, category-level weak-supervision, and prediction smoothness performs best, achieving \(\mbox{68.73}\) mAP. This amounts to a relative improvement of \(\mbox{11.5\%}\) over the best baselines. Additionally, the RNN improves detection accuracy on most individual categories (Table [1](#table:per_category_results){reference-type="ref" reference="table:per_category_results"}). [\[table:per_category_results\]]{#table:per_category_results label="table:per_category_results"} [\[table:final_mAP\]]{#table:final_mAP label="table:final_mAP"} ## Subjective Evaluation We provide a subjective evaluation of the proposed RNN model in Figure [\[fig:subjective1\]](#fig:subjective1){reference-type="ref" reference="fig:subjective1"}. Top and bottom rows in every pair of sequences correspond to *pseudo-labels* and results from our approach respectively. While only the last frame in each sequence has associated ground truth, we can observe that the RNN produces more accurate and more consistent predictions across time frames. The predictions are consistent with respect to classification, localization and confidence scores. In the first example, the RNN consistently detects the *dog* throughout the sequence, even though the *pseudo-labels* for the first two frames were wrong (*bird*). In the second example, *pseudo-labels* were *motorbike*, *person*, *bicycle* and even *none* at different time-steps. However, our approach consistently predicted *motorbike*. The third example shows that the RNN consistently predicts both of the cars while the *pseudo-labeler* detects only the smaller car in two frames within the sequence. The last two examples show how the RNN increases its confidence scores, bringing out the positive detection for *cat* and *car* respectively both of which fell below the detection threshold of the *pseudo-labeler*. ## Areas For Improvement The YOLO scheme for unifying classification and localization imposes strong spatial constraints on bounding box predictions since each grid cell can have only one class. This restricts the set of possible predictions, which may be undesirable in the case where many objects are in close proximity. Additionally, the rigidity of the YOLO model may present problems for the refinement RNN, which encourages smoothness of predictions across the sequence of frames. Consider, for example, an object which moves slightly but transits from one grid cell to another. Here smoothness of predictions seems undesirable. Figure [\[fig:failure_cases\]](#fig:failure_cases){reference-type="ref" reference="fig:failure_cases"} shows some failure cases. In the first case, the *pseudo-labeler* classifies the instances as *dogs* and even as *birds* in two frames whereas the ground truth instances are *horses*. The RNN cannot recover from the incorrect pseudo-labels. Strangely, the model increases the confidence score marginally for a different wrong category *cow*. In the second case, possibly owing to motion and close proximity of multiple instances of the same object category, the RNN predicts the correct category but fails on localization. These point to future work to make the framework robust to motion. The category-level weak supervision in the current scheme assumes the presence of all objects in nearby frames. While for short snippets of video this assumption generally holds, it may be violated in case of occlusions, or sudden arrival or departure of objects. In addition, our assumptions regarding the desirability of prediction smoothness can be violated in the case of rapidly moving objects. # Related Work {#sec:prior-art} Our work builds upon a rich literature in both image-level object detection,video analysis, and recurrent neural networks. Several papers propose ways of using deep convolutional networks for detecting objects. Some approaches classify the proposal regions into object categories and some other recent methods unify the localization and classification stages. Kalogeiton *et al*identifies domain shift factors between still images and videos, necessitating video-specific object detectors. To deal with shift factors and sparse object-level annotations in video, researchers have proposed several strategies. Recently, proposed both transfer learning from the image domain to video frames and optimizing for temporally consistent object proposals. Their approach is capable of detecting both moving and static objects. However, the object proposal generation step that precedes classification is slow. Prest *et al*, utilize weak supervision for object detection in videos via category-level annotations of frames, absent localization ground truth. This method assumes that the target object is moving, outputting a spatio-temporal tube that captures this most salient moving object. This paper, however, does not consider context within video for detecting multiple objects. A few recent papers identify the important role of context in visual recognition. For object detection in images, Bell *et al*use spatial RNNs to harness contextual information, showing large improvements on PASCAL VOC and Microsoft COCO object detection datasets. Their approach adopts proposal generation followed by classification framework. This paper exploits spatial, but not temporal context. Recently, Kang *et al*introduced tubelets with convolutional neural networks (T-CNN) for detecting objects in video. T-CNN uses spatio-temporal tubelet proposal generation followed by the classification and re-scoring, incorporating temporal and contextual information from tubelets obtained in videos. T-CNN won the recently introduced ImageNet object-detection-from-video (VID) task with provided densely annotated video clips. Although the method is effective for densely annotated training data, it's behavior for sparsely labeled data is not evaluated. By modeling video as a time series, especially via GRU or LSTM RNNs, several papers demonstrate improvement on visual tasks including video classification, activity recognition, and human dynamics. These models generally aggregate CNN features over tens of seconds, which forms the input to an RNN. They perform well for global description tasks such as classification but require large annotated datasets. Yet, detecting multiple generic objects by explicitly modeling video as an ordered sequence remains less explored. Our work differs from the prior art in a few distinct ways. First, this work is the first, to our knowledge, to demonstrate the capacity of RNNs to improve localized object detection in videos. The approach may also be the first to refine the object predictions of frame-level models. Notably, our model produces significant improvements even on a small dataset with sparse annotations. # Conclusion We introduce a framework for refining object detection in video. Our approach extracts contextual information from neighboring frames, generating predictions with state of the art accuracy that are also temporally consistent. Importantly, our model benefits from context frames even when they lack ground truth annotations. For the recurrent model, we demonstrate an efficient and effective training strategy that simultaneously employs localization-level strong supervision, category-level weak-supervision, and a penalty encouraging smoothness of predictions across adjacent frames. On a video dataset with sparse object-level annotation, our framework proves effective, as validated by extensive experiments. A subjective analysis of failure cases suggests that the current approach may struggle most on cases when multiple rapidly moving objects are in close proximity. Likely, the sequential smoothness penalty is not optimal for such complex dynamics. Our results point to several promising directions for future work. First, recent state of the art results for video classification show that longer sequences help in global inference. However, the use of longer sequences for localization remains unexplored. We also plan to explore methods to better model local motion information with the goal of improving localization of multiple objects in close proximity. Another promising direction, we would like to experiment with loss functions to incorporate specialized handling of classification and localization objectives.
{'timestamp': '2016-07-20T02:04:30', 'yymm': '1607', 'arxiv_id': '1607.04648', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04648'}
# Introduction The wave-particle duality is an alternative statement of the complementarity principle, and it establishes the relation between nature of quantum entities. It can be illustrated in a two-way interferometer, where the apparatus can be set to observe behavior when a single path is or wave-like behavior, when the impossibility to define a path is shown by the interference. A modern approach to the wave-particle duality includes quantitative relations between quantities that represent the possible *a priori* knowledge of the which-way information () and the "quality" of the interference fringes (). Several publications in the literature contributed to the formulation of the quantitative analysis of the wave-particle duality. For a bipartite system give an extra which-way (path) information about the interferometric possibilities. The quantitative relations for systems composed by two particles were studied in. Therefore, the behavior of such quantities, in various regimes and situations, is essential to answer fundamental and/or technological questions of the quantum theory. an example is the so-called *quantum eraser*, Since its proposal has both theoretically and experimentally (see for example Refs. ). In , the authors explore the quantum eraser problem in model. In this work, an increase of visibility is achieved by performing appropriate projective measurements. An intrinsic relation between the complementarity quantities and the performed measurements is outlined: since were made in order to obtain an of the , the remaining quantities (Entanglement as measured by the concurrence, and the predictability) must obey a "complementary" behavior. In that case, and predictability increases, and entanglement decreases, since the measurements are made in order to the quantum eraser. In Reference only the maximization of the visibility was considered, in the present work we extend the analysis and consider maximization of predictability, visibility and concurrence. Also, in the previous work, only one value of the coupling constant was considered. In this contribution we consider a second coupling regime that allows for the comparison between stronger and weaker interactions. Some questions may arise from the analysis presented in: how is the behavior of the , predictability and Entanglement, for different strengths of the coupling between the cavities and the \(N\) atoms? if one measure the qubits in order to maximize another complementarity quantity? For finite \(N\), could resemble the reservoir (dissipative) limit? Moreover, one can think about a three-part control scheme: initially parts \(A\) and \(B\) possesses a maximally entanglement state, constituted by two qubits \(q_A\) and \(q_B\), respectively. A third part \(R\) may have, in principle, *full* control of a group of \(N\)-qubits (each one we call as \(q_i\)); i.e. \(R\) may control: (i) the initial state of each qubit \(q_i\), (ii) the interaction strength between \(q_i\) and \(q_B\) and (iii) the measurement basis where each \(q_i\) could be projected by \(R\). Here we will focus in the control of item (iii), therefore the initial state of all \(q_i\) and the coupling strenght will be fixed for each realization of the scheme. Thus, part \(R\) is able to control which complementarity quantity of part \(A\) they (\(A\) and \(R\)) would like to maximize. For instance, if \(A\) and \(R\) desire that \(q_A\) is in a superposition state, \(R\) can choose which basis he/she will project each qubit in order to accomplish the task (quantum eraser task ). However, now \(R\) and \(A\) are able to choose another complementarity quantity: if they would like to obtain and/or maintain an Entangled state between \(A\) and \(B\), \(R\) may project each \(q_i\) in a basis chosen in order obtain an state nearly maximally Entangled (the same idea follows for the predictability). More than that, since part \(R\) can adjust the strenght of the interaction between \(q_B\) and \(q_i\), he/she can study what is the best option of coupling to do each task (together with the freedom to choose the basis of projection). In that way, parts \(A\), \(B\) and \(R\) are able to study in details the behavior of the complementarity quantities, for a variety of conditions. In the present work we answer the questions and provide a useful tool to implement the control scheme mentioned above, considering a similar model two entangled qubits, \(q_{A}\) and \(q_{B}\), and a third system (\(R\)) which is composed by \(N\) qubits. It is well known that \(R\) will play the role of a reservoir. As it is possible to measure each qubit of system \(R\) after the interaction, we can control the evolution of \(q_{A}\) and \(q_{B}\), induced by the interaction with \(R\), by selecting an adequate sequence of results of measurements performed in the qubits of \(R\). Such control would allow us to make \(q_{A}\) and \(q_{B}\) approach a chosen asymptotic state. This scheme can be implemented in cavity-QED system, where \(q_{A}\) and \(q_{B}\) would be cavity modes, prepared in an entangled state with one excitation, and \(N\) two level atoms, interacting with the cavities one at the time, would play the role of the qubits that compose the system \(R\). We consider the complementarity quantities concurrence, predictability and to guide the manipulation over \(R\) and to quantify the information present in each subsystem. We also consider two regimes for the strength of the coupling constant \(g\) between \(q_B\) and each qubit of \(R\), \(g= \frac{1}{4}\) and \(g= 4\). The paper is organized as : in section [2](#model){reference-type="ref" reference="model"} we the principal quantities studied: , concurrence and predictability. the distinguishability between different parts of the global system. In subsection [2.1](#digression){reference-type="ref" reference="digression"} we briefly review the case where \(q_A+q_B\) interacts in a dissipative reservoir, and how the complementarity quantities behave in this case. Section [3](#results){reference-type="ref" reference="results"} shows how we implement the projective measurements in \(R\), and present our results and discussions for the behavior of the complementarity quantities and for the variation of distinguishability (after and before the measurements). In section [4](#conclusion){reference-type="ref" reference="conclusion"} we conclude our work. # Model and definitions {#model} Let us consider that initially qubits \(q_A\) and \(q_B\) were prepared in the entangled state \(\ket{\psi(0)} = \frac{1}{\sqrt 2} (\ket{0_A 1_B} + \ket{1_A 0_B})\) and a third system \(R\) composed by , prepared all in the ground state \(\ket 0\). Each qubit of \(R\) interacts one at a time with \(q_B\) and after a sequence of these interactions, the information, initially stored in \(q_A\) and \(q_B\), will be distributed over the and qubits \(q_A\) and \(q_B\). As an example of our interaction model, consider the following dynamics governing the interaction of an atom (between a total of \(N\) atoms) and a cavity (\(q_B\)) (see Figure [\[scheme\]](#scheme){reference-type="ref" reference="scheme"}). The Hamiltonian that gives the interaction between the \(k\)-th atom and \(q_B\) is \(\hat{H}^{(k)}= \omega \hat{b}^{\dagger }\hat{b}+ \frac{\omega}{2}\hat{\sigma}_{z}^{(k)}+ g (\hat{b}^{\dagger}\hat{\sigma}_{-}^{(k)}+\hat{b}\hat{\sigma}_{+}^{(k)}),\) where \(\hat{b}^{\dagger }\) (\(\hat{b}\)) corresponds to the creation (annihilation) operator for \(q_B\), \(\omega\) their transition frequency, \(\hat{\sigma}_{z}^{(k)}=|1^{(k)}\rangle\langle 1^{(k)}|-|0^{(k)}\rangle\langle 0^{(k)}|\), \(\hat{\sigma}_{-}^{(k)}=|0^{(k)}\rangle\langle 1^{(k)}|\), \(\hat{\sigma}_{+}^{(k)}=|1^{(k)}\rangle\langle 0^{(k)}|\), \(g\) the coupling constant for the interaction between the \(k\)-th qubit of \(R\) and \(q_B\). As the initial state has one excitation and the Hamiltonian preserves the excitation number, the states of mode \(B\) can be written in the basis \(\lbrace\vert 0 \rangle, \vert 1 \rangle\rbrace\). Although constant in each preparation, we let the parameter \(g\) free in order to quantify the strength of the interaction, since we will analyze different coupling regimes and the correspondent behavior of the Complementarity quantities. In this model, \(\left|0^{(k)} \right\rangle\) and \(\left|1^{(k)}\right\rangle\) stand for the levels \(0\) and \(1\) of the \(k\)-th interacting atom, respectively. After \(n\) qubits of \(R\) have interacted with \(q_B\) (note the difference between \(N\), the total number of qubits that are able to interact, and \(n\), the number of qubits that will interact at a given time) the global system is left in the state \[\begin{aligned} \ket{\psi^{(n)}} &=& \frac{1}{\sqrt 2} \Big( a^n \ket{0_{A}} \ket{0_{res}} \ket{1_{B}} + \Gamma \ket{0_{A}} \ket{1_{res}} \ket{0_{B}} + \nonumber \\ && + \ket{1_{A}} \ket{0_{res}} \ket{0_{B}} \Big), \label{psin}\end{aligned}\] where \(a = \cos\left( \frac{g T}{N} \right)\) and \(b =-i \sin \left( \frac{g T}{N} \right)\), assuming the same interaction time between each qubit of \(R\) and \(q_{B}\), given by \(\Delta t=\frac{T}{N}\). To simplify the notation we define a normalized state with one excitation in subsystem \(R\): \(\ket{1_{res}}= \frac{1}{\Gamma} \left( a^{n-1} b \ket{n} + \ldots + a b \ket{2} + b\ket{1} \right)\) \(\Gamma^{2} = 1-a^{2n}\) \(\ket{i}= \ket{0_1 0_2 \ldots 0_{i-1} ~1_i ~ 0_{i+1} \ldots 0_n}\) represents a state with an excitation in the \(i\)-th qubit and \(\ket{0_{res}} = \ket{0_1 0_2 \ldots 0_n}\). For a general pure two qubit state: \(\ket{\Psi} = \gamma_1 \ket{00} + \gamma_2 \ket{01} + \gamma_3 \ket{10} + \gamma_4 \ket{11},\) the complementary quantities are defined in the following way. The concurrence, which is related to the quantum correlation between the parts, is given by \(C(\ket{\Psi}) = 2 \abs{\gamma_1 \gamma_4-\gamma_2 \gamma_3}\). The coherence between two orthogonal states gives the visibility, defined by \(V = 2 \abs{\bra{\Psi}\sigma_{+}\ket{\Psi}} = 2 \abs{\gamma_1 \gamma_3^* + \gamma_2 \gamma_4^*}\). Besides, the predictability measures the knowledge if one of the parts is in state \(\ket{0}\) or \(\ket{1}\), \(P= \abs{\bra{\Psi}\sigma_{z}\ket{\Psi}} = \abs{\abs{\gamma_3}^2+\abs{\gamma_4}^2-\left( \abs{\gamma_1}^2 + \abs{\gamma_2}^2 \right)}\). The distinguishability, or the "measure of the possible which-path information that one can obtain" in an interferometer setup, is given by \(D = \sqrt{C^2 + P^2}\). ## Continuous Limit-A digression {#digression} Defining \(k = g^2 \frac{T}{N},\) one can write \(a = \cos \left(\sqrt{\frac{k T}{N}} \right),\) where \(T\) is the total time of interaction between \(q_B\) and \(R\) . We consider that the interaction time between each qubit of \(R\) and \(q_B\) is equal, given by \(\Delta t=\frac{T}{N}\). We also assume that \(0<\Delta t<\pi/2g.\) After \(N\) interactions the reduced state in the subsystem \(q_A\) is a statistical mixture \(\rho_{A}=\frac{1}{2}\left(\vert 0_{A} \rangle \langle 0_{A} \vert + \vert 1_{A} \rangle \langle 1_{A} \vert\right)\), therefore \(V_{q_A} = 0\) and \(P_{q_A} = 0\). The concurrence can be calculated from the reduced state of the subsystem \(q_A, q_B\) and is given by \(C_{q_A, q_B} = a^N\). The limit \(N \rightarrow \infty\) (and consequently \(g \rightarrow \sqrt{\frac{N}{T}} \rightarrow \infty\); \(\Delta t \rightarrow dt\)) is well known in the Literature and it gives the reservoir limit (at a given temperature implicitly defined in \(k\)) of a qubit interacting with a Markovian pure dissipative reservoir. The term \(a^n\) in [\[psin\]](#psin){reference-type="eqref" reference="psin"} is in this limit \(\lim_{N \rightarrow \infty} a^N = \lim_{N \rightarrow \infty} \cos^N \sqrt{\frac{k T}{N}} = \mathrm{e}^{-kT/2}\), and consequently the concurrence \(C_{q_A, q_B} = \mathrm{e}^{\frac{-k T}{2}}\) decays exponentially with \(T\). # Results ## Complementarity quantities versus coupling intensity Similar to what was done in Ref. , let us now consider that, after \(n\) interactions, the \(i\)-th qubit of \(R\) is projected in the state: \(\ket{M_i} = \alpha_i \ket{0_i} + \beta_i \ket{1_i},\) where \(\alpha_i = \cos \theta_i\) e \(\beta_i = \mathrm{e}^{i \phi_i} \sin \theta_i\) (this measure can be done experimentally, see for example ). The vector state \(\ket{M_i}\) is an eigenstate of the operator with \(\vec{n}=\left(\sin2\theta_{i} ~ \cos2\phi_{i},~ \sin2\theta_{i}~\sin2\phi_{i},~\cos2\theta_{i}\right)\) and , the Pauli matrices. One can, in principle, choose in which base (\(\theta_i\) and \(\phi_i\)) the global state will be measured. projective measurements performed on the state [\[psin\]](#psin){reference-type="eqref" reference="psin"}, the projector is given by \(\Pi = \Pi_1 \otimes \ldots \otimes \Pi_n,\) with \(\Pi_i = \mathbb{I}_1 \otimes \ldots \otimes \mathbb{I}_{i-1} \otimes \ket{M_i}\bra{M_i} \otimes \ldots \mathbb{I}_n.\) Notice that the projective measure \(\Pi\) acts only on the subsystem \(R\). After \(n\) projective measurements the normalized global state vector is given by: \[\begin{aligned} \ket \psi^{(n,M)} &=& \frac{1}{N}\Big( \gamma_1 \ket{0_A} \ket M \ket{0_B} + \gamma_2 \ket{0_A} \ket{M} \ket{1_B} \nonumber \\ &&+ \gamma_3 \ket{1_A} \ket M \ket{0_B} \Big), \label{rhoreducedM}\end{aligned}\] where \(\ket M = \ket{M_1} \ldots \ket{M_n}\), \(N = \sqrt{\abs{\gamma_1}^2 + \abs{\gamma_2}^2 + \abs{\gamma_3}^2},\) and \[\begin{aligned} \gamma_1 &=& \frac{1}{\sqrt 2} \left(b \sum_{i = 1}^{n} \left[ a^{i-1} \frac{\beta_i}{\alpha_i} \left( \prod_{j=1}^{n} \alpha_j \right) \right] \right),\nonumber \\ \gamma_2 &=& \frac{1}{\sqrt 2} \left(a^n \prod_{i=1}^{n} \alpha_i \right), \nonumber \\ \gamma_3 &=& \frac{1}{\sqrt 2} \left( \prod_{i=1}^{n} \alpha_i \right).\end{aligned}\] The information carried by the qubits of \(R\) are now embodied in the measurement outcomes \(\theta_i\) and \(\phi_i\). The complementarity quantities after the measurements are given by: \[\begin{aligned} V^{(n,M)}_{q_{A}} &=& \frac{2 \abs{\gamma_1 \gamma_3^*}}{N^2} \nonumber \\ P^{(n,M)}_{q_{A}} &=& \frac{\abs{\abs{\gamma_3}^2-\abs{\gamma_1}^2-\abs{\gamma_2}^2}}{N^2} \nonumber \\ C^{(n,M)}_{q_{A},q_{B}} &=& \frac{2 \abs{\gamma_2 \gamma_3}}{N^2}. \nonumber \label{complementarity} \end{aligned}\] Since the reduced state is pure [\[rhoreducedM\]](#rhoreducedM){reference-type="eqref" reference="rhoreducedM"}, the closure relation for complementarities : \[\left(C^{(n,M)}_{q_{A},q_{B}}\right)^2 +\left(P^{(n,M)}_{q_{A}}\right)^2 +\left(V^{(n,M)}_{q_{A}}\right)^2 = 1. \label{complem1}\] quantities depend explicit on the coefficients \(\theta_i\) and \(\phi_i\) of \(\ket M_i\). Concerning the Complementarity quantities, one can project the global state so that \(V^{(n,M)}_{q_{A}}, P^{(n,M)}_{q_{A}}\) or \(C^{(n,M)}_{q_{A}, q_B}\) acquire the maximum allowed values, after \(n\) measurements on the qubits of \(R\) have interacted with \(q_B\). In , the authors studied a similar maximization procedure, only for the \(V^{(n,M)}_{q_{A}}\) to produce a multipartite quantum eraser the coefficients \(\alpha_i\) and \(\beta_i\) were chosen to obtain an in the , maintaining a standard value for the coupling parameter (\(g T = 2 \pi\)). Here we are interested in how each of the Complementarity quantities behaves, between \(q_B\) and the \(i\)-th of \(R\), while making projective measurements in each qubit of \(R\). if the function to be maximized is the concurrence \(C^{(n,M)}_{q_{A}, q_B}\), for example, the procedure gives the values of \(\theta_i\) and \(\phi_i\) that provides the maximum value of \(C^{(n,M)}_{q_{A}, q_B}\), after \(n\) qubits have interacted with \(q_B\); then, in the possession of \(\theta_i\) and \(\phi_i\), one can evaluate \(V^{(n,M)}_{q_{A}}\) and \(P^{(n,M)}_{q_{A}}\). The same procedure is carried out in order to maximize the or the predictability. Therefore, we have all the Complementarity quantities for each function to be maximized: \(V^{(n,M)}_{q_{A}}, P^{(n,M)}_{q_{A}}\) or \(C^{(n,M)}_{q_{A}, q_B}\). In Figures [\[maxvis\]](#maxvis){reference-type="ref" reference="maxvis"}, [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"} and [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"} we show the Complementarity quantities, for the maximization of \(V^{(n,M)}_{q_{A}}, P^{(n,M)}_{q_{A}}\) and \(C^{(n,M)}_{q_{A}, q_B}\), respectively. We also present the behaviour for different couplings (each Figure (a) depicts \(g = 4\) and each Figure (b) \(g = \frac{1}{4}\)). The solid curve is related to the \(\ankb{Visibility}{visibility}\); the same follows for the \(concurrence\) ( dashed curve) and the \(predictability\) (black dotted curve). Note in Figures [\[maxvis\]](#maxvis){reference-type="ref" reference="maxvis"}a and [\[maxvis\]](#maxvis){reference-type="ref" reference="maxvis"}b, the is increasing function of \(n\), when the measurements are made to maximize \(V^{(n,M)}_{q_{A}}\) if one measurements in order to maximize another Complementarity quantity, the does not increase, and remains zero, as one can see in Figures [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"} and [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}. [Differently from the visibility, always when the predictability P is maximized it achieves the maximum value for some finite value of \(n\) (Figures [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"}a and [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"}b). Moreover, it achieves a maximum valeu also when other quantities are maximized (Figure [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}a). Using an interferometric analogy, predictability is the which-way information that is available in the interferometric system, while gives the quality of the interference pattern and concurrence is a measure of entanglement between \(q_a\) (main system) and \(q_B\) (which-way detector). Therefore, in Figures [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"}a, [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"}b and [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}a, since the system loses entanglement, with no acquisition of visibility, the predictability must increase.]{style="color: 0.00,0.00,0.00"} This feature can be understood in our approach, since the projective measurements will inherently modify the global system. In Table [\[table\]](#table){reference-type="ref" reference="table"} we show the states after \(n\) interactions (\(n = 1, 2\) and \(10\)), and after performing the maximization procedures. For instance, performing a maximization of \(P_{q_A}^{(n,M)}\), the state \(\ket{\psi}^{(n,M)}\) for \(g T = 2 \pi \times 4\) tends to the state \(\ket{1_A 0_B}\). [The dashed curves in Figures [\[maxvis\]](#maxvis){reference-type="ref" reference="maxvis"}, [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"} and [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"} show the concurrence as a function of \(n\). For \(g T = \frac{2 \pi}{4}\), if the function to be maximized is the concurrence itself \(C^{(n,M)}_{q_{A}, q_B}\) (Figure [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}b) , it is possible to maintain the state almost maximally Entangled--\(C^{(n,M)}_{q_{A}, q_B} \sim 1\)--by choosing the proper values of \(\theta_i\) and \(\phi_i\). However, if the coupling is increased--Figures [\[maxvis\]](#maxvis){reference-type="ref" reference="maxvis"}a, [\[maxpre\]](#maxpre){reference-type="ref" reference="maxpre"}a, or [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}a--even if the projective measures were made to maximize the concurrence (Figure [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}a), Entanglement decreases to zero. This behavior is similar to two entangled qubits, where \(q_B\), is coupled to a thermal reservoir (red solid curves), but in our case we have a finite number of interacting qubits with \(q_B\) (where the maximum number of interacting qubits is \(N = 20\)). how the initial information (given by the distinguishability between \(q_A\) and the \(i\)-th qubit) is distributed over the \(N\) qubits . An interesting aspect concerning Figure [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}b, for \(g T = \frac{2 \pi}{4}\), is that one can see an approximately steady behavior of the concurrence \(C^{(n,M)}_{q_{A}, q_B}\), near the initial value \(C^{(0)}_{q_{A}, q_B} = 1\). It is possible, therefore, to maintain the system \(q_A + q_B\) in an approximately maximal Entangled state, notwithstanding the qubits of \(R\) became dynamically correlated with \(q_A\) and \(q_B\) (Equation [\[rhon\]](#rhon){reference-type="eqref" reference="rhon"}). One can see from Table [\[table\]](#table){reference-type="ref" reference="table"}, for \(g T = \frac{2 \pi}{4}\) and \(n=10\), the state resemble the initial maximally entangled state corroborating Figure [\[maxcon\]](#maxcon){reference-type="ref" reference="maxcon"}b.]{style="color: 0.00,0.00,0.00"}
{'timestamp': '2017-01-26T02:07:48', 'yymm': '1607', 'arxiv_id': '1607.04617', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04617'}
null
null
null
null
null
null
# Introduction {#sec:intro} Compact stars have a large number of pulsation modes that have been extensively studied since the seminal work of Chandrasekhar on radial oscillations. In general, these modes are very difficult to observe in the electromagnetic spectrum; therefore most efforts have concentrated on gravitational wave asteroseismology in order to characterise the frequency and damping times of the modes that emit gravitational radiation. In particular, various works focused on the oscillatory properties of pure hadronic stars, hybrid stars and strange quark stars trying to find signatures of the equation of state of high density neutron star matter (see and references therein). More recently, compact star oscillations have attracted the attention in the context of Soft Gamma ray Repeaters (SGRs), which are persistent X-ray emitters that sporadically emit short bursts of soft \(\gamma\)-rays. In the quiescent state, SGRs have an X-ray luminosity of \(\sim 10^{35}\) erg/s, while during the short \(\gamma\)-bursts they release up to \(10^{42}\) erg/s in episodes of about 0.1 s. Exceptionally, some of them have emitted very energetic giant flares which commenced with brief \(\gamma\)-ray spikes of \(\sim 0.2\) s, followed by tails lasting hundreds of seconds. Hard spectra (up to 1 MeV) were observed during the spike and the hard X-ray emission of the tail gradually faded modulated at the neutron star (NS) rotation period. The analysis of X-ray data of the tails of the giant flares of SGR 0526-66, SGR 1900+14 and SGR 1806-20 revealed the presence of quasi-periodic oscillations (QPOs) with frequencies ranging from \(\sim\) 18 to 1840 Hz. There are also candidate QPOs at higher frequencies up to \(\sim 4\) kHz in other bursts but with lower statistical significance; in fact, according to a more recent analysis only one burst shows a marginally significant signal at a frequency of around 3706 Hz. Several characteristics of SGRs are usually explained in terms of the *magnetar* model, assuming that the object is a neutron star with an unusually strong magnetic field (\(B \sim 10^{15}\) G). In particular, giant flares are associated to catastrophic rearrangements of the magnetic field. Such violent phenomena are expected to excite a variety of oscillation modes in the stellar crust and core. In fact, recent studies have accounted for magnetic coupling between the crust and the core, and associate QPOs to global magneto-elastic oscillations of highly magnetized neutron stars. There has also been interest in the possible excitation of low order \(f\)-modes because of their strong coupling to potentially detectable gravitational radiation. In the present paper we focus on radial oscillations of neutron stars permeated by ultra-strong magnetic fields. These modes might be relevant within the magnetar model because they could be excited during the violent events associated with gamma flares. Since they have higher frequencies than the already known QPOs, they cannot be directly linked to them at present. However, it is relevant to know all the variety of pulsation modes of strongly magnetized neutron stars because the number of observations is still small and new features could emerge in future flares' data. On the other hand, in the case of rotating objects we can expect some amount of gravitational radiation from even the lowest (\(l = 0\)) quasi-radial mode making them potentially relevant for gravitational wave astronomy. # Equations of state {#sec2} ## Hadronic phase under a magnetic field {#sec:A} In this section we present an overview of the hadronic equations of state (EOS) used in this work. We describe hadronic matter within the framework of the relativistic non-linear Walecka (NLW) model. In this model we employ a field-theoretical approach in which the baryons interact via the exchange of \(\sigma-\omega-\rho\) mesons in the presence of a magnetic field \(B\) along the \(z-\)axis. The total lagrangian density reads: \[\label{lt} \mathcal{L}_{H}=\sum_{b}\mathcal{L}_{b}+\mathcal{L}_{m}+\sum_{l}\mathcal{L}_{l}+\mathcal{L}_B\.\] where \(\mathcal{L}_{b}\), \(\mathcal{L}_{m}\), \(\mathcal{L}_{l}\) and \(\mathcal{L}_{B}\) are the baryons, mesons, leptons and electromagnetic field Lagrangians, respectively, and are given by \[\begin{aligned} \mathcal{L}_{b} &=& \overline{\psi}_{b}\left(i\gamma_{\mu}\partial^{\mu}-q_{b}\gamma_{\mu}A^{\mu}-m_{b}+g_{\sigma b}\sigma \right. \nonumber \\ && \left.-g_{\omega b}\gamma_{\mu}\omega^{\mu}-g_{\rho b}\tau_{3b}\gamma_{\mu}\rho^{\mu}\right)\psi_{b}\,, \\ \mathcal{L}_{m} &=& \tfrac{1}{2}(\partial_{\mu}\sigma\partial^{\mu}\sigma-m_{\sigma}^{2}\sigma^{2})-U(\sigma)+ \tfrac{1}{2}m_{\omega}^{2}\omega_{\mu}\omega^{\mu} \nonumber \\ &&-\tfrac{1}{4}\Omega_{\mu\nu}\Omega^{\mu\nu}+ \tfrac{1}{2}m_{\rho}^{2}\vec{\rho}_{\mu}\cdot\vec{\rho}_{\mu}-\tfrac{1}{4}P^{\mu\nu}P_{\mu\nu} \,, \\ \mathcal{L}_{l} &=& \overline{\psi}_{l}\left(i\gamma_{\mu}\partial^{\mu}-q_{l}\gamma_{\mu}A^{\mu}-m_{l}\right)\psi_{l} \,, \\ \mathcal{L}_{B} &=&-\tfrac{1}{4}F^{\mu\nu}F_{\mu\nu} \. \end{aligned}\] where he \(b\)-sum runs over the baryonic octet \(b\equiv N~(p,~n),~\Lambda,~\Sigma^{\pm,0},~\Xi^{-,0}\), \(\psi_{b}\) is the corresponding baryon Dirac field, whose interactions are mediated by the \(\sigma\) scalar, \(\omega_{\mu}\) isoscalar-vector and \(\rho_{\mu}\) isovector-vector meson fields. The baryon charge, baryon mass and isospin projection are denoted by \(q_{b}\), \(m_{b}\) and \(\tau_{3b}\), respectively, and the masses of the mesons are \(m_{\sigma}= 512~\) MeV, \(m_{\omega}=783~\)MeV and \(m_{\rho}=770~\)MeV. The strong interaction couplings of the nucleons with the meson fields are denoted by \(g_{\sigma N}=8.910\), \(g_{\omega N}=10.610\) and \(g_{\rho N}=8.196\). We consider that the couplings of the hyperons with the meson fields are fractions of those of the nucleons, defining \(g_{iH}=X_{iH}g_{iN}\), where the values of \(X_{iH}\) are chosen as \(X_{\sigma H}=0.700\) and \(X_{\omega H}=X_{\rho H}=0.783\). The term \(U(\sigma)=\frac{1}{3}\,bm_{n}(g_{\sigma N}\sigma)^{3}-\frac{1}{4}\,c(g_{\sigma N}\sigma)^{4}\) denotes the scalar self-interactions, with \(c=-0.001070\) and \(b=0.002947\). The mesonic and electromagnetic field tensors are given by their usual expressions \(\Omega_{\mu\nu}=\partial_{\mu}\omega_{\nu}-\partial_{\nu}\omega_{\mu}\), \({\bf P}_{\mu\nu}=\partial_{\mu}\vec{\rho}_{\nu}-\partial_{\nu}\vec{\rho}_{\mu}-g_{\rho b}(\vec{\rho}_{\mu}\times\vec{\rho}_{\nu})\) and \(F_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}\). The \(l\)-sum runs over the two lightest leptons \(l\equiv e,\mu\) and \(\psi_{l}\) is the lepton Dirac field. The symmetric nuclear matter properties at saturation density adopted in this work are given by the GM1 parametrization, with compressibility \(K=300\) MeV, binding energy \(B/A=-16.3\) MeV, symmetry energy \(a_{sym}=32.5\) MeV, slope \(L=94\) MeV, saturation density \(\rho_{0}= 0.153\) fm\(^{-3}\) and nucleon mass \(m=938\) MeV. The following equations present the scalar and vector densities for the charged and uncharged baryons, respectively: \[\begin{aligned} &\label{densities1}\rho_{b}^{s}=\frac{|q_{b}|B \bar{m}_{b} }{2\pi^{2}}\sum_{\nu}^{\nu_{\mathrm{max}}}\sum_{s}\frac{\bar{m}_{b}^{c}}{\sqrt{ \bar{m}_{b}^2 + 2\nu |q_{b}|B}} \ln\bigg|\frac{k_{F,\nu,s}^{\,b}+E_{F}^{\,b}}{\bar{m}_{b}^{c}} \bigg|,\\ &\label{densities2}\rho_{b}^{v}=\frac{|q_{b}|B}{2\pi^{2}}\sum_{\nu}^{\nu_{\mathrm{max}}}\sum_{s}k_{F,\nu,s}^{\,b},\\ &\label{densities3}\rho_{b}^{s}=\frac{\bar{m}_{b}}{4\pi^{2}}\sum_{s}\bigg[E_{F}^{\,b}k_{F,s}^{\,b}-\bar{m}_{b}^{2}\ln\bigg|\frac{k_{F,s}^{\,b}+E_{F}^{\,b}}{\bar{m}_{b}}\bigg|\bigg] ,\\ &\label{densities4}\rho_{b}^{v}=\frac{1}{2\pi^{2}}\sum_{s}\bigg[\frac{1}{3}(k_{F,s}^{\,b})^{3}\bigg], \end{aligned}\]
{'timestamp': '2016-07-19T02:02:47', 'yymm': '1607', 'arxiv_id': '1607.04707', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04707'}
# Introduction Ultra-cold molecules can play important roles in studies of many-body quantum physics , quantum logic operations , and ultra-cold chemistry . In our recent studies of LiRb, motivated largely by its large permanent dipole moment , we have explored the generation of these molecules in a dual species MOT . In particular, we have found that the rate of generation of stable singlet ground state molecules and first excited triple state molecules through photoassociation, followed by spontaneous emission decay, can be very large . There have been very few experimental studies of triplet states in LiRb , in part because they are difficult to access in thermally-distributed systems. Triplet states of bi-alkali molecules are important to study for two reasons: first, Feshbach molecules, which are triplet in nature, provide an important association gateway for the formation of stable molecules ; also, photoassociation (PA) of trapped colliding atoms is often strongest for triplet scattering states. Mixed singlet-triplet states are usually required to transfer these molecules to deeply bound singlet states. We show an abbreviated set of potential energy curves (PEC), as calculated in Ref. , in Fig. [\[fig:PEC\]](#fig:PEC){reference-type="ref" reference="fig:PEC"}. The d\(^3 \Pi\)-D\(^1 \Pi\) complex in LiRb, asymptotic to the Li 2p \(^2P_{3/2, 1/2}\) + Rb 5s \(^2S_{1/2}\) free atom state, has several features that can promote its utility in stimulated-Raman-adiabatic-passage (STIRAP) and photoassociation. First, the *ab inito* calculations of Ref.  predict mixing between low vibrational levels of the \(d \ ^3\Pi_1\) and the D\(^1 \Pi\) states. Second, both legs of a STIRAP transfer process from loosely bound triplet-character Feshbach molecules to the rovibronic ground state can be driven with commercially-available diode lasers. And third, similar deeply bound \(^3 \Pi\) resonances have been successfully used for short-range PA in RbCs  . While an interesting discovery on its own, spontaneous decay of these states after PA can populate the \(a \ ^3\Sigma^+ \ v^{\prime \prime}=0\) state; one RbCs team  found spontaneous decay of these states even populated the \(X \ ^1\Sigma^+ \ v^{\prime \prime}=0\) state. In the present work, we study the \(d \ ^3\Pi_{\Omega}\) states of LiRb, from the asymptote to the most bound vibrational level. We have found signatures of state mixing between low-lying vibrational levels of the d\(^3 \Pi_1\) and D\(^1 \Pi\) levels. We have determined the term energies of the different spin-orbit components of the \(d \ ^3\Pi\) state, as well as their vibrational energies. We have also observed alternation of the intensities of the rotational lines, a possible indication of a \(p\)-wave resonance in the scattering state, and determined the rotational constants of the lowest vibrational levels. # Experiment We have previously described the details of our experimental apparatus , and provide here only a brief summary. We trap \(\sim 5 \times 10^7\) Li atoms and \(\sim 2 \times 10^8\) Rb atoms in a dual species magneto-optical trap (MOT), \(\lesssim\)`<!-- -->`{=html}1 mK in temperature and 1 mm in diameter . Our Rb MOT is a spatial dark spot MOT . We photoassociate Li and Rb atoms to form LiRb molecules using either a 300 mW cw Ti:Sapphire laser or a 150 mW cw external cavity diode laser. After spontaneous decay to a distribution of vibrational levels of the \(a \ ^3\Sigma^+\) state, we use two-color resonantly-enhanced-multi-photon-ionization (RE2PI) to ionize the molecules. The lasers that we use to drive the RE2PI process are a Nd:YAG-pumped, pulsed dye laser (PDL) for the first photon, tunable in the wavelength range between 667 nm-750 nm (14950-13300 cm\(^{-1}\) frequency range), and part of the 532 nm pump laser for the second photon. The repetition rate of this system is 10 Hz, and it delivers \(\sim\)`<!-- -->`{=html}1.5 mJ/pulse of dye energy in a 4 mm diameter beam and a larger \(\sim\)`<!-- -->`{=html}2 mJ/pulse 532 nm beam to the MOT region. When the frequency of the dye laser is resonant with a transition from an initial state (populated through spontaneous decay from the PA state) to an intermediate bound state (in this experiment the \(d \ ^3\Pi\) state), absorption of a dye photon and a 532 nm photon ionizes the molecule. We detect the molecular LiRb\(^+\) ions using a time-of-flight spectrometer and a microchannel plate detector. In this paper we will adopt the following notation: \(v^{\prime \prime}\) and \(J^{\prime \prime}\) denote the vibrational and rotational levels of the \(a \: ^3 \Sigma ^+\) and \(X \ ^1\Sigma^+\)states, \(v\) and \(J\) (without a prime) denote the vibrational and rotational levels of the PA resonances (and for these vibrational numbers, we count down from the asymptote using negative integers), and \(v^{\prime}\) and \(J^{\prime}\) denote vibrational and rotational labeling of other excited electronic states. We used two techniques, RE2PI and depletion spectroscopy, to measure the \(d \ ^3\Pi\) bound states. In RE2PI spectroscopy, we tune the PA laser to either the \(v=-11 \ J=1\) or \(v=-8 \ J=1\) lines of the \(2(0^-)\) long range state, from which spontaneous decay leads primarily to the vibrational levels of the \(a \ ^3\Sigma^+\) state . We count the number of ions detected over the course of 100 laser pulses, and tune the laser frequency \(\nu_c\) of the PDL in 0.35 cm\(^{-1}\) increments. We record the number of ions detected, normalized by the number of laser pulses, as a function of the PDL frequency \(\nu_c\). In order to reach the full range of vibrational levels of the \(d \ ^3\Pi\) states, we used two different laser dyes in the PDL. An LDS 698 dye covered the 13950-14950 cm\(^{-1}\) range, and an LDS 750 dye covered from 13300 to 13950 cm\(^{-1}\). These dyes are difficult to work with because of short lifetimes and low power output. The LDS 750 dye in particular was very troublesome: it has a lifetime \(\le\)`<!-- -->`{=html}8 hours, produces low power (\(\le\)`<!-- -->`{=html}0.5 mJ/pulse for much of its range) and because it has a very broad pulse width (i.e. lots of spontaneous emission) the baseline noise of our RE2PI spectra is enhanced over what we have observed with other dyes. The \(2(0^-) \ v=-11 \ J=1\) PA line at \(\nu_a = 12516.89\) cm\(^{-1}\) is relatively weak, but it decays almost exclusively to a single vibrational level (\(v^{\prime \prime}=11\)) of the \(a \ ^3\Sigma^+\) state. This facilitates straight-forward identification of the vibrational levels of the intermediate state. Unfortunately, several vibrational levels of the \(d \ ^3\Pi\) state do not appear in this spectrum, presumably due to poor Franck-Condon overlap with the \(v^{\prime \prime}=11\) state. This problem was even more evident when using the LDS 750 dye in the PDL. For this reason, we collected several RE2PI spectra using the stronger \(2(0^-) \ v=-8 \ J=1\) PA resonance \(\nu_a = 12557.60\) cm\(^{-1}\). This line decays to a wider spread of vibrational levels, giving more complete coverage of the vibrational lines of the \(d \ ^3\Pi\) states, but also making our analysis more difficult, due to the increased congestion of the spectra and frequent overlap between individual lines. To explore the deeply-bound levels of the \(d \ ^3\Pi\) states, we used a depletion spectroscopy technique. In these measurements, we used the 150 mW ECDL tuned to the \(2(0^-) \ v=-5 \ J=1\) PA resonance at \(\nu_a = 12575.05\) cm\(^{-1}\) . Spontaneous decay of this state populates the \(a \ ^3\Sigma^+ \ v^{\prime \prime}=13\) state. We tune the PDL laser frequency to the \((3) \ ^3\Pi_{0} \ v^{\prime}=6 \leftarrow a \ ^3\Sigma^+ \ v^{\prime \prime}=13\) one-color resonantly-enhanced two-photon ionization (REMPI) transition at 17736.6 cm\(^{-1}\) . We then tune the frequency of the Ti:Sapphire laser into resonance with bound-to-bound transitions from the \(a \ ^3\Sigma^+ \ v^{\prime \prime}=13\) state to ro-vibrational levels in the \(d \ ^3\Pi\) state. Exciting these transitions depletes the population of the \(a \ ^3\Sigma^+ \ v^{\prime \prime}=13\) state, causing the REMPI signal to decrease. # RE2PI Measurements We show an example of a RE2PI spectrum in Fig. [\[fig:v11progression\]](#fig:v11progression){reference-type="ref" reference="fig:v11progression"}. Transitions observed in this spectrum are \(d \ ^3\Pi_{\Omega} \ v^{\prime} \leftarrow a \ ^3\Sigma^+ \ v^{\prime \prime}=11\). We have marked the transitions to the \(\Omega=2\) progression with black solid lines, \(\Omega=1\) with blue dashed lines, and \(\Omega=0\) with green dot-dashed lines. \(\Omega\) is the total electronic angular momentum, orbital \(L\) + spin \(S\), projected onto the internuclear axis. The numerical label for each peak is the vibrational number \(v^{\prime}\) of the \(d \ ^3\Pi_{\Omega}\) state. We have also marked three lines in this spectrum corresponding to transitions to the \(D \ ^1\Pi\) state with red dotted lines. From the spectrum of these \(d \ ^3\Pi\) states, we observe the typical hierarchy of line spacings: the vibrational splitting is large (on the order of 100 cm\(^{-1}\) for low vibrational quantum number \(v^\prime\), and decreasing with increasing \(v^\prime\)), and the spin-orbit splitting between different \(\Omega\) states is smaller (on the order of 30 cm\(^{-1}\)). The rotational splitting for low \(J^\prime\) (on the order of 0.1 cm\(^{-1}\)) is too small to be resolved in these RE2PI spectra since the spectral resolution of the PDL is \(\sim\)`<!-- -->`{=html}0.5 cm\(^{-1}\). The appearance of transitions belonging to vibrational levels of the \(D \ ^1\Pi\) electronic state in Fig. [\[fig:v11progression\]](#fig:v11progression){reference-type="ref" reference="fig:v11progression"} is evidence of mixing between the \(D \ ^1\Pi\) and the \(d \ ^3\Pi_{1}\) potentials near an avoided crossing between the two states. The energy of these \(D \ ^1\Pi\) states is known from Refs. . State mixing gives these states partial character of each electronic state, which in this case manifests itself through strong transitions from a triplet state (i.e. \(a \ ^3\Sigma^+ \ v^{\prime \prime}=11\)) to singlet states (\(D \ ^1\Pi \ v^{\prime}\) = 6, 7, and 8). This state mixing also adds \(D ^1\Pi\) character to the \(d \ ^3\Pi_{1}\) states, so one should expect the nearby \(d \ ^3\Pi_{1}\) states to appear in the singlet spectra. This expectation is borne out in the spectrum shown in Fig. [\[fig:v10progression\]](#fig:v10progression){reference-type="ref" reference="fig:v10progression"}. This spectrum is a REMPI scan generated in our system after photoassociating ultracold LiRb molecules through a 2(1)-4(1) mixed state at \(\nu_a = 12574.85\) cm\(^{-1}\) , which spontaneously decays to vibrational levels of the \(X \ ^1\Sigma^+\) ground electronic state. The spectrum in Fig. [\[fig:v10progression\]](#fig:v10progression){reference-type="ref" reference="fig:v10progression"} primarily shows transitions to low-lying \(D \ ^1\Pi\) vibrational levels from \(X \ ^1\Sigma^+ \ v^{\prime \prime}=10\). We also observe in this spectrum \(d \ ^3\Pi_{1} \ v^{\prime}=4\), 5, and 6 \(\leftarrow X \ ^1\Sigma^+ \ v^{\prime \prime}=10\) transitions. We chose \(X \ ^1\Sigma^+ \ v^{\prime \prime}=10\) because it is strongly populated by spontaneous decay of the 2(1)-4(1) PA resonance and transitions to deeply bound \(D \ ^1\Pi\) vibrational levels are clearly identified. We can estimate the degree of mixing based on the relative strength of the different REMPI peaks. The \(D \ ^1\Pi \ v^\prime = 7 \leftarrow X \ ^1\Sigma^+ \ v^{\prime \prime}=10\) transition is twice as strong as the \(d \ ^3\Pi_1 \ v^\prime = 5 \leftarrow X \ ^1\Sigma^+ \ v^{\prime \prime}=10\) transition so there is twice as much singlet character to \(D \ ^1\Pi \ v^\prime = 7\) as \(d \ ^3\Pi_1 \ v^\prime = 5\). Following the same procedure that we have used in the past , we can estimate the interaction strength to be about 7 cm\(^{-1}\). Interestingly, this rough estimate is consistent with the following simple perturbative argument. The spin-orbit interaction responsible for the state mixing can be estimated as about one half the spin orbit mixing in atomic rubidium , or about 120 cm\(^{-1}\). The Franck-Condon factor (FCF) between \(D \ ^1\Pi \ v^\prime = 7\) and \(d \ ^3\Pi_1 \ v^\prime = 5\), as calculated by LEVEL 8.0  using the PEC from , is about 0.08 and thus the strength of interaction between these states should be approximately 10 cm\(^{-1}\). We have applied this perturbative analysis to each of the vibrational levels of the \(d \ ^3\Pi_{1}\) state (not including \(v^{\prime}=5\)), and find that each contains some small component of \(D \ ^1\Pi\) perturber state, on the order of 10% or smaller. This is too small to be seen in the spectra of Fig. [\[fig:v11progression\]](#fig:v11progression){reference-type="ref" reference="fig:v11progression"}, but could be sufficient to be useful in a Raman or STIRAP transfer of population to low lying levels of the electronic ground state in the future. Many of the RE2PI spectra that we collected are less clear than that shown in Fig. [\[fig:v11progression\]](#fig:v11progression){reference-type="ref" reference="fig:v11progression"}. In particular, the peaks in the RE2PI spectra near the Rb 5S + Li 2P asymptote are strong, but line congestion becomes significant, and clear identification of the lines in this region becomes difficult. These assignments could probably be improved using a spectroscopic technique that is capable of higher spectral resolution, such as photoassociative spectroscopy, but this was beyond the scope of the present work. Assigning peaks in the RE2PI spectra was equally difficult for deeply bound vibrational states (i.e. \(v^{\prime}\) \(\le\) 4). In fact we were unable to observe a clear cut off in our RE2PI data corresponding to \(v^{\prime}=0\). We attribute this to difficulties with the LDS 750 dye, specifically the large spontaneous emission content in the pulse. To rectify this problem, we turned to a second set of measurements, based upon depletion spectroscopy. # Depletion Measurements We used depletion spectroscopy to identify the lowest two vibrational levels of the \(d \ ^3\Pi\) state. Compared to RE2PI, depletion spectra are sparse and have narrow peak widths, in this case \(\sim\)`<!-- -->`{=html}1 GHz, a typical linewidth for this type of measurement at this intensity . Additionally, depletion spectra allow us to extract some rotational constants of the lines, a very useful tool for comparing experiment to theory. We show depletion spectra for \(v^{\prime}=0\) in Fig. [\[fig:depletion\]](#fig:depletion){reference-type="ref" reference="fig:depletion"}. To assign these data care must be taken with selection rules for radiative transitions in molecules. The two that apply here are: \(\Delta J= 0, \ \pm1\) and \(-\leftrightarrow +\), that is positive symmetry states (with respect to coordinate inversion) must transition to negative symmetry states and vise versa. The initial state in this depletion transition is a \(^3\Sigma^+\) state, which is a strict Hund's case (b) state. As such, its rotational energy is determined by quantum number \(K\), which designates the total angular momentum of the molecule apart from spin, rather than the total angular momentum (including spin) quantum number \(J\). For this \(a ^3\Sigma^+\)state, the electronic spin is \(S\)=1, and levels with \(J=K+1, K\) and \(K-1\), are nearly degenerate for \(K\ge 1\). Additionally \(K\) determines the symmetry of the state. This is summarized in Fig. [\[fig:parity\]](#fig:parity){reference-type="ref" reference="fig:parity"}, adapted and modified from Ref. . The logic that leads to our assignments in Fig. [\[fig:depletion\]](#fig:depletion){reference-type="ref" reference="fig:depletion"} goes as follows. We start with the depletion data on transitions to \(\Omega=1\) shown in Fig. [\[fig:depletion\]](#fig:depletion){reference-type="ref" reference="fig:depletion"}(c), which is extensive enough to show that we do not populate \(J^{\prime \prime}=2\) of the \(a \ ^3\Sigma^+\) state; that is, we see no depletion signal corresponding to a transition to \(J^{\prime} = 3\). Since we do see transitions to \(J^{\prime} = 1\) and \(2\), we know that we populate some mix of \(J^{\prime \prime}=0\) and \(1\) belonging to either \(K^{\prime \prime}=0\), 1, or 2. \(K^{\prime \prime}=0\) and \(K^{\prime \prime}=2\) have the same symmetry and some of the same rotational numbers; this implies that any PA state that could decay to \(K^{\prime \prime}=0\) or 2 could decay to the other as well. Because the energies of \(K^{\prime \prime}=0\) and \(K^{\prime \prime}=2\) differ from one another, but we see no additional structure in the depletion spectra, we infer that the transitions seen in Fig. [\[fig:depletion\]](#fig:depletion){reference-type="ref" reference="fig:depletion"}(c) originate from a single state, that is \(K^{\prime \prime}=1\). The spacing between the two peaks in this spectrum should be \(4B_0\), where \(B_0\) is the rotational constant of the \(d \ ^3\Pi \ v^\prime = 0\) state, allowing us to determine \(B_0 = 4.59\) GHz. This rotational constant agrees with the prediction from LEVEL 8.0 with PECs from Ref. , which confirms our assignments and guides our interpretation of the \(\Omega=0\) data shown in Fig. [\[fig:depletion\]](#fig:depletion){reference-type="ref" reference="fig:depletion"}(d), in which we see three transitions. The first two are spaced by \(6B_0\) (27.5 GHz), implying that these peaks are transitions to the positive symmetry levels of the \(\Omega=0^+\) electronic state, \(J^{\prime}=0 \leftarrow K^{\prime \prime}=1\) and \(J^{\prime}=2 \leftarrow K^{\prime \prime}=1\). The absence of a peak for \(J^{\prime}=1\) is consistent with the selection rule \(-\not\leftrightarrow-\). The remaining transition in \(\Omega=0\) must be the only allowed transition to \(\Omega=0^-\), that is \(J^{\prime}=1 \leftarrow K^{\prime \prime}=1\). There is only one transition in \(\Omega=2\), which is trivial to identify as \(J^{\prime}=2 \leftarrow K^{\prime \prime}=1\). This picture is consistent with what we know about our PA state. That is, we PA into the \(J=1\) level of a \(2(0^-)\) state, which has positive symmetry, and the allowed decay paths are to \(K^{\prime \prime}=1\) (with \(J^{\prime \prime}=0\), \(1\), and \(2\), each with negative symmetry, of which we populate \(0\) and \(1\)). Of note, we chose to PA to \(J=1\) because it is the strongest PA resonance in the \(2(0^-) \ v=-5\) progression. Our depletion data confirms that the parity of this PA state is even, which is suggestive of a scattering state \(p\)-wave shape resonance . Similar shape resonances have been seen in other bi-alkali's . We used these data to determine the spin-orbit splitting between the different \(\Omega\) progressions deep in the \(d \ ^3\Pi\) well, and followed these progressions back to the asymptote in our RE2PI data. Most importantly, these data provide accurate locations of \(v^{\prime}=0\) and \(1\), \(J^{\prime}\) for future work on short range PA . The depletion data is limited by the shot noise in our ion counting. For each data point we integrate 200 shots and usually count \(\sim\)`<!-- -->`{=html}350 ions. To get 2\(\sigma\) resolution the smallest depletion feature that we can be confident in has to be a 10% decrease. Here our on-resonance depletion signal drops by around 30% which is 6\(\sigma\) and statistically significant. Unfortunately, our Ti:Sapphire struggles to tune to wavelengths much shorter than 740 nm, which limited our depletion spectra to the lowest two vibrational levels only. Despite these short comings, depletion spectroscopy gives us accurate identifications of these bottom two vibrational levels of the \(d \ ^3\Pi\) electronic state and unambiguously determines the spin orbit splitting. # Discussion We determine the vibrational binding energies and rotational constants of the states seen in depletion spectroscopy, which we tabulate in Table. [\[tab:depletionAssignments\]](#tab:depletionAssignments){reference-type="ref" reference="tab:depletionAssignments"}. We list in Table [\[tab:masterAssignments\]](#tab:masterAssignments){reference-type="ref" reference="tab:masterAssignments"} the assignments and energy of each of the \(d \ ^3\Pi\) states that we observe through RE2PI and depletion spectroscopy. We also include in this table the energy difference between adjacent states, which aids in the assignment of the lines. The theoretical vibrational levels and spin-orbit splittings that we used to guide our work and for comparison of results come from *ab initio* calculations by Korek et al.  with aid from LEVEL 8.0 . We found good overall agreement with these *ab initio* results. The spin-orbit splittings for \(\Omega=0\) to \(\Omega=1\), predicted to be 21 cm\(^{-1}\), are measured here to be 37 cm\(^{-1}\); for \(\Omega=1\) to \(\Omega=2\), they are predicted to be 38 cm\(^{-1}\), and we found them to be 33 cm\(^{-1}\). For the spin-orbit splitting between the \(\Omega=0^{+}\) to \(\Omega=0^{-}\) states, however, we observe 0.9 cm\(^{-1}\), significantly less than the predicted 36 cm\(^{-1}\). Our depletion data is unambiguous in establishing the \(\Omega=0^{+}\) to \(\Omega=0^{-}\) splitting, and a small \(\Omega=0^{+}\) to \(\Omega=0^{-}\) splitting is consistent with our observations in the \((3) \ ^3\Pi\) state . We show the vibrational spacing, \(\Delta E = E_{v+1}-E_v\) vs \(v\), of the different series in Fig. [\[fig:vibrationalSplitting\]](#fig:vibrationalSplitting){reference-type="ref" reference="fig:vibrationalSplitting"}. These data are in reasonable agreement with the predicted vibrational splittings although there appears to be a nearly uniform difference of a few cm\(^{-1}\). We found that the depth of the \(d \ ^3\Pi\) potential (exp. value) is less than that predicted (th. value). We looked extensively for another vibrational level below our assigned \(v^{\prime}=0\) level. We covered \(\pm\) 10 cm\(^{-1}\) around the expected vibrational location with our depletion spectra, but found no indication of a depletion resonance. Our extracted molecular constants are listed in Tab. [\[tab:molecularConstants\]](#tab:molecularConstants){reference-type="ref" reference="tab:molecularConstants"}. These provide an easy estimation of the spectral structure of the \(d \ ^3\Pi\) states as well as a quick comparison to theoretical predictions. As borne out in Fig. [\[fig:vibrationalSplitting\]](#fig:vibrationalSplitting){reference-type="ref" reference="fig:vibrationalSplitting"}, there is good agreement between our fitted harmonic constant, \(\omega_e\), and the predictions. However, there is considerably less agreement between our extracted term energy, \(T_e\), and the predictions for reasons discussed previously. Additionally, it is important to note that when we fitted the experimental data to determine \(T_e\), \(\omega_e\), \(x_e\) and \(y_e\) we used only \(v^{\prime}\)=0-19. This increased the accuracy of the fit so that for these vibrational levels our molecular constants reproduce our data with a standard deviation of 2 cm\(^{-1}\). We believe most of the deviations are caused by experimental uncertainties on determining the frequencies of the peaks, as well as small perturbations to state locations caused by spin-orbit mixing. # Conclusion In this work we have found and identified the \(v=0\) to \(v=22\) states of the \(d \ ^3\Pi_{\Omega}\) electronic state in LiRb. We explored singlet-triplet mixing to evaluate the possibility of using the \(d \ ^3\Pi\)-\(D \ ^1\Pi\) mixed states as the intermediate state in a STIRAP transfer from a weakly bound triplet state to deeply bound singlet states. We know from heat pipe spectra  and other REMPI data  that the transitions from deeply bound \(D \ ^1\Pi \ v^{\prime}\) to deeply bound \(X \ ^1\Sigma^+\) states are strong. Our depletion data shows that transitions from weakly bound triplet molecules to deeply bound \(d \ ^3\Pi_1 \ v^\prime\) states are also strong. Finally, our RE2PI data demonstrate that there is about 10% mixing between most of these singlet and triplet states. We suggest using the \(d \ ^3\Pi_1 \ v^\prime = 0 \ J^\prime = 1\) state as the intermediate state for a STIRAP transfer in future work. The laser wavelengths for this transfer would be 740 nm, achievable with Ti:Sapphire or diode lasers, and 516 nm, which is accessible with green laser diodes. Using calculated FCFs and an estimate of the transition dipole moment of a few times \(e a_0\), the 'up' transition would be around \(4 \times 10^{-2}\) \(e a_0\) and the 'down' transition would be around \(10^{-2}\) \(e a_0\) which is competitive with the transfer strength used by the KRb JILA team . Using the binding energy of the lowest several vibrational levels we have assessed their potential for short-range photoassociation. There are only six possible PA transitions we can observe with our current Ti:Sapphire laser, the same six we saw in depletion. However, our depletion data provides the perfect stepping stone for this type of work, providing very trusted locations of vibrational levels (to within 0.5 cm\(^{-1}\), the uncertainty in the binding energy of \(a \ ^3\Sigma^+ \ v^{\prime \prime}=13\)). Of particular interest to us is the combination of short range PA and mixing between \(d \ ^3\Pi\)-\(D \ ^1\Pi\). This provides an avenue through simple spontaneous decay to the rovibronic ground state although we will need a different laser to access these PA resonances. We are happy to acknowledge useful conversations with Jesús Pérez-Rı́os, and university support of this work through the Purdue OVPR AMO incentive grant. And we would like to acknowledge the work done by S. Dutta and J. Lorenz in building the LiRb machine.
{'timestamp': '2016-07-18T02:09:11', 'yymm': '1607', 'arxiv_id': '1607.04608', 'language': 'en', 'url': 'https://arxiv.org/abs/1607.04608'}

Dataset Card for "chemnlp-arxiv"

More Information needed

Downloads last month
0
Edit dataset card