Filename
stringlengths 22
64
| Paragraph
stringlengths 8
5.57k
|
---|---|
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | Fig. 11 demonstrates this spatial filtering property with an observation corresponding to the sky shown in Fig. 9. To obtain these data, the AAVS0.5 and MWA tiles are electron- ically pointed to the direction of a Hydra A (Dec. −12◦06(cid:48), RA 09:18:05 (J2000)), a compact radio galaxy with ∼ 0.14◦ angular extent [35]. Each data point is a visibility measurement of the correlated flux density for a given baseline (more on this in the next Subsection). |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | We see in Fig. 11 that the known flux density (310 Jy) of Hydra A dominates the visibilities for baselines of ∼ 30 − 150λ. For baselines shorter than ∼ 30λ, the spatially- extended Galactic noise is a significant contributor which causes the mean visibility amplitude to rise. Hence, for the purpose of telescope calibration and the AAVS0.5 characteri- zation, we exclude measurements from baselines of less than 30λ (vertical dashed line). The decreasing trend in visibility amplitude for baselines longer than ∼ 150λ is due to partial spatial filtering of Hydra A itself, the treatment of which will be discussed in Sec. V. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | Figure 12. Signal diagram of a radio interferometer. The system is composed of four parts: antennas, receiver, correlator and averaging. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | In radio interferometry, we infer the sensitivity of an AUT by measuring and normalizing the second-order statistics of the outputs of the interferometer. For brevity, we review only the salient features of this topic and refer interested readers to [36] for the full treatment. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | For N > 3, this problem is overdetermined. Consequently, each gi Ki may be solved via a standard least-squares method 9. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | where K = Ae/2k, B is the receiver noise bandwidth, k is the Boltzmann constant, gi and gj are the voltage gains of receivers i and j, respectively; SC is the correlated flux density which is less than or equal to the total flux density ST of the source. If this source is unresolved (not spatially filtered) by the ij baseline then SC = ST . |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | For N baselines, we again obtain N (N − 1)/2 pairs of such equations which can be solved for each SEFD through least- squares [39]. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | The measurement of A/T through cross-correlation requires a compact astronomical source that is sufficiently bright to obtain a good S/N calibration gain solution from a short (∼minutes) scan. Thus the measurement of tile gain is re- stricted to the path in (θ, φ) traced by calibrator sources moving across the beam of the AUT as the sky apparently rotates. Furthermore we seek to make measurements close to zenith, the direction of maximum gain. For a given source, the smallest zenith angle occurs at the meridian, a great circle passing through the zenith and the celestial poles (i.e. Az=0◦or Az=180◦). |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | Tab. II lists the southern calibrator sources used for sensitiv- ity measurement, selected to balance angular extent, brightness and zenith angle when transiting the meridian. They are not the brightest sources visible at the MRO, but other sources transit at larger zenith angles or are of considerable spatial-extent and difficult to model. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | Basic aperture synthesis imaging via the Fourier transform is another tool to verify calibration as it highlights errors in calibration that are not readily apparent in the visibility domain [40]. The MWA tiles and the AAVS0.5 forms a collection of 2-antenna cross-correlation interferometers (i.e. baselines) of various separation lengths to approximate a filled aperture; by extension of (5), there is, to a good approximation, a 2-D Fourier relationship10 between the visibilities and the image of the incident radiation. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | Fig. 13 shows a dual-polarization image of the calibrator source Hydra A. Although the full set of MWA–MWA and MWA–AAVS0.5 baselines were used for calibration, only the 127 baselines involving the AAVS0.5 were used to create the image, thus verifying the AAVS0.5 functionality as an interferometer component. At this frequency, these baselines are not of sufficient length to resolve the spatial structure in Hydra A. Other sources in the sky close (within ∼ 2◦) to Hydra A are significantly weaker but they align with known positions, demonstrating successful calibration. The X and Y-polarization images are similar, demonstrating consistency between the two signal chains. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | The measurement of the AAVS0.5 sensitivity was conducted with the MWA over several nights during 2014 and early 2015. The observations were made with the MWA tiles and the AAVS0.5 pointing at a zenith angle on the meridian such that the calibrator source passes through beam maximum. The MWA is limited to observing a total of 30.72 MHz of bandwidth at any one time. To sample the full 75–300 MHz frequency range of the MWA, we use two approaches: (i) stepping across the band over multiple nights and (ii) ob- serving widely separated spot frequencies during one night’s observation. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | Fig. 14 shows A/T measurements made with Hydra A over the period 14–21 January 2015, for the X (E-W) and Y (N-S) antenna polarizations. The observing strategy involved 6×2-minute snapshots of different 30.72 MHz bands to cover the full MWA frequency range. Observing each snapshot on a different night ensures the A/T measurement across the band is accurate to an angular resolution of 0.5◦ (the apparent motion of the calibrator source). The lack of data at ∼138 MHz and ∼240–285 MHz is due to the persistent satellite-based RFI at these frequencies. Fig. 14 also shows results expected from simulation for soil with 2% and 10% moisture (Sect. III). These represent a reasonable range in the likely soil moisture levels experienced at the MRO. |
Processed_Characterization_of_a_Low-Frequency_Radio_Astronom.txt | trends in Fig. 14 show good agreement between simulated and measured results. The degradation in A/T between 100–150 MHz is due to the log- periodic antenna interaction with the ground. Design optimiza- tion to minimize this undulation is being explored. |
Processed_QCD_modeling_of_hadron_physics.txt | We review recent developments in the understanding of meson properties as solutions of the Bethe–Salpeter equation in rainbow-ladder truncation. Included are recent results for the pseudoscalar and vector meson masses and leptonic decay constants, ranging from pions up to c¯c bound states; extrapolation to b¯b states is explored. We also present a new and improved calculation of Fπ(Q2) and an analysis of the πγγ transition form factor for both π(140) and π(1330). Lattice-QCD results for propagators and the quark-gluon vertex are analyzed, and the effects of quark-gluon vertex dressing and the three-gluon coupling upon meson masses are considered. |
Processed_QCD_modeling_of_hadron_physics.txt | The Dyson–Schwinger equations [DSEs] are the equations of motion of a quantum field theory. They form an infinite hierarchy of coupled inte- gral equations for the Green’s functions (n-point functions) of the theory. Bound states (mesons, baryons) appear as poles in the Green’s functions. Thus, a study of the poles in n-point functions us- ing the set of DSEs will tell us something about hadrons. For recent reviews on the DSEs and their use in hadron physics, see Refs. [1,2,3]. |
Processed_QCD_modeling_of_hadron_physics.txt | renormalized according to S(p)−1 = i /p + m(µ) at a sufficiently large spacelike µ2, with m(µ) the current quark mass at the scale µ. Both the prop- agator, S(p), and the vertex, Γi µ depend on the quark flavor, although we have not indicated this explicitly. The renormalization constants Z2 and Z4 depend on the renormalization point and on the regularization mass-scale, but not on flavor: in our analysis we employ a flavor-independent renormalization scheme. |
Processed_QCD_modeling_of_hadron_physics.txt | where Dµν(q = k gluon propagator, and Γi ized dressed quark-gluon vertex. The notation k Λ d4k/(2π)4. For divergent integrals R stands for a translationally invariant regularization is neces- sary; the regularization scale Λ is to be removed at the end of all calculations, after renormaliza- tion, and will be suppressed henceforth. |
Processed_QCD_modeling_of_hadron_physics.txt | at P 2 = ˜ko − The properly normalized BSA Γ(po, pi; P ) [or equivalently, χ(po, pi; P )] completely describes 2Here and subsequently, q refers to dressed quark (quasi- particle) states. |
Processed_QCD_modeling_of_hadron_physics.txt | the meson as a q ¯q bound state. Different types of mesons, such as pseudoscalar or vector mesons, are characterized by different Dirac structures. The ground state in any particular spin-flavor channel corresponds to the largest eigenvalue λ in that channel being one. The first excited state is determined by the next-largest eigenvalue λ be- ing one, and so on for higher excited states. |
Processed_QCD_modeling_of_hadron_physics.txt | This truncation is the first term in a system- atic expansion [4] of the quark-antiquark scat- tering kernel K; asymptotically, it reduces to leading-order perturbation theory. Furthermore, these two truncations are mutually consistent in the sense that the combination produces vector and axial-vector vertices satisfying their respec- tive Ward identities. In the axial case, this en- sures that in the chiral limit the ground state pseudoscalar mesons are the massless Goldstone bosons associated with chiral symmetry break- ing [5,6]. in combination with impulse approximation, elec- tromagnetic current conservation [7]. We will come back to this point later, when we discuss electromagnetic processes. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 1. The effective ladder-rainbow kernel from the Maris–Tandy [MT] model [8] compared to LO perturbative QCD and to that deduced from quenched lattice QCD data on the gluon and quark propagators [9]. |
Processed_QCD_modeling_of_hadron_physics.txt | The ultraviolet behavior of the effective run- ning coupling is dictated by the one-loop renor- malization group equation; the infrared behav- ior of the effective interaction is modeled, and constrained by phenomenology. Here, we are us- ing a 2-parameter model for the effective inter- action [8], fitted to give a value for the chiral condensate of (240 MeV)3 and fπ = 131 MeV. This effective running coupling, with ΛQCD = 0.234 GeV and Nf = 4, is shown in Fig. 1, and agrees perfectly with pQCD for q2 > 25 GeV2. The model is finite in the infrared region, and it is very similar to an effective interaction that was recently deduced [9] from quenched lattice QCD results for the gluon and quark propaga- tors. Indications are that unquenching does not change the infrared behavior of the running cou- pling much [10], and thus we believe that this 2- parameter model is a realistic parametrization of the effective quark-quark interaction in the space- like region. |
Processed_QCD_modeling_of_hadron_physics.txt | can also see that in the infrared region the dynam- ical mass function of the u and d quarks becomes very similar to that of chiral quarks. This is a di- rect consequence of DχSB, and leads to a dynam- ical mass function of several hundred MeV for the light quarks in the infrared region, even though the corresponding current quark masses are only a few MeV. Thus we connect the current quark mass of perturbative QCD with a “constituent- like” quark mass at low energies. |
Processed_QCD_modeling_of_hadron_physics.txt | These predictions for the quark mass func- tion have been confirmed in lattice simulations of QCD [14,15,16]. Pointwise agreement for a range of quark masses requires this interaction to be flavor-dependent [9], suggesting that dress- ing the quark-gluon vertex Γi ν(q, p) is important. The consequences of a dressed vertex for the me- son BSEs are currently being explored [17]. This will be discussed in more detail in Sec. 5. |
Processed_QCD_modeling_of_hadron_physics.txt | where the invariant amplitudes E, F , G and H are Lorentz scalar functions of k2 and k P , and they also depend on the choice for η. The natural choice for mesons with equal-mass constituents (like a pion) is η = 1 2 , though physical observables are of course independent of η. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 3. Pion and excited pion mass and decay constant, as function of the current quark mass (normalized by the realistic u/d quark masses); for comparison we have also included the vector meson mass. |
Processed_QCD_modeling_of_hadron_physics.txt | describing pions: for a realistic description of pi- ons as q ¯q bound states one necessarily has to have a (strongly) momentum dependent quark mass function. |
Processed_QCD_modeling_of_hadron_physics.txt | Table 1 DSE results [8] for the meson masses and decay constant, together with experimental data [20]. |
Processed_QCD_modeling_of_hadron_physics.txt | where T i µ(k, P ) are eight independent transverse Dirac tensors. Again, the invariant amplitudes f i are Lorentz scalar functions of k2 and k P , and depend on η. |
Processed_QCD_modeling_of_hadron_physics.txt | The BSE is usually solved in the rest-frame of the meson. To be explicit, the most convenient frame is that characterized by Pµ = (i M, 0, 0, 0). However, this is not required: One of the advan- tages of the DSE approach to hadron physics is its explicit Poincar´e covariance. This makes it a particularly useful tool for studying electromag- netic form factors and other processes. No mat- ter what frame one chooses for calculating say the pion electromagnetic form factor, at least one of the pions is moving. |
Processed_QCD_modeling_of_hadron_physics.txt | and the integral equation has to be solved in the three independent variables k2, α, and β. With current computer resources, this can be done without further approximations, and the re- sults, shown in Fig. 4, are indeed independent of the meson 3-momentum, illustrating that this ap- proach is indeed Poincar´e covariant. We can now use this same approach to calculate meson form factors in an explicitly covariant manner. |
Processed_QCD_modeling_of_hadron_physics.txt | In order to study electromagnetic processes, we need to know the dressed quark-photon coupling, in addition to the meson BSAs and the dressed quark propagators. |
Processed_QCD_modeling_of_hadron_physics.txt | The q ¯qγ vertex is the solution of the renormal- ized inhomogeneous BSE with the same kernel K as the homogeneous BSE for meson bound states. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 4. Pion and ρ mass and decay constant calculated in a moving frame, as function of the meson 3-momentum. |
Processed_QCD_modeling_of_hadron_physics.txt | The most general form of the quark-photon vertex Γµ(po, pi; Q) requires a decomposition into twelve Dirac structures. Four of these covariants represent the longitudinal components which are completely specified by the WTI in terms of the (inverse) quark propagator and they do not con- tribute to elastic form factors. The transverse vertex can be decomposed into eight Dirac ten- sors T i µ(p; Q) with the corresponding amplitudes being Lorentz scalar functions. |
Processed_QCD_modeling_of_hadron_physics.txt | where ΓV µ is the vector meson BSA, and fV the electroweak decay constant. The fact that the dressed q ¯qγ vertex exhibits these vector meson poles lies behind the success of naive vector- meson-dominance [VMD] models; the effects of intermediate vector meson states on electromag- netic processes can be unambiguously incorpo- rated by using the properly dressed q ¯qγ vertex rather than the bare vertex γµ [22]. |
Processed_QCD_modeling_of_hadron_physics.txt | The ladder BSE kernel is indeed independent of the meson momentum P , and with quark propa- gators dressed in rainbow approximation and the dressed quark-photon vertex calculated in ladder approximation, the both the vector WTI and the differential Ward identity are satisfied [22]; thus impulse approximation for the quark-photon ver- tex in combination with rainbow-ladder trunca- tion is consistent in the sense that the correct electric charge of the meson is produced indepen- dent of the model parameters and the resulting meson electromagnetic current is conserved. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 5. Our results for Q2 Fπ(Q2), with es- timates of our numerical error bars, compared to a VMD model and to experimental data from CERN [24] and JLab [25]. |
Processed_QCD_modeling_of_hadron_physics.txt | with the propagators solutions of the quark DSE in rainbow truncation. The pion BSAs and the q ¯qγ vertex are solutions of their respective BSEs in ladder truncation, using exactly the same mo- mentum frame as used in the form factor calcu- lation. That does mean re-calculating the pion BSAs for every value of the photon Q2, but the advantage is that one does not have to do any interpolation or extrapolation on the numerical solutions of the BSE. Our results for spacelike Q2 are shown in Fig. 5. They are in excellent agree- ment with the available experimental data [24,25] and with our previous results3. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 6. Pion charge radius as function of the quark mass within our DSE model, and as ex- tracted from lattice simulation [26,27]. |
Processed_QCD_modeling_of_hadron_physics.txt | We have calculated Fπ(Q2) for a range of cur- rent quark masses. For all masses considered, the form factor behaves like a monopole, with a monopole mass that is slightly smaller than the ρ-meson mass. A similar behavior has been found in quenched lattice simulations [26,27]. |
Processed_QCD_modeling_of_hadron_physics.txt | radius, defined by 6F ′(0), is also in excellent agreement with r2 = the experimental value. Our calculated charge ra- dius as a function of the quark mass is shown in Fig. 6, together with the charge radius obtained from lattice simulations. For comparison, we also show the charge radius one would obtain from a naive VMD model, using the corresponding (cal- culated) vector meson mass. Again, both our nu- merical result for r2 π and that of lattice simula- tions are represented reasonably well by a VMD model for the charge radius. |
Processed_QCD_modeling_of_hadron_physics.txt | in Ref. [28] it was demonstrated that the dressed quark core generates most of the pion charge ra- dius, and that pion loops contribute less than 15% to r2 π at the physical value of the pion mass. For larger values of the current quark mass (and thus of the pion mass), we expect these corrections to become negligible. Also for larger spacelike val- ues of Q2 the effects from meson loops decrease, and for Q2 > 1 GeV2 we expect the contribution of such loops to be negligible. |
Processed_QCD_modeling_of_hadron_physics.txt | where the internal momenta p, q are determined in terms of the integration momentum k and the external momenta Q1, Q2 by momentum conser- vation; furthermore, the pion is on-shell, but the photons not necessarily. |
Processed_QCD_modeling_of_hadron_physics.txt | is within 2% of the experimental width of 7.8 eV. The corrections due to finite pion mass are small. 4This expression is conventionally given in terms of ˆfπ = fπ /√2, the pion decay constant in the convention where its value is 92 MeV, rather than the convention fπ = 131 MeV used throughout this work. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 7. Symmetric πγγ transition form factor G(Q2, Q2) for both the ground state pion and its first radial excitation. Adapted from [21]. |
Processed_QCD_modeling_of_hadron_physics.txt | The axial anomaly is preserved by the ladder- rainbow truncation of the DSEs combined with an impulse approximation for the π0γγ vertex because the relevant manifestations of electro- magnetic gauge invariance and chiral symmetry are present [7,29]. We do indeed reproduce the above value for this form factor, as can be seen from Fig. 7. Our results also agree with the experimental data for the transition form factor Gπγ⋆γ(Q2, 0) [30,31]. |
Processed_QCD_modeling_of_hadron_physics.txt | On the other hand, the radially excited pion effectively decouples from the axial anomaly [21], and its coupling to two on-shell photons is not constrained by the anomaly. A naive application of Eq. (32) to the first radially excited pion would suggest an enormously large value of 625 GeV−1 for its coupling G(0, 0) to two photons, because its decay constant is very small, ˆfπ1 ≈ − 0.0016 GeV, and in the chiral limit this decay constant van- ishes, as is evident from Eq. (14) and Fig. 3. How- ever, numerically we find that for the excited pion this coupling is about 0.7 GeV−1. Furthermore, the chiral limit πγγ form factor is at low-Q2 al- most identical to the curves shown in Fig. 7, both for the ground state and for the excited pion. |
Processed_QCD_modeling_of_hadron_physics.txt | quark mass, this form factor indeed behaves like Eq. (36), and it approaches zero from below as Q2 , since its decay constant is nega- tive. In the chiral limit however, the decay con- stant ˆfπ1 is zero, as dictated by Eq. (14), and evident in Fig. 3. It turns out that in that case the form factor vanishes like 1/Q4, with a positive coefficient: the form factor has a zero-crossing in the spacelike region. The coefficient for this 1/Q4 behavior depends on the details of the model and its parameters, but it is not simply related to ˆfπ1. The leading order behavior of Eq. (36) however is model-independent. |
Processed_QCD_modeling_of_hadron_physics.txt | Let us first address the question of the quark mass dependence of the pseudoscalar and vector meson properties. In Fig. 9 we show the meson mass and decay constants as function of mq(19), the current quark mass at the renormalization point µ = 19 GeV, both for heavy-light mesons (light meaning u/d quarks and s quarks) and for heavy-heavy mesons. |
Processed_QCD_modeling_of_hadron_physics.txt | From this figure it is apparent that all of the meson masses scale approximately linear with the current quark mass above about mq ≈ ≈ 6ms, with ms being the strange quark mass. The pseudoscalar meson masses show a clear curva- ture below this mass scale, as one would expect based on the Gell-Mann–Oakes–Renner relation. Also the vector meson masses show a slight cur- vature at small current quark masses. The re- sults for the D, Ds, ηc, and J/Ψ mesons are in remarkably good agreement with experiments: the calculated masses with mc(19) = 0.88 GeV are within 2% of the experimental values. Com- pared to the light current quark masses, we find mc/mu = 238 and mc/ms = 10.5. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 9. Meson mass as function of mq. Top: our calculation (with estimates of numerical er- rors) up to mc plus the fit Eq. (37); bottom: ex- trapolated to mb using the same fit; the vertical dashed lines indicate mu/d, ms, mc, and mb. |
Processed_QCD_modeling_of_hadron_physics.txt | as indicated by the lines in Fig. 9. Here mi are the current quark masses of the constituents (at µ = 19 GeV); the chiral limit masses are M PS 0 = 0.75 GeV; and the remaining fit parameters are aPS 1 = 3.24 GeV, and a2 = 1.12 both for pseudoscalar and vector mesons. |
Processed_QCD_modeling_of_hadron_physics.txt | In Fig. 10 shows the corresponding leptonic de- cay constants. At small current quark masses, all of the decay constants increase with the quark mass, but they tend to level off between the s and the c quark mass, both for the heavy-light and for the heavy-heavy mesons. This is consistent with the expected behavior in the heavy-quark limit: namely a decrease of the decay constant with in- creasing meson mass (or equivalently, increasing 1/√M , at least for heavy-quark mass) like f the heavy-light mesons. Our current calculations suggest that the onset for this asymptotic behav- ior could be as low as the charm quark mass for the q¯u mesons. However, our results for both fD and fDs are about 20% below their experi- mental values, fD = 0.222 0.020 GeV [33] and fDs = 0.266 0.032 GeV [20] respectively, and our result [34] for the decay constant of the J/ψ is about 25% below its experimental value, ex- tracted from the e+e− decay [20]. This could in- dicate that the rainbow-ladder truncation is not reliable in the charm quark region, and/or our model for the effective interaction is not applica- ble to charm quarks, despite the fact that the meson masses are in good agreement with ex- periments. Since the decay constants depend on the norm of the BSAs, which in turn depends on the derivative of the quark propagators, it is not a surprise that the decay constants are more sensitive to details of the model than the meson masses, and are thus better indicators for defi- ciencies in the modeling. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 10. Meson decay constant (with estimates of our numerical errors) as function of mq, up to the charm quark mass region; the vertical dashed lines indicate mu/d, ms, and mc. |
Processed_QCD_modeling_of_hadron_physics.txt | the experimental value [18]. Furthermore, these heavier mesons, as well as the scalar mesons [37], tend to be more sensitive to details of the inter- action than the ρ and π. |
Processed_QCD_modeling_of_hadron_physics.txt | This suggests a deficiency of the model and/or of the rainbow-ladder truncation for these higher- mass states. A possible explanation for this de- ficiency lies in the fact that these higher-mass states have a different spin structure and (espe- cially in case of the radially excited pion) very different BS wave functions, and therefore they are sensitive to different aspects of the BS kernel K. Again, we should study the effects from con- tributions beyond the rainbow-ladder truncation in more detail in order to resolve these issues. |
Processed_QCD_modeling_of_hadron_physics.txt | tributions to a number of observables are possi- ble with a better understanding of the infrared structure of the vertex. These diverse model in- dications include: an enhancement in the quark condensate [10,17]; an increase of about 300 MeV in the b1/h1 axial vector meson mass [36]; and about 200 MeV of attraction in the ρ/ω vector meson mass [17]. |
Processed_QCD_modeling_of_hadron_physics.txt | Over the last five years, the gluon 2-point function has become fairly well known through explicit lattice-QCD calculations [42] and also through approximate solutions of the gauge sec- tor DSEs [2,10,43]. A strictly defined rainbow truncation of the quark DSE for calculation of the dressed quark propagator would proceed from this input. However, the modern information on the dressed gluon propagator shows a strong in- frared suppression and it is not possible to ob- tain a realistic value for the condensate in rainbow truncation [44]. In this sense it is clear that the empirically successful rainbow-ladder kernels developed earlier implicitly include effects of quark-gluon vertex dressing for the quark DSE. Although such an effective kernel shows a bare vertex Dirac matrix structure, the infrared pa- rameterization produces a strength and depen- dence upon gluon momentum that is over and above that of the gluon propagator. |
Processed_QCD_modeling_of_hadron_physics.txt | This point is illustrated in Fig. 11 [9], where quenched lattice data on the gluon and quark propagators are used to deduce a phenomenolog- ical quark-gluon vertex by means of the DSE. is modeled by The DSE kernel of Ref. 4πα(q2) Dfree µν (q) in Eq. (9), µν (q) where Dlat µν (q) represents the quenched lattice re- sult for the Landau gauge gluon propagator, and V (q2, mq) is a phenomenological representation of vertex dressing whose ultraviolet behavior is determined by the requirement that the resulting running coupling reproduces the one-loop renor- malization group behavior of QCD. Thus, the ver- tex dressing has been mapped onto a single ampli- tude corresponding to the canonical Dirac matrix γµ, and depends on the gluon momentum and the current quark mass. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 11. Dynamical quark mass functions ob- tained with the DSE-Lat model together with the quenched lattice data [15]. Adapted from [9]. |
Processed_QCD_modeling_of_hadron_physics.txt | compared to the MT model in Fig. 1. The param- eters are determined by requiring that the DSE solutions reproduce the quenched lattice data [15] for S(p) in the available domain p2 < 10 GeV2 In this sense, and m(µ = 2 GeV) < 200 MeV. the DSE-Lat model represents quenched dynam- ics. It is found that the necessary vertex dress- ing is a strong but finite enhancement. As one would expect, the vertex dressing V (q2, mq) de- creases with increasing mq to represent the effect of quark propagators internal to the vertex. |
Processed_QCD_modeling_of_hadron_physics.txt | The model easily reproduces mπ with a current mass that is within acceptible limits. However the = (0.19 GeV)3 is resulting chiral condensate ¯qq h a factor of 2 smaller than the empirical value 0.01 GeV)3. This is attributed to the (0.24 quenched approximation in the lattice data. |
Processed_QCD_modeling_of_hadron_physics.txt | Let us denote the dressed quark-gluon vertex for gluon momentum k and quark momenta p and p+k by ig λ (g2), i.e., to 2 Γσ(p + k, p). Through O σ + ΓNA one loop, we have Γσ = Z1F γσ + ΓA σ , see Fig. 12. Here Z1F is the vertex renormalization constant to ensure Γσ = γσ at the renormaliza- tion scale µ. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 12. The one-loop corrections to the quark- gluon vertex dressing. Left: the Abelian-like term σ ; Right: the non-Abelian term ΓNA ΓA σ . |
Processed_QCD_modeling_of_hadron_physics.txt | is CF = c over the color-octet ΓA σ term: single gluon exchange between a quark and antiquark has relatively weak repulsion in the color-octet channel, compared to the strong attraction in the color-singlet channel. |
Processed_QCD_modeling_of_hadron_physics.txt | the ladder-rainbow kernel; hence the natural ex- tension is g2 D0(q2) 4παeff(q2)/q2 and the ex- ternal quark-gluon vertex is taken to be bare. |
Processed_QCD_modeling_of_hadron_physics.txt | it takes full advantage of the simpli- fication, Eq. (38), that happens for zero momen- tum gluons: replace g2 D0(q2) by 4παeff(q2)/q2. This term is similar to a contribution to the derivative of the quark self-energy but the differ- ences are important; at one-loop this term pro- vides the explicitly non-Abelian contributions to the Slavnov-Taylor identity [45]. The justifica- tions for this nonperturbative vertex model are consistency and simplicity; no new parameters are introduced. |
Processed_QCD_modeling_of_hadron_physics.txt | Another useful comparison is the correspond- ing vertex in an Abelian theory like QED; it is given by the differential Ward identity, Eq. (25). With S−1(p) = iγ p A(p2) + B(p2), this leads to A′/2, and the correspondance λWI 2 = 3 = B′, where f ′ = ∂f (p2)/∂p2. The Abelian λWI Ansatz [46], while clearly inadequate for the dom- inant amplitude λ1 below 1.5 GeV, does repro- duce the DSE-Lat results for both λ2 and λ3. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 13. The amplitudes of the dressed quark- gluon vertex at k = 0 for mq(2 GeV) = 60 MeV. Quenched lattice data [38] are compared to the re- sults of the DSE-Lat model [9], and for λ1(p) the Abelian Ansatz is also shown. Adapted from [9]. |
Processed_QCD_modeling_of_hadron_physics.txt | There is a known constructive scheme [4] that defines a diagrammatic expansion of the BSE ker- nel corresponding to any diagrammatic expansion of the quark self-energy such that the vector and axial-vector WTIs are preserved. Among other things, this guarantees the Goldstone boson na- ture of the flavor non-singlet pseudoscalars inde- pendently of model details [5]. |
Processed_QCD_modeling_of_hadron_physics.txt | This has recently been exploited to produce the first study [17] of how an infinite sub-class of quark-gluon vertex diagrams, including at- traction from three-gluon coupling, contributes to vector and pseudoscalar meson masses. The symmetry-preserving BSE kernel obtained from the self-energy has an infinite sub-class of dia- grams beyond ladder; it is manageable here be- cause of the algebraic structure afforded from use of a modification of the Munczek-Nemirovsky δ- function model [47]. It is verified that the pseu- doscalar q ¯q meson remains a Goldstone boson at any order of truncation. As shown in Fig. 14, the ladder-rainbow truncation has at most a 8% re- pulsive error as mq is varied up to the b quark region. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 14. The relative error of the ladder- rainbow truncation relative to the completely re- summed model of vertex dressing and correspond- ing BSE kernel. Adapted from Ref. [17]. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 15. The two leading-order (1-loop) vertex corrections to the rainbow DSE (top) and the cor- responding 5 additions to the ladder BSE kernel (bottom) to preserve the relevant Ward identities. Quark and gluon lines are dressed. |
Processed_QCD_modeling_of_hadron_physics.txt | pect the qualitative aspects of these observations to be confirmed whenever a more realistic study becomes feasible. |
Processed_QCD_modeling_of_hadron_physics.txt | If one goes beyond the rainbow-ladder trunca- tion for the propagators, the meson BSAs and the quark-photon vertex, one has to go beyond impulse approximation for the meson form factor to ensure current conservation [23]. For exam- ple, following the general procedure of Ref. [4], one-loop dressing of the quark-gluon vertex in the DSE requires 5 additions to the ladder BSE ker- nel to preserve the relevant WTIs, see Fig. 15. |
Processed_QCD_modeling_of_hadron_physics.txt | Figure 16. Top: The 4 contributons to the BSA normalization condition from the additions of Fig. 15 to the DSE dynamics. Bottom: The corresponding 4 corrections to the impulse ap- proximation to the form factor. Slashes denote derivatives as described in the text. |
Processed_QCD_modeling_of_hadron_physics.txt | mation to maintain current conservation [23], and they are displayed in the bottom row of Fig. 16. The transversality of the photon-meson coupling is also maintained by this truncation. |
Processed_QCD_modeling_of_hadron_physics.txt | We note that this procedure systematically ex- poses Fock-space components of the meson that are included in the exact q ¯q scattering kernel of the BSE but which necessarily show up explicitly as corrections to the impulse approximation. |
Processed_QCD_modeling_of_hadron_physics.txt | which can be obtained by solving an inhomoge- neous BSE [22]. The rainbow-ladder truncation, in combination with impulse approximation for the form factor, guarantees current conservation. Our result for Fπ(Q2) is in remarkable agreement with the data from TJNAF [25]. |
Processed_QCD_modeling_of_hadron_physics.txt | The results for the kaon form factors are also in good agreement with the data [23], as are the results for the radiative [30,31] and strong de- cays [48] of the vector mesons, without any re- adjustments of the model parameters. Hadronic processes involving four (or more) external par- ticles can also be described in this framework, though one has to go beyond impulse approxi- mation. Our results for π-π scattering [49] agree with predictions from chiral symmetry, and this method has been applied to the anomalous γπππ coupling [50] as well. This approach unambigu- ously incorporates bound state effects in pro- cesses that receive contributions from off-shell in- termediate mesons, such as σ and ρ mesons in case of π-π scattering [49], and VMD effects in the case of electromagnetic interactions. |
Processed_QCD_modeling_of_hadron_physics.txt | Within the same framework, one can consider baryons as bound states of a quark and a diquark, the latter in a color anti-triplet configuration. It has been shown [51] that one can obtain a rea- sonable description of the ground state octet and decouplet baryons using both scalar and axial- vector diquarks. There has also been significant progress in understanding the nucleon form fac- tors using this approach [52]. |
Processed_QCD_modeling_of_hadron_physics.txt | This work was supported by the US De- partment of Energy, contract No. DE-FG02- 00ER41135 and by the National Science Foun- dation, grant No. PHY-0301190. The work bene- fited from the facilities of the NSF Terascale Com- puting System at the Pittsburgh Supercomputing Center. We thank C.D. Roberts for very useful discussions and are grateful to P. Bowman and J.I. Skullerud for providing lattice-QCD results. Appreciation is extended to D. Leinweber, the Light Cone organizing committee, and the staff and members of the CSSM, University of Ade- laide for hospitality and support. |
Processed_QCD_modeling_of_hadron_physics.txt | Skullerud, A.G. Williams and C. Parrinello [UKQCD Collaboration], Phys. Rev. D 60, 094507 (1999) [Erratum-ibid. D 61, 079901 (2000)] [arXiv:hep-lat/9811027]. |
Processed_Developing_the_GOTO_telescope_control_system.txt | The Gravitational-wave Optical Transient Observer (GOTO) is a wide-field telescope project focused on detecting optical counterparts to gravitational-wave sources. The GOTO Telescope Control System (G-TeCS) is a custom robotic control system which autonomously manages the GOTO telescope hardware and nightly operations. Since the commissioning the GOTO prototype on La Palma in 2017, development of the control system has focused on the alert handling and scheduling systems. These allow GOTO to receive and process transient alerts and then schedule and carry out observations, all without the need for human involvement. GOTO is ultimately anticipated to include multiple telescope arrays on independent mounts, both on La Palma and at a southern site in Australia. When complete these mounts will be linked to form a single multi-site observatory, requiring more advanced scheduling systems to best optimise survey and follow-up observations. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | Abstract: Site measurements were collected at Mount John University Observatory in 2005 and 2007 using a purpose-built scintillation detection and ranging system. C 2 N (h) profiling indicates a weak layer located at 12 – 14 km above sea level and strong low altitude turbulence extending up to 5 km. During calm weather conditions, an additional layer was detected at 6 – 8 km above sea level. V (h) profiling suggests that tropopause layer velocities are nominally 12 – 30 ms−1, and near-ground velocities range between 2 – 20 ms−1, dependent on weather. Little seasonal variation was detected in either C 2 N (h) and V (h) profiles. The average coherence length, r0, was found to be 7 ± 1 cm for the full profile at a wavelength of 589 nm. The average isoplanatic angle, θ0, was 1.0 ± 0.1 arcsec. The mean turbulence altitude, h0, was found to be 2.0 ± 0.7 km above sea level. No average in the Greenwood frequency, fG, could be established due to the gaps present in the V (h) profiles obtained. A modified Hufnagel-Valley model was developed to describe the C 2 N (h) profiles at Mount John, which estimates r0 at 6 cm and θ0 at 0.9 arcsec. A series of V (h) models were developed, based on the Greenwood wind model with an additional peak located at low altitudes. Using the C 2 N (h) model and the suggested V (h) model for moderate ground wind speeds, fG is estimated at 79 Hz. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | Unlike r0, θ0 is dependent on h5/3 indicating that weak high altitude layers have a significant impact on θ0. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | Astronomical images taken by ground based telescopes are subject to distortion caused by atmospheric turbu- lence. Adaptive optics (AO) provides a real-time solu- tion to compensate for an aberrated wavefront through the use of deformable optics in a closed-loop system. Accurate measurements and models of atmospheric tur- bulence are an essential tool in the design and optimi- sation of an AO system (Avila et al. 2001). |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | where V (h) is the average horizontal wind velocity as a function of altitude h. fG determines how quickly an AO system must respond to adequately compensate for the aberrations induced by atmospheric turbulence. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | The 1-m McLellan telescope at Mount John Uni- versity Observatory (MJUO), located at Tekapo, New Zealand, is used for a variety of different astronomi- cal research and is known to regularly experience poor seeing (> 2 arcsec angular resolution) by observers (A. Gilmore (MJUO) 2006, private communication). This work is part of a feasibility study on installing an AO system to improve photometric images with the CCD photometer head and to improve light through- put into the HERCULES ´echelle spectrograph (Hearn- shaw et al. 2002) currently installed on the 1-m tele- scope. The elevation of MJUO is 1024 m. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | the techniques used to measure C 2 N (h) and V (h) pro- files and the purpose-built system. Sections 3 and 4 discuss the data collected and the trends noted in the profiles obtained. Section 5 introduces the models de- veloped to describe atmospheric turbulence at MJUO. Concluding remarks are in section 6. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | SCIDAR (SCIntillation Detection And Ranging) is a remote sensing technique that has been used at many different sites around the world to characterise opti- cal turbulence (Avila et al. 2001; Garc´ıa-Lorenzo et al. 2009; Masciadri et al. 2010; Prieur et al. 2001; Tokovinin et al. 2005). It uses the spatio-temporal covariance functions obtained from a sequence of short exposure images of the scintillation pattern seen at a telescope pupil to infer the C 2 N (h) and V (h) profiles present above a site (Kl¨uckers et al. 1998). |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | SCIDAR measurements are commonly taken using a double star system, as indicated in Figure 1. Light from each star passes through the same region of a tur- bulent layer forming identical, but separated, scintil- lation patterns. The distance between the two scintil- lation patterns is directly proportional to the angular separation of the double star system, φ, and the height of the turbulent layer above the measurement plane, hi. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | scintillation patterns seen at the telescope aperture. In doing this, a clear picture of optical turbulence in the free at- mosphere can be obtained. As scintillation is propor- tional to h5/6 (Roddier 1981), any scintillation result- ing from near-ground turbulence (NGT) is not readily detectable. Using a simple lens change, the measure- ment plane can be shifted to a virtual plane located at d below the telescope. If hL is the height of the layer above the telescope then the height of the layer above the measurement plane becomes hi = |hL − d|, where d is negative due to sign conventions. This increased propagation distance allows for scintillation from NGT to be adequately measured. This version of SCIDAR is known as generalised SCIDAR (Kl¨uckers et al. 1998). The University of Canterbury SCIDAR system (UC-SCIDAR) is a purpose-built instrument designed to measure C 2 N (h) and V (h) profiles (Johnston et al. 2004, 2005; Mohr at al. 2006, 2008a, 2008b). UC- SCIDAR saw first light on the McLellan 1-m tele- scope at MJUO in late 2003, at an approximate cost of USD$4000. To keep costs low the system design utilised primarily off-the-shelf components. Over the years, the system has evolved through several itera- tions. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | Figure 1: The concept of double star SCIDAR. Light from each star passes through the same tur- bulent region forming identical scintillation pat- terns separated by a distance proportional to the double star separation φ and the height of the tur- bulent layer above the measurement plane hi = |hL − d|. Due to sign conventions d is negative. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | Figure 2: (a) Physical and (b) optical layout of the UC-SCIDAR instrument. See Section 2 for a detailed description of the system. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | Kodak KAI-0340DM sensor which has a 640 x 480 grid of 7.4 µm square pixels. UC-SCIDAR has been config- ured to capture images at full resolution with a frame rate of 60 Hz. Operational exposure times range from 0.5 to 5 ms. The lens used provides a nominal spatial sampling, ∆r, of 1/125 m.pix−1 when the 1-m tele- scope is operating at a focal ratio of F/13.5. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | A second identical CCD camera is mounted in the side path (C2). As this channel is typically used for generalised SCIDAR, a f10mm achromat lens (L2) is mounted a focal length of L1 away from C2. This pro- vides a measurement plane at approximately 3.5 km below the telescope. Figure 3 shows a typical pupil- plane and generalised SCIDAR scintillation image ob- tained using UC-SCIDAR. |
Processed_Optical_Turbulence_Measurements_and_Models_for_Mou.txt | where ∆m is the magnitude difference between the double star components. CB(ρ, φ, ∆t) describes a se- ries of triplets where each corresponds to a different turbulent layer. |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.