1: %\documentclass[apj]{emulateapj}
2: \documentclass[12pt,preprint]{aastex}
3: \usepackage{apjfonts}
4:
5:
6: %% VINCENT: COMMENT THIS OUT BEFORE SUBMITTING. This allows us
7: %% to use colorful language in our intra-paper discourse.
8: %\usepackage[usenames,dvips]{color}
9: %\newcommand\qc[1]{{\color{red} \bf #1}} %Question comment ( \qc{question} )
10: %\newcommand\ac[1]{{\color{blue} \bf #1}} %Answer comment ( \ac{answer} )
11: %\newcommand\aeb[1]{{\color{OliveGreen} #1}} % My additions,
12: % conveniently colorized for you!
13: % To remove comments, comment out above lines and uncomment out below
14: \newcommand\qc[1]{}
15: \newcommand\ac[1]{}
16: %% END
17:
18: \def\Rs{R_{\rm S}}
19:
20: \addtolength{\topmargin}{0.6truein}
21: %\newcommand{\nod}{\nodata}
22: %\newcommand{\lt}{\ensuremath{< 10}}
23:
24: %% Something to make greek letters bold
25: \newcommand\bmath[1] {\mbox{\boldmath$\rm #1$}}
26:
27:
28:
29: %%% Abbreviations
30:
31: % Standard abbreviations
32: \newcommand\cf{{cf.}} %.......confer
33: \newcommand\eg{{e.g.}} %......exempli gratia
34: \newcommand\ie{{i.e.}} %......id est
35: \newcommand\etal{{et al.}} %..and others
36: \newcommand\etc{{etc}} %......et cetera
37:
38: % Unit abbreviations
39: %%% Temperature
40: \newcommand\K{{\rm K}} %...........Kelvin
41: %%% Time
42: \newcommand\s{{\rm s}} %...........seconds
43: \newcommand\ns{{\rm n}\s} %........nanoseconds
44: \newcommand\mus{\mu\s} %...........microseconds
45: \newcommand\ms{{\rm m}\s} %........milliseconds
46: \newcommand\ks{{\rm k}\s} %........kiloseconds
47: \newcommand\hr{{\rm hr}} %.........hours
48: \newcommand\yr{{\rm yr}} %.........years
49: \newcommand\Myr{{\rm M}\yr} %......megayears
50: \newcommand\Gyr{{\rm G}\yr} %......gigayears
51: %%% Frequency
52: \newcommand\Hz{{\rm Hz}} %.........Hertz
53: \newcommand\muHz{\mu\Hz} %.........microhertz
54: \newcommand\mHz{{\rm mHz}} %.......millihertz
55: \newcommand\MHz{{\rm MHz}} %.......Megahertz
56: \newcommand\GHz{{\rm GHz}} %.......Gigahertz
57: \newcommand\THz{{\rm THz}} %.......Terahertz
58: %%% Length
59: \newcommand\m{{\rm m}} %...........meters
60: \newcommand\nm{{\rm n}\m} %........nanometers
61: \newcommand\mum{\mu\m} %...........micrometers
62: \newcommand\mm{{\rm m}\m} %........millimeters
63: \newcommand\cm{{\rm c}\m} %........centimeters
64: \newcommand\km{{\rm k}\m} %........kilometers
65: \newcommand\pc{{\rm pc}} %.........parsecs
66: \newcommand\kpc{{\rm k}\pc} %......kiloparsecs
67: \newcommand\Gpc{{\rm G}\pc} %......gigaparsecs
68: \newcommand\au{{\rm au}} %.........astronomical units
69: %%% Mass
70: \newcommand\g{{\rm g}} %...........grams
71: \newcommand\mug{\mu\g} %...........micrograms
72: \newcommand\mg{{\rm m}\g} %........milligrams
73: \newcommand\kg{{\rm k}\g} %........kilograms
74: \newcommand\Ms{M_\odot} %...........solar masses
75: %%% Energy
76: \newcommand\eV{{\rm eV}} %.........electron volts
77: \newcommand\keV{{\rm k}\eV} %......kiloelectron volts
78: \newcommand\MeV{{\rm M}\eV} %......megaelectron volts
79: \newcommand\GeV{{\rm G}\eV} %......gigaelectron volts
80: \newcommand\TeV{{\rm T}\eV} %......teraelectron volts
81: \newcommand\PeV{{\rm P}\eV} %......petaelectron volts
82: \newcommand\XeV{{\rm X}\eV} %......exaelectron volts
83: \newcommand\ZeV{{\rm Z}\eV} %......zetaelectron volts
84: \newcommand\erg{{\rm erg}} %.......ergs
85: \newcommand\J{{\rm J}} %...........joules
86: %%% Magnetic Fields
87: \newcommand\G{{\rm G}} %...........gauss
88: \newcommand\nG{{\rm n}\G} %........nanogauss
89: \newcommand\muG{\mu\G} %...........microgauss
90: \newcommand\mG{{\rm m}\G} %........milligauss
91: \newcommand\kG{{\rm k}\G} %........kilogauss
92: \newcommand\T{{\rm T}} %...........tesla
93: \newcommand\muT{\mu\T} %...........microtesla
94: \newcommand\mT{{\rm m}\T} %........millitesla
95: %%% Angular Resolution
96: \newcommand\mas{{\rm mas}} %.......milli-arcseconds
97: \newcommand\muas{\mu{\rm as}} %....micro-arcseconds
98: \newcommand\nas{{\rm nas}} %.......nano-arcseconds
99: \newcommand\e{{\rm e}}
100:
101:
102:
103: \begin{document}
104: %\journalinfo{Target: Astrophysical Journal}
105: %
106: \slugcomment{Version 4.01: \today}
107: %
108: % 12345678901234567890123456789012345678901234
109: \shorttitle{Sgr A* Flare Detection with Millimeter VLBI}
110: %
111: \shortauthors{Doeleman et al.}
112: %
113: \title{Methods for Detecting Flaring Structures in Sagittarius A* with High Frequency VLBI}
114: %
115: \author{
116: %
117: Sheperd S.~Doeleman\altaffilmark{1},
118: %
119: Vincent L.~Fish\altaffilmark{1},
120: %
121: Avery E.~Broderick\altaffilmark{2},
122: %
123: Abraham Loeb\altaffilmark{3}, \&
124: %
125: Alan E.~E.~Rogers\altaffilmark{1}}
126: %
127: \altaffiltext{1}{Massachusetts Institute of Technology, Haystack
128: Observatory, Route 40, Westford, MA 01886.}
129: %
130: \altaffiltext{2}{Canadian Institute for Theoretical Astrophysics,
131: University of Toronto, 60 St.\ George St., Toronto, ON, M5S 3H8
132: Canada.}
133: %
134: \altaffiltext{3}{Institute for Theory and Computation, Harvard
135: University, Center for Astrophysics, 60 Garden St., Cambridge, MA
136: 02138.}
137:
138: \begin{abstract}
139: The super massive black hole candidate, Sagittarius~A*, exhibits variability
140: from radio to X-ray wavelengths on time scales that correspond to < 10
141: Schwarzschild radii. We survey the potential of millimeter-wavelength VLBI to
142: detect and constrain time variable structures that could give rise to such
143: variations, focusing on a model in which an orbiting hot spot is embedded in an
144: accretion disk. Non-imaging algorithms are developed that use interferometric
145: closure quantities to test for periodicity, and applied to an ensemble of
146: hot-spot models that sample a range of parameter space. We find that
147: structural periodicity in a wide range of cases can be detected on most
148: potential VLBI arrays using modern VLBI instrumentation. Future enhancements
149: of mm/sub-mm VLBI arrays including phased array processors to aggregate VLBI
150: station collecting area, increased bandwidth recording, and addition of new
151: VLBI sites all significantly aid periodicity detection. The methods described herein
152: can be applied to other models of Sagittarius~A*, including jet outflows and
153: Magneto-Hydrodynamic accretion simulations.
154: \end{abstract}
155: \keywords{black hole physics --- Galaxy: center --- techniques:
156: interferometric --- submillimeter --- accretion, accretion disks}
157:
158: \section{Introduction}
159: \label{introduction}
160:
161: Observations of the compact radio/IR/X-ray source Sagittarius~A*
162: (Sgr~A*) make the most compelling case for the existence of super massive
163: black holes. Both speckle imaging and adaptive optics work in the
164: near-infrared (NIR) band shows that multiple stars orbit the position
165: of Sgr~A* \citep{schodel03,ghez05} These orbits are consistent with a
166: central mass of $\sim4\times10^6 ~ M_\sun$ contained within 45~AU -
167: the closest approach of any star. Radio interferometric proper motion
168: measurements \citep{backer99,reid04} limit the motion of Sgr~A* to
169: $<15$~km\,s$^{-1}$, implying that Sgr~A* must trace at least 10\% of
170: the mass determined from stellar orbits. Very long baseline interferometry (VLBI) at 1.3~mm
171: wavelength \citep{doeleman08} has resolved Sgr~A*, and measures an intrinsic
172: size of $<0.3$~AU \citep[assuming a distance of 8.0~kpc, from][]{reid93}, or 4
173: times the Schwarzschild radius of the central mass ($\Rs \approx 10~\mu$as).
174: Assuming Sgr~A* marks the position of the black hole, the implied density,
175: using the proper motion lower limit on the mass and the VLBI size, is
176: $>9.3\times10^5~M_\sun\,{\mathrm{AU}}^{-3}$. Almost any conceivable
177: aggregation of matter would, at this density, quickly collapse to a black hole
178: \citep{maoz98}.
179:
180: The 1.3~mm VLBI result confirms the existence of structures within Sgr~A* on
181: size scales commensurate with the innermost accretion region, and matches size
182: scales inferred from light curve monitoring over a broad wavelength range.
183: Sgr~A* exhibits variability on time scales of minutes to hours in the radio,
184: millimeter, NIR, and X-ray bands \citep[e.g.,][]{baganoff01, aschenbach04,
185: genzel03, ghez04, belanger06, meyer06, yusefzadeh06, hornstein07, marrone07},
186: and flare rise times in the X-ray and NIR correspond to light-crossing times of
187: $<12~\Rs$. Models that produce flaring X-ray flux density via Synchrotron Self
188: Compton scattering of IR photons require emission regions of diameter $<10~\Rs$
189: \citep{marrone07}.
190:
191: The resolution of millimeter- and submillimeter-wavelength VLBI is
192: well matched to the scale of inner disk physics. Baselines from
193: Hawaii or Western Europe to Chile provide fringe spacings as small as
194: $30~\mu$as ($3~\Rs$) at 230~GHz and $20~\mu$as at 345~GHz. Millimeter VLBI
195: thus has the potential to detect signatures of hot spot and jet models proposed
196: to explain the rapid variability of Sgr~A* as well as strong general
197: relativistic effects, such as the black hole silhouette or shadow
198: \citep{falcke00,broderick06b,huang07,markoff07}. Extending the VLBI technique
199: to short (0.85~mm, 1.3~mm) wavelengths is essential for this work due to the
200: interstellar scattering towards Sgr~A*, which broadens radio images with a
201: $\lambda^2$ dependence \citep{backer78}. VLBI at 7~mm and 3~mm wavelengths
202: \citep{rogers94,doeleman01,bower04,shen05} limits the intrinsic size of Sgr~A*
203: to be $<2$~AU and $<1$~AU respectively, but VLBI at these wavelengths is
204: strongly influenced by scattering effects. For $\lambda < 1.3$~mm the scattering size is
205: less than the fringe spacings on the longest possible baselines. Recent 1.3~mm
206: VLBI results \citep{doeleman08}, coupled with ongoing technical advances to
207: reach 0.85~mm, strongly suggest that it is not a question of \emph{if} but of
208: \emph{when} VLBI will directly probe Sgr~A* on event horizon scales.
209:
210: Claims of observed periodicity in IR and X-ray light curves
211: \citep{belanger06, meyer06, eckart06} during Sgr~A* flares, can
212: potentially be explained in the context of hot spots orbiting the
213: black hole at a few times $\Rs$. It has been proposed that the
214: fastest periodicity can be used to constrain the spin of the black
215: hole, since the period of the innermost stable circular orbit (ISCO)
216: is much shorter for a maximally rotating Kerr black hole than for a
217: nonrotating Schwarzschild black hole \citep{genzel03}. Indeed, several
218: authors have argued that the black hole must be rotating based on
219: observed rapid X-ray and infrared periodicities
220: \citep{aschenbach04,belanger06,meyer06}. However, it has also been
221: proposed that the flaring activity can be explained by
222: magnetohydrodynamic turbulence along with density fluctuations
223: \citep{goldston05,chan06}. Rossby wave instabilities may naturally
224: produce periodicities on the order of tens of minutes, in which case
225: it is not necessary to appeal to a nonzero black hole spin to explain
226: the fast quasi-periodic flares seen at multiple wavelengths
227: \citep{tagger06,falanga07}. Regardless of the source of variability
228: at millimeter wavelengths, VLBI has the potential to confirm
229: conclusively its association with the inner disk of Sgr~A*, to probe
230: the size of the region of variability (since an interferometer acts as
231: a spatial filter on the emission), and to extract periodicity.
232:
233: Preliminary studies involving the analysis of simulated data from
234: expected models are important for a number of reasons. Such studies
235: will highlight the abilities and limitations of millimeter VLBI in
236: regards to detecting the signatures of the physical processes in the
237: accretion disk surrounding the black hole in Sgr~A*. Critical
238: resources (such as stable frequency standards, high-bandwidth
239: recording equipment, and phased array processors) are likely to be
240: limited initially, and telescope upgrades (such as surface accuracy
241: improvement, expanded IF bandwidth, additional receiver bands, and
242: simultaneous dual-polarization capability) must necessarily be prioritized.
243: Simulated observational data can help assess the tradeoffs that must be
244: considered for optimization of Galactic Center VLBI observations. The ultimate
245: goal is to explore the potential of black hole parameter estimation by present
246: and future millimeter VLBI observations.
247:
248: In this manuscript, we explore the observational signatures of an orbiting hot
249: spot embedded in a quiescent disk around Sgr~A*. We consider a non-imaging
250: approach to analyzing millimeter VLBI data for several reasons. First, one
251: fundamental assumption of Earth rotation aperture synthesis is that the source
252: structure is not changing. Since orbital periods in the hot spot models are
253: much shorter than the rotation of the Earth, this assumption is clearly
254: violated. Second, phase-referencing the data is presently not feasible at
255: millimeter wavelengths since the phase path through the atmosphere changes on a
256: time scale which is faster than the time needed to move the antennas between
257: the reference source and Sgr A*. We note, though, that this problem could be
258: circumvented at connected-element arrays where some antennas could be dedicated
259: to simultaneously observing a reference source while others observe Sgr A*, but
260: this is also currently limited by the low SNR that can be achieved in the
261: coherence time of the atmosphere. Potential arrays for millimeter VLBI will
262: have a small number of telescopes, initially as few as three, possibly with
263: vastly different sensitivies. Low expected signal-to-noise ratios (SNRs) on
264: some baselines combined with the few antennas available will prevent adequate
265: self-calibration via closure relations. Third, even if the visibility data
266: could be properly calibrated and the source structure were not changing over
267: the observation, the $(u,v)$-plane would be sparsely populated with noisy data
268: points, resulting in poor image fidelity, as shown by the simulations of
269: \citet{miyoshi04}. At least initially, it will be more productive to analyze
270: the data by model fitting in the visibility domain rather than in the image
271: domain.
272:
273:
274: \section{Models of Sgr A*}
275: \label{models}
276:
277: Models for the flaring emission of Sgr~A* at millimeter wavelengths
278: necessarily require a number of components, most succinctly decomposed
279: into models for the quiescent emission and models for the
280: short-timescale dynamical phenomena responsible for the flare. Any
281: such model has a number of existing observational constraints that it
282: must meet, including reproducing the observed spectra \& polarization
283: properties of the quiescent \& flaring emission as well the dynamical
284: properties of the flare light curves. Here we describe a set of flare
285: models involving orbiting hot spots embedded within a large-scale
286: accretion flow that are consistent with all existing observations,
287: based upon those described in \citet{broderick06b}.
288:
289: \subsection{Quiescent Emission}
290:
291: As implied by its spectrum, Sgr~A* is only starting to become optically thin at
292: millimeter wavelengths. Due to relativistic effects this does not
293: happen isotropically \citep[e.g.,][]{broderick06a}. As a consequence,
294: the opacity of the underlying accretion flow is important for both
295: imaging the black hole's silhouette and for the variability arising
296: from hot spots on compact orbits.
297:
298: Despite being faint compared to the Eddington luminosity for
299: a $4\times10^6\,\Ms$ black hole, Sgr~A* is still considerably bright,
300: emitting a bolometric luminosity of approximately $10^{36}\,\erg\,
301: \s^{-1}$. As a consequence it has been widely accepted that Sgr~A*
302: must be accretion powered, implying a minimum accretion rate of at
303: least $10^{-10}\,\Ms\, \yr^{-1}$.
304:
305: It is presently unclear how this emission is produced. This is
306: evidenced by the variety of models that have been proposed to explain
307: the emission characteristics of Sgr A*
308: \citep[e.g.,][]{narayan98,blandford99,falcke00,yuan02,yuan03,loeb07}.
309: Models in which the emission arises directly from the accreting
310: material have been subsumed into the general class of Radiatively
311: Inefficient Accretion Flows (RIAF), defined by the generally weak
312: coupling between the electrons, which radiate rapidly, and the ions,
313: which efficiently convert gravitational potential energy into heat
314: \citep{narayan98}. This coupling may be sufficiently weak to allow
315: accretion flows substantially in excess of the $10^{-10}\,\Ms\,
316: \yr^{-1}$ required to explain the observed luminosity with a canonical
317: radiative efficiency.
318:
319: Nevertheless, following the detection of polarization from Sgr A*
320: above $100\,\GHz$ \citep{aitken00,bower01,bower03,marrone06}, and
321: subsequent measurements of the Faraday rotation measure
322: \citep{macquart06,marrone07b}, the accretion rate near the black hole
323: has been inferred to be significantly less than the Bondi rate,
324: implying the existence large-scale outflows \citep{agol00,quataert00}.
325: Therefore, in the absence of an unambiguous theory, we adopt a simple,
326: self-similar model for the underlying accretion flow which includes
327: substantial mass-loss.
328:
329: Following \citet{yuan03}, this model is
330: characterized by a Keplerian velocity distribution, a population of
331: thermal electrons with density and temperature
332: \begin{equation}
333: n_{e,\rm th} = n^0_{e,\rm th} \left(\frac{r}{\Rs}\right)^{-1.1} \e^{-z^2/2 \rho^2}
334: \quad {and} \quad
335: T_{e} = T^0_{e} \left(\frac{r}{\Rs}\right)^{-0.84}\,,
336: \end{equation}
337: respectively, a population of non-thermal electrons
338: \begin{equation}
339: n_{e,\rm nth} = n^0_{e,\rm nth} \left(\frac{r}{\Rs}\right)^{-2.9} \e^{-z^2/2 \rho^2}\,,
340: \end{equation}
341: and spectral index $\alpha_{\rm disk} = 1.25$ (defined as
342: $S\propto\nu^{-\alpha_{\rm disk}}$), and a toroidal magnetic field in
343: approximate ($\beta=10$) equipartition with the ions (which produce
344: the majority of the pressure), i.e.,
345: \begin{equation}
346: \frac{B^2}{8\pi}
347: =
348: \beta^{-1} n_{e,\rm th} \frac{m_p c^2 \Rs}{12 r}\,.
349: \end{equation}
350: In all of these expressions the radial structure was taken directly
351: from \citet{yuan03} and the vertical structure was determined by
352: assuming the disk height is comparable to the polar radius, $\rho$.
353: To correct for the fact that \citet{yuan03} was a Newtonian study, we
354: determine the three coefficients ($n^0_{e,\rm th}$, $T^0_e$ and
355: $n^0_{e,\rm nth}$) by fitting the the radio, submillimeter and
356: near-infrared spectrum of Sgr~A*. For every inclination and black
357: hole spin presented here this was possible with extraordinary accuracy
358: (reduced $\chi^2<1$ in all cases and $\lesssim0.2$ for many), implying
359: that this model is presently significantly under-constrained by the
360: quiescent spectrum alone. It is also capable of producing the Faraday
361: rotation measures observed, and thus the polarimetric properties of
362: Sgr~A*.
363:
364: The primary emission mechanism is synchrotron, arising from both the
365: thermal and non-thermal electrons. We model the emission from the
366: thermal electrons using the emissivity described in \citet{yuan03},
367: appropriately altered to account for relativistic effects \citep[see,
368: e.g.,][]{broderick04}. Since we perform polarized radiative
369: transfer via the entire complement of Stokes parameters, we employ the
370: polarization fraction for thermal synchrotron as derived in
371: \citet{petrosian83}. In doing so we have implicitly assumed that the
372: emission due to thermal electrons is isotropic, which while generally
373: not the case is unlikely to change our results significantly. For the
374: non-thermal electrons we follow \citet{jones77} for a power-law
375: electron distribution, cutting the electron distribution off below a
376: Lorentz factor of $10^2$ and corresponding to a spectral index of
377: $\alpha_{\rm disk}=1.25$, both roughly in agreement with the
378: assumptions in \citet{yuan03}. For both the thermal and non-thermal
379: electrons the absorption coefficients are determined directly via
380: Kirchoff's law.
381:
382:
383: \subsection{Flares}
384:
385: We model the flare emission by a localized over-density in the
386: non-thermal electron distribution. This naturally explains the
387: short-timescale variability and potential periodicity claimed in some
388: flares \citep{genzel03,belanger06}. Such a feature is also a natural
389: consequence of dissipation in black hole accretion flows.
390:
391: Sgr~A*'s quiescent radio spectrum appears to require a population of
392: non-thermal electrons. If strong magnetic turbulence is present,
393: driven by, e.g., the magnetorotational instability, the production of
394: non-thermal electrons at strong shocks and magnetic reconnection
395: events is unavoidable. Generally, we expect that the production of
396: non-thermal electrons will be most prominent in the innermost regions
397: of the accretion flow, where the magnetic turbulence is strongest.
398:
399: It is important to note that the region producing the flare need not
400: be dynamically important. Within the context of a RIAF model for the
401: accretion flow onto Sgr~A*, the pressure is overwhelmingly dominated
402: by the ions. This is a direct consequence of the assumed weak
403: coupling between the electrons and ions, and thus the low luminosity
404: of the accretion flow. For typical RIAF accretion rates
405: ($10^{-8}\,\Ms\, \yr^{-1}$), the luminosity would need to be increased
406: by orders of magnitude before the non-thermal electrons become
407: dynamically significant. For the observed flares, which typically do
408: not increase the NIR luminosities by more than a single order of
409: magnitude, this means that the accelerated electrons will be frozen
410: into the accretion flow. In this case, the size of the emitting
411: region, $2 \Delta r$, is determined by the scale of the magnetic
412: turbulence.
413:
414: The dominant constraints upon the lifetimes of hot spots in the
415: accretion flow are Keplerian shear and synchrotron cooling. Hot spots
416: will shear apart on roughly $r/\Delta r$ orbital periods, and thus
417: small spots will last many orbits. Unlike shear, cooling is not
418: achromatic, with the flare cooling more rapidly at higher frequencies.
419: The cooling time is approximately
420: \begin{equation}
421: \tau_{\rm c}
422: \simeq
423: 3 \left(\frac{\lambda}{1\,\mm}\right)^{1/2} \left(\frac{B}{30\,
424: \G}\right)^{-3/2}\,\hr\,,
425: \label{eqn-cooling}
426: \end{equation}
427: which should be compared to the period at the ISCO, ranging from
428: $30\,\min$ for a non-rotating black hole to $4\,\min$ for a maximally
429: rotating black hole. Thus we expect that hot spots will typically
430: survive many orbits at millimeter and submillimeter wavelengths.
431:
432: Our hot-spot model consists of a locally spherically symmetric,
433: Gaussian over-density of non-thermal electron. Explicitly, given
434: $\Delta x^\mu = x^\mu-x_{\rm spot}^\mu$ and a scale, $R_{\rm spot}=0.75\Rs$,
435: the hot-spot density is
436: \begin{equation}
437: n_{e,\rm spot} = n^0_{e,\rm spot}
438: \exp\left[-\frac{\Delta x^\mu \Delta x_\mu + \left(u_{\rm spot}^\mu
439: \Delta x_\mu\right)^2}{2 R_{\rm spot}^2} \right],
440: \end{equation}
441: where the hot-spot four-velocity, $u_{\rm spot}^\mu$ is assumed to be
442: the same as that of the underlying disk (which we have chosen to be
443: Keplerian). Our description of the hot spot is completed by the
444: spectral index of the power-law distribution of electrons,
445: $\alpha_{\rm spot}=1.3$, taken from observations of NIR
446: flares \citep{eckart04}. Like the disk, the hot-spot radiates
447: primarily via synchrotron.
448:
449:
450:
451: \subsection{Generating Images}
452:
453: The method by which images of the flaring disk are produced is
454: discussed at length in \citet{broderick03} \citep[see
455: also,][]{broderick06a,broderick06b}. As a result, we only briefly
456: summarize the procedure here.
457:
458: Null geodesics are constructed by integrating a Hamiltonian
459: formulation of the geodesic equations:
460: \begin{equation}
461: \frac{dx^\mu}{d\eta} = f(x^\sigma) \, k^\mu
462: \quad{\rm and}\quad
463: \frac{dk_\mu}{d\eta} = - \left.\frac{f(x^\sigma)}{2}\frac{\partial k^\nu
464: k_\nu}{\partial x^\mu}\right|_{k_\alpha}\,,
465: \end{equation}
466: where $f(x^\sigma)$ is an arbitrary function, corresponding to the
467: freedom inherent in the affine parametrization, $\eta$. In order to
468: regularize the affine parametrization near the horizon, we choose
469: \begin{equation}
470: f(x^\sigma) = r^2\sqrt{1-\frac{r}{r_{\rm h}}},
471: \quad{\rm where}\quad
472: r_{\rm h } = \frac{\Rs}{2}\left( 1 + \sqrt{1-a^2} \right)
473: \end{equation}
474: is the horizon radius ($a$ is the dimensionless black hole
475: spin). It can be explicitly shown that this does
476: reproduce the null geodesics \citep{broderick03}.
477:
478: The relativistic generalization of the radiative transfer problem is
479: most easily obtained by directly integrating the Boltzmann equation
480: directly \citep{lindquist66,broderick06a}. In this case, it is the
481: photon distribution function, $N_\nu\propto I_\nu/\nu^3$ (which has
482: the virtue of being a Lorentz scalar), that is evolved. In the case
483: of polarized transfer, it is possible to define the covariant
484: analogues of the Stokes parameters, ${\bf N}_\nu = \left( N_\nu,
485: N^Q_\nu, N^U_\nu, N^V_\nu \right)$, in terms of a parallel propagated
486: tetrad \citep{broderick04}. In terms of these, the radiative transfer
487: equation takes on its standard form:
488: \begin{equation}
489: \frac{d{\bf N}_\nu}{d\eta}
490: =
491: {\bf \bar{j}}_\nu - \bmath{ \bar{\alpha}}_\nu {\bf N}_\nu\,,
492: \end{equation}
493: where ${\bf \bar{j}}$ and $\bmath{\bar{\alpha}_\nu}$ are the
494: appropriately generalized emission and absorption coefficients, and
495: may be trivially related to the same quantities in the local plasma
496: frame \citep{lindquist66,broderick04}.
497:
498: Images are produced by tracing a collection of initially parallel null
499: geodesics from pixels on a distant plane backwards in time, towards
500: our model of Sgr~A*, terminating the ray when it has been captured by
501: the black hole, escaped the system, or accrued an optical depth
502: greater than $10$. Along each ray we integrate the polarized
503: radiative transfer, obtaining ${\bf N}_\nu$ at the original plane. We
504: construct 100 such images with resolutions of $128\times128$ pixels
505: for each hot-spot orbit. This procedure is repeated for each
506: frequency of interest, for which we keep the underlying physical model
507: fixed. Thus, for each spin/inclination pair, the relationship between
508: 230~GHz and 345~GHz images is dictated by the spectral properties of
509: the source.
510:
511:
512: \section{VLBI Analysis}
513: \label{VLBI}
514:
515: \subsection{Techniques}
516: \label{techniques}
517:
518: Typical VLBI analysis techniques employ an iterative
519: ``self-calibration'' loop, whereby a sky brightness model, the VLBI
520: data, and a series of complex gains for each VLBI site are brought
521: into convergence \citep{cornwell81,cornwell82}. This process relies
522: on the assumption that all array calibration can be expressed as
523: station-based gains, which is equivalent to requiring that all
524: calibration errors 'close', or cancel when computed over suitable
525: closed loops of VLBI baselines. In almost all cases, this assumption
526: is valid, and closure quantities can be constructed from the data,
527: which contain structural information on the observed source, but that
528: are largely immune to station-based phase and gain errors
529: \citep{jennison58,twiss60}. However, closure quantities alone are
530: insufficient to determine baseline phases for a VLBI array. The
531: number of independent closure phases computed over closed triangles of
532: VLBI stations, for example, grows as $\onehalf \,(N-1)\,(N-2),$ where
533: $N$ is the number of antennas, while the number baseline phases in the
534: array will be $\onehalf \, N \,(N-1).$ The fraction of visibility
535: phase information available from closure phases is thus $(N-2)/N$.
536: For millimeter VLBI $N$ will be small (initially $N = 3$ or $4$; \S
537: \ref{antennas}), and closure quantities will not be sufficiently
538: numerous to allow for full calibration of the data. However, closure
539: quantities are robust observables and can therefore be used for model
540: fitting even when baseline-based visibilities are contaminated with
541: station-based phase and gain errors. Indeed, closure quantities have
542: been successfully used for model fitting \citep[e.g.,][]{rogers74},
543: including recent experiments to place limits on the apparent size and
544: structure of Sgr~A* at wavelengths as short as 3~mm
545: \citep{doeleman01,bower04,shen05,markoff07}.
546:
547: Closure phase and amplitude in the weak-detection limit are discussed
548: in detail by \citet{rogers95}. We summarize the most relevant
549: information below.
550:
551: The closure phase is the sum of the baseline phases along a triangle
552: of antennas. It is independent of instrumental and atmospheric
553: complex gain terms. The closure phase of a symmetric distribution of
554: emission is always zero or 180\degr\ \citep[e.g.,][]{monnier07}.
555: Deviations from these values are indicative of asymmetries (about the
556: origin) in the source structure on the size scales probed by the
557: baselines of the triangle. A static asymmetric source structure will
558: show slow variations in closure phase over the course of observations
559: due to the rotation of the Earth, which changes the projected length
560: and orientation of baselines. If the source structure is changing, as
561: would be the case for an orbiting hot spot, the closure phase on some
562: triangles will change as well. Because the time scale of hot spot
563: orbits near the black hole is on the order of tens of minutes
564: (depending on the mass and spin of the black hole and the orbital
565: radius of the hot spot), closure phase variability will be much more
566: rapid for the case of a hot spot embedded in a disk as compared to a
567: quiescent disk alone. Since a hot spot may survive for
568: several orbits before cooling or shearing (\S \ref{models}), closure
569: phases can exhibit approximate perodic behavior over several cycles.
570: Thus, closure phases are appropriate observables for detecting
571: periodic source structure changes on short time scales. The SNR of
572: the closure phase is dominated by the lowest SNR of the visibility
573: data along the three segments.
574:
575: The closure amplitude is constructed as the ratio of products of
576: visibility amplitudes ($A$) along a quadrangle of antennas. For four
577: antennas $a$, $b$, $c$, and $d$, two independent closure amplitudes
578: can be constructed:
579: \begin{equation}
580: A_{abcd} \equiv \frac{|A_{ab}| |A_{cd}|}{|A_{ac}| |A_{bd}|}
581: \quad \mathrm{and} \quad
582: A_{adbc} \equiv \frac{|A_{ad}| |A_{bc}|}{|A_{ac}| |A_{bd}|}.
583: \end{equation}
584: For an array of $N$ antennas, there are $\onehalf\,N\,(N-3)$ independent
585: closure amplitudes. As for closure phase, the closure amplitude is
586: unaffected by gain calibration errors on each of the antennas.
587: Deviations of the closure amplitude from unity are usually indicative
588: of resolved, asymmmetric source structure. Large excursions in
589: closure phase and amplitude are often correlated, since small changes
590: with time in the complex visibility, due to Earth rotation or source
591: structure changes, have the largest effects on the phase when the
592: visibility amplitude is near zero. Since a small visibility amplitude
593: results in a very small closure amplitude if it appears in the
594: numerator or a very large closure amplitdue if it appears in the denominator.
595: For this reason, closure amplitudes are generally presented on logarithmic
596: scales.
597:
598: In the low-signal regime, the closure amplitude is a biased quantity.
599: The detected visibility amplitude on a baseline is the modulus of the
600: vector (amplitude and phase) sum of the signal and the noise and is by
601: definition nonnegative. When the SNR is small, the visibility
602: amplitude is dominated by the noise amplitude and is therefore larger
603: than the signal amplitude. Thus, if the source is not detected on
604: baseline $ab$ but is detected on baselines $cd$, $ac$, and $bd$, the
605: closure amplitude $A_{abcd}$ will on average be larger than the value
606: predicted by a noiseless model of the source structure. The closure
607: phase does not suffer from a similar bias.
608:
609:
610: \subsection{Antennas}
611: \label{antennas}
612:
613: \input{tab1.tex}
614:
615: \begin{figure}
616: \resizebox{\hsize}{!}{\includegraphics{f1a.eps}\includegraphics{f1b.eps}}
617: \caption{Locations of candidate telescopes for millimeter-wavelength
618: VLBI as viewed from the declination of Sgr~A*.
619: \label{fig-locations}
620: \notetoeditor{For the reader's convenience, Figure 1 should be placed
621: near Table 1, preferably on the same page}
622: }
623: \end{figure}
624:
625: Whereas centimeter-wavelength astronomy can be performed from
626: virtually any location with a clean spectrum, potential sites for
627: (sub)millimeter-wavelength astronomy are limited by the need to be
628: above most of the water vapor content in the atmosphere.
629: Consequently, possible arrays for millimeter VLBI are sparser
630: than for centimeter VLBI. In this section, we summarize the best
631: prospects for millimeter VLBI among existing telescopes and those that
632: may come on line in the near future.
633:
634: \emph{Hawaii}: The Caltech Submillimeter Observatory (CSO), James
635: Clerk Maxwell Telescope (JCMT), and Submillimeter Array (SMA) are all located atop Mauna
636: Kea. Each telescope can observe in both the 230 and 345~GHz bands.
637: The JCMT has a single-polarization 230~GHz receiver and a
638: dual-polarization 345~GHz receiver. The SMA presently does not
639: support simultaneous dual-polarization observations but is expected to
640: do so eventually at 345~GHz. Development of instrumentation to phase
641: together the CSO, JCMT and several SMA dishes is underway, and will
642: result in an effective 23~m aperture, which we refer to as ``Hawaii,''
643: by early 2009.
644:
645: \emph{CARMA}: The Combined Array for Research in Millimeter-wave
646: Astronomy (MWA) consists of six 10.4~m antennas and nine 6.1~m
647: antennas. The ability to phase together up to 8 CARMA dishes
648: for VLBI is planned, and will result in an effective 27~m aperture.
649: CARMA can presently observe in the 230~GHz band. A future upgrade to
650: include 345~GHz is planned, but it is unclear when this band will be
651: available for observations. We also consider a single 10.4~m dish at
652: 230~GHz (``CARMA 1'') in \S \ref{results-arrays}.
653:
654: \emph{SMTO}: The Arizona Radio Observatory Submillimeter Telescope (SMT) on Mt.
655: Graham in Arizona is a 10~m dish capable of observing at both 230 and 345~GHz.
656:
657: \emph{LMT}: The Large Millimeter Telescope (LMT), presently under
658: construction, will be a 50~m dish capable of observing at both the 230
659: and 345~GHz bands. When complete, it will be the most sensitive
660: millimeter telescope in the Northern hemisphere. In anticipation that
661: the LMT collecting area will be installed in phases, we conservatively adopt
662: an effective aperture of 32~m.
663:
664: \emph{Chile}: Several millimeter telescopes are available in Chile.
665: The Atacama Submillimeter Telescope Experiment (ASTE) is a 10~m dish
666: with a single-polarization receiver at 345~GHz. The Atacama
667: Pathfinder Experiment (APEX) is a 12~m dish with a double-sideband
668: (DSB) at 345~GHz. Future two sideband (2SB) heterodyne receivers
669: capable of observing at 230~GHz and 345~GHz are under construction.
670: We refer to using one of these facilities as ``Chile 1.'' ALMA will
671: be composed of a large number of 12~m dishes with 2SB receivers at
672: both 230 and 345~GHz. We also consider the possibility of a
673: 10-element phased ALMA station as the Chilean site (``Chile 10''),
674: since our models do not produce much expected signal on the long
675: baselines to Chile (\S \ref{methods}).
676:
677: \emph{Pico Veleta}: The 30~m Institut de Radioastronomie
678: Millim\'{e}trique (IRAM) telescope on Pico Veleta (PV) can observe at
679: 230~GHz.
680:
681: \emph{Plateau de Bure}: The IRAM PdB Interferometer consists of six
682: 15~m telescopes equipped with 230~GHz receivers. An upgrade to add
683: 345~GHz capability is presently under construction and is expected to
684: be available in late 2008. The six telescopes can currently be phased
685: up over a 256~MHz bandwidth into a single 37~m-equivalent aperture.
686: Extension to higher bandwidths will require instrumentation
687: development similar to what will be deployed to phase up the antennas
688: on Mauna Kea.
689:
690: A summary of telescope capabilities is given in Table~\ref{tab-ants}.
691: Sensitivity is given in terms of the system equivalent flux density
692: (SEFD), which is equal to $2 k T_\mathrm{sys}/A_\mathrm{eff}$, where
693: $k$ is Boltzmann's constant and the effective area $A_\mathrm{eff}$ is
694: the product of the geometric area and the aperture efficiency.
695: Assumed $T_\mathrm{sys}$ values toward Sgr~A* include a typical
696: expected atmospheric contribution, which can be quite large for
697: Northern hemisphere telescopes due to the low elevation of the
698: Galactic Center. Since the atmospheric contribution is highly
699: weather-dependent, actual observations may achieve significantly
700: different values of the SEFD at each telescope. SEFD values include a
701: 10\% phasing loss factor for phased arrays. All facilities can in
702: principle provide an IF bandwidth of at least 4~GHz, although
703: accessing the full bandwidth at some stations may be problematic.
704: Thus, 32~Gbit\,s$^{-1}$, corresponding to 2-bit Nyquist samples of a
705: 4~GHz bandwidth in two orthogonal polarizations, is the maximum
706: possible recording rate that we will consider. At present
707: 16~Gbit\,s$^{-1}$ digital back ends are still in the planning stage,
708: so initial observations will likely employ a smaller bandwidth.
709: Except as noted, all telescopes can observe two polarizations
710: simultaneously.
711:
712: The European telescopes (PV and PdB) and the North American telescopes
713: (Hawaii, CARMA, SMTO, and LMT) have no mutual visibility of Sgr~A*,
714: except for approximately 1~hr of overlap above 10\degr\ elevation on
715: the PV-LMT baseline. Thus, there are only three possible types of
716: subarrays with at least three elements possible among these millimeter
717: facilities: North America only, North America plus Chile, and Europe
718: plus Chile (Fig.~\ref{fig-locations}). For our simulations, we take our 230~GHz array to consist
719: of Hawaii, CARMA, SMTO, LMT (32~m), either APEX or a 10-element phased
720: ALMA, PV, and PdB. Early science at 345~GHz will likely consist of
721: the single triangle of Hawaii, SMTO, and a Chilean telescope.
722: However, we consider the same set of telescopes as at 230~GHz in order
723: to illustrate what future 345~GHz upgrades might accomplish.
724:
725:
726: \subsection{Methods}
727: \label{methods}
728:
729: Simulated data were obtained using task UVCON in the Astronomical
730: Image Processing System (AIPS). Synthetic data were produced using an
731: averaging interval of 10~s. Typical coherence times at 230~GHz are 10~s, but
732: can be as low as 2-4~s and, under good weather conditions, as long as 20~s
733: \citep{doeleman02}. At the ALMA site in Chile, the measured coherence time of
734: the atmosphere is > 10 seconds 60\% of the time at 230GHz and 45\% of the time
735: at 345GHz \citep{holdaway97}. Since AIPS cannot directly handle time-varying
736: source structure, for a hot spot model we simulate data from 100 static,
737: equally-spaced time slices in a single hot spot orbit and construct a data set.
738: All data assume 2-bit sampling at the Nyquist rate. Thus, the recording rate
739: (in Gbit\,s$^{-1}$) is four times the observing bandwidth (in GHz). For
740: continuum observations at a constant sampling rate, 2-bit quantization achieves
741: sensitivity levels very nearly approaching that of 1-bit quantization at half
742: the observing bandwidth \citep{thompson01}, which is limited by the hardware at
743: certain telescopes (\S \ref{antennas}).
744:
745: Observations of total intensity (i.e., Stokes $I$) theoretically
746: obtain identical noise levels regardless of whether the data are taken
747: at full bandwidth in single-polarization mode or at half bandwidth in
748: dual-polarization mode, provided that circularly-polarized feeds are
749: used. The hot spot models do not produce circular polarization
750: (Stokes $V$) but do produce large amounts of linear polarization
751: (Stokes $Q$ and $U$), especially on small spatial scales.
752: Circularly-polarized feeds measures Stokes visibilities $I \pm V$ and
753: therefore are identical for our models. Dual-polarization
754: observations are essential if linearly-polarized feeds are used, since
755: the parallel-hand data are sensitive to the Stokes visibilities $I \pm
756: P$, where $P$ is a linear combination of Stokes $Q$ and $U$ (depending
757: on feed orientation). In practice the observer will usually prefer to
758: observe in dual-polarization mode when available even using circular
759: feeds, since cross-hand correlation products provide polarimetric
760: information as well. The polarimetric products are of special
761: interest, since they may show asymmetries not seen in total intensity
762: \citep[e.g.,][]{bromley01}. However, we focus on total intensity
763: methods and results in this work and defer polarimetric considerations
764: to a future manuscript.
765:
766: In the subsequent discussion, we consider a suite of models
767: parameterized by black hole spin ($a = 0$ nonrotating or $a = 0.9$
768: highly rotating), disk orientation, and hot spot orbital radius ($r$).
769: For each spin, the radius of the ISCO ($r_{ISCO} = 6 \, R_G$ for $a =
770: 0$ and $2.32 \, R_G$ for $a = 0.9$)
771: and one larger radius are chosen; the larger radius for $a = 0.9$ is
772: chosen to have the same period as the $a = 0$, ISCO model. Disk
773: models assume that the spin axes of the black hole and accretion disk
774: are aligned, with the major axis of the projected disk aligned
775: east-west except in Model C, in which the disk and hot spot are
776: aligned north-south instead. Flux densities have been scaled to match
777: connected-element interferometric observations.
778: Contributions from the quiescent disk are assumed to be approximately
779: 3~Jy, depending slightly on the specific model parameters chosen, as
780: listed in Table~\ref{tab-models}. A single prograde orbiting hot spot
781: contributes a variable flux density component, depending on orbital
782: phase and the specific model, consistent with submillimeter
783: observations \citep{marrone07b}. The hot
784: spot contributes some flux density even at the minimum in the light
785: curve, an effect that is most pronounced in the $i = 30$\degr\ models.
786: All models are convolved with the expected interstellar scattering given by
787: \citep{bower06}. While these model parameters do not span the entire range of
788: possibilities of the Sgr A* disk system, they do show how changes in model
789: parameters affect the observable quantities. Non-imaging VLBI methods are
790: applicable to all reasonable models for the Sgr A* system. Figure~\ref{fig-modelb} shows
791: a series of images from Model B at both 230~GHz and 345~GHz.
792:
793: \input{tab2.tex}
794:
795: \begin{figure*}
796: \resizebox{0.19\hsize}{!}{\includegraphics{f2a.eps}}
797: \resizebox{0.19\hsize}{!}{\includegraphics{f2b.eps}}
798: \resizebox{0.19\hsize}{!}{\includegraphics{f2c.eps}}
799: \resizebox{0.19\hsize}{!}{\includegraphics{f2d.eps}}
800: \resizebox{0.19\hsize}{!}{\includegraphics{f2e.eps}}\\
801: \resizebox{0.19\hsize}{!}{\includegraphics{f2f.eps}}
802: \resizebox{0.19\hsize}{!}{\includegraphics{f2g.eps}}
803: \resizebox{0.19\hsize}{!}{\includegraphics{f2h.eps}}
804: \resizebox{0.19\hsize}{!}{\includegraphics{f2i.eps}}
805: \resizebox{0.19\hsize}{!}{\includegraphics{f2j.eps}}\\
806: \resizebox{0.19\hsize}{!}{\includegraphics{f2k.eps}}
807: \resizebox{0.19\hsize}{!}{\includegraphics{f2l.eps}}
808: \resizebox{0.19\hsize}{!}{\includegraphics{f2m.eps}}
809: \resizebox{0.19\hsize}{!}{\includegraphics{f2n.eps}}
810: \resizebox{0.19\hsize}{!}{\includegraphics{f2o.eps}}\\
811: \resizebox{0.19\hsize}{!}{\includegraphics{f2p.eps}}
812: \resizebox{0.19\hsize}{!}{\includegraphics{f2q.eps}}
813: \resizebox{0.19\hsize}{!}{\includegraphics{f2r.eps}}
814: \resizebox{0.19\hsize}{!}{\includegraphics{f2s.eps}}
815: \resizebox{0.19\hsize}{!}{\includegraphics{f2t.eps}}
816: \caption{Images of Model B. Columns show orbital phases 0, 0.2, 0.4,
817: 0.6, and 0.8 from left to right. From top to bottom, rows show the
818: input model at 230~GHz, the same model after convolution with
819: expected interstellar scattering at 230~GHz, the model at 345~GHz,
820: and the same model after scattering at 345~GHz.
821: \label{fig-modelb}
822: }
823: \end{figure*}
824:
825: The angular resolution available to potential millimeter VLBI arrays
826: is well matched to potentially interesting scales for observing Sgr~A*
827: (Fig.~\ref{fig-uvplt}). The longest baselines, Hawaii-Chile and
828: PdB-Chile, provide fringe spacings of 30--35~$\mu$as at 230~GHz and
829: 19--23~$\mu$as at 345~GHz, slightly larger than the expected
830: interstellar scattering and only several times $\Rs$. Our
831: models do not produce much detectable signal on these small angular
832: scales (Fig.~\ref{fig-uv}), but it is possible that smaller hot spots
833: or disk instabilities (not modelled) will produce greater amplitudes
834: on small angular scales. The shortest baselines, SMTO-CARMA and
835: PV-PdB, provide fringe spacings of 230--1000~$\mu$as at 230~GHz and
836: 160--700~$\mu$as at 345~GHz.
837:
838: \begin{figure}
839: \resizebox{\hsize}{!}{\includegraphics{f3.eps}}
840: \caption{Possible $(u,v)$ tracks for millimeter VLBI. Tracks are
841: labelled by baseline (\emph{A}: ALMA/APEX/ASTE, \emph{B}: Plateau de
842: Bure, \emph{C}: CARMA, \emph{H}: Hawaii, \emph{L}: LMT, \emph{S}:
843: SMTO, \emph{V}: Pico Veleta). Unlabelled tracks correspond to the
844: baseline indicated by $(-u,-v)$. Axes are in units of
845: gigawavelengths at $\nu = 230$~GHz.
846: \label{fig-uvplt}}
847: \end{figure}
848:
849: \begin{figure}
850: \resizebox{\hsize}{!}{\includegraphics{f4.eps}}
851: \caption{Plot of expected visibility amplitude as a function of
852: baseline length for two frames of Model B230. Symbols show
853: the noiseless visibility amplitudes that would be obtained if the
854: source structure were frozen near minimum (black) and maximum
855: (color) flux in the hot spot orbit. At maximum flux, points are
856: color-coded and labelled by baseline as in Figure~\ref{fig-uvplt}.
857: Visibility amplitude falls off rapidly with baseline length, and the
858: fractional variability on long baselines can be significantly
859: different than that seen at zero spacing.
860: \label{fig-uv}}
861: \end{figure}
862:
863: \section{Results}
864: \label{results}
865:
866: \subsection{Closure Phases and Amplitudes}
867:
868: \begin{figure*}
869: \resizebox{\hsize}{!}{\rotatebox{-90}{\includegraphics[]{f5_gimp.eps}}}
870: \caption{Closure phases on selected triangles at 230~GHz. The solid
871: line (red in the online edition) shows the predicted closure phase
872: in the absence of noise. Each point indicates 10~s of
873: coherently-integrated data. The fourth column shows the same
874: triangle as the third column but with a higher data rate and
875: substitution of Chile~10 for Chile~1. The same 2-hour period,
876: corresponding to 4.5 periods (14.8 periods for Model E) is
877: shown in all panels excepting PV-PdB-Chile~10.
878: \label{fig-phase230}}
879: \end{figure*}
880:
881: Figure~\ref{fig-phase230} shows predicted closure phases at 230~GHz on
882: selected triangles. The closure phase responses are highly dependent
883: on the physical parameters of the Sgr~A* system, but several patterns
884: emerge. Data from the smallest triangles (i.e., those composed of
885: only North American telescopes) achieve a high SNR even at the present
886: maximum recording rate of 4~Gbit\,s$^{-1}$. Models with a greater
887: north-south extent (e.g., Models A and C), produce a larger closure
888: phase signature on the small triangles. The SNR on all triangles to
889: Chile~1 is low owing to the small flux on small angular scales.
890: Higher bit rates and the substitution of phased ALMA for Chile~1
891: greatly increase detectability, as shown by the third and fourth
892: columns, which correspond to the expected array and recording
893: capabilities in the next few years. Net offsets from zero closure
894: phase on some triangles result from apparent asymmetric structure in
895: the disk caused by lensing and opacity effects.
896:
897: Figure~\ref{fig-phase345} shows modelled closure phases at 345~GHz.
898: The only available triangle in the near future at 345~GHz will be
899: Hawaii-SMTO-Chile~1, which produces rather low SNR closure phases at
900: 4~Gbit\,s$^{-1}$. Of the remaining telescopes, both CARMA and Plateau
901: de Bure plan for 345~GHz receivers in the future. In addition to
902: these, we show triangles including the LMT and Pico Veleta, neither of
903: which has planned 345~GHz capability, to demonstrate what might be
904: seen in the event of future upgrades to those facilities. At both
905: 230~GHz and 345~GHz, closure phases on several baseline triangles show
906: clear evidence for periodicity associated with hot-spot orbits.
907:
908: \begin{figure*}
909: \resizebox{\hsize}{!}{\rotatebox{-90}{\includegraphics[]{f6_gimp.eps}}}
910: \caption{Closure phases on selected triangles at 345~GHz. See
911: Figure~\ref{fig-phase230} for details.
912: \label{fig-phase345}}
913: \end{figure*}
914:
915: Four telescopes are required to obtain a closure amplitude, which
916: effectively means that closure amplitudes can only be measured on
917: Western hemisphere arrays. Figures~\ref{fig-amp230} and
918: \ref{fig-amp345} show closure amplitudes on selected quadrangles at
919: 230 and 345~GHz, respectively. Periodicity associated with the
920: orbiting hot-spot is evident on most quadrangles. In some models,
921: closure amplitudes including the Hawaii-Chile~1 baseline show bias (as
922: described in \S \ref{techniques}) due to the low visibility amplitude
923: on this baseline. The substitution of phased ALMA for Chile~1
924: combined with higher recording rates suffice to clearly detect Sgr~A*
925: on this baseline, resulting in unbiased closure amplitudes. It is
926: possible (and necessary for model fitting) to de-bias closure
927: amplitudes by correcting the visibility amplitudes for the expected
928: noise levels, but the procedure can be difficult when the SNR on a
929: baseline is low \citep[for more details, see][]{trotter98}. In any
930: case, the presence of bias does not hinder detection of periodicity.
931:
932:
933: \begin{figure*}
934: \resizebox{\hsize}{!}{\rotatebox{-90}{\includegraphics[]{f7_gimp.eps}}}
935: \caption{Closure amplitudes on selected quadrangles at 230~GHz. See
936: Figure~\ref{fig-phase230} for details.
937: \label{fig-amp230}}
938: \end{figure*}
939:
940: \begin{figure*}
941: \resizebox{\hsize}{!}{\rotatebox{-90}{\includegraphics[]{f8_gimp.eps}}}
942: \caption{Closure amplitudes on selected quadrangles at 345~GHz. See
943: Figure~\ref{fig-phase230} for details.
944: \label{fig-amp345}}
945: \end{figure*}
946:
947:
948: \subsection{Autocorrelation Functions}
949: \label{ACF}
950:
951: Signatures of time periodic structure associated with orbiting
952: hot-spots can be derived from autocorrelations of closure quantity
953: time series. The autocorrelation function of a time series of $n$
954: closure amplitudes $A$ on a quadrangle of telescopes is given by
955: \begin{equation}
956: \mathrm{ACF}_A(k) \equiv \frac{1}{(n-k)\,\sigma^2}
957: \sum\limits_{i=1}^{n-k}\left[\left(\log A_i-\mu\right)\left(\log
958: A_{i+k}-\mu\right)\right],
959: \end{equation}
960: where $\mu$ and $\sigma^2$ are the mean and variance of the
961: distribution of the logarithm of the closure amplitudes, respectively.
962: The logarithm is used in preference to the closure amplitude itself
963: due to the tendency for the closure amplitude to obtain both very
964: small (near 0) and very large values. We also define a variant of the
965: autocorrelation function for closure phases, $\phi$:
966: \begin{equation}
967: \mathrm{ACF}_\phi(k) \equiv \frac{1}{n-k}
968: \sum\limits_{i=1}^{n-k}\cos\left(\phi_i-\phi_{i+k}\right).
969: \end{equation}
970: This definition has the advantage that it handles phase-wrap ambiguity
971: gracefully, since $\cos$ is a periodic function. The normalizations
972: are such that $\mathrm{ACF}(k) = 1$ when $k$ is an integer multiple of
973: the period for a noiseless periodic function. In practice, the peaks
974: of an ACF (other than the trivial peak at $k = 0$, where
975: $\mathrm{ACF}(0) = 1$ by definition) of closure phases or amplitudes
976: fall off with lag rather than returning precisely to unity, because
977: the projected baseline geometries change with Earth rotation. In the
978: case of long periodicities ($\gtrsim 2$~hr), this effect can be large
979: enough to obscure any periodicity at all on some
980: triangles/quadrangles. While the duration of some SgrA* flare events
981: exceed this time interval, there are claims of modulation within some
982: NIR flares with characteristic time scales of $\sim17$~min
983: \citep{genzel03}. From an observational perspective, periodicity is
984: not convincingly detected until at least two full periods (preferably
985: more) have been observed. The longest mutual visibility appears on
986: the SMTO-LMT-Chile triangle, which can see Sgr~A* for less than 7 full
987: hours. Triangles including Hawaii or Europe have significantly
988: smaller windows of mutual visibility. The distribution of millimeter
989: telescopes is not optimal for detecting slow periodic variability in
990: Sgr~A*.
991:
992: The largest nontrivial peak in the ACF indicates the period of the hot
993: spot orbit, as indicated in Figure~\ref{fig-acfs}, excluding the slow
994: periodic case. The ACF correctly identifies the period in all models
995: at all recording rates provided that the average SNR $\gtrsim 1$ and
996: that the triangle contains at least one baseline long enough to be
997: sensitive to the changing source structure. On the small triangles,
998: $\mathrm{ACF}_\phi(k)$ shows little variation as a function of lag
999: $k$, since the phase autocorrelation function of a nearly-constant
1000: function is itself nearly constant. However, small peaks are visible
1001: in the ACF on the Hawaii-SMTO-CARMA and Hawaii-CARMA-LMT triangles in
1002: Figure~\ref{fig-acfs}. The large SNR on the small triangles, due to
1003: the fact that the shorter baselines do not resolve out much of the
1004: source flux density, permits detectability for reasonable bandwidths.
1005:
1006: In addition to the peak at the period, the ACF has peaks at integer
1007: multiples of the period. In the weak-detection limit, this may be an
1008: important discriminator indicating that the detected period is real.
1009: While purely random noise may produce peaks in the ACF, there is no
1010: reason why it should produce periodic peaks. At the other end of time
1011: scales, intraperiod sub-peaks in the ACF may be produced depending on
1012: the details of the source structure and array geometry. Excluding
1013: long-period orbits, it is usually clear which peak indicates the
1014: orbital period of the hot spot, since the sub-peaks are of smaller
1015: amplitude. Nevertheless, it is possible that pathological cases exist
1016: in which the observer may misidentify the orbital period. For
1017: instance, two identical hot spots at the same radius but separated by
1018: 180\degr\ in azimuth would produce strong peaks in the ACF at both
1019: integer and half-integer multiples of the period, which would lead to
1020: a conclusion that the orbital period is half of its true value. A
1021: similar ambiguity could arise if the observed variability is produced,
1022: for instance, by the rotation of a pattern produced by a two-armed
1023: Rossby wave instability \citep[e.g.,][]{falanga07}.
1024:
1025: The key point about the ACF plots is that for the same model,
1026: \emph{all triangles and quadrangles indicate the same period}. It
1027: will be important to observe with as many telescopes as possible
1028: simultaneously, since the extra information provided by the additional
1029: telescopes may be important for detecting or confirming
1030: marginally-detected variability. Three antennas provide a single
1031: closure phase, for a total of 1 ACF; four antennas provide three
1032: independent closure phases and two independent closure amplitudes, for
1033: a total of 5 independent ACFs; and five antennas provide six closure
1034: phases and five closure amplitudes, for a total of 11 independent
1035: ACFs. Even if the source is not detected on one long baseline,
1036: resulting in a biased closure amplitude, the ACF may still clearly
1037: indicate periodicity (e.g., the Hawaii-Chile~1-CARMA-SMTO panel for
1038: Model A230 in Fig.~\ref{fig-acfs}).
1039:
1040: \begin{figure*}
1041: \resizebox{\hsize}{!}{\rotatebox{-90}{\includegraphics{f9.eps}}}
1042: \caption{Autocorrelation function plots of selected triangles and
1043: quadrangles. Representative models are shown. Black, red, and blue
1044: lines indicate ACFs for noiseless data, 16~Gbit\,s$^{-1}$, and
1045: 4~Gbit\,s$^{-1}$, respectively. Panels marked with a red star show
1046: $\mathrm{ACF}_\phi$ on a different ordinate scale, as indicated to
1047: the left in that row; other panels in the same row are on the scale
1048: show in the upper left panel. The dotted line shows the period of
1049: the hot spot orbit.
1050: \label{fig-acfs}}
1051: \end{figure*}
1052:
1053:
1054: \subsection{Detecting Periodicity with Likely Arrays}
1055: \label{results-arrays}
1056:
1057: We consider potential observing arrays that may be employed in
1058: observations of Sgr~A*. The minimum number of telescopes required to
1059: produce a closure quantity is 3, and the maximum number of telescopes
1060: located at substantially different sites with mutual visibility of
1061: Sgr~A* is 5. At 230~GHz, we consider western hemisphere arrays
1062: consisting of the US triangle; US and Chile-1; US, Chile-1, and the
1063: LMT; and the same array with Chile-10 instead. We also consider
1064: European-Chile arrays. Since PV is not expected to have 345~GHz
1065: capability in the near future, we expect that Chile-10 will exist
1066: before 345~GHz observations are possible on the Europe-Chile triangle.
1067: Likewise, we expect that western hemisphere arrays will include
1068: Chile-10 before the LMT is available at 345~GHz, since 345~GHz
1069: capability is not currently planned at the LMT.
1070:
1071: For each array, we generate $\mathrm{ACF}_\phi$ for each triangle of
1072: three antennas and $\mathrm{ACF}_A$ for each quadrangle of four
1073: antennas. The ACFs can be combined to produce a single combined ACF
1074: of the entire array. For optimum detectability, it is necessary to
1075: weight the individual ACFs by the square of their effective
1076: $\mathrm{SNR}_\mathrm{ACF} \equiv X/\xi$, where $X$ and $\xi$ quantify
1077: the effective signal and noise in the ACF, respectively. For
1078: definiteness, we take $\xi$ as the rms deviation in the ACF after a
1079: 5-channel boxcar-smoothed SNR has been subtracted. We define $X
1080: \equiv \max(\mathrm{ACF}(k)) - \max(\min(\mathrm{ACF}(k)),0), k \neq
1081: 0$ as a measure of the range of the ACF. The quantity $X$ reflects
1082: the contrast of the ACF and is especially critical for proper
1083: weighting of the phase ACFs, since on many triangles
1084: $\mathrm{ACF}_\phi$ is approximately constant though with small
1085: periodic peaks (Fig.~\ref{fig-acfs}). Effectively, $X/\xi$ quantifies
1086: the significance of a peak in the ACF. In practice, $X$ is frequently
1087: small for $\mathrm{ACF}_\phi$ on certain triangles, although the very
1088: high SNR of the closure phases on the small SMTO-CARMA-LMT and
1089: Hawaii-SMTO-CARMA triangles (which do not resolve out much of the
1090: total flux) often compensates by producing a very clean ACF (i.e.,
1091: $\xi$ is small as well). In order to identify the period in an
1092: algorithmic way, we created a ``folded'' version of the composite
1093: array ACF, defined as
1094: \begin{equation}
1095: \mathrm{Fold}(k) \equiv \frac{1}{\lfloor
1096: m/k\rfloor}\sum\limits_{i=1}^{\lfloor m/k\rfloor}
1097: \mathrm{ACF}(i\cdot k),
1098: \end{equation}
1099: where $m$ is the number of points in the ACF. Folding the ACF
1100: effectively suppresses the trivial peak at zero lag. In the absence
1101: of noise, the peak of $\mathrm{Fold}(k)$ is the period.
1102:
1103: In order to test for the significance of periodicity detection, we ran
1104: Monte Carlo simulations for each model and array at bit rates from 1
1105: to 32~Gbit\,s$^{-1}$ by powers of two. We ran 10000 simulations of
1106: the data with different noise instantiations and obtained the value of
1107: $k$ maximizing $\mathrm{Fold}(k)$. The distribution of
1108: $\mathrm{Fold}(k)$ provides information on the probability of false
1109: detection of periodicity. The results are summarized in
1110: Table~\ref{tab-results}, which lists the number of trials for which
1111: $\max(\mathrm{Fold}(k))$ misidentifies the period, rounded to the
1112: nearest 10~s, by more than one minute. Four and a half orbital
1113: periods of simulated data were used, corresponding to 2~hr for Models
1114: A-D and 37~min for Model E.
1115:
1116: \clearpage \input{tab3.tex} \clearpage
1117:
1118: It is clear that the inclusion of a fourth or fifth telescope in the western
1119: hemisphere array produces a large improvement in periodicity detectability.
1120: This is due both to the much larger number of closure quantities that can be
1121: averaged together to produce a detection and also to the fact that a 4- or
1122: 5-element array will necessarily probe a larger range of spatial scales than
1123: possible 3-element arrays, which is important because it is not clear a priori
1124: which triangle will be best matched to the angular resolution of the variable
1125: emission in Sgr~A*. Additional bandwidth is important as well, but less so
1126: than additional telescopes in the observing array. The Hawaii-SMTO-CARMA
1127: array, for example, is insensitive to periodic structure in several models,
1128: even at high recording rates. But inclusion of a Chilean telescope, and/or the
1129: LMT, produces a robust detection of periodicity at modest recording rates. It
1130: is especially worth noting that a bit rate of 8~Gbit\,s$^{-1}$, corresponding
1131: to 2~GHz total bandwidth (e.g., 1~GHz in each of two orthogonal polarizations)
1132: is sufficient to produce clear periodicity detections (with error rate $< 1
1133: \times 10^{-4}$) in nearly all models provided that at least four telescopes
1134: are used in the array. For the Europe-Chile triangle, where no clear fourth
1135: telescope is presently available, the use of high bit rates and phased ALMA
1136: will be critical for periodicity detection.
1137:
1138: Array options at 345~GHz are relatively limited, since neither the LMT
1139: nor PV have planned 345~GHz capability. We include the Europe-Chile
1140: triangle to show what might be expected if PV is ever upgraded to
1141: include a 345~GHz receiver. The largest simultaneous array that can
1142: be deployed at 345~GHz in the near future consists of the three US
1143: telescopes and a Chilean station. Should phased ALMA not become
1144: available, an array consisting of Hawaii, SMTO, CARMA, Chile~1, and
1145: the LMT with a 345~GHz receiver (but not otherwise optimized for
1146: 345~GHz performance) would suffice to detect periodicity in Models A-E
1147: at bit rates of 2~Gbit\,s$^{-1}$ or higher.
1148:
1149: Observations have recently been taken with an array consisting of the
1150: JCMT, SMTO, and CARMA-1 at 230~GHz \citep{doeleman08}. We run
1151: simulations of this array, which is usable already, as well as the
1152: same three telescopes with the addition of Chile-1, since APEX is
1153: likely to have 230~GHz capability in the near future. We find that
1154: periodicity may be just marginally detectable at the current maximum
1155: bit rate of 4~Gbit\,s$^{-1}$ if the source geometry in Sgr~A* is
1156: favorable. Larger bandwidths are critical for detecting periodicity
1157: on these arrays (and may not suffice if only the US telescopes are
1158: used, depending on source geometry). It is likely that phased-array
1159: capability at Hawaii and CARMA will become available on the same time
1160: scale as higher bit rate capability. At 345~GHz, an array consisting
1161: of the JCMT, SMTO, and Chile-1 (APEX or ASTE) would not suffice to
1162: detect periodicity except at very high bit rates ($\geq
1163: 32$~Gbit\,s$^{-1}$, depending on the model).
1164:
1165:
1166: \subsection{Long-Period Models}
1167:
1168: \begin{figure}
1169: \resizebox{\hsize}{!}{\includegraphics{f10.eps}}
1170: \caption{Closure phases and amplitudes for Model F at 230~GHz (period
1171: 2.8~hr). The colored lines indicate the closure quantities that
1172: would be obtained in the absence of noise. The abscissa shows the
1173: time range over which Sgr~A* is above $5\degr$ elevation at at least
1174: three western hemisphere telescopes. Due to changing baseline
1175: orientations, closure phases and amplitudes do not approximately
1176: repeat, although significant deviations are seen simultaneously on
1177: most subarrays. Very small spikes seen in some of the lines are
1178: artifacts of the modelling.
1179: \label{fig-modelF}
1180: }
1181: \end{figure}
1182:
1183: Detecting periodicity in the closure quantities will be more difficult
1184: if the hot spot orbital period is long (e.g., several hours). The
1185: change in baseline orientation is significant enough within the
1186: 167-minute orbital period in Model F that the closure quantities do
1187: not approximately repeat, as shown in Figure~\ref{fig-modelF}.
1188: Consequently, the corresponding ACFs on most triangles and quadrangles
1189: fall off with increasing lag, so the period cannot be determined by
1190: finding the largest peak in the ACF. Closure phases and amplitudes
1191: show periods of relative quiescence punctuated by periods of large
1192: simultaneous variability on most subarrays.
1193:
1194: The mutual visibility on most triangles and quadrangles is comparable
1195: to twice the orbital period of the long-period models, so periodic
1196: behavior in closure phases and amplitudes may not be convincing in
1197: demonstrating source structure periodicity. For instance, it may not
1198: be clear whether closure phases and amplitudes such as those shown in
1199: Figure~\ref{fig-modelF} are due to a hot spot in a large orbit or due
1200: to two unrelated flaring episodes of an aperiodic nature. The LMT may
1201: be critical for establishing long-period periodicity, since its
1202: inclusion lengthens the total time range for observing, increasing the
1203: chance that a third period will be detected. While present
1204: millimeter-VLBI arrays are thus not optimal for detecting long-period
1205: orbiting hot spots, observational evidence suggests that hot spot
1206: periods may be significantly shorter than that in Model~F. VLBI
1207: measurements of the position of the centroid of emission from Sgr~A*
1208: presently constrain the orbital period of hot spots to be $\lesssim
1209: 120$~min if the hot spot flux dominates the disk flux, with
1210: progressively longer periods allowed as the hot-spot-to-disk flux
1211: ratio decreases \citep{reid08}.
1212:
1213: However, there are fundamental constraints upon the existence of such
1214: long-period hot spots. The synchrotron cooling time at millimeter
1215: wavelengths is roughly $3\,\hr$ (see eq.\,[\ref{eqn-cooling}]), and
1216: thus in the absence of a continuous mechanism for injecting energetic
1217: electrons in the hot spot, we would not expect to see single hot spots
1218: persist longer than this. Additionally, the available energy
1219: (magnetic \& hydrodynamic) that may be tapped to generate substantial
1220: emission decreases rapidly with radius, implying that the brightest
1221: (and therefore dominant) events will preferentially occur at small
1222: radii, with correspondingly short dynamical timescales.
1223:
1224:
1225:
1226:
1227: \section{Discussion}
1228: \label{discussion}
1229:
1230: \subsection{Observing Strategies and Telescope Prioritization}
1231:
1232: Ultimately, observations are limited by the telescopes that are
1233: available. Individual observers often have little influence over
1234: telescope construction priorities. Nevertheless, it is important to
1235: consider the relative advantages of planned and potential instruments.
1236: Future observers may be limited by scarce resources, such as DBEs and
1237: recording equipment, instead of by available telescopes.
1238:
1239: Array phasing will be crucial for maximizing the potential of
1240: millimeter VLBI arrays. Phased versions of CARMA and the millimeter
1241: telescopes on Mauna Kea are necessary in order to increase the
1242: detectability of very weak signals on the long baselines. However,
1243: further observations on the Hawaii-CARMA-SMTO triangle should not be
1244: deferred due to the lack of a phased-array processor at Hawaii and/or
1245: CARMA. In some models, periodicity could be detected on the triangle
1246: consisting of the JCMT, a single CARMA dish, and the SMTO at high
1247: bandwidth. Most of our models also indicate that periodicity may be
1248: detectable on the Hawaii-CARMA-Chile~1 triangle, although the use of
1249: phased ALMA rather than a single dish as the Chilean telescope may
1250: also be important for detecting periodicity if the black hole is
1251: highly rotating or if hot spot flux density signatures are typically
1252: smaller than assumed in this work.
1253:
1254: Given the schedule of proposed telescope upgrades, observations
1255: utilizing closure techniques in the near future are most likely to use
1256: the Hawaii-CARMA-SMTO triangle at 230~GHz and the Hawaii-SMTO-Chile
1257: triangle at 345~GHz, with the possibility that the JCMT and/or CARMA-1
1258: may necessarily be used in place of Hawaii and CARMA for the earliest
1259: observations. As illustrated in Figure~\ref{fig-phase230} (and
1260: Table~\ref{tab-results}), the Hawaii-CARMA-SMTO triangle at 230~GHz
1261: may not provide adequate spatial resolution to detect an orbiting hot
1262: spot, depending on the parameters of the Sgr~A* system. At the other
1263: extreme, the currently available array of Hawaii, SMTO, and Chile-1 at
1264: 345~GHz may resolve out most of the emission (Fig.~\ref{fig-uv}) and
1265: is unlikely to be useful for periodicity detection by itself at data
1266: rates less than about 16~Gbit\,s$^{-1}$. We therefore conclude that
1267: observations of Sgr~A* in the near future to detect periodicity should
1268: be taken at 230~GHz.
1269:
1270: It will be important to include a fourth antenna in the array at each
1271: frequency. As demonstrated in \S \ref{ACF}, the additional closure
1272: phases and amplitudes obtained by a 4-element array may be important
1273: for detecting periodicity. Even if periodicity can be marginally
1274: detected by a 3-element array, the extra 4 independent closure
1275: quantities provided by inclusion of a fourth telescope will vastly
1276: increase the significance of periodicity detection. Early
1277: observations at 230~GHz should include APEX or a single ALMA dish, if
1278: at all possible, in order to provide a fourth station in the array
1279: (otherwise assumed to include Hawaii, SMTO, and CARMA). A key benefit
1280: of a 4-station US-Chile array, beyond the much larger number of
1281: closure quantities provided, is that the constituent baselines will be
1282: sensitive to a large range of spatial scales, making it less likely
1283: that periodicity will be missed due to a mismatch between the angular
1284: scale of source structure variability and the angular resolution of
1285: the array.
1286:
1287: The LMT will be a critical telescope for two reasons. First, it fills
1288: in an important hole in the $(u,v)$ plane, allowing for a large range
1289: of angular scales to be probed by the full array. Indeed, the LMT
1290: produces excellent closure phase data on some triangles for every
1291: model considered in this work. Second, inclusion of the LMT allows
1292: for simultaneous visibility of Sgr~A* from up to 5 telescopes. The
1293: utility of the LMT derives especially from its location and its size.
1294: Even an incomplete LMT (e.g., a 32 m dish, as assumed in this work)
1295: whose surface has not yet been fully tuned to maximize aperture
1296: efficiency would be highly useful for observations of the Galactic
1297: center at 230~GHz. Thus, we conclude that millimeter-wavelength VLBI
1298: arrays observing Sgr~A* should include the LMT as soon as is
1299: practical.
1300:
1301: There is also a need for 345~GHz capability at a greater number of
1302: telescopes. CARMA would be especially useful due both to its large
1303: effective collecting area as well as its short baselines to the SMTO
1304: and Hawaii. The same comments that apply to the LMT in its planned
1305: 230~GHz band also apply to a future 345~GHz system, if ever planned.
1306: Expanding PV to include 345~GHz would allow for observations on the
1307: Europe-Chile triangle, which would provide only one additional closure
1308: phase and is therefore not a high priority for periodicity detection.
1309: However, the long Europe-Chile baselines cover an otherwise unsampled
1310: region of the $(u,v)$ plane and may therefore prove to be important
1311: for eventual modelling of the Sgr~A* quiescent disk.
1312:
1313: An alternative observing strategy for the Chilean site, if phased ALMA
1314: is unavailable, would be to use multiple individual telescopes.
1315: Baselines from the Chilean telescopes (APEX, ASTE, or a single ALMA
1316: dish) to North American or European telescopes would effectively be
1317: redundant with each other in the $(u,v)$ plane but would offer
1318: independent data, thus increasing the number of closure quantities
1319: available. Baselines between Chilean telescopes would be too short to
1320: resolve the emission from Sgr~A* but would provide a valuable
1321: simultaneous (near-)zero-spacing flux measurement, allowing estimation
1322: of the fraction of flux resolved out by the short PV-PdB or SMTO-CARMA
1323: baselines.
1324:
1325:
1326: \subsection{Bandwidth Considerations}
1327:
1328: Our simulations indicate that the periodicity of reasonable hot spot
1329: models is detectable on most millimeter VLBI triangles/quadrangles at
1330: data rates from 1 to 32~Gbit\,s$^{-1}$. However, we note several
1331: caveats whose applicability depends on the exact physical parameters
1332: of the Sgr~A* system. It may be difficult to detect periodicity on
1333: the smallest triangles due to insufficient spatial resolution.
1334: Conversely, the largest triangles may resolve out much of the
1335: emission. Very high data rates may be required to detect periodicity
1336: on long triangles to Chile~1. During any particular track of
1337: observations, it is possible that no hot spot will be present, or that
1338: the flux density of the hot spot may be substantially smaller than
1339: assumed here.
1340:
1341: We have assumed throughout that the atmospheric coherence time is 10~s. In
1342: poorer weather conditions, the coherence time may be significantly shorter.
1343: Since the SNR of a coherently-integrated visibility grows as $(Bt)^{1/2}$,
1344: where $B$ is the bandwidth, the SNR for a single visibility integrated over a
1345: 2.5~s coherent time interval will be half that of a visibility integrated over
1346: 10~s. So long as the signal is strong enough compared to the noise (i.e., the
1347: SNR $\gg 1$ in a single coherent integration), there is no net loss of signal
1348: because the decrease in SNR is exactly compensated by the increase in the
1349: number of data points obtained. The SNR of the closure phase can then be built
1350: up over a longer period of time, if desired, by averaging consecutive closure
1351: phases (provided that the noiseless closure phase is not changing rapidly over
1352: the time scale of integration). Periodicity may still be weakly detected if
1353: the average SNR exceeds unity over only a portion of the hot spot orbit, as
1354: could be the case if the visibility amplitude on the weakest baseline changes
1355: by a factor of several over an orbit.
1356:
1357: Due to this effective SNR cutoff, it will be important to obtain data
1358: at the maximum bandwidth (or recording rate) possible at the time of
1359: observations. Initial observations may be bandwidth-limited to 1~GHz
1360: (4~Gbit\,s$^{-1}$). If subsequent observations are limited by average
1361: recording rate rather than IF bandwidth, it will be advantageous to
1362: use burst-mode recording, if available. In the limit of marginal
1363: detections, it is far preferable to have a reduced number of good data
1364: points rather than a full set of poor data.
1365:
1366:
1367: \subsection{Black Hole Parameter Estimation}
1368:
1369: The observation of periodicity in the closure quantities would provide
1370: important evidence, independent of observations of periodicity in
1371: flare light curves, that at least some subset of Sgr~A*'s flares
1372: are due to bright orbiting structures. This is critical to
1373: efforts to use such structures ( e.g., hot spots) to infer the
1374: properties of the black hole spacetime.
1375:
1376: In principle, combined with the flare amplitude (as measured via the unresolved
1377: light curves), the degree of variability in the closure quantities is
1378: indicative of the size of the hot-spot orbit. Combined with the period, this
1379: provides a straightforward way in which to estimate the spin. However, the
1380: precision with which submillimeter closure quantities can constrain the spin
1381: has yet to be determined, and will likely have to await a detailed
1382: parameter-space study. On the topic of black hole spin, it should be pointed
1383: out that arguments for a non-zero spin of the SgrA* black hole can be made
1384: based on the observed intrinsic size of Sgr~A* at 1.3~mm \citep{doeleman08},
1385: and from NIR variability results \citep{genzel03}.
1386:
1387: Our methods are generalizable to any mechanism of flare production.
1388: There will almost surely be asymmetric structure on scales of a few to
1389: a few tens of $\Rs$ due to general relativistic effects from the
1390: accretion disk. Regardless of whether flares are produced by orbiting
1391: or spiralling hot spots, jets, magnetohydrodynamic instabilities, or
1392: some other mechanism, there will also be asymmetric structures in the
1393: inner disk region on scales of a few $\Rs$. Our models
1394: demonstrate that millimeter-wavelength VLBI arrays will be sensitive
1395: to changes in the source structure no matter how these asymmetries are
1396: oriented relative to each other on the sky.
1397:
1398: \subsection{Summary of Simulation Findings}
1399:
1400: \begin{itemize}
1401:
1402: \item The currently useable 230~GHz array consisting of SMTO, CARMA-1, and JCMT
1403: at 4~Gbit\,s$^{-1}$ cannot detect periodicity in any of the models tested.
1404: \item Phasing connected element arrays (CARMA, Hawaiian telescopes, ALMA) to
1405: form large effective apertures, significantly improves the probability of
1406: detecting the hot spot period. When the CARMA array is phased, and the
1407: Hawaiian telescopes are coherently summed, periodicity in two out of five of
1408: the 230~GHz models can be detected. When 10 elements of ALMA are phased
1409: together, periodicity can be reliably extracted in all of the 230~GHz models
1410: tested.
1411: \item Higher recording bandwidth increases signal-to-noise on all baselines,
1412: thereby improving periodicity detection, and making the array more robust
1413: against loss of VLBI signal coherence due to atmospheric turbulence. In three
1414: out of five 230~GHz models tested, a recording bandwidth of 16~Gbit\,s$^{-1}$
1415: allows detection of periodicity using just a three-station VLBI array.
1416: \item Adding a fourth or fifth telescope to the array enables detection of the
1417: hot-spot period \emph{in every} 230~GHz model tested. The added baseline
1418: coverage in larger arrays allows more complete sampling of spatial scales in
1419: SgrA*.
1420: \item Inclusion of the LMT in 230~GHz arrays will be very important as
1421: baselines to the LMT fill critical voids in baseline coverage. When long
1422: baselines between the US and Chile heavily resolve SgrA*, baselines between
1423: each of these regions and the LMT can still provide high signal-to-noise
1424: detections that connect all telescopes in the array.
1425: \item A 345~GHz capability at more sites is needed. The currently available
1426: 345~GHz array, consisting of SMTO, Hawaii, and Chile has long baselines that
1427: will resolve much of the flux density of SgrA*. Enhancing either CARMA or the
1428: LMT with low-noise 345~GHz receivers, puts periodicity detection of all the
1429: tested 345~GHz models within reach when ALMA comes on line.
1430:
1431: \end{itemize}
1432:
1433: \subsection{Conclusions}
1434:
1435: Motivated by recent detection of intrinsic structure within SgrA* on $<4\Rs$
1436: scales, this work explores the feasibility of detecting time variable structure
1437: and periodicity in the context of flare models in which a hot-spot orbits the
1438: central black hole. Algorithms are described that use interferometric closure
1439: quantities, direct VLBI observables which reflect asymmetries in source
1440: structure that can be tracked on time scales that are short compared to
1441: presumed hot-spot orbital periods. We find that periodicity from these models
1442: over a representative range of parameters can be reliably extracted using mm
1443: and submm VLBI arrays that are planned over the next 5 years. Thus, mm and sub-mm
1444: VLBI has matured to the level where one can envisage studying fundamental black
1445: hole parameters, accretion physics, and General Relativity in the strong field
1446: regime at meaningful angular resolutions.
1447:
1448: The techniques and concepts described here are applicable to a broad range of
1449: Sgr~A* models, including emission due to jets and outflows, MHD simulations,
1450: and adiabadically expanding flaring structures. Closure amplitude analysis,
1451: for example, is capable of tracking the size of SgrA* over time and sensitively
1452: testing for expansion during a flare event. Short wavelength VLBI, when
1453: coupled with wide spectrum simultaneous monitoring, can thus make detailed
1454: tests of emission models for Sgr~A* in which low frequency synchrotron photons
1455: are up-scattered to produce X-rays.
1456:
1457: The three fundamental technical efforts to improve mm and submm arrays include
1458: increasing the number of VLBI sites, achieving the capability to phase
1459: connected element arrays into a large effective aperture, and increasing the
1460: VLBI recording data rate. Projects in each of these areas, supported by an
1461: international collaboration, are underway, and it is expected that observations
1462: on VLBI arrays at 1.3~mm and 0.85~mm wavelength with dramatically improved
1463: sensitivities will begin by 2009. The increase of bandwidth, while of
1464: general importance to lowering VLBI detection thresholds, may be most useful by
1465: enabling full polarization observations. The techniques developed in this work
1466: can be extended to non-imaging VLBI polarimetry analysis, allowing tests for
1467: small scale polarization structure to be carried out. The prospect of using
1468: VLBI to probe such polarized sub-structure emphasizes the power of the
1469: technique for studying Sgr~A*.
1470:
1471: \acknowledgments
1472:
1473: The high-frequency VLBI program at Haystack Observatory is funded
1474: through a grant from the National Science Foundation.
1475:
1476: \begin{thebibliography}{}
1477:
1478: \bibitem[Agol(2000)]{agol00} Agol, E.\ 2000, \apjl, 538, L121
1479:
1480: \bibitem[Aitken et al.(2000)]{aitken00} Aitken, D.~K., et al.\ 2000,
1481: \apj, 534, L173
1482:
1483: \bibitem[Aschenbach et al.(2004)]{aschenbach04} Aschenbach, B.,
1484: Grosso, N., Porquet, D., \& Predehl, P.\ 2004, \aap, 417, 71
1485:
1486: \bibitem[Backer(1978)]{backer78} Backer, D.~C.\ 1978, \apjl, 222, L9
1487:
1488: \bibitem[Backer \& Sramek(1999)]{backer99} Backer, D.~C., \& Sramek,
1489: R.~A.\ 1999, \apj, 524, 805
1490:
1491: \bibitem[Baganoff et al.(2001)]{baganoff01} Baganoff, F.~K., et al.\
1492: 2001, \nat, 413, 45
1493:
1494: \bibitem[B\'{e}langer et al.(2006)]{belanger06} B\'{e}langer, G.,
1495: Terrier, R., de Jager, O.~C., Goldwurm, A., \& Melia, F.\ 2006,
1496: Journal of Physics Conference Series, 54, 420
1497:
1498: \bibitem[Blandford \& Begelman(1999)]{blandford99} Blandford, R.~D. \&
1499: Begelman, M.~C.\ 1999, \mnras, 303, L1
1500:
1501: \bibitem[Bower et al.(2004)]{bower04} Bower, G.~C., Falcke, H.,
1502: Herrnstein, R.~M., Zhao, J.-H., Goss, W.~M., \& Backer, D.~C.\ 2004,
1503: Science, 304, 704
1504:
1505: \bibitem[Bower et al.(2001)]{bower01} Bower, G.~C., Wright, M.~C.~H.,
1506: Falcke, H. \& Backer, D.~C.\ 2001, \apj, 555, 103
1507:
1508: \bibitem[Bower et al.(2003)]{bower03} Bower, G.~C., Wright, M.~C.~H.,
1509: Falcke, H. \& Backer, D.~C.\ 2003, \apj, 588, 331
1510:
1511: \bibitem[Bower et al.(2006)]{bower06} Bower, G.~C., Goss, W.~M.,
1512: Falcke, H. \& Backer, D.~C., \& Lithwick, Y.\ 2006, \apjl, 648, L127
1513:
1514: \bibitem[Broderick (2006)]{broderick06} Broderick, A.~E.\ 2006,
1515: \mnras, 366, L10
1516:
1517: \bibitem[Broderick \& Blandford(2003)]{broderick03} Broderick, A. \&
1518: Blandford, R.\ 2003, \mnras, 342, 1280
1519:
1520: \bibitem[Broderick \& Blandford(2004)]{broderick04} Broderick,
1521: A.~E. \& Blandford, R.~D.\ 2004, \mnras, 349, 994
1522:
1523: \bibitem[Broderick \& Loeb(2005)]{broderick05} Broderick, A.~E. \&
1524: Loeb, A.\ 2005, \mnras, 363, 353
1525:
1526: \bibitem[Broderick \& Loeb(2006a)]{broderick06a} Broderick, A.~E. \&
1527: Loeb, A.\ 2006a, \apjl, 636, L109
1528:
1529: \bibitem[Broderick \& Loeb(2006b)]{broderick06b} Broderick, A.~E. \&
1530: Loeb, A.\ 2006b, \mnras, 367, 905
1531:
1532: \bibitem[Bromley et al.(2001)]{bromley01} Bromley, B.~C., Melia, F.,
1533: \& Liu, S.\ 2001, \apjl, 555, L83
1534:
1535: \bibitem[Chan et al.(2006)]{chan06} Chan, C.-k., Liu, S., Fryer,
1536: C.~L., Psaltis, D., Ozel, F., Rockefeller, G., \& Melia, F.\ 2006,
1537: ArXiv Astrophysics e-prints, arXiv:astro-ph/0611269
1538:
1539: \bibitem[Cornwell(1982)]{cornwell82} Cornwell, T.\ 1982, Synthesis
1540: Mapping, 13
1541:
1542: \bibitem[Cornwell \& Wilkinson(1981)]{cornwell81} Cornwell, T.~J., \&
1543: Wilkinson, P.~N.\ 1981, \mnras, 196, 1067
1544:
1545: \bibitem[Doeleman et al.(2002)]{doeleman02} Doeleman, S., et
1546: al.\ 2002, Proceedings of the 6th EVN Symposium, 223
1547:
1548: \bibitem[Doeleman et al.(2001)]{doeleman01} Doeleman, S.~S., et
1549: al.\ 2001, \aj, 121, 2610
1550:
1551: \bibitem[Doeleman et al.(2008)]{doeleman08} Doeleman, S.~S., et
1552: al.\ 2008, \nat, submitted
1553:
1554: \bibitem[Eckart et al.(2006)]{eckart06} Eckart, A., Sch{\"o}del, R.,
1555: Meyer, L., Trippe, S., Ott, T. \& Genzel, R.\ 2006, \aap, 455, 1.
1556:
1557: \bibitem[Eckart et al.(2004)]{eckart04} Eckart, A., et al.\ 2004, \aap,
1558: 427, 1
1559:
1560: \bibitem[Falanga et al.(2007)]{falanga07} Falanga, M., Melia, F.,
1561: Tagger, M., Goldwurm, A., \& B{\'e}langer, G.\ 2007, \apjl, 662, L15
1562:
1563: \bibitem[Falcke \& Markoff(2000)]{facke00a} Falcke, H. \& Markoff,
1564: S.\ 2000, \aap, 362, 113
1565:
1566: \bibitem[Falcke et al.(2000)]{falcke00} Falcke, H., Melia, F., \&
1567: Agol, E.\ 2000, \apjl, 528, L13
1568:
1569: \bibitem[Genzel et al.(2003)]{genzel03} Genzel, R., Sch{\"o}del, R.,
1570: Ott, T., Eckart, A., Alexander, T., Lacombe, F., Rouan, D., \&
1571: Aschenbach, B.\ 2003, \nat, 425, 934
1572:
1573: \bibitem[Ghez et al.(2005)]{ghez05} Ghez, A.~M., Salim, S., Hornstein,
1574: S.~D., Tanner, A., Lu, J.~R., Morris, M., Becklin, E.~E., \&
1575: Duch{\^e}ne, G.\ 2005, \apj, 620, 744
1576:
1577: \bibitem[Ghez et al.(2003)]{ghez03} Ghez, A.~M.\, et al.\ 2003, \apjl,
1578: 586, L127
1579:
1580: \bibitem[Ghez et al.(2004)]{ghez04} Ghez, A.~M., et al.\ 2004, \apjl,
1581: 601, L159
1582:
1583: \bibitem[Goldston et al.(2005)]{goldston05} Goldston, J.~E., Quataert,
1584: E., \& Igumenschev, I.~V.\ 2005, \apj, 621, 785
1585:
1586: \bibitem[Holdaway(1997)]{holdaway97} Holdaway, M.~A.\ 1997, MMA Memo Series \#169.
1587:
1588: \bibitem[Hornstein et al.(2007)]{hornstein07} Hornstein, S.~D.,
1589: Matthews, K., Ghez, A.~M., Lu, J.~R., Morris, M., Becklin, E.~E.,
1590: Rafelski, M., \& Baganoff, F.~K.\ 2007, \apj, 667, 900
1591:
1592: \bibitem[Huang et al.(2007)]{huang07} Huang, L., Cai, M., Shen, Z.-Q.,
1593: \& Yuan, F.\ 2007, \mnras, 379, 833
1594:
1595: \bibitem[Jennison(1958)]{jennison58} Jennison, R.~C.\ 1958, \mnras,
1596: 118, 276
1597:
1598: \bibitem[Jones \& O'Dell(1977)]{jones77} Jones, T.~W. \& O'Dell,
1599: S.~L.\ 1977, \apj, 214, 522
1600:
1601: \bibitem[Lindquist(1966)]{lindquist66} Lindquist, R.~W.\ 1966, Annals
1602: of Physics, 37, 487
1603:
1604: \bibitem[Lo et al.(1998)]{lo98} Lo, K.~Y., Shen, Z.-Q., Zhao, J.-H.,
1605: \& Ho, P.~T.~P.\ 1998, \apjl, 508, L61
1606:
1607: \bibitem[Loeb \& Waxman(2007)]{loeb07} Loeb, A. \& Waxman, E.\ 2007,
1608: JCAP, 3, 11
1609:
1610: \bibitem[Macquart et al.(2006)]{macquart06} Macquart, J.-P. et
1611: al.\ 2006, \apj, 646, 111
1612:
1613: \bibitem[Maoz(1998)]{maoz98} Maoz, E.\ 1998, \apjl, 494, L181
1614:
1615: \bibitem[Markoff et al.(2007)]{markoff07} Markoff, S., Bower, G.~C.,
1616: \& Falcke, H.\ 2007, \mnras, 379, 1519
1617:
1618: \bibitem[Marrone et al.(2007)]{marrone07} Marrone, D.~P.\ 2007,
1619: astro-ph/0712.2877
1620:
1621: \bibitem[Marrone et al.(2006)]{marrone06} Marrone, D.~P., Moran,
1622: J.~M., Zhao, J.-H. \& Rao, R.\ 2006, Journal of Physics Conference
1623: Series, 54, 354
1624:
1625: \bibitem[Marrone et al.(2007)]{marrone07b} Marrone, D.~P., Moran,
1626: J.~M., Zhao, J.-H. \& Rao, R.\ 2007, \apj, 654, 57
1627:
1628: \bibitem[Melia et al.(2001)]{melia01} Melia, F., Bromley, B.~C., Liu,
1629: S., \& Walker, C.~K.\ 2001, \apjl, 554, L37
1630:
1631: \bibitem[Meyer et al.(2006)]{meyer06} Meyer, L., Sch{\"o}del, R.,
1632: Eckart, A., Karas, V., Dov{\v c}iak, M., \& Duschl, W.~J.\ 2006,
1633: \aap, 458, L25
1634:
1635: \bibitem[Miyoshi et al.(2004)]{miyoshi04} Miyoshi, M., Ishitsuka,
1636: J.~K., Kameno, S., Shen, Z., \& Horiuchi, S.\ 2004, Progress of
1637: Theoretical Physics Supplement, 155, 186
1638:
1639: \bibitem[Monnier(2007)]{monnier07} Monnier, J.~D.\ 2007, New Astronomy
1640: Review, 51, 604
1641:
1642: \bibitem[Narayan et al.(1998)]{narayan98} Narayan, R., Mahadevan, R.,
1643: Grindlay, J.~E., Popham, R.~G. \& Gammie, C.\ 1998, \apj, 492, 554
1644:
1645: \bibitem[Petrosian \& McTiernan(1983)]{petrosian83} Petrosian, V. \&
1646: McTiernan, J.~M.\ 1983, Phys. Fluids, 26, 3023
1647:
1648: \bibitem[Quataert \& Gruzinov(2000)]{quataert00} Quataert, E. \&
1649: Gruzinov, A.\ 2000, \apj, 545, 842
1650:
1651: \bibitem[Reid(1993)]{reid93} Reid, M.~J.\ 1993, \araa, 31, 345
1652:
1653: \bibitem[Reid et al.(2008)]{reid08} Reid, M.~J., Broderick, A.~E.,
1654: Loeb, A., Honma, M., \& Brunthaler, A.\ 2008, \apj, submitted,
1655: arXiv:0801.4505
1656:
1657: \bibitem[Reid \& Brunthaler(2004)]{reid04} Reid, M.~J., \& Brunthaler,
1658: A.\ 2004, \apj, 616, 872
1659:
1660: \bibitem[Rogers et al.(1995)]{rogers95} Rogers, A.~E.~E.,
1661: Doeleman, S.~S., \& Moran, J.~M.\ 1995, \aj, 109, 1391
1662:
1663: \bibitem[Rogers et al.(1974)]{rogers74} Rogers, A.~E.~E., et
1664: al.\ 1974, \apj, 193, 293
1665:
1666: \bibitem[Rogers et al.(1994)]{rogers94} Rogers, A.~E.~E., et
1667: al.\ 1994, \apjl, 434, L59
1668:
1669: \bibitem[Sch{\"o}del et al.(2003)]{schodel03} Sch{\"o}del, R., Ott,
1670: T., Genzel, R., Eckart, A., Mouawad, N., \& Alexander, T.\ 2003,
1671: \apj, 596, 1015
1672:
1673: \bibitem[Sch\"{o}del et al.(2002)]{schodel02} Sch\"{o}del, R., et
1674: al.\ 2002, \nat, 419, 694
1675:
1676: \bibitem[Shen et al.(2005)]{shen05} Shen, Z.-Q., Lo, K.~Y., Liang,
1677: M.-C., Ho, P.~T.~P., \& Zhao, J.-H.\ 2005, \nat, 438, 62
1678:
1679: \bibitem[Tagger \& Melia(2006)]{tagger06} Tagger, M., \& Melia,
1680: F.\ 2006, \apjl, 636, L33
1681:
1682: \bibitem[Thompson et al.(2001)]{thompson01} Thompson, A.~R., Moran,
1683: J.~M., \& Swenson, G.~W., Jr.\ 2001, Interferometry and Synthesis in
1684: Radio Astronomy (New York: Wiley)
1685:
1686: \bibitem[Trotter et al.(1998)]{trotter98} Trotter, A.~S., Moran,
1687: J.~M., \& Rodr\'{\i}guez, L.~F.\ 1998, \apj, 493, 666
1688:
1689: \bibitem[Twiss et al.(1960)]{twiss60} Twiss, R.~Q., Carter, A.~W.~L.,
1690: \& Little, A.~G.\ 1960, The Observatory, 80, 153
1691:
1692: \bibitem[Yuan et al.(2002)]{yuan02} Yuan, F., Markoff, S. \& Falcke,
1693: H.\ 2002, \aap, 383, 854
1694:
1695: \bibitem[Yuan et al.(2003)]{yuan03} Yuan, F., Quataert, E. \& Narayan,
1696: R.\ 2003, \apj, 598, 301
1697:
1698: \bibitem[Yusef-Zadeh et al.(2006)]{yusefzadeh06} Yusef-Zadeh, F.,
1699: Roberts, D., Wardle, M., Heinke, C.~O., \& Bower, G.~C.\ 2006, \apj,
1700: 650, 189
1701:
1702: \end{thebibliography}
1703:
1704: \end{document}
1705: