Interferometric Imaging Directly with Closure Phases and Closure Amplitudes

, , , , , and

Published 2018 April 10 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Andrew A. Chael et al 2018 ApJ 857 23 DOI 10.3847/1538-4357/aab6a8

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/857/1/23

Abstract

Interferometric imaging now achieves angular resolutions as fine as ∼10 μas, probing scales that are inaccessible to single telescopes. Traditional synthesis imaging methods require calibrated visibilities; however, interferometric calibration is challenging, especially at high frequencies. Nevertheless, most studies present only a single image of their data after a process of "self-calibration," an iterative procedure where the initial image and calibration assumptions can significantly influence the final image. We present a method for efficient interferometric imaging directly using only closure amplitudes and closure phases, which are immune to station-based calibration errors. Closure-only imaging provides results that are as noncommittal as possible and allows for reconstructing an image independently from separate amplitude and phase self-calibration. While closure-only imaging eliminates some image information (e.g., the total image flux density and the image centroid), this information can be recovered through a small number of additional constraints. We demonstrate that closure-only imaging can produce high-fidelity results, even for sparse arrays such as the Event Horizon Telescope, and that the resulting images are independent of the level of systematic amplitude error. We apply closure imaging to VLBA and ALMA data and show that it is capable of matching or exceeding the performance of traditional self-calibration and CLEAN for these data sets.

Export citation and abstract BibTeX RIS

1. Introduction

Synthesis imaging for interferometry is an ill-posed problem. An interferometer measures a set of complex visibilities that sample the Fourier components of an image. Standard deconvolution approaches to imaging, such as the CLEAN algorithm (Högbom 1974), begin with an inverse Fourier transform of the sampled visibilities and then proceed to deconvolve artifacts introduced by the sparse sampling in the Fourier domain. To use CLEAN and other traditional imaging algorithms, interferometric visibilities must be calibrated for amplitude and phase errors. However, at high frequencies, the atmospheric coherence time can be as short as seconds, introducing rapid phase variations and effectively eliminating the capability to measure the absolute interferometric phase. Amplitude calibration also becomes more difficult at high frequencies, and pointing errors due to small antenna beam sizes can introduce large, time-varying errors in visibility amplitudes. While amplitude gain errors typically have longer characteristic timescales than phase errors, some very long baseline interferometry (VLBI) instruments such as the Event Horizon Telescope (EHT; Doeleman et al. 2009) use phased arrays as single stations, which can introduce rapid amplitude variations from fluctuations in the individual station's phasing efficiency.

A key simplification for interferometric calibration arises due to the fact that most calibration errors can be decoupled into station-based gain errors (e.g., Hamaker et al. 1996; Thompson et al. 2017). For an interferometric array consisting of N sites, there are $N(N-1)/2$ visibilities at each time but only N unknown gains. Hence, the calibration is overconstrained, and combinations of visibilities can be formed that are unaffected by calibration errors. For example, a closure phase is the phase of a product of three visibilities around a triangle, which cancels out the station-based phase errors on each individual visibility (Jennison 1958; Rogers et al. 1974). Likewise, the closure amplitude is a combination of four visibility amplitudes that cancels out amplitude gain errors in a specified ratio (Twiss et al. 1960; Thompson et al. 2017). Both of these quantities provide access to information about the source image that is unaffected by calibration assumptions. Despite the challenges in absolute calibration of a VLBI array, closure quantities provide robust measurements of certain relative quantities, which carry information about source structure that is only limited by the level of thermal noise.

The standard algorithm used for interferometric imaging is CLEAN (Högbom 1974; Clark 1980), which deconvolves a dirty image produced by an inverse Fourier transform by decomposing it into point sources. When the calibration is uncertain, the usual approach is to iterate between imaging with CLEAN and deriving new calibration solutions using information from the previous image—a so-called "self-calibration" or "hybrid-mapping" loop (e.g., Wilkinson et al. 1977; Readhead & Wilkinson 1978; Readhead et al. 1980; Schwab 1980; Cornwell & Wilkinson 1981; Pearson & Readhead 1984; Walker 1995; Cornwell & Fomalont 1999; Thompson et al. 2017). The results and time to convergence of this approach depend on many assumptions made in the course of this hybrid process, including the initial source model used for self-calibration, the choice of which regions to clean in a given iteration, the method used for deriving complex gains from a given image, and the choice of how frequently to recalibrate the data. The sensitivity of the final image to these assumptions cannot be directly inferred from the result.

In contrast with CLEAN's approach of deconvolving the dirty image into point sources, another family of methods (most famously, the maximum entropy method, or MEM; see, e.g., Narayan & Nityananda 1986) for interferometric imaging directly solves the source image pixels by fitting them to data constrained by additional convex regularization terms, such as entropy, sparsity, or smoothness (e.g., Frieden 1972; Gull & Daniell 1978; Cornwell & Evans 1985; Briggs 1995). Like CLEAN, these MEM-like methods can be used to produce images from complex visibilities in conjunction with a self-calibration loop. In contrast to CLEAN, however, these approaches can also be used directly with other data products derived from complex visibilities, as they only rely on comparing the data computed from the reconstructed image to the specified measurements. In other words, these methods never need to perform an inverse Fourier transform from calibrated input data. Consequently, these approaches can use closure quantities directly as the fundamental data product, bypassing the self-calibration loop entirely. The field of optical interferometry, for example, has pioneered the use of imaging directly from the measured visibility amplitudes and closure phases, bypassing the corrupted visibility phase (Buscher 1994; Baron et al. 2010; Thiébaut 2013; Thiébaut & Young 2017). Recently, several other methods have built on these techniques in preparing imaging algorithms for EHT data, fitting some combination of closure phases and visibility amplitudes directly while using different regularizing functions (Bouman et al. 2015; Akiyama et al. 2017b).

In this paper, we take the next step and present a method to reconstruct images directly using only closure amplitudes and closure phases. Our reconstructions require no assumptions about absolute phase or amplitude calibration beyond stability during the integration time used to obtain the visibilities. However, we find that a single round of self-calibration to the final image and reimaging with complex visibilities can produce even better results (Section 5.2). To make closure-only imaging computationally efficient, we derive analytic gradients of the data χ2 terms for closure quantities, which greatly improves the speed of our algorithm. When using these analytic gradients, closure-only imaging of VLBI data does not require significantly more computational time than standard imaging with complex visibilities, and it is still feasible on a personal computer for large data sets (e.g., those of connected-element interferometers such as ALMA).

We begin, in Section 2, by reviewing the fundamental properties of interferometric visibilities and closure quantities. Next, in Section 3, we discuss imaging via regularized maximum likelihood and demonstrate how to efficiently implement closure-only imaging in this framework. In Section 4, we detail our implementation of closure-only imaging in the eht-imaging software library (Chael et al. 2018),6 our methods for simulating data with gain and phase errors, and our techniques for evaluating the fidelity of the reconstructed images. In Section 5, we show the results of applying our method to both simulated EHT data and real data sets from the VLBA and ALMA. In Section 6, we discuss the general properties of closure-only imaging, and in Section 7, we summarize our results.

2. Visibilities and Closure Quantities

2.1. Interferometric Visibilities

The van Cittert–Zernike theorem identifies the visibility Vij measured by a baseline ${{\boldsymbol{b}}}_{{ij}}$ between stations i and j as a Fourier component of the source image intensity distribution $I(x,y)$ (Thompson et al. 2017, hereafter TMS):

Equation (1)

Here x and y are real space angular coordinates and u and v are the coordinates of the given baseline vector ${{\boldsymbol{b}}}_{{ij}}$ projected in the plane perpendicular to the line of sight and measured in wavelengths. Since $I(x,y)$ is a real number, the visibility is conjugate-symmetric in the Fourier plane, $\tilde{I}(-u,-v)={\tilde{I}}^{* }(u,v)$. When Ns stations can observe the source, the number of independent instantaneous visibilities is given by the binomial coefficient

Equation (2)

To fill in samples of the Fourier plane from the small number ${N}_{\mathrm{vis}}$ available at a single instant in time, interferometric observations typically use the technique of "Earth rotation aperture synthesis." As the Earth rotates, the projected baseline coordinates (u, v) trace out elliptical curves in the Fourier domain, providing measurements of new visibilities.

The identification of measured visibilities with Fourier components of the image is complicated by several factors. First, thermal noise from the telescope receiver chains, Earth's atmosphere, and the astronomical background is added to the measured visibility. This thermal noise, epsilonij, is Gaussian with a time- and baseline-dependent standard deviation, which depends on the telescope sensitivities, bandwidth, and integration time. Second, each station transforms the measured incoming polarized waveform according to its own (time-dependent) 2 × 2 Jones matrix that adjusts the level of the measured signal amplitude and mixes the measured polarizations (e.g., Hamaker et al. 1996; TMS). For the purposes of this paper, we ignore polarization and consider each station as contributing a single (time-dependent) complex gain ${G}_{i}{e}^{i{\phi }_{i}}$ to the visibility.

The phase error ϕi results from uncorrected propagation delays and clock errors. In particular, atmospheric turbulence contributes a rapidly varying stochastic term to each ϕi, which generally varies more quickly than the amplitude gain term, Gi, which arises from uncertainty in the conversion of the correlation coefficients measured on each baseline to units of flux density. In effect, the visibility amplitude is first measured in the units of the noise and then scaled to physical units from knowledge of the telescope noise properties.

Including all of these corrupting factors, the full complex visibility is

Equation (3)

where the gain amplitudes, phases, and thermal noise all vary in time.

Note that Equation (3) represents all systematic errors (e.g., those other than thermal noise) as station-based effects. In practice, effects such as polarization leakage and bandpass errors will also contribute small baseline-based effects that can bias closure quantities. However, these errors are generally slowly varying and can be removed with a priori calibration.

2.2. Closure Phases and Closure Amplitudes

Two types of "closure quantities" can be formed from complex visibilities that are insensitive to the particular station-based complex gain terms. While these quantities are robust to the presence of arbitrarily large complex gains on the visibilities, they contain less information about the source than the full set of complex visibilities. Furthermore, because closure quantities mix different Fourier components, they can be difficult to interpret physically.

First, multiplying three complex visibilities around a triangle of baselines eliminates the complex gain phase terms (Jennison 1958; Rogers et al. 1974; TMS). For any three stations, the visibility bispectrum is

Equation (4)

While the bispectral amplitude $| {V}_{{\rm{B}}}| $ is affected by the amplitude gain terms in Equation (3), the phase of the bispectrum, or closure phase ψ, is preserved under any choice of station-based phase error. The closure phase is a robust interferometric observable: apart from thermal noise, the measured closure phase is the same as the closure phase of the observed image. In the limit of low signal-to-noise ratio (S/N), both the bispectrum amplitude and the phase are biased by thermal noise and should be debiased before use in imaging (Wirnitzer 1985; Gordon & Buscher 2012).

The total number of closure phases at a moment in time is equal to the number of triangles that can be formed from sites in the array, $\left(\genfrac{}{}{0em}{}{{N}_{s}}{3}\right)$. However, not all of these closure phases are independent, as some can be formed by adding or subtracting other closure phases in the set. The total set of independent closure phases can be obtained by selecting an antenna as a reference and choosing only the triangles that include that antenna (Twiss et al. 1960; TMS). The total number of such independent closure phases is

Equation (5)

Here ${N}_{\mathrm{cl}\mathrm{phase}}$ is less than the number of visibilities at a given time, Equation (2), by the fraction $1\mbox{--}2/{N}_{s}$.

Second, on any set of four stations, closure amplitudes are formed by taking ratios of visibility amplitudes so as to cancel all the amplitude gain terms in Equation (3). Up to inverses, the baselines among any set of four stations can form three quadrangles with three corresponding closure amplitudes:

Equation (6)

Since the product of the three closure amplitudes in Equation (6) is unity, only two of the set are independent. The total number of closure quadrangles is $3\left(\genfrac{}{}{0em}{}{{N}_{s}}{4}\right)$, but the number of independent closure amplitudes is (TMS)

Equation (7)

Here ${N}_{\mathrm{cl}\mathrm{amp}}$ is equal to the total number of visibilities minus the number of unknown station gains. At any given time, the number of closure amplitudes is less than the number of visibilities by a fraction $1\mbox{--}2/({N}_{s}-1)$. Like the visibility amplitude and bispectrum, closure amplitudes are biased by thermal noise (TMS).

The robustness of closure phases and amplitudes to calibration errors comes with the loss of some information about the source. For instance, closure phases are insensitive to the absolute position of the image centroid, and closure amplitudes are insensitive to the total flux density. These can be constrained separately, either through arbitrary choices (e.g., centering the reconstructed image) or through additional data constraints (e.g., specifying the total image flux density through a separate measurement).

2.3. Redundant and Trivial Closure Quantities

Some VLBI arrays include multiple stations that are geographically colocated. For instance, the EHT includes multiple sites on Maunakea (the SMA and the JCMT), as well as multiple sites in the Atacama desert in Chile (the ALMA array and the APEX telescope). Practically, any two sites that form a baseline that does not appreciably resolve any source structure can be considered colocated.

These "redundant" sites can be used to form closure quantities. In the case of closure phase, the added triangles provide no new source information. Specifically, any triangle $\{{{\boldsymbol{b}}}_{12},{{\boldsymbol{b}}}_{23},{{\boldsymbol{b}}}_{31}\}$ that includes two colocated sites {1, 2} will include one leg that measures the zero-baseline visibility, which has zero phase: ${V}_{12}=\tilde{I}(0,0)$ is the integrated flux density of the source (see Equation (1)). The remaining two long legs from the pair of colocated sites to the third site will have ${{\boldsymbol{b}}}_{23}=-{{\boldsymbol{b}}}_{31}$ and, consequently, ${V}_{23}={V}_{31}^{* }$. Thus, the bispectrum will be a positive real number, and the closure phase must be zero, regardless of the source structure. These trivial triangles are not useful for imaging but provide valuable tests of the closure-phase statistics and systematic bias (see, e.g., Fish et al. 2016). In short, redundant sites can provide additional redundant closure-phase triangles, which can be averaged to reduce thermal noise, but they do not give new measurements of source structure.

Redundant sites also give rise to trivial closure amplitudes, which have a value of unity regardless of the source. However, redundant sites also yield new closure amplitudes that are nontrivial and provide additional information on the source structure. For instance, one can measure the normalized visibility amplitude, $| V({\boldsymbol{u}})/V(0)| $, as a closure quantity on any baseline joining two sets of colocated sites (Johnson et al. 2015). In the limiting case where every site in an array has a redundant companion, the complete source visibility amplitude information could be recovered through closure amplitudes, except for a single degree of freedom for the total flux density.

Figure 1 shows examples of the trivial and nontrivial closure amplitudes for an array with partial redundancy. As these examples illustrate, redundant sites can significantly inform and improve calibration and imaging. Figure 2 shows the number of closure amplitudes and phases for the EHT with and without redundant sites, both including and excluding trivial additions. The two redundant sites of the 2017 EHT array more than double the amount of information contained in the set of closure amplitudes over the same array without these sites.

Figure 1.

Figure 1. Example closure amplitudes for a portion of the EHT. Solid red lines connecting sites denote visibilities in the numerator of the closure amplitude; dashed blue lines denote visibilities in the denominator. An array containing redundant sites (such as SMA/JCMT and ALMA/APEX in the EHT) will produce new trivial closure amplitudes, which are equal to unity (plus thermal noise), and new nontrivial closure amplitudes, which yield new information about the source. Without redundant sites, there would be no closure amplitudes from this portion of the array.

Standard image High-resolution image
Figure 2.

Figure 2. (Left) Number of independent closure phases for the 2017 EHT over 24 hr GMST while observing Sgr A*. The blue line shows the total number of independent closure phases in the array containing redundant stations, the black line shows the number of independent closure phases that measure source structure, and the red line shows the number of independent closure phases in the array when the redundant sites are excluded. Redundant sites do not add any closure-phase information to the array apart from decreasing the overall thermal noise. (Right) Total independent (blue) and nontrivial (black) closure amplitudes over 24 hr for the EHT, including redundant sites. Unlike for closure phases, adding only two redundant sites significantly increases the amount of information contained in the set of independent closure amplitudes compared to the same array without these sites (red), because not all closure amplitudes containing a baseline between two colocated sites are trivial (see Figure 1).

Standard image High-resolution image

2.4. Thermal Noise on Closure Quantities

The thermal noise ${\epsilon }_{{ij}}$ on the baseline ij in Equation (3) is a circularly symmetric complex Gaussian random variable with zero mean that is independently sampled for each visibility measurement. The standard deviation ${\sigma }_{{ij}}$ of the thermal noise on this baseline is determined according to the standard radiometer equation (TMS):

Equation (8)

In Equation (8), SEFDi and SEFDj are the "system equivalent flux densities" of the two telescopes, where, for a telescope with system temperature Tsys and effective area Aeff, the SEFD is $2{k}_{{\rm{B}}}{T}_{\mathrm{sys}}/{A}_{\mathrm{eff}}$, with kB as the Boltzmann constant. The observing bandwidth of the visibility measurement is Δν, and Δt is the integration time. The factor of 1/η in Equation (8) is due to quantization losses in the signal digitization; for two-bit quantization, η = 0.88 (TMS).

When the S/N is high, the visibility amplitudes will also be Gaussian-distributed with a standard deviation σ given by Equation (8). At lower S/N > 1, the distribution of the amplitude becomes non-Gaussian, and the estimate of the visibility amplitude taken directly from the norm of the complex visibility is biased upward by the noise. To first order, we debias the amplitudes with the equation (TMS)

Equation (9)

For the purposes of this paper, whenever measured visibility amplitudes are used, e.g., in the computation of a closure amplitude or χ2 statistic, we assume that they have already been debiased by Equation (9).

Turning to the closure quantities, in the high-S/N limit, the baseline-based thermal noise on the closure amplitudes and phases introduced in Section 2 will also be Gaussian-distributed. To first order, the standard deviation of the complex noise on the bispectrum ${V}_{{\rm{B}}}={V}_{1}{V}_{2}{V}_{3}$ due to the thermal noise on the three-component visibilities (σ1, σ2, σ3) is

Equation (10)

Then, in the high-S/N regime, the standard deviation on the closure phase, σψ, is

Equation (11)

Similarly, the standard deviation σC of the thermal noise of the closure amplitude $| {V}_{C}| =| {V}_{1}{V}_{2}| /| {V}_{3}{V}_{4}| $ is, to leading order in the inverse S/N,

Equation (12)

Note that the expressions for σψ and σC depend only on the measured S/Ns of the visibilities, and so they too are independent of calibration.

At moderately low S/N, the Gaussianity of the thermal noise on phase and amplitude breaks down, as does the appropriateness of using the measured S/N $| {V}_{i}| /{\sigma }_{i}$ as an estimate of the true S/N when estimating σψ and σC. Because the measured phase is unbiased by thermal noise and wraps at 2π, the true σψ is smaller than the estimate in Equation (11) in the low-S/N limit. While the χ2 approach above could be extended to the exact log-likelihood for closure phases with low S/N, low-S/N closure phases are not prone to extreme outliers, so the Gaussian approximation is reasonable to use over a broad range of S/N.

The distribution for the reciprocal visibility amplitude, which appears in the denominator of Equation (12) for σC, takes on an extreme tail at low S/N that extends to positive infinity. This tail causes a large positive bias in the measured closure amplitudes, as well as a severely non-Gaussian distribution. Fitting to log closure amplitudes has the dual benefits of mitigating the tail of the reciprocal amplitude distribution and symmetrizing the numerator and denominator. In this case, where the numerator and denominator of the closure amplitude are symmetric, debiasing the component amplitudes with Equation (9) corrects the estimate of the closure amplitude to first order. While detailed analysis of the statistics of closure quantities will be explored in a forthcoming work (L. Blackburn et al. 2018, in preparation), we have generally found log closure amplitudes to be more robust observables for imaging (see Section 5).

3. Imaging with Regularized Maximum Likelihood

3.1. Imaging Framework

The standard methods of interferometric imaging are based on the CLEAN algorithm (Högbom 1974; Clark 1980). CLEAN operates on the so-called "dirty image" obtained by directly taking the Fourier transform of the sparsely sampled visibilities and attempts to deconvolve the "dirty beam" that results from the incomplete sampling of the Fourier domain. To perform the initial transform, CLEAN requires well-calibrated complex visibilities. When a priori calibration is ineffective, which is often the case at high frequencies when atmospheric phase terms vary rapidly, the visibilities must be "self-calibrated." Although self-calibration is used most frequently with CLEAN, it can be used in conjunction with any imaging method that requires calibrated complex visibilities (such as MEM).

The self-calibration procedure starts from an initial model image and solves for the set of time-dependent complex gains in Equation (3), either by fixing a sufficient set of amplitudes or phases directly from the image and solving for the rest analytically (Wilkinson et al. 1977; Readhead et al. 1980) or by finding a set that minimizes the sum of the squares of the differences between the measured and model visibilities (Schwab 1980; Cornwell & Wilkinson 1981). Self-calibration is often performed by first solving only for the phases of the complex gains and correcting the amplitudes at a later stage (Walker 1995; Cornwell & Fomalont 1999). At each round of self-calibration, the estimated inverse gain terms are then applied to the measured visibilities, and the imager (usually CLEAN) is run again to obtain a new source model. These steps are repeated many times until convergence. There are several assumptions in this procedure that may affect the final image or the time to convergence of the algorithm. Most critical are the choice of the initial source model (often taken as a point source) and of where to clean the image in each iteration (the so-called "clean boxes"). These choices enforce assumptions about the source flux distribution early on in the self-calibration process that then propagate to later rounds via the self-calibrated complex visibilities.

In contrast, the various methods of interferometric imaging explored in this paper all fall under the category of regularized maximum-likelihood algorithms. Regularized maximum-likelihood methods search for some image that maximizes the sum of a −χ2 "data term" and a "regularizer" function that prefers images with certain features when the data are not sufficient to constrain the structure by themselves. These methods can often be interpreted in a Bayesian framework, where the data term is identified with a log-likelihood and the regularizer term with a log-prior. Regularized maximum-likelihood methods require only a forward Fourier transform from trial images to the visibility domain. Consequently, they can fit directly to data terms derived from the visibilities even if the visibilities are corrupted by gain and phase errors.

In astronomy, the most familiar of these methods is MEM (see, e.g., Frieden 1972; Gull & Daniell 1978; Cornwell & Evans 1985; Narayan & Nityananda 1986). While traditional MEM uses calibrated complex visibilities as its fundamental data product, MEM and other, more general regularized maximum-likelihood methods have been developed using other regularizers, such as the 1 norm (Honma et al. 2014) or a Gaussian patch prior (Bouman et al. 2015). Other algorithms have gone beyond complex visibilities as the fundamental data product to produce images directly from the image bispectrum. Development of imaging algorithms that use different fundamental data products from complex visibilities has been particularly fruitful in optical interferometry, where the absolute visibility phase is almost never accessible (e.g., Buscher 1994; Baron et al. 2010; Thiébaut 2013; Thiébaut & Young 2017), although it has also been explored in the context of VLBI (Lu et al. 2014; Bouman et al. 2015; Akiyama et al. 2017b). Regularized maximum-likelihood methods have also been extended to polarization (Ponsonby 1973; Nityananda & Narayan 1983; Holdaway & Wardle 1990; Chael et al. 2016; Coughlan & Gabuzda 2016; Akiyama et al. 2017a), mitigation of interstellar scattering (Johnson 2016), and dynamical imaging to reconstruct movies of time-variable sources (Bouman et al. 2017; Johnson et al. 2017).

To make our discussion concrete, image reconstruction via regularized maximum likelihood seeks to find an image ${\boldsymbol{I}}$ that minimizes an objective function $J({\boldsymbol{I}})$. In this paper, we consider only square images of dimension m × m. We represent the image ${\boldsymbol{I}}$ as a one-dimensional vector of length M = m2. If we consider N observed visibilities, the corresponding sampled Fourier components, or trial image visibilities, $\tilde{V}$, of the trial image vector ${\boldsymbol{I}}$ are ${\boldsymbol{V}}^{\prime} ={\boldsymbol{AI}}$, where ${\boldsymbol{A}}$ is an N × M matrix with entries

Equation (13)

Here $({x}_{j},{y}_{j})$ are the angular coordinates (in rad) of the jth pixel, and $({u}_{i},{v}_{i})$ are the angular frequencies of the ith visibility measurement. The direct-time Fourier transform (DTFT) represented by Equation (13) is often the fastest way to compute trial visibilities for sparse arrays observing with narrow fields of view, like the EHT. For large images or large numbers of visibilities, the DTFT is slow and prohibitively expensive in terms of computer memory. In this regime, we must use the fast Fourier transform (FFT) to obtain the trial visibilities. Algorithms like the nonequispaced fast Fourier transform (NFFT; e.g., Readhead & Wilkinson 2009) are particularly useful for this purpose. In its simplest form, the NFFT takes the FFT of the trial image and interpolates the result to the irregularly sampled (u, v) points. To compensate for inaccuracies in the interpolation procedure, the NFFT both zero-pads the input image and multiplies the pixels by a scaling function (the inverse Fourier transform of a convolution kernel in the Fourier domain).

In the most general case, where we may have multiple data terms and regularizers informing our reconstruction, the objective function $J({\boldsymbol{I}})$ to be minimized is

Equation (14)

In the above expression, ${\chi }_{D}^{2}$ are the data terms or χ2 goodness-of-fit functions corresponding to the data product ${\boldsymbol{d}}$. If the data product ${\boldsymbol{d}}$ is normally distributed, these are proportional to the log-likelihoods representing the log probability that the data could be observed given an underlying image ${\boldsymbol{I}}$. For data products whose distributions are not Gaussian (like closure phases and amplitudes), χ2D is usually an approximation to the likelihood. The SR are regularizing functions (which we want to maximize), which provide missing information on the image characteristics to constrain the space of possible images given our measured data. While relatively new to radio interferometry and VLBI, reconstructions using Equation (14) with multiple data terms and regularizers are common in optical interferometry (see, e.g., Buscher 1994; Baron et al. 2010; Thiébaut 2013; Thiébaut & Young 2017).

The "hyperparameters" αD and βR control the relative weighting of the different data and regularizer terms. Because the location of the global minimum of J(${\boldsymbol{I}}$) is unaffected by changes of scale, one hyperparameter can be set to unity or some other arbitrary value without changing the solution. Furthermore, when interpreting the ${\chi }_{D}^{2}$ data terms as log-likelihoods, the data term weights αD should ideally be determined by the number of data points of each type. For example, using the reduced χ2 we define in Section 3.2, if we set one data term with N1 measurements to α1, the remaining data terms i > 1 with Ni measurements should all be set as

Equation (15)

In practice, we find that with multiple rounds of imaging, heavily weighting a single data term away from the log-likelihood weighting in Equation (15) can aid initial convergence. We then restore the ideal weighting in Equation (15) in later rounds of imaging.

In practice, the hyperparameters αD and βR are usually adjusted manually to yield reconstructions that converge to the expected values of χ2 (Cornwell & Evans 1985). Recently, Akiyama et al. (2017b) determined hyperparameters self-consistently using cross-validation. In this method, images are reconstructed with different combinations of the hyperparameters using different data sets where a portion of the data is held in reserve. The set of hyperparameters that produces the image most compatible with the data held in reserve is then used in the final reconstruction.

3.2. Data Terms for Robust Imaging

Having defined the general form of the objective function, we turn now to the different choices of the data χ2 term that can be used in interferometric (total intensity) imaging. The simplest choice is the χ2 of the measured visibilities ${\boldsymbol{V}}$. If there are N total measured visibilities Vj, with associated (real) thermal noise rms values σj, then the reduced χ2 is

Equation (16)

where ${V}_{j}^{{\prime} }$ are the sampled visibilities corresponding to the trial image ${\boldsymbol{I}}$.

If the visibility phases are significantly corrupted by atmospheric turbulence, a χ2 term that uses only the visibility amplitudes can be used:

Equation (17)

Because the closure phase is robust to station-based phase errors such as those introduced by atmospheric turbulence, a χ2 defined on the bispectrum can be used instead of Equation (16). We define NB as the number of independent bispectrum measurements and ${\sigma }_{{\rm{B}}}^{2}$ as the estimate of the variance on each complex bispectrum measurement (Equation (10)). Then,

Equation (18)

where ${V}_{{\rm{B}}j}^{{\prime} }$ is the sampled bispectrum value corresponding to the trial image ${\boldsymbol{I}}$.

We can also define a data term purely using Nψ measured closure phases, ψ (typically Nψ = NB, but we may, e.g., drop trivial closure phases from the fit). Defining ${\sigma }_{\psi }^{2}$ as their estimated closure-phase variances using Equation (11), a natural choice of a χ2 term that automatically respects 2π phase wraps in the difference of measured and trial image closure phases ψ is

Equation (19)

where the ${\psi }_{j}^{{\prime} }$ are the sampled closure phases corresponding to the trial image.

Similarly, a data term that uses only the closure amplitudes $| {V}_{C}| $ is

Equation (20)

where there are a total of NC measured independent closure amplitudes $| {V}_{{Cj}}| $, $| {V}_{{Cj}}^{{\prime} }| $ are the corresponding sampled closure amplitudes of the trial image, and ${\sigma }_{{Cj}}^{2}$ are the estimated variances of the measured closure amplitudes from Equation (12).

As discussed in Section 2.4, because closure amplitudes are formed from the quotient of visibility amplitudes, the noise on the closure amplitudes (Equation (6)) may be highly non-Gaussian. The logarithm of the closure amplitude will remain approximately Gaussian at lower S/N, so the χ2 of the logarithm of the closure amplitudes may be a better choice than Equation (20) in practice. In this case, the χ2 term is

Equation (21)

where we used the fact that, to lowest order, the variance on the logarithm of a quantity x is ${\sigma }_{\mathrm{log}(x)}^{2}={\sigma }_{x}^{2}/{x}^{2}$.

3.3. Data Term Gradients

When using gradient descent algorithms to minimize the objective function (Equation (14)), providing an analytic expression for the gradient of the objective function with respect to the image pixel values greatly increases the speed of the algorithm by bypassing the expensive step of estimating gradients numerically. When using a DTFT, the number of computations to evaluate the gradient of a χ2 term numerically via finite differences is roughly ${ \mathcal O }({M}^{2}\times N)$ (where M is the total number of image pixels and N is the number of measurements). When using an FFT, the scaling is roughly ${ \mathcal O }(M\times (M\mathrm{log}M+N))$. In contrast, when using the analytic gradients that we derive below, the corresponding scalings for DTFT and FFT are ${ \mathcal O }(M\times N)$ and ${ \mathcal O }(M\mathrm{log}M+N)$, respectively. In practice, for typical reconstructions, such as those we will show later, analytic gradients improve the imaging speed by a factor comparable to the number of free parameters.

The gradient of the simplest χ2 term, using complex visibilities (Equation (16)), is

Equation (22)

The gradients for the other χ2 terms given in Section 3.2 are presented in the Appendix.

Note that the visibility χ2 gradient, Equation (22), is the adjoint DTFT of the weighted data residuals. In fact, all of the data term gradients considered here can be written as an adjoint DTFT of appropriately weighted residual quantities (see Appendix).

3.4. Regularizer Terms

To facilitate comparisons across the different data terms in Section 3.2, we fixed the regularizer terms SR in the objective function, Equation (14), to be identical for all the reconstructions displayed in this paper. For each reconstruction, we chose to use four regularizer terms.

The first regularizer is a simple "entropy" (Frieden 1972; Gull & Daniell 1978; Narayan & Nityananda 1986) that rewards pixel-to-pixel similarity to a "prior image" with pixel values Pi:

Equation (23)

For the second regularizer, we use one of two forms of a "smoothness" constraint that pushes the final image to favor pixel-to-pixel smoothness. The first is an isotropic total variation regularizer (TV) (or 2 norm on the image gradient) that favors piecewise-smooth images with flat regions separated by sharp edges (Rudin et al. 1992),

Equation (24)

where the two sums are taken over the two image dimensions and the image pixels ${I}_{l,m}$ are now indexed by their position (l, m) in the 2D m × m grid. It should be noted that the total variation in Equation (24) is not everywhere differentiable, so care must be taken when using it in imaging. Thiébaut & Young (2017) presented a differentiable hyperbolic form of an edge-preserving smoothness regularizer (Charbonnier et al. 1997) that approximates TV when the image is far from being smooth (STV is large).

In the reconstructions in Section 5.1, however, we instead use a "total squared variation" regularizer that favors smooth edges and may be more appropriate for astronomical image reconstruction (see the forthcoming work by Kuramochi et al. 2017):

Equation (25)

The third and fourth regularizers constrain image-averaged properties. First, because closure amplitudes are independent of the normalization of the image, we include a constraint on the total image flux density,

Equation (26)

where the sum is over the M pixels in the image and F is the total source flux density, considered to be known a priori (e.g., by a simultaneous measurement of the source by a flux-calibrated single station). Next, because the closure phase does not constrain the position of the image centroid, we also include a regularizing constraint to center the image in the chosen field of view,

Equation (27)

where $({x}_{i},{y}_{i})$ is the coordinate of the ith pixel and the desired image centroid position is $({\delta }_{x},{\delta }_{y})$. In this paper, we use coordinates where $({x}_{i},{y}_{i})=(0,0)$ is in the center of the frame and set $({\delta }_{x},{\delta }_{y})=(0,0)$. When only closure phases and closure amplitudes are used in the reconstruction, both the centroid and the total flux density are completely unconstrained by data. Thus, in this case, almost any amount of weight on ${S}_{\mathrm{tot}\mathrm{flux}}$ and ${S}_{\mathrm{centroid}}$ should guide the final image to a centered image with the specified total flux, and the precise weighting of these terms relative to the data is not as significant in informing the final image as the relative weighting of STV or STSV and ${S}_{\mathrm{entropy}}$.

While the four above regularizers are used for all of the data sets imaged in this paper, we adjusted their relative weighting (the βR terms in Equation (14)), as well as the prior image used in Equation (23), based on the data set considered. However, when comparing images produced with different data terms from Section 3.2, we were consistent in using the same prior image and relative regularizer weightings in the different reconstructions to produce fair comparisons. Furthermore, when imaging synthetic data sets in Section 5.1, we used the same combination of regularizer weights (see Table 3). The prior images in Equation (23) are different from data set to data set in this section, but in every case, we use a relatively uninformative prior consisting of a Gaussian with a size of roughly half of the reconstruction field of view, and the weighting of Equation (23) is small.

4. Implementation

4.1. Imaging Methods

We implemented the imaging framework described in Section 3, including all of the data terms introduced in Section 3.2 in the eht-imaging software library (Chael et al. 2018), originally developed for polarimetric VLBI imaging (Chael et al. 2016). To minimize the objective function, Equation (14), using different combinations of data terms and regularizers, the imaging routines in eht-imaging use the limited-memory BFGS algorithm (Byrd et al. 1995) as implemented in the Scipy package (Jones et al. 2001). The L-BFGS algorithm is a quasi-Newton gradient descent method that uses the analytic forms of the data term gradients presented in Section 3.3 to progress toward a minimum in the objective function.

As described in Chael et al. (2016), our imaging algorithm ensures a positive flux in each pixel by performing a change of variables ${I}_{i}=\exp {\xi }_{i}$, where $-\infty \lt {\xi }_{i}\lt \infty $. When imaging in the log intensity domain, the gradients in Section 3.3 must be multiplied by $\exp {\xi }_{i}$. We also use the continuous image representation introduced in Bouman et al. (2015), where each array of pixel intensities is taken to represent a continuous function formed by convolving a comb of Dirac delta functions with a pixel "pulse" function. Introducing a continuous image representation multiplies the visibilities of the discrete image array by a taper given by the Fourier transform of the pulse function, removing spurious high-frequency structure introduced by the regular pixel spacing. For this paper, we used a triangular pulse function in both dimensions with width 2Δ, where Δ is the image pixel spacing.

Finally, to aid in convergence and help the minimizer avoid local minima in the objective function, we run each imager multiple times for each data set, substituting a blurred version produced by convolving the result of the previous run with a circular Gaussian as the next initial image. This procedure smooths out initial spurious high-frequency artifacts that the imager will not remove on its own, given a lack of data constraints. Each time we restart the imager, we also adjust the various hyperparameters αD and βR in Equation (14). The prescriptions for each data set are presented below, but in general, our approach is to generally increase the weight on the smoothness regularizer term to suppress the emergence of spurious high-frequency artifacts. We also usually begin by weighting the closure-phase data term more heavily in the reconstruction than is supported by the log-likelihood interpretation (Equation (15)), since we find that minimizing the closure-phase χ2 is more helpful in constraining the overall image structure in early rounds of imaging. As we progress to later rounds, we restore the relative data term weighting to that given by Equation (15).

On small data sets or reconstructions with a small field of view, as is the case for simulated EHT data (Section 5.1), DTFTs are sufficient to compute the data terms in Section 3.2. For larger data sets, such as those produced by the VLBA and especially ALMA (Section 5.2), the DTFT matrix Aij becomes prohibitively large to store in memory and prohibitively slow at extracting visibilities from the trial image at each step. In this regime, we use the NFFT package (Readhead & Wilkinson 2009) accessible in Python via the pyNFFT wrapper. 7 The forward NFFT is used for the computation of the irregularly sampled visibilities from the regularly sampled input image, and the adjoint NFFT is used to compute image domain gradient components from data term residuals (see the Appendix).

4.2. Simulated Data

To test the effects of using the different data terms in Section 3.2 in VLBI imaging, we simulated VLBI observations from model images and applied different amounts of uncertainty in the complex station gains ${G}_{i}{e}^{i{\phi }_{i}}$ (Equation (3)). We then produced images using different data term combinations and the regularizers as described in Section 4.1 and compared our results with the true source images.

To generate synthetic data with different degrees of gain error, we generated time-dependent station-based complex gains sampled from known underlying distributions. Because the atmospheric coherence time, which determines the additional phase ϕi added at each station, is much shorter than a typical observing cadence at 1.3 mm, we sampled these phases from a uniform distribution over −π < ϕi < π at each time, independent of the uncertainty in the amplitude.

Our prescription for the amplitude gain terms consisted of a random time-independent offset and a fluctuating part,

Equation (28)

where Xi and Yi are real Gaussian random variables with zero mean, but Xi is drawn only once per telescope per observation and Yi is drawn independently at each time when u, v points are sampled. For simplicity, we chose to use identical standard deviations for the underlying Gaussian distributions of Xi and Yi and call this standard deviation our level of gain error.

We applied our different sampled sets of station-based gains computed at different levels of amplitude error to the ideal visibilities plus identical Gaussian thermal noise using Equation (3). To preserve the S/N, the reported noise standard deviation terms ${\sigma }_{{ij}}$ from Equation (8) were multiplied by the same gain factors Gi and Gj.

We also included the effects of varying elevation and opacity τ on our S/N at each site. The opacity attenuates the measured perfect visibility Vij (before adding thermal noise) by a factor $\sqrt{{e}^{-{\tau }_{i}/\sin {\theta }_{i}}{e}^{-{\tau }_{j}/\sin {\theta }_{j}}}$, where θi and θj are the elevation angles of the source at the different telescopes. This reduces the S/N by the same factor $\sqrt{{e}^{-{\tau }_{i}/\sin {\theta }_{i}}{e}^{-{\tau }_{j}/\sin {\theta }_{j}}}$. This factor can be corrected for by multiplying the measured visibility (including thermal noise) by its inverse using the measured opacity, keeping the reduced S/N constant. In general, the imperfect measurement of opacities introduces an additional source of amplitude gain error. For the purposes of this paper, when we simulate data, we assume the perfect measurement of opacities and set all zenith opacities τi = 0.15.

4.3. Image Evaluation

To evaluate the fidelity of images reconstructed from the data, we followed Chael et al. (2016) and Akiyama et al. (2017b) in using a simple normalized root-mean-square error (NRMSE) fidelity metric. The NRMSE is a point-to-point metric that evaluates images based on pixel-to-pixel similarities rather than common large-scale features. Given two images ${\boldsymbol{A}}$ and ${\boldsymbol{B}}$ with M pixels each, the NRMSE of image ${\boldsymbol{A}}$ relative to ${\boldsymbol{B}}$ is

Equation (29)

Several factors complicate the simple application of Equation (29) in evaluating our reconstructed images. First, often the true source image will contain fine-scale features that are at too high a resolution for any image reconstruction algorithm to capture given the longest projected baseline in the u, v plane. To prevent NRMSE from unduly penalizing reconstructions that successfully reconstruct the lower-resolution features in the data, we convolve both the true and reconstructed image with a Gaussian kernel to blur out high-frequency structure. Since we expect our algorithms to provide some "super-resolution" above the scale corresponding to the longest projected baseline, we choose to blur the images with a Gaussian that has the same proportions as the interferometer "clean" beam—the Gaussian fitted to the central lobe of the Fourier transform of the u, v coverage—but we scale the beam size by a factor of 1/3.

A second complication arises because images reconstructed without calibrated visibility phases are not sensitive to the true position of the image centroid in the field of view, so reconstructed images may be offset from the true source location. In addition, the number of pixels and field of view in the reconstructed image may be different from those in the true source image. Therefore, when comparing images, we first resample our reconstructions onto the same grid as the model image using cubic spline interpolation and then find the overall centroid shift of the reconstruction that produces the maximal cross-correlation before computing the NRMSE with Equation (29).

5. Results

5.1. Results: Simulated EHT Images

We simulated data on EHT baselines from several 230 GHz model images at the positions of the EHT's primary science targets: Sgr A* (R.A.: 17h45m40fs04, decl.:−29°0'28farcs12) and M87 (R.A.: 12h30m49fs42, decl.: +12°23'28farcs04). Our model images were generated by performing general relativistic ray tracing and radiative transfer on the density and temperature distributions from general relativistic magnetohydrodynamic (GRMHD) simulations of hot supermassive black hole accretion disks (Chan et al. 2015; Mościbrodzka et al. 2016; Gold et al. 2017). We also simulated data from a 7 mm VLBA image of the quasar 3C273 (Jorstad & Marscher 2016) rescaled to a smaller field of view of 250 μas, which we placed at the sky location of Sgr A*.

The EHT's station parameters are listed in Table 1. In addition to the full EHT array described in Table 1, we also generated data on a reduced array without the "redundant" sites—JCMT and APEX—that are located at the same location as the more sensitive SMA and ALMA, respectively. The $u,v$ coverage maps for the 2017 EHT when observing Sgr A* and M87 are displayed in Figure 3.

Figure 3.

Figure 3. (Left) EHT 2017 $u,v$ coverage for Sgr A*. The "redundant" (JCMT and APEX) sites make practically no unique contributions to the $u,v$ coverage or nominal resolution aside from adding an effective zero baseline. However, these sites add closure amplitudes that are essential for closure-amplitude imaging to approach the fidelity of imaging with visibility amplitudes. (Right) EHT 2017 $u,v$ coverage for M87.

Standard image High-resolution image

Table 1.  EHT 2017 Station Parameters

Facility Location Diameter (m) SEFD (Jy) X (m) Y (m) Z (m)
JCMT Maunakea, USA 15 6000 −5464584.68 −2493001.17 2150653.98
SMA Maunakea, USA 7(×6) 4900 −5464555.49 −2492927.99 2150797.18
SMT Arizona, USA 10 5000 −1828796.2 −5054406.8 3427865.2
APEX Atacama Desert, Chile 12 3500 2225039.53 −5441197.63 −2479303.36
ALMA Atacama Desert, Chile 40(×12) 90 2225061.164 −5440057.37 −2481681.15
SPT South Pole 10 5000 0.01 0.01 −6359609.7
LMT Sierra Negra, Mexico 50 600 −768715.63 −5988507.07 2063354.85
IRAM Pico Veleta, Spain 30 1400 5088967.75 −301681.186 3825012.206

Note. Current EHT sites are the Atacama Large Millimeter/Submillimeter Array (ALMA), the Large Millimeter Telescope (LMT), the Submillimeter Array (SMA), the Submillimeter Telescope (SMT), the Institut de Radioastronomie Millimétrique (IRAM) telescope on Pico Veleta (PV), the IRAM Plateau de Bure Interferometer (PdB), and the South Pole Telescope (SPT). Note that the PdB did not participate in the 2017 EHT observations.

Download table as:  ASCIITypeset image

In all cases, we use an integration time Δt = 30 s and a bandwidth Δν = 2 GHz, with scans taken every 5 minutes for a full 24 hr rotation of the Earth. The zenith opacity was set to τ = 0.15 at all sites with no uncertainties in the opacity calibration. We did not include the effects of either refractive or diffractive interstellar scattering in our simulated Sgr A* data (see, e.g., Fish et al. 2014; Johnson 2016).

To test the quality of our different imaging methods with different levels of gain uncertainty, we produced one data set for each image with only thermal noise, then generated random gain terms at seven different levels of uncertainty—0%, 5%, 10%, 25%, 50%, 75%, and 100%—as described in Section 4.2.

We reconstructed each data set on a 128 × 128 pixel grid using four different data term combinations: bispectrum (Equation (18)), visibility amplitude and closure phase (Equations (17) and (19)), closure amplitude and closure phase (Equations ((19) and 20)), and log closure amplitude and closure phase (Equations (21) and (19)). In all of our reconstructions, we used all four regularizer terms introduced in Section 3.4 (using "total squared variation" as our smoothness regularizer; Equation (25)). To ensure consistency in our comparisons, we followed the same imaging procedure for all images, arrays, and methods. For each data set, we only changed the image field of view and corresponding initial image, which is also used as the prior in the Sentropy regularizer. The initial/prior image was a circular Gaussian in all cases. The total fluxes, fields of view, and initial image FWHMs are given in Table 2.

Table 2.  Initial/Prior Image Parameters

Image $u,v$ Coverage FOV (μas) Gaussian FWHM Flux (Jy)(μas)
Figure 4 Sgr A* 135 60 2
Figure 5 M87 155 60 2
Figure 6 Sgr A* 255 80 2
Figure 7 Sgr A* 375 25 2

Download table as:  ASCIITypeset image

The parameters that specify our imaging procedure are listed in Table 3. As mentioned in Section 4.1, we image each data set in multiple rounds, blurring out the final image from a given round to serve as the initial image in the next. The FWHM of the circular Gaussian blurring kernel used is given as a fraction fblur of the nominal array resolution. The other imaging parameters listed in Table 3 include the data term and regularizer hyperparameters, αD and βR. For the data terms, in each case, α1 refers to the amplitude term (bispectrum, visibility amplitude, closure amplitude, or log closure amplitude), and α2 is the hyperparameter for the closure-phase term, if present (all methods except bispectrum). We parameterize α2 by stating its ratio fα2 with the correct log-likelihood ratio given by Equation (15). That is, if there are N1 measurements of the first (amplitude) data product and N2 measurements of the second (phase) data product,

Equation (30)

Table 3.  Imaging Parameters

Round ${f}_{\mathrm{blur}}$ ${\beta }_{\mathrm{entropy}}$ ${\beta }_{\mathrm{TSV}}$ ${\beta }_{\mathrm{tot}\mathrm{flux}}$ ${\beta }_{\mathrm{centroid}}$ ${\alpha }_{1}$ ${f}_{\alpha 2}$ ${N}_{\mathrm{iter}}$ ${\epsilon }_{\mathrm{conv}}$
1 N/A 1 1 100 100 100 2 50 10−10
2 0.75 1 50 50 50 100 0.75 150 10−10
3 0.5 1 100 10 10 100 0.5 200 10−10
4 0.33 1 500 1 1 100 1 200 10−10

Download table as:  ASCIITypeset image

Finally, we also list the maximum number of imager steps allowed in each round, Niter, and the convergence criterion, epsilonconv, for the minimum allowed fractional change in the objective function and gradient magnitude between imager steps.

Our results are displayed in Figures 47. In each figure, we show the initial model image, the initial model image blurred with a "clean" beam scaled to 1/3 of its fitted value, and the reconstructions from each method for each level of gain uncertainty, all blurred with the same beam. We also display a plot showing the NRMSE (Equation (29)) for each method as a function of the level of gain error in the underlying data set.

Figure 4.

Figure 4. (Top left) The 230 GHz image from a GRMHD simulation of Sgr A*(Gold et al. 2017). (Top middle) Same image blurred with the effective beam (filled ellipse), 1/3 the size of the fitted CLEAN beam (open ellipse). The image was observed at the sky location of Sgr A* using EHT 2017 baselines, and images were reconstructed with each method using the parameters in Table 3. (Top right) Curves of NRMSE (Equation (29)) vs. gain error for each reconstruction method. (Bottom) Individual reconstructions from each method (y-axis) at each level of gain error (x-axis), blurred with the same beam as the model in the top middle panel. The images and NRMSE curves show that, except at the lowest levels of amplitude gain error, the closure-only results are as faithful to the model as the reconstructions that use either the bispectrum or visibility amplitudes and closure phases. Furthermore, the results of the closure-only methods are insensitive to the level of amplitude gain error, while the reconstructions using visibility amplitude information fail completely starting at the 25% level of gain error.

Standard image High-resolution image
Figure 5.

Figure 5. Reconstructions of a 230 GHz image from a GRMHD simulation of the M87 jet (Mościbrodzka et al. 2016). As in the Sgr A* image in Figure 4, closure-only methods produce results that are as good or better than the bispectrum or visibility amplitude + closure-phase methods in all but the zero gain error case, and the closure-only results are consistent at all levels of gain error.

Standard image High-resolution image
Figure 6.

Figure 6. Reconstructions of a 230 GHz image from a GRMHD simulation of the Sgr A* accretion flow (Chan et al. 2015). Both the closure amplitude and log closure amplitude reconstructions performed consistently at all levels of gain error.

Standard image High-resolution image
Figure 7.

Figure 7. The 43 GHz VLBA image of 3C273 from Jorstad & Marscher (2016), rotated and scaled to a 250 μas field of view. Simulated data were generated using the 230 GHz EHT 2017 Sgr A* $u,v$ coordinates and sensitivities (Figure 3). Unlike the other images in this section, this image is displayed with a log scale, and the NRMSE was computed from the log of the image. The closure-only reconstructions again capture the overall jet structure at all levels of amplitude gain error. With no gain error, imaging directly with closure amplitudes (or log closure amplitudes) instead of visibility amplitudes provides less dynamic range, as is evident from the spurious low-luminosity off-axis features in the closure-only reconstructions, likely resulting from a local minimum in the objective function.

Standard image High-resolution image

Our results indicate that, as long as some redundant sites are included to constrain the reconstruction with "trivial" closure phases and amplitudes, closure-only imaging of EHT data can achieve fidelities nearly as good as bispectral or amplitude + closure phase imaging. As the level of amplitude gain error increases, the fidelity of the results produced using the bispectrum or visibility amplitudes drops quickly, while closure-only imaging is completely insensitive to gain error.

Figures 46 show that imaging with closure amplitudes directly can produce results that are more faithful to the underlying image than reconstructing the image with log closure amplitudes. However, we have found that imaging with the closure amplitudes often takes much longer to converge and is more sensitive to the particular choices of data term weight and initial field of view.

Finally, for the narrow, high dynamic range scaled 3C273 image in Figure 7, we computed the NRMSE using the logarithm of the image. This results in a range of NRMSE values for the bispectrum and visibility amplitude + closure phase images that is substantially lower than those in Figures 46; however, visual inspection of the images shows that in this case, as in Figures 4 and 5, imaging methods that rely on calibrated amplitudes perform significantly worse with increasing gain error and completely fail with amplitude gain error levels >25%. In contrast, the closure-only methods have consistent performance at all levels of amplitude gain. However, the final dynamic range achieved in the closure-only reconstructions is worse than in the images produced with visibility amplitudes with zero gain error, as is evident from the spurious low-luminosity features in the closure-only reconstructions in Figure 7. These features parallel to the jet axis result from being trapped in a local minimum of the objective function, which is invariant to overall image shifts. Since there are no data constraints on certain spatial frequencies due to sparse coverage, these Fourier components can be made large through periodic structure without increasing χ2. Defining a masked region along the jet axis outside which the flux is zero (analogous to a CLEAN box) may help remove these features.

We also compared reconstructions using data from the full EHT 2017 array and the 2017 array without "redundant" sites. Figure 8 shows that in both cases, closure-only methods converge to the same reconstruction for all values of systematic gain error. However, without redundant sites, the results are substantially less accurate, while the closure-only reconstructions with redundant sites included in the data set approach the fidelity of images produced with gain-calibrated amplitudes. "Redundant" sites contribute important short baselines that combine into nontrivial closure amplitudes and act to further constrain the underlying image (Section 2.3). In other words, the closure-only images approach the bispectrum/amplitude + closure phase images in quality as the number of closure amplitudes increases, even if some of those closure quantities contain zero baselines from colocated sites.

Figure 8.

Figure 8. (Top) Image fidelity with the EHT 2017 array. The left panel shows NRMSE curves of image fidelity for reconstructions of the model in Figure 4 with different levels of gain error. The curves are styled consistently with those in Figures 47. The right image is the log closure amplitude + closure phase reconstruction produced at 100% gain error. (Bottom) Image fidelity with the EHT 2017 array without redundant stations (JCMT, APEX). The reconstructions from the data without including the redundant stations are still insensitive to different levels of gain error, but their overall fidelity is worse compared with those produced from data including these redundant stations.

Standard image High-resolution image

5.2. Results: VLBA and ALMA Images

To test its performance on real observations, we applied our closure-only imaging algorithms to millimeter-wavelength interferometric data sets from the VLBA and ALMA. In both cases, the number of visibilities and closure quantities greatly exceeds the number produced by the sparse EHT 2017 array, so we used NFFTs to speed up the imaging procedure. Our first example is a VLBA observation of M87 at 7 mm wavelength. In this case, and for other images with jets or narrow structure (see Figure 7), we have found that the major difficulty in closure-only imaging is convergence in the minimization of the objective function (Equation 14). When we initialize to an uninformative image, the algorithm has difficulty converging to an image that has a reduced χ2 near 1 in either the closure phases or the log closure amplitudes.

To mitigate this problem while still preserving the benefits of only using closure quantities, we have found that initially including the visibility amplitudes in the minimization, even with a low weight and no calibration, can significantly aid the initial convergence. To avoid any bias from self-calibration in the data, we first applied a "null" calibration to the M87 data by assuming a single, constant SEFD (see Equation (8)) for all sites and times. Next, we imaged the data using closure quantities and visibility amplitudes, which were down-weighted by a factor of 10 relative to the closure quantities. We then performed another two rounds toward convergence, initializing to the previous image convolved with a circular Gaussian matching the nominal array resolution but with visibility amplitudes this time down-weighted by a factor of 100. Finally, we performed two additional rounds of imaging using only closure quantities.

Figure 9 compares the reconstructed image to an image reconstructed using CLEAN and iterative self-calibration (Walker et al. 2016, 2018). We also derived a table of complex gains from a single iteration of self-calibration to the final closure image to test how different our self-calibration solution would be from that obtained with CLEAN. The self-calibrated gains are significantly different than our initialized "null" calibration solution; after normalizing to the median gain (effectively fixing the total flux density), although 50% of the visibilities had residual gain corrections of less than 3%, 10% of the visibilities had residual gain corrections of more than 30%. This result justifies post hoc our choice to use visibility amplitudes in the initial minimization steps. The majority of uncalibrated amplitudes have low errors compared to the final self-calibrated set, so they are useful in aiding convergence; however, relying primarily on closure amplitudes ensures a final image that is less affected by the large gain errors present on some baselines.

Figure 9.

Figure 9. Application of closure-only imaging to a VLBA observation of M87 at 7 mm wavelength observed on 2007 May 9 (for details, see Walker et al. 2016, 2018). (Left) CLEAN image made using iterative imaging and self-calibration. (Center) Image reconstructed using closure-only imaging with a weak visibility amplitude constraint to aid initial convergence. (Right) Image reconstructed using complex visibilities after self-calibrating to the closure-only image. To simplify the comparison between these approaches, each image has been convolved with the same CLEAN restoring beam and is rescaled to have the same total flux density as the CLEAN image. Contours in all panels are at equal levels, starting at 9.7 mJy mas−2 (=1 mJy beam−1) and increasing by factors of 2.

Standard image High-resolution image

For comparison, we also applied our self-calibration solution to the data and produced an image with self-calibrated complex visibilities, minimizing Equation (14) with a standard complex visibility χ2 term (Equation (16)). The result is displayed in the third panel of Figure 9. All three methods in Figure 9 give results that are broadly consistent, demonstrating the potential of closure imaging to obtain images that are comparable to those obtained by multiple rounds of finely tuned CLEAN and self-calibration from an expert user.

A general characteristic of closure imaging is its tendency to avoid high-frequency artifacts when highly converged; by removing spurious features from CLEAN images, closure methods could be useful in aiding in the physical interpretation of VLBI images. However, note that the CLEAN reconstruction in Figure 9 recovers more extended structure along the jet. This is likely because the CLEAN reconstructions were done using a multiscale approach (Wakker & Schwarz 1988; Cornwell 2008); a similar multiscale approach could likewise be used to improve closure imaging and is a key goal for improving the methods presented in this paper and applying them to further data sets.

With 64 telescopes, ALMA has baseline coverage that much more densely fills the u, v plane than the EHT and VLBA observations considered above. We imaged a 2014 ALMA observation of HL Tau taken at 1.3 and 0.87 mm (ALMA Partnership et al. 2015) using our log closure amplitude and closure phase method, described in Sections 3.2 and 4.1. We first averaged the data in 5 minute intervals. As in Figure 9, we used down-weighted visibility amplitudes in the initial steps of the minimization to aid in convergence and removed them in the final runs of the imager.

The results are displayed in Figure 10. Closure imaging is able to replicate the overall structure of the published CLEAN images, including all of the gaps in the protoplanetary disk identified by the original reconstruction. Critically, our closure-imaging algorithm does not yet include multiscale imaging (Wakker & Schwarz 1988; Cornwell 2008), which was necessary to produce the ALMA CLEAN image. After producing an image from closure quantities, we self-calibrated the data to the resulting image (center panel of Figure 10) and imaged the data again directly using the resulting complex visibilities (minimizing Equation (14) with Equation (16)). The resulting image has a higher resolution than the closure-only image alone, with sharper and more distinct gaps apparent in the disk (right panel of Figure 10). Furthermore, particularly in the 0.87 mm image, the final reconstruction lacks the prominent periodic dark spots present in the CLEAN image that are likely caused by prominent dirty-beam sidelobes, which were also ameliorated in recently reprocessed CLEAN + self-calibration images by Akiyama et al. (2016).

Figure 10.

Figure 10. (Top) The 1.3 mm band 6 ALMA image of the protoplanetary disk around HL Tau, comparing the CLEAN reconstruction from ALMA Partnership et al. (2015) with our reconstructions. The left panel shows the CLEAN image with a field of view of 1farcs8. The center panel shows an image of the same data produced by directly fitting to log closure amplitudes and closure phases, with down-weighted visibility amplitudes used in the initial steps to aid convergence. Closure-only imaging produces an image that is consistent with the CLEAN result, despite not using any multiscale imaging, but the overall resolution is lower. The right panel shows an image produced from complex visibilities using a strong total variation regularizer after self-calibrating the data to the center closure-only image. After self-calibration, complex visibility imaging with total variation produces a sharp image with distinct disk gaps. (Bottom) The 0.87 mm band 7 ALMA images, produced using the same imaging parameters as the 1.3 mm images. The 0.87 mm image obtained after closure-only imaging and one round of self-calibration eliminates prominent CLEAN artifacts (dark spots) present in the original image. Our 0.87 mm image is similar to recently reprocessed images using CLEAN and a modified self-calibration loop (Akiyama et al. 2016).

Standard image High-resolution image

6. Discussion

The results in Sections 5.1 and 5.2 demonstrate that closure amplitudes and phases can be directly used in interferometric imaging to produce images that are insensitive to phase and amplitude calibration errors. Traditional self-calibration and imaging loops require many iterations of CLEAN imaging and fitting complex gains to visibilities (e.g., Wilkinson et al. 1977; Readhead & Wilkinson 1978; Readhead et al. 1980; Schwab 1980; Cornwell & Wilkinson 1981; Pearson & Readhead 1984; Walker 1995; Cornwell & Fomalont 1999). These loops contain many tunable parameters, including the choice of initial source model, the strategy for independent or concurrent calibration of amplitude and phase gains, the CLEAN convergence criterion, the choice of taper and weighting for the CLEAN visibilities, and the scales and regions to clean in each CLEAN iteration.

Our closure-imaging method does not remove all tunable parameters from the model, but imaging with closure quantities alone necessarily produces results that are less biased by calibration assumptions. Images from closure imaging can stand on their own as minimal-assumption estimates of the source structure; alternatively, results from closure-only imaging may be used as a well-motivated self-calibration model or initial source image for other imaging pipelines using calibrated data. On the ALMA and VLBA data sets in Section 5.2, we found that just one round of self-calibration to an image produced with closure quantities can be used to produce smooth high-resolution images from the resulting complex visibilities that match the best iterative, multiscale CLEAN + self-calibration results.

The most significant challenge that we have encountered in closure-only imaging is difficulty in the early convergence and a tendency to quickly get stuck in wildly incorrect local minima. Counterintuitively, this tendency seems to be more of a problem in data sets with more interferometric baselines. This limitation may arise because the energy landscape represented by the closure amplitude terms (Equations (20) and (36)) becomes increasingly complicated with more correlated closure data. When using simulated data from the sparse EHT array (Section 5.1), using closure quantities alone with a reasonable Gaussian prior and several imaging iterations is enough to guide the algorithm to converge on a reasonable image. However, imaging the real data sets from ALMA and the VLBA in Section 5.2 using closure quantities alone with an uninformative initial model was difficult. For these cases, we found that adding a weak data constraint using uncalibrated visibility amplitudes (Equation (17)) helped guide the minimizer to the region of a good local minimum. This constraint can be as low as 1%–10% of the closure amplitude χ2 term and still produce excellent results; in practice, the amplitude error bars can also be increased to represent an estimate of the systematic amplitude error and further down-weight this term (K. Akiyama et al. 2018, in preparation). As the imaging proceeds, we remove the amplitude constraint and ultimately allow the final image to be guided only by the closure amplitudes and phases. Given the robustness of the results in Figures 9 and 10 to different choices of regularizer and data weights in the presence of a weak amplitude constraint, we see significant promise for this method to eventually allow for unsupervised closure imaging that can blindly produce a calibration-free image from decent initial data without user intervention.

In the eht-imaging library, we have developed a flexible framework where images can be easily produced from the same data set using different data and regularizer terms. We can also use eht-imaging to self-calibrate data, to generate synthetic data from images with realistic thermal errors and calibration uncertainties, and for the general plotting, analysis, and comparison of interferometer data. Within this framework, it is easy to experiment with different arbitrary combinations of data terms and implement new imaging methods, such as polarimetric imaging (Chael et al. 2016), imaging in the presence of refractive scattering (Johnson 2016), and producing continuous movies from multi-epoch observations (Johnson et al. 2017).

7. Conclusions

We have presented a framework for interferometric imaging using regularized maximum likelihood with arbitrary data products and its implementation in the software library eht-imaging. This work builds on decades of past work in applying regularized maximum-likelihood approaches to interferometric imaging and is particularly inspired by the simultaneous minimization of multiple data terms pioneered in optical interferometric imaging (see, e.g., Thiébaut 2013; Thiébaut & Young 2017). This work extends that framework by imaging data directly with closure amplitudes (or their logarithms) for the first time, rather than relying on amplitude self-calibration.

In Section 3.3 and the Appendix, we give analytic expressions for the gradients of various data χ2 terms, including those for closure phases, closure amplitudes, and log closure amplitudes. The most powerful feature of this framework is its ability to produce images using closure quantities directly, making it possible to produce images directly from uncalibrated data.

Using our method of closure-only imaging, self-calibration in imaging can be bypassed entirely, producing an image that will contain minimal calibration assumptions and will not depend on the choice of initial self-calibration model or other assumptions made in the self-calibration loop. In Section 5.1, we showed that this strategy performs well on simulated EHT data of Sgr A* and M87. Images produced using only closure quantities have consistent fidelity at all levels of amplitude gain or miscalibration. Furthermore, when redundant sites are included in the array, the overall fidelity of the results approaches that of images made with perfectly calibrated data using conventional algorithms.

In Section 5.2, we showed that closure imaging can also produce high-quality images for VLBA and ALMA data sets at millimeter wavelengths, giving results that are of comparable quality to expert reconstructions with multiscale CLEAN and self-calibration. Results from closure-only imaging can also be used to self-calibrate data and initialize additional imaging. We found that for the ALMA data sets considered, just one round of self-calibration and complex visibility imaging after closure-only imaging produces further refined results with fewer suspicious features that may be attributable to artifacts from CLEAN.

Techniques involving calibration-insensitive closure quantities like those presented in this paper can help push interferometric imaging to more and more challenging regimes, including higher frequencies. While applicable to all interferometric astronomical data, our techniques are especially valuable at millimeter and submillimeter wavelengths, where calibration uncertainties are a large and variable component of the error budget. In Section 5.2, we found that including a soft constraint from uncalibrated visibility amplitudes can dramatically aid in the convergence of closure imaging. Adding more data terms, this method can be easily generalized for polarimetric imaging (Chael et al. 2016; Akiyama et al. 2017a), spectral index maps, simultaneous multiband images, scattering mitigation (Johnson 2016), and dynamical movies of multi-epoch data (Bouman et al. 2017; Johnson et al. 2017).

We extend particular thanks to the anonymous referee, whose suggestion of the NFFT library and unusually thorough and thoughtful comments substantially improved this work. We thank Craig Walker and Michael Janßen for their helpful conversations, comments, and suggestions. We thank Roman Gold, Monica Mościbrodzka, and Chi–Kwan Chan for providing the model images used in Section 5.1 and Svetlana Jorstad and Alan Marscher for the 43 GHz 3C273 image used in the same section. We additionally thank Craig Walker for providing the 7 mm VLBA data used in Section 5.2. We thank the National Science Foundation (AST-1440254, AST-1614868, AST-1312651) and the Gordon and Betty Moore Foundation (GBMF-5278) for financial support of this work. This work was supported in part by the Black Hole Initiative at Harvard University, which is supported by a grant from the John Templeton Foundation. KA is financially supported by the Jansky Fellowship of the National Radio Astronomy Observatory.

This study makes use of 43 GHz VLBA data from the VLBA-BU Blazar Monitoring Program (VLBA-BU-BLAZAR; http://www.bu.edu/blazars/VLBAproject.html), funded by NASA through the Fermi Guest Investigator Program. The VLBA is operated by the Long Baseline Observatory. The Long Baseline Observatory is a facility of the National Science Foundation, operated under cooperative agreement by Associated Universities, Inc. This paper makes use of the following ALMA data: ADS/JAO.ALMA#2011.0.00015.SV. ALMA is a partnership of the ESO (representing its member states), NSF (USA), and NINS (Japan), together with the NRC (Canada), NSC and ASIAA (Taiwan), and KASI (Republic of Korea), in cooperation with the Republic of Chile. The Joint ALMA Observatory is operated by the ESO, AUI/NRAO, and NAOJ. The National Radio Astronomy Observatory is a facility of the National Science Foundation, operated under cooperative agreement by Associated Universities, Inc.

Appendix:

We present here the expressions for the gradients of the various data terms presented in Section 3.2 that we use in our imaging software. The equations below assume a DTFT matrix Aij (see Equation (13)); the conjugate transpose matrix ${A}_{{ij}}^{\dagger }$ gives the adjoint DTFT matrix. (Note that since the visibility data is sparsely sampled, ${A}^{\dagger }A\ne 1$.)

The gradient of the complex visibility χ2 term (Equation (16)), already presented in the main text as Equation (22), is

Equation (31)

The gradient of the visibility amplitude χ2 (Equation (17)) is

Equation (32)

For the bispectrum ${\chi }^{2}$ (Equation (18)), the gradient is

Equation (33)

The closure-phase χ2 (Equation (19)) has a gradient

Equation (34)

And finally, the gradient of the closure amplitude χ2 term (Equation (20)) is

Equation (35)

and for the log closure amplitudes (Equation (21)), it is

Equation (36)

Footnotes

Please wait… references are loading.
10.3847/1538-4357/aab6a8