Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Digital-holographic detection in the off-axis pupil plane recording geometry for deep-turbulence wavefront sensing

Open Access Open Access

Abstract

This paper uses wave-optics and signal-to-noise models to explore the estimation accuracy of digital-holographic detection in the off-axis pupil plane recording geometry for deep-turbulence wavefront sensing. In turn, the analysis examines three important parameters: the number of pixels across the width of the focal-plane array, the window radius in the Fourier plane, and the signal-to-noise ratio. By varying these parameters, the wave-optics and signal-to-noise models quantify performance via a metric referred to as the field-estimated Strehl ratio, and the analysis leads to a method for optimal windowing of the turbulence-limited point spread function. Altogether, the results will allow future research efforts to assess the number of pixels, pixel size, pixel-well depth, and read-noise standard deviation needed from a focal-plane array when using digital-holographic detection in the off-axis pupil plane recording geometry for estimating the complex-optical field when in the presence of deep turbulence and detection noise.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

Corrections

19 January 2018: A correction was made to the copyright.

1. INTRODUCTION

Digital-holographic detection, in practice, provides us with an estimate of the complex-optical field. As such, it can be applied to a variety of applications, which require the propagation of coherent light through distributed-volume atmospheric aberrations or “deep turbulence.” These applications include atmospheric characterization [1,2], adaptive-optics phase compensation [3,4], remote sensing [5,6], and free-space laser communications [7,8] (to name a few). With these applications in mind, this paper varies the parameters associated with digital-holographic detection in the off-axis pupil plane recording geometry (PPRG) to investigate its use as a deep-turbulence wavefront sensor (WFS). The goal throughout is to examine the parameters needed to produce worthwhile estimates of the amplitude and wrapped phase associated with the complex-optical field.

In practice, several different digital-holographic recording geometries exist, such as the off-axis image plane and on-axis phase shifting recording geometries [912]; however, this paper focuses on the off-axis PPRG. As shown in Fig. 1(a), the off-axis PPRG gets its namesake from placing the focal-plane array (FPA) in a pupil plane. Recall that in order to obtain an estimate of the complex-optical field, we need to first interfere a signal beam with a reference beam, so that we can record a digital hologram with the FPA. Using signal-processing techniques, the off-axis PPRG ultimately allows for the isolation of the desired complex-optical field, all from a single digital hologram. These advantages make the off-axis PPRG an ideal candidate for use in deep-turbulence applications, which involve high-resolution, low-latency wavefront sensing.

 figure: Fig. 1.

Fig. 1. Above is a diagram of digital-holographic detection in the off-axis PPRG. In (a), the coherent light from a master-oscillator (MO) laser actively illuminates a distant cooperative object (e.g., an unresolved ball bearing) creating a point-source beacon that propagates through deep turbulence to the entrance-pupil plane of an imaging system. Through collimation, the entrance-pupil plane creates a signal beam, which we interfere with a reference beam in the exit-pupil plane of the imaging system. Note that the reference beam manifests itself from an off-axis local oscillator (LO), which we derive from the MO laser. Also note that the exit-pupil plane is where we place the FPA, so that we can record a digital hologram. In (b), we can then use signal-processing techniques to manipulate the digital hologram into a wrapped-phase estimate.

Download Full Size | PDF

Current wavefront-sensing techniques suffer from degraded performance when in the presence of deep turbulence. For example, if the turbulence is strong enough and the propagation distance is long enough, the distributed-volume phase aberrations can cause the coherent light from a point-source beacon to interfere with itself [cf. Fig. 1]. Known as scintillation, this self interference is one cause of low signal-to-noise ratios (SNRs) over the field-of-view of the WFS, as well as branch points, which result from total-destructive interference. Branch points, in practice, also cause branch cuts to arise in the phase function. These 2π phase discontinuities get mapped to the null space of least-squares phase unwrapping and reconstruction algorithms [13]; thus, traditional wavefront-sensing techniques (e.g., the Shack–Hartmann WFS [14]) do not estimate the correct phase function when in the presence of deep turbulence. It is important to note that branch-point tolerant phase unwrapping and reconstruction algorithms do exist within the open literature [3,15]; however, the performance of these algorithms needs to be quantified with hardware solutions [16].

With the harmful effects of scintillation in mind, another cause of low SNRs is the extinction of light as it propagates through the atmosphere. Extinction, in general, results from molecular and aerosol absorption and scattering and ultimately leads to transmission losses through Beer’s law [17,18]. Together, the low-SNR effects of scintillation and extinction plague our ability to perform deep-turbulence wavefront sensing using traditional wavefront-sensing techniques. This outcome reinforces the need for a more robust WFS when in the presence of deep turbulence (i.e., scintillation and extinction).

With the use of a strong reference beam (e.g., one that is a large fraction of the FPA’s pixel-well depth), digital-holographic detection overcomes the low-SNR challenges caused by scintillation and extinction. In practice, a strong reference beam can boost the overall power of the signal beam to be well above the read-noise floor of the FPA. When faced with low SNRs, this boost allows us to approach the shot-noise limit (depending on the parameters of the FPA), which is a fundamental limit that does not allow for any more sensitivity with respect to detection. In regards to branch points, digital-holographic detection also grants access to a direct estimate of the wrapped phase; thus, we can directly sense the correct phase function when in the presence of deep turbulence.

In what follows, this paper quantifies the abilities of digital-holographic detection in the off-axis PPRG to accurately estimate the complex-optical field when in the presence of deep turbulence and detection noise. With that said, Section 2 gives an overview of the wave-optics and signal-to-noise models (both analytical and numerical) that we will use within the analysis. It also develops several parameters that we will use in an associated trade study. Section 3 then provides the results from this trade study with appropriate discussion. Following that, Section 4 provides a conclusion for this paper.

Before moving on to the next section, it is worth mentioning that this paper builds upon the analysis contained in a recent conference proceeding by Banet et al. [19] and is a “companion paper” to an analysis presented by Spencer et al. [10] with respect to digital-holographic detection in the off-axis image plane recording geometry (IPRG). Specifically, this paper formulates an analytical wave-optics model, which describes digital-holographic detection in the off-axis PPRG. It also includes detection noise in the numerical wave-optics model. In so doing, this paper validates the use of a closed-form expression for the SNR and formulates a method for optimal windowing when in the presence of deep turbulence and detection noise. Thus, this paper significantly expands upon the analysis presented in [19], which did not develop an analytical wave-optics model nor include the effects of detection noise in its numerical wave-optics model. The analysis also shows that, given a point-source beacon, the off-axis PPRG is less prone to pixel saturation because of focal-plane hot spots caused by deep-turbulence scintillation. This outcome results from the use of a stronger reference beam (e.g., one that is a larger fraction of the FPA’s pixel-well depth when compared with digital-holographic detection in the off-axis IPRG [10]). In comparison with the off-axis IPRG, we can better approach the shot-noise limit when using digital-holographic detection in the off-axis PPRG. This last point is important to remember when faced with low SNRs, for example, when investigating the branch-point problem with hardware solutions [16]. As such, it is our belief that the trade-space analysis contained in this paper will better enable future research efforts to design and implement digital-holographic detection in the off-axis PPRG for the purpose of exploring the branch-point problem.

2. WAVE-OPTICS AND SIGNAL-TO-NOISE MODELS

In this section, we develop wave-optics and signal-to-noise models (both analytical and numerical) for digital-holographic detection in the off-axis PPRG [cf. Fig. 1]. It is important to note that, throughout this paper, the computational simulations use MATLAB along with WaveProp [20] and AOTools [21], which are MATLAB toolboxes written by the Optical Sciences Company. For additional insight into these wave-optics and signal-to-noise models, we encourage the reader to review [9,10,19].

A. Analytical Wave-Optics Model

For the analytical wave-optics model, let’s start at the exit-pupil plane described in Fig. 1(a). Here, we can describe the signal complex-optical field US(x2,y2) or “signal beam” that is incident on the FPA as

US(x2,y2)=US(x2,y2)cyl(x22+y22D2),
where US(x2,y2) is the signal complex-optical field that is incident on the exit-pupil plane,
cyl(x2+y2)={10x2+y2<0.50.5x2+y2=0.50x2+y2>0.5
is a cylinder function, and D2 is the exit-pupil diameter. Next, we can describe the reference complex-optical field UR(x2,y2) or “reference beam” that is incident on the FPA as
UR(x2,y2)=ARexp[j2πxRx2MTλz]exp[j2πyRy2MTλz],
where AR is a complex constant; MT is the transverse magnification between the entrance- and exit-pupil planes, such that
MT=D2D1;
D1 is the entrance-pupil diameter; λ is the wavelength; z is the propagation distance between the object plane and entrance-pupil plane; and (xR,yR) are the spatial coordinates of the off-axis LO, which gives rise to the reference beam.

From the signal and reference beams [cf. Eqs. (1) and (3)], we can represent the real-valued hologram irradiance IH(x2,y2) that is incident on the FPA as

IH(x2,y2)=|US(x2,y2)+UR(x2,y2)|2=|US(x2,y2)|2+|UR(x2,y2)|2+US(x2,y2)UR*(x2,y2)+UR(x2,y2)US*(x2,y2),
where the superscript * denotes complex conjugate. Now, let’s assume that the FPA has N×M pixels separated by pixel pitches of xs and ys along the x and y axes, respectively. Let us also assume that the FPA has pixel widths of wx and wy along the x and y axes, respectively. In turn, the per-pixel average hologram irradiance I^H(nxs,mys) becomes [22,23]
I^H(nxs,mys)=[IH(x2,y2)×1wxrect(x2nxswx)1wyrect(y2myswy)]dx2dy2,
where n is an integer from 1 to N, m is an integer from 1 to M, and
rect(x)={10|x|<0.50.5|x|=0.50|x|<0.5
is a rectangle function. We can rewrite the per-pixel average hologram irradiance, I^H(nxs,mys), in terms of a per-pixel mean number of photoelectrons [24,25], m¯H(nxs,mys), viz.
m¯H(nxs,mys)=ηThνI^H(nxs,mys)wxwy,
where η is the quantum efficiency of the FPA, T is the integration time, h is Planck’s constant, and ν is the temporal frequency of the signal and reference beams in units of inverse seconds.

Provided Eqs. (5)–(8), the continuous photoelectron density dH(x2,y2), in units of photoelectrons (pe) per square meter follows as [22,23]

dH(x2,y2)=m¯H(x2,y2)×1xscomb(x2xs)1yscomb(y2ys)×rect(x2Nxs)rect(y2Mys),
where
m¯H(x2,y2)=ηThν[IH(x2,y2)×rect(x2x2wx)rect(y2y2wy)]dx2y2=ηThνIH(x2,y2)**rect(x2wx)rect(y2wy)
is the continuous form of Eq. (8) given Eq. (6) and the fact that rect(x) is an even function;
1|w|comb(xw)=n=δ(xnw)
is a scaled comb function;
δ(x)=1|w|limw0p(xw)
is a Dirac-delta function; and p(x) is a pulse-like function [e.g., rect(x)]. The reader should note that the ** operator in Eq. (10) denotes 2D convolution, viz.
V(x,y)**W(x,y)=V(x,y)W(xx,yy)dxdy,
where x and y are dummy variables of integration.

To go to the Fourier plane, we must use a 2D Fourier transform, such that

V˜(νx,νy)=V(x,y)ej2π(xνx+yνy)dxdy.
Per se, we obtain the following relationship from Eq. (9):
d˜H(νx,νy)=ηThνα˜(νx,νy)I˜H(νx,νy)**wxcomb(xsνx)wycomb(ysνy)**Nxssinc(Nxsνx)Myssinc(Mysνy),
where
α˜(νx,νy)=sinc(wxνx)sinc(wyνy)
and sinc(x)=sin(πx)/(πx) is a sinc function. In Eq. (15), I˜H(νx,νy) follows as [cf. Eqs. (1), (3), and (5)]
I˜H(νx,νy)=Γ˜(νx,νy)+|AR|2δ(νx)δ(νy)+AR*U˜S(νx+xRMTλz,νy+yRMTλz)+ARU˜S*(νxxRMTλz,νyyRMTλz),
where
Γ˜(νx,νy)=U˜S(νx,νy)**U˜S*(νx,νy).
Note that the first term in Eq. (17) is an autocorrelation of the transformed signal beam centered at D.C. Also centered at D.C. are separable Dirac-delta functions of strength |AR|2. The final two terms are the transformed signal beam and its complex conjugate, each scaled by AR and shifted diagonally in opposing directions. This scaling and shifting are what allows us to isolate the third term on the right-hand side of Eq. (17), as we will see in the coming analysis. Also note that, when in the presence of deep turbulence, the transformed signal beam is nothing more than a scaled and shifted turbulence-limited amplitude spread function (ASF) [26], given a point-source beacon [cf. Fig. 1].

Moving forward, let us convert from spatial frequencies to spatial coordinates, so that νx=x0/(MTλz) and νy=y0/(MTλz). In effect, this choice scales the Fourier plane so that it is equivalent to the complex-optical field after propagation from the entrance-pupil plane back to the object plane. Let us also assume a uniform pixel pitch, so that pxs=ys=wx=wy. Thus, NPN=M is the number of FPA pixels across the exit-pupil diameter D2.

From Eqs. (15) and (17), we arrive at the following relationship:

d˜H(x0MTλz,y0MTλz)=ηThνα˜(x0MTλz,y0MTλz)×{1(MTλz)2Γ˜(x0MTλz,y0MTλz)+|AR|2δ(x0MTλz)δ(y0MTλz)+AR*U˜S(x0+xRMTλz,y0+yRMTλz)+ARU˜S*(x0xRMTλz,y0yRMTλz)}**p2(MTλz)2comb(px0MTλz)comb(py0MTλz)**NP2p2(MTλz)2sinc(NPpx0MTλz)sinc(NPpy0MTλz).
We can simplify Eq. (19) even further by noticing that the quantity, MTλz/p, is equivalent to the Fourier-plane side length, LF, such that
LF=MTλzp=D2D1λzp=λzD1NP=ρINP.
Here, ρI=λz/D1 is approximately half the diffraction-limited bucket diameter DI=2.44λz/D1. This simplification yields the following relationship:
d˜H(x0MTλz,y0MTλz)=ηThνα˜(x0MTλz,y0MTλz)×{1(MTλz)2Γ˜(x0MTλz,y0MTλz)+|AR|2δ(x0MTλz)δ(y0MTλz)+AR*U˜S(x0+xRMTλz,y0+yRMTλz)+ARU˜S*(x0xRMTλz,y0yRMTλz)}**1ρI2NP2comb(x0ρINP)comb(y0ρINP)**1ρI2sinc(x0ρI)sinc(y0ρI).

In units of pe, we can see that the terms within the curly braces in Eq. (21) repeat themselves at regular intervals of the Fourier-plane side length LF in both the x and y directions. This outcome is the result of 2D convolution with the separable comb functions [cf. Eq. (11)]. The second 2D convolution with the separable sinc functions serves to smooth out these repeated terms, whereas the envelope function α˜(νx,νy) [cf. Eq. (16)] serves to amplitude modulate these repeated terms. Before moving on in the analysis, the reader should note that this result is physically relevant, as the sampling theorem dictates that a sampled function [e.g., the continuous photoelectron density dH(x2,y2) in Eq. (9)] becomes periodic upon finding its spectrum [22,23].

Provided Eqs. (19)–(21), we can use a window function w(x0,y0) to isolate the turbulence-limited ASF [i.e., the third term within the curly braces of Eq. (21)]. In general,

w(x0,y0)=cyl(x02+y022ρW),
where ρW is the radius of w(x0,y0) in the Fourier plane. At its maximum value in the Fourier plane, ρWLF/4, so that w(x0,y0) does not appreciably sample the smoothed Dirac-delta term in Eq. (21).

In using Eqs. (21) and (22), we must satisfy Nyquist sampling with the FPA pixels, so that the repeated terms within the Fourier plane do not overlap and cause significant aliasing. Here, the Nyquist rate is LF and the Nyquist interval is 1/LF [22]. If we also notice that we can express the Fourier-plane side length LF as

LF=MTr0pλzr0=QPλzr0,
where QP is the number of FPA pixels across the Fried coherence diameter r0 after it has been demagnified, and, if we assume that the bandwidth of the transformed signal beam is approximately
DT=2.44λzr0,
where DT is the turbulence-limited bucket diameter [9], then we can set the bandwidth of the transformed signal beam equal to LF/2 (i.e., the maximum possible diameter of the window function) and solve for the minimum value of QP. In turn,
QP2λzr02.44λzr0,
yielding
QP4.88.
The reader should note that if we set QP=4.88, then we arrive at the minimum number of FPA pixels NP across the exit-pupil diameter D2 required to largely avoid aliasing.

Provided Eqs. (22)–(26), we can use Eq. (21) to obtain an estimate of the turbulence-limited ASF. The reader should note once more that the final 2D convolution with the separable sinc functions in Eq. (21) acts to smooth out the Fourier plane. With that said, this diffraction-limited smoothing is often negligible when in the presence of deep turbulence. In the limit that ρI0 and QP4.88, we can make use of the convolution-sifting property of the Dirac-delta function and neglect the 2D convolutions in Eq. (21). As a result, we can shift the Fourier plane, so that the origin is located at

(xRMTλz,yRMTλz).
From here, we can write the following relationship:
ψ˜^S(x0MTλz,y0MTλz)d˜H(x0xRMTλz,y0yRMTλz)w(x0,y0)ηThνα˜(x0xRMTλz,y0yRMTλz)AR*AS=ψ˜S(x0MTλz,y0MTλz)w(x0,y0),
where ψ˜^S(νx,νy) is an estimate of the normalized turbulence-limited ASF and ψ˜S(νx,νy)=U˜S(νx,νy)/AS is the normalized turbulence-limited ASF. In using Eq. (28), we need to assume that the amplitude of the reference beam is much, much greater than the amplitude of the signal beam (i.e., |AR||AS|). This strong-reference-beam assumption, along with the shifted origin, ensures that the window function w(x0,y0) effectively isolates the turbulence-limited ASF from all the other terms contained in the Fourier plane [cf. Eq. (17)].

To obtain an estimate of the normalized signal beam ψ^S(MTx1,MTy1), which is incident on the entrance-pupil plane, we need to use a 2D inverse Fourier transform, such that

V(x,y)=V˜(νx,νy)ej2π(xνx+yνy)dνxdνy.
With this in mind, let us convert from spatial coordinates back to spatial frequencies, so that x0=MTλzνx and y0=MTλzνy. From Eqs. (22) and (28), we then obtain the following relationship:
ψ^S(MTx1,MTy1)ψS(MTx1,MTy1)**πρW2(λz)2somb(2ρWx12+y12λz),
where ψS(MTx1,MTy1)=US(MTx1,MTy1)/AS is the normalized signal beam that is incident on the entrance-pupil plane,
somb(x2+y2)=2J1(πx2+y2)πx2+y2
is a sombrero function, and J1(πx2+y2) is a Bessel function of the first kind. Before moving onto the next section, it is important to note that the 2D convolution with the sombrero function in Eq. (30) acts to smooth out the normalized-signal-beam estimate. Accordingly, for small values of λz/(2ρW), the smoothing becomes minimized; however, for large values of λz/(2ρW), the smoothing becomes more pronounced. In turn, there is a distinct trade space found in using Eq. (30). We will explore this trade space when in the presence of deep turbulence and detection noise in the analysis to come.

B. Numerical Wave-Optics Model

For the numerical wave-optics model or “model” for short, we used MATLAB along with WaveProp and AOTools to simulate complex-optical fields via discretized N×N grids. Starting at the simulated object plane [cf. Fig. 1(a)], the model used a narrow sinc function modulated by a raised-cosine envelope in a 4096×4096 grid to emulate a point-source beacon of wavelength λ=1μm. In order to avoid the detrimental effects of aliasing, the sampling of this function and the object-plane side length was set, so that, after propagation from the object plane to the entrance-pupil plane, the illuminated region of interest was half the user-defined, entrance-pupil plane side length (i.e., the simulations satisfied Fresnel scaling, such that N=S0S1/(λZ), where S0 and S1=16D1 are the object and entrance-pupil plane side lengths, respectively). The model then used the split-step beam propagation method to propagate the light from the point-source beacon a distance of Z=7.5km through 10 equally spaced Kolmogorov phase screens, which simulated the distributed-volume phase aberrations indicative of deep turbulence [26]. With 10 phase screens, the continuous and discrete calculations for the turbulence scenarios presented in Table 1 were within 1% error.

Tables Icon

Table 1. Turbulence Scenarios Used in the Trade-Space Analysis

With Table 1 in mind, a great deal of work has been done to characterize regimes in which deep-turbulence conditions exist. For example, in terms of optics, we often characterize turbulence with an index-of-refraction structure parameter Cn2, which varies both seasonally and diurnally due to temperature fluctuations in the atmosphere. Given Cn2, we can determine the Fried coherence diameter r0, which gives us a metric for “seeing” [18,2629]. This is said because increasing the size of the entrance-pupil diameter D1 beyond r0 yields diminishing returns in terms of imaging resolution. For horizontal-propagation paths (like those studied in this paper), the Fried coherence diameters for a plane wave and a spherical wave follow as

r0pw=0.185(λ2ZCn2)3/5
and
r0sw=0.33(λ2ZCn2)3/5,
respectively. It is important to note that, when D1/r0sw>4, higher-order aberrations beyond tilt start to limit the achievable imaging resolution. In this paper, D1=30cm.

We can also determine the log-amplitude variance σχ2 from the index-of-refraction structure parameter Cn2. In practice, σχ2 gives us a gauge for the amount of scintillation that the propagated light has experienced [18,2629]; as such, we can also use it as a gauge for the density of branch points. For horizontal-propagation paths, the log-amplitude variances for a plane wave and a spherical wave follow as

σχpw2=0.307k7/6Z11/6Cn2
and
σχsw2=0.124k7/6Z11/6Cn2,
respectively, where k=2π/λ is the angular wavenumber. For all intents and purposes, branch points start to accumulate in the phase function when σχsw2>0.25. In the next section, we will use the log-amplitude variances and Fried coherence diameters presented in Table 1 to parameterize the trade-space analysis.

Given Eqs. (32)–(35), at the simulated entrance-pupil plane, a collimating lens of diameter D1=30cm removed the parabolic phase from the propagated light, and the model created a “truth field” US(x1,y1) by cropping out the center 256×256 grid points. With that said, the model accounted for the demagnification between the simulated entrance- and exit-pupil planes via interpolation. As shown in Figs. 2(a) and 2(b), the model created a signal beam US(x2,y2) by interpolating US(x1,y1) to have the correct number of FPA pixels NP across the exit-pupil diameter D2, resulting in an NP×NP grid. Upon interpolating, the model interfered US(x2,y2) with a reference beam UR(x2,y2) to obtain a hologram irradiance IH(x2,y2) [cf. Eqs. (1), (3), and (5)]. Note that the model adjusted the LO coordinates (xR,yR) in Eq. (3) to be 1/4 the simulated Fourier plane side length LF [cf. Eq. (20)]. In addition, the model set the amplitude of the reference beam |AR| to produce a per-pixel mean number of reference photoelectrons equal to 75% of the pixel-well depth, as opposed to 25% of the pixel-well depth in [10] (i.e., m¯R=75,000pe in this study for the off-axis PPRG and m¯R=25,000pe in a companion study for the off-axis IPRG). Also note that we observed no saturated pixels as a result of this choice. The model then scaled the amplitude of the signal beam |AS| to produce a per-pixel mean number of signal photoelectrons, such that m¯S=0.28328.3pe. By varying m¯S, the model allowed us to examine performance when in the presence of varying levels of detection noise.

 figure: Fig. 2.

Fig. 2. In (a) and (b), we show the normalized irradiance of the signal beam, whereas in (c) and (d), we show the resultant digital holograms for two different values of the number of detector pixels, NP, across the exit-pupil diameter, D2 (white circles). The black-dashed circles have diameters equal to the demagnified spherical-wave Fried coherence diameter, that is, MTr0sw for turbulence Scenario 5 in Table 1. It is important to note that we display the associated turbulence-limited sampling quotient for a spherical wave QPsw in the bottom-left corner of each subplot.

Download Full Size | PDF

With regard to detection noise, the model included both shot noise and read noise within the computational simulations. Shot noise followed a Poisson distribution, whereas read noise followed a Gaussian distribution. We set the read-noise standard deviation to 100 pe and the pixel-well depth to 100×103pe. As such, the standard deviation of the shot noise varied within the computational simulations and was the dominate source of detection noise.

Following the above steps, we obtained our digital holograms [cf. Figs. 2(c) and 2(d)]. From here, the model performed signal-processing techniques to go to the simulated Fourier plane, isolate the turbulence-limited ASF, and obtain an estimate of the complex-optical field in the simulated entrance-pupil plane, that is, the “estimated field” U^S(x1,y1). The reader can find a pictorial representation of these signal-processing techniques in Fig. 1(b). For all intents and purposes, these signal-processing techniques took place in five steps as follows.

  • 1. The model padded the digital holograms with zeros to contain twice as many pixels across one side (e.g., a 256×256 padded grid for NP=128).
  • 2. The model took a 2D fast Fourier transform (FFT) of the zero-padded digital holograms to go to the Fourier plane.
  • 3. In the simulated Fourier plane, the model used circular window functions of varying radii to isolate the turbulence-limited ASF.
  • 4. The model zero-padded the turbulence-limited ASF, so that the padded array had twice as many grid points across one side of the grid as the number of grid points across the truth field (i.e., 512 grid points).
  • 5. The model took a 2D inverse FFT of the zero-padded turbulence-limited ASF to back out the estimated field at the simulated entrance-pupil plane.

Within these steps, we can infer several important conclusions about the number of FPA pixels, NP, across the exit-pupil diameter, D2, and the radius, ρW, of the window function in the simulated Fourier plane. First, as evidenced in Fig. 3, varying NP alters the side length of the simulated Fourier plane per Eq. (20), which shows that the Fourier-plane side length LF is directly proportional to NP. While NP determines LF, the value of ρW actually determines the portion of the turbulence-limited point spread function (PSF) (i.e., the irradiance associated with the turbulence-limited ASF) to be inverse Fourier transformed. Thus, ρW determines what spatial-frequency content makes it into the estimated field U^S(x1,y1) and the sampling of U^S(x1,y1).

 figure: Fig. 3.

Fig. 3. This figure displays the simulated Fourier plane for two different values of the number of detector pixels, NP, across the exit-pupil diameter, D2. From Eq. (21), the autocorrelation term and the Dirac-delta term appear at the center of each plot; however, the strengths of these terms are negligible because we used a strong reference beam and subtracted off the average photoelectron count from each digital hologram. With that said, the turbulence-limited PSF and its complex conjugate are readily visible in each plot, and we isolate the correct turbulence-limited PSF with a window function. Note that the white circles represent the window function, and the green-dotted circles demarcate the turbulence-limited bucket diameter for a spherical wave DTsw for Scenario 5 in Table 1. Also note that the Fourier-plane side length, LF, and radius, ρW, of the window function vary from subplot to subplot. Here, we express the size of ρW and LF in terms of the approximate diffraction-limited bucket radius ρI.

Download Full Size | PDF

Provided the truth field US(x1,y1) and the estimated field U^S(x1,y1), the model used a metric dubbed the field-estimated Strehl ratio SF to quantify performance. In practice,

SF=|E{US(x1,y1)U^S*(x1,y1)}|2E{|US(x1,y1)|2}E{|U^S(x1,y1)|2},
where E{} is the expected-value operator. This metric is analogous to a Strehl ratio, which, in general, provides a normalized measure for performance [9,10,19]. In Eq. (36), if U^S(x1,y1)=US(x1,y1), then SF=1. Else if U^S(x1,y1)US(x1,y1), then SF<1. Thus, Eq. (36) is in line with the general understanding of a Strehl ratio and provides a normalized measure for the estimation accuracy of digital-holographic detection in the off-axis PPRG.

C. Signal-to-Noise Models

As previously stated, the numerical wave-optics model included both shot noise and read noise within the computational simulations. Here, we assumed that the shot noise resulted from the random arrival times of the photons that were incident on the FPA and that the read noise resulted from the read-out integrated circuitry of the FPA. The reader should note once more that the modeled shot noise followed a Poisson distribution, whereas read noise followed a Gaussian distribution.

For a Poisson-distributed random process, the mean is equal to the variance [25]. In this paper, the mean is equal to the sum of the per-pixel mean number of photoelectrons from the signal and reference beams, m¯S and m¯R, respectively. Given that m¯Rm¯S, we can assume that the mean number of photoelectrons varies little from pixel to pixel because m¯R is at 75% of the pixel-well depth. In turn, the Poisson-distributed shot noise follows a Gaussian distribution (to a good approximation) with variance m¯S+m¯R. Armed with these assumptions, we can add the variances for each Gaussian-distributed random process and arrive at the total noise variance σn2, such that

σn2=m¯S+m¯R+σr2,
where σr2=10,000pe2 is the variance of the read noise.

In terms of obtaining a closed-form expression for our analytical SNR, we can determine the SNR, in general, as the expected value of the signal power in the Fourier plane divided by the variance of the noise in the Fourier plane [30]. This ratio gives the following relationship:

SNR=E{|U˜S|2}Var{U˜N},
where U˜S denotes the complex-optical field associated with the windowed turbulence-limited ASF in the Fourier plane, U˜N denotes the complex-optical field associated with the noise in the Fourier plane, and Var{} denotes the variance operator. From Eq. (21), we can determine the total signal power associated with the turbulence-limited PSF in the Fourier plane as m¯Sm¯R. If the radius, ρW, of our window function meets or exceeds the turbulence-limited bucket radius, DT/2, then we can assume that we will capture most of the signal power associated with the turbulence-limited PSF. By Parseval’s theorem [22], we can also assume that the total noise variance in the windowed Fourier plane is equal to γ times the original noise variance in a pupil plane, where γ is the ratio of the window area to the Fourier-plane area. Therefore, our analytical SNR becomes
SNRA=m¯Sm¯Rγσn2=m¯Sm¯Rγ(m¯S+m¯R+σr2),
where
γ=πρW2ρI2NP2
is, again, the ratio of the window area to the Fourier-plane area.

In terms of obtaining our numerical SNR, we first determined the power associated with the signal and noise in the windowed area of the simulated Fourier plane. Next, we determined the power associated with just the noise in the windowed area of the simulated Fourier plane. The reader should note that we determined this noise-only power by taking a 2D FFT of the reference beam that was incident of the FPA without the signal beam present. This step was only valid if m¯Rm¯S. Put another way,

Var{U˜N}m¯R+σr2.
The reader should also note that, in practice, we could experimentally determine this noise-only power by taking the 2D FFT of the reference beam that is incident on and recorded by the FPA. To obtain the signal-only power, we then subtracted the noise-only power from the power associated with the signal and noise in the windowed area of the Fourier plane.

With all of this in mind, we can determine the numerical SNR as the ratio of the expected value of the signal-only power to the variance of the noise in the Fourier plane. In mathematical form, our numerical SNR becomes

SNRN=E{|U˜^S+N|2|U˜^N|2}Var{U˜^N},
where U˜^S+N denotes the estimate of the complex-optical field associated with the windowed turbulence-limited ASF and noise in the Fourier plane, and U˜^N denotes the estimate of the complex-optical field associated with the windowed noise in the Fourier plane. Before moving on to the next section, it is important to note that we used our numerical SNR to validate the use of the closed-form expression for our analytical SNR. For this purpose, Fig. 4 presents percentage-error results as a function of the analytical SNR. In Fig. 4, we averaged the results obtained from 20 independent realizations of Scenarios 1 and 5 in Table 1 and 20 independent realizations of detection noise. In general, the closed-form expression for SNR [cf. Eq. (39)] is valid and accurate and we will use it in the trade-space analysis that follows.

 figure: Fig. 4.

Fig. 4. This plot displays the percentage errors between the analytically and numerically calculated SNRs for 10 different SNRs as given by Eq. (39) for turbulence Scenarios 1 and 5 in Table 1. We calculated the SNRs by using the maximum allowed radius, ρW, of the window function in the Fourier plane (i.e., ρW was set to one fourth of the Fourier-plane side length LF). To vary the SNRs, we adjusted the per-pixel mean number of photoelectrons from the signal beam. Here, the error bars depict the standard-deviation offsets for 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations.

Download Full Size | PDF

3. RESULTS AND DISCUSSION

In this section, we present the results of a trade-space analysis that

  • 1. varies the number of FPA pixels NP across the exit-pupil diameter D2;
  • 2. varies the radius ρW of the window function relative to the approximate diffraction-limited bucket radius ρI in the Fourier plane; and
  • 3. varies the validated closed-form expression for SNR [cf. Eq. (39)] given maximum sampling of the turbulence-limited PSF (i.e., where ρW=ρINP/4).
With that said, this section assesses the performance of digital-holographic detection in the off-axis PPRG via the field-estimated Strehl ratio SF [cf. Eq. (36)] for the turbulence scenarios outlined in Table 1. In the trade-space trends that follow, we average the results for 20 independent realizations of turbulence and 20 independent realizations of detection noise.

In Fig. 5(a), we see that the performance remains stagnant for values of NP that exceed 32pixels for each turbulence scenario. While this result might seem strange, recall that NP only affects the Fourier-plane side length LF and is completely decoupled from the size of the window function that encompasses the turbulence-limited PSF. Figure 5(b) shows that the performance increases as the SNR increases for each value of NP, but that the performance curves approach a higher asymptotic value when NP=96 and NP=192 than when NP=16. This outcome is most likely due to aliasing that occurs when QP<4.88, as discussed in Section 2.A.

 figure: Fig. 5.

Fig. 5. In this plot, we see the trade-space trends that arise from varying the number of FPA pixels, NP, across the exit-pupil diameter, D2, in (a) and the SNR in (b) for the turbulence scenarios in Table 1. All of the solid colored lines are averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations. The dashed lines are standard-deviation offsets from these averages. In both plots, the radius, ρW, of the window function in the Fourier plane is always four times the approximate diffraction-limited bucket radius, ρI.

Download Full Size | PDF

In Fig. 6(a), the performance reaches a different distinct maximum for each turbulence scenario, and there appears to be an optimal value of ρW. As the strength of turbulence increases, the high-spatial-frequency content contained in the turbulence-limited PSF increases and necessitates the use of a wider window function in the Fourier plane to better encompass this high-spatial-frequency content. Thus, the optimal value of ρW increases for increasing turbulence strengths.

 figure: Fig. 6.

Fig. 6. In this plot, we see the trade-space trends that arise from varying the radius ρW of the window function in the Fourier plane in (a) and the SNR in (b) for the turbulence scenarios in Table 1. All of the solid colored lines are averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations. The dashed lines are standard-deviation offsets from these averages. It is important to note that the peaks in the curves in (a) are pushed out to higher values of ρW/ρI as the strength of turbulence increases (ρW/ρI=31.5 for Scenario 1 and ρW/ρI=38.0 for Scenario 5). Again, ρI is the approximate diffraction-limited bucket radius. In both plots, the number of FPA pixels, NP, across the diameter of the exit pupil, D2, is 256.

Download Full Size | PDF

Figure 6(b) shows that smaller values of ρW provide better results in low SNR scenarios and that larger values of ρW provide better results as the SNR increases. This is said because larger values of ρW (which correspond to larger window functions) cause more of the Fourier plane, and thus the noise in the Fourier plane, to be encompassed. Therefore, smaller window functions are more robust to noise in the Fourier plane because they encompass a tighter region of the turbulence-limited PSF and less of the regions where the signal floor is comparable with the noise floor in the Fourier plane.

Figures 7(a)7(d) affirm the behaviors found in Fig. 6 by looking at the wrapped-phase estimate for varying values of ρW. The lowest value of ρW gives a wrapped-phase estimate that is smooth when compared with the wrapped-phase truth because of the 2D convolution with the sombrero function in Eq. (30). As the value of ρW increases, the wrapped-phase estimates look more like the wrapped-phase truth, as seen in Fig. 7(c), but eventually become degraded by high-frequency noise, as seen in Fig. 7(d).

 figure: Fig. 7.

Fig. 7. Here, we show the wrapped-phase truth in (a) and the wrapped-phase estimates in (b)–(d) determined via a single independent realization of digital-holographic detection in the off-axis PPRG. The number of FPA pixels, NP, across the exit-pupil diameter, D2, is 256 for these results, and the turbulence strength corresponds to Scenario 5 in Table 1. Note that the black-dashed circle in (a) has a diameter equal to the Fried coherence diameter for a spherical wave r0sw.

Download Full Size | PDF

The final point of discussion lies in the analysis of the performance peaks when varying the value of ρW in Fig. 6(a). With that said, we can explain this behavior when the noise floor in the Fourier plane is compared with the turbulence-limited PSF without noise. Figure 8 displays the azimuthally averaged and normalized turbulence-limited PSFs, with and without noise, as well as the simulated and expected noise-only power, as per Eq. (37). Upon examination of the turbulence-limited PSFs without noise (blue curves in Fig. 8), the reader can see that the crossing points with the noise-only power (red curves in Fig. 8) occur at nearly the same radial values as do the peaks in SF in Fig. 6(a). The end result is that the optimal window radius occurs right as the turbulence-limited PSF dips below the noise floor in the Fourier plane. As seen in Fig. 7, this minimizes the smoothing caused by convolution with the sombrero function [cf. Eq. (30)] while omitting high-frequency noise. Figure 8 verifies the claim that increasing the turbulence strength increases the high-spatial-frequency content contained within the turbulence-limited PSF and necessitates a wider window function in the Fourier plane to achieve optimal performance.

 figure: Fig. 8.

Fig. 8. This figure displays the azimuthally averaged and normalized [with respect to the maximum of (a)] turbulence-limited PSFs with and without noise as well as the simulated-noise power and expected-noise power σn2 in the Fourier plane for Scenarios 1 and 5 [cf. Table 1] in (a) and (b), respectively. Note that the values of ρ/ρI at the crossings of the blue and red curves for each turbulence scenario are roughly equal to the values of ρW/ρI at the peaks in performance for Fig. 6(a) (ρW/ρI=31.5 for Scenario 1 and ρW/ρI=38.0 for Scenario 5). Also note that the y axes for both plots are normalized and displayed on a log scale to help visualize the high-spatial-frequency content of the turbulence-limited PSFs. Both plots show the averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations.

Download Full Size | PDF

With all of this in mind, we extended the analysis of the optimal window radius to account for all of the turbulence scenarios in Table 1. Figure 9 ultimately shows that we have developed a method for optimal windowing of the turbulence-limited PSF. Here, the differences between the window radii associated with maximum performance (like those seen in Fig. 6) and the crossover radii where the turbulence-limited PSF dips below the noise floor in the Fourier plane (as seen in Fig. 8) are small. Across all scenarios, the differences waver around a mean value of 0.0498 with a standard deviation of 0.6449 in units of length normalized to the approximate diffraction-limited bucket radius ρI. These results indicate good agreement between the window radii (associated with maximum performance) and the crossover radii (where the turbulence-limited PSF dips below the noise floor in the Fourier plane).

 figure: Fig. 9.

Fig. 9. This figure displays the differences between the window radii associated with maximum performance and the crossover radii where the turbulence-limited PSF dips below the noise floor in the Fourier plane for all of the turbulence scenarios in Table 1. As a function of SNR, all of the points waver around a mean value of 0.0498 with a standard deviation of 0.6449 in units of length normalized to the approximate diffraction-limited bucket radius ρI (which is a small quantity compared with the radius ρW of the window functions in the Fourier plane). The crossing points of the turbulence-limited PSFs with the noise floor were determined by taking the average of the first 10 points that were within 0.0005pe2 of the expected-noise power σn2. The data show the averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise.

Download Full Size | PDF

4. CONCLUSION

From the results presented in this paper, we find that, when using the appropriate parameters, digital-holographic detection in the off-axis PPRG is a robust and effective method for sensing the complex-optical field when in the presence of deep turbulence and detection noise. The results show that varying the number of FPA pixels across the exit-pupil diameter directly affects the side length of the Fourier plane, which contains the turbulence-limited PSF. Increasing the number of FPA pixels across the exit-pupil diameter past the point where aliasing becomes mitigated also offers no appreciable change in performance. It is important to note that the radius of the window function in the Fourier plane determines what spatial-frequency content makes it into the estimated field and the sampling of this estimated field. With this in mind, the results show that the window radii of maximum performance correspond to where the turbulence-limited PSF crosses the noise floor in the Fourier plane. This outcome provides us with a method for optimal windowing of the turbulence-limited PSF. In addition, our numerically calculated SNRs match well with our analytically calculated SNRs. This outcome means that future research efforts can calculate experimental SNRs with high accuracy via a closed-form SNR expression (provided that the strength of the reference beam is much, much greater than the strength of the signal beam and that these future research efforts have accurate equipment specifications).

Similar to the companion paper by Spencer et al. [10], this paper shows that digital-holographic detection offers a distinct way forward (with respect to wavefront sensing) when faced with the low SNRs brought about by deep-turbulence scintillation and extinction. We make this claim because digital-holographic detection uses a reference beam to boost the signal beam well above the read-noise floor that is created by the FPA. In practice, this strong reference beam allows us to reach a shot-noise-limited detection regime, and, in contrast with the companion paper by Spencer et al. [10], this paper shows that we are able to set the per-pixel reference beam strength to be much higher in the off-axis PPRG than in the off-axis IPRG (75% of the pixel-well depth as opposed to 25% in [10]) given a point-source beacon. Overall, the off-axis PPRG is less prone to pixel saturation because of focal-plane hot spots caused by deep-turbulence scintillation; thus, we can better approach the shot-noise limit when in the face of low SNRs. Given that digital-holographic detection in the off-axis PPRG also provides us with access to the complex-optical field directly, we can better explore the associated branch-point problem.

REFERENCES

1. M. A. Vorontsov, S. L. Lachinova, and A. K. Majumdar, “Target-in-the-loop remote sensing of laser beam and atmospheric turbulence characteristics,” Appl. Opt. 55, 5172–5179 (2016). [CrossRef]  

2. G. DiComo, M. Helle, J. Peñano, A. Ting, A. Schmitt-Sody, and J. Elle, “Implementation of a long range, distributed-volume, continuously variable turbulence generator,” Appl. Opt. 55, 5192–5197 (2016). [CrossRef]  

3. M. J. Steinbock, M. W. Hyde, and J. D. Schmidt, “LSPV+7, a branch-point-tolerant reconstructor for strong turbulence adaptive optics,” Appl. Opt. 53, 3821–3831 (2014). [CrossRef]  

4. A. T. Watnik and P. S. Lebow, “Limits of bootstrapping in a weak-signal holographic conjugator,” Appl. Opt. 53, 3841–3847 (2014). [CrossRef]  

5. J. C. Marron, R. L. Kendrick, N. Seldomridge, T. D. Grow, and T. A. Höft, “Atmospheric turbulence correction using digital-holographic detection: experimental results,” Opt. Express 17, 11638–11651 (2009). [CrossRef]  

6. J. R. Kraczek, P. F. McManamon, and E. A. Watson, “High resolution non-iterative aperture synthesis,” Opt. Express 24, 6229–6239 (2016). [CrossRef]  

7. J. A. Louthain and J. D. Schmidt, “Synergy of adaptive thresholds and multiple transmitters in free-space optical communication,” Opt. Express 18, 8948–8962 (2010). [CrossRef]  

8. J. A. Tellez and J. D. Schmidt, “Multiple transmitter performance with appropriate amplitude modulation for free-space optical communication,” Appl. Opt. 50, 4737–4745 (2011). [CrossRef]  

9. M. F. Spencer, I. V. Dragulin, D. S. Cargill, and M. J. Steinbock, “Digital holography wave-front sensing in the presence of strong atmospheric turbulence and thermal blooming,” Proc. SPIE 9617, 961705 (2015). [CrossRef]  

10. M. F. Spencer, R. A. Raynor, M. T. Banet, and D. K. Marker, “Deep-turbulence wavefront sensing using digital-holographic detection in the off-axis image plane recording geometry,” Opt. Eng. 56, 031213 (2016). [CrossRef]  

11. J. D. Barchers and T. A. Rhoadarmer, “Evaluation of phase-shifting approaches for a point-diffraction interferometer with the mutual coherence function,” Appl. Opt. 41, 7499–7509 (2002). [CrossRef]  

12. T. C. Poon and J. P. Liu, Introduction to Modern Digital Holography (Cambridge University, 2014).

13. D. L. Fried, “Branch point problem in adaptive optics,” J. Opt. Soc. Am. A 15, 2759–2768 (1998). [CrossRef]  

14. J. D. Barchers, D. L. Fried, and D. J. Link, “Evaluation of the performance of Hartmann sensors in strong scintillation,” Appl. Opt. 41, 1012–1021 (2002). [CrossRef]  

15. M. F. Spencer and T. J. Brennan, “Branch-cut accumulation using LSPV+7,” in Imaging and Applied Optics (Optical Society of America, 2017), paper PTh2D.2.

16. M. F. Spencer, R. A. Raynor, T. A. Rhoadarmer, and D. K. Marker, “Deep-turbulence simulation in a scaled-laboratory environment using five phase-only spatial light modulators,” in 18th Coherent Laser Radar Conference (2015).

17. P. E. Nielson, Effects of Directed Energy Weapons (Directed Energy Professional Society, 2009).

18. G. P. Perram, S. J. Cusumano, R. L. Hengehold, and S. T. Fiorino, An Introduction to Laser Weapon Systems (Directed Energy Professional Society, 2010).

19. M. T. Banet, M. F. Spencer, R. A. Raynor, and D. K. Marker, “Digital holography wavefront sensing in the pupil-plane recording geometry for distributed-volume atmospheric aberrations,” Proc. SPIE 9982, 998208 (2016). [CrossRef]  

20. T. J. Brennan, P. H. Roberts, and D. C. Mann, WaveProp: A Wave Optics Simulation System for use with MATLAB, User’s Guide Version 1.3 (Optical Sciences, 2010).

21. T. J. Brennan and P. H. Roberts, AOTools: The Adaptive Optics Toolbox for use with MATLAB, User’s Guide Version 1.4 (Optical Sciences, 2010).

22. J. D. Gaskill, Linear Systems, Fourier Transforms, and Optics (Wiley, 1978).

23. J. S. Tyo and A. S. Alenin, Field Guide to Linear Systems in Optics (SPIE, 2015).

24. B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics, 2nd ed. (Wiley, 2007).

25. E. L. Dereniak and G. D. Boreman, Infrared Detectors and Systems (Wiley, 1996).

26. J. D. Schmidt, Numerical Simulation of Optical Wave propagation (SPIE, 2010).

27. R. K. Tyson and B. W. Frazier, Field Guide to Adaptive Optics (SPIE, 2004).

28. L. C. Andrews, Field Guide to Atmospheric Optics (SPIE, 2004).

29. L. C. Andrews and R. L. Phillips, Laser Beam Propagation through Random Media, 2nd ed. (SPIE, 2005).

30. G. R. Osche, Optical Detection Theory for Laser Applications (Wiley, 2002).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1.
Fig. 1. Above is a diagram of digital-holographic detection in the off-axis PPRG. In (a), the coherent light from a master-oscillator (MO) laser actively illuminates a distant cooperative object (e.g., an unresolved ball bearing) creating a point-source beacon that propagates through deep turbulence to the entrance-pupil plane of an imaging system. Through collimation, the entrance-pupil plane creates a signal beam, which we interfere with a reference beam in the exit-pupil plane of the imaging system. Note that the reference beam manifests itself from an off-axis local oscillator (LO), which we derive from the MO laser. Also note that the exit-pupil plane is where we place the FPA, so that we can record a digital hologram. In (b), we can then use signal-processing techniques to manipulate the digital hologram into a wrapped-phase estimate.
Fig. 2.
Fig. 2. In (a) and (b), we show the normalized irradiance of the signal beam, whereas in (c) and (d), we show the resultant digital holograms for two different values of the number of detector pixels, N P , across the exit-pupil diameter, D 2 (white circles). The black-dashed circles have diameters equal to the demagnified spherical-wave Fried coherence diameter, that is, M T r 0 s w for turbulence Scenario 5 in Table 1. It is important to note that we display the associated turbulence-limited sampling quotient for a spherical wave Q P s w in the bottom-left corner of each subplot.
Fig. 3.
Fig. 3. This figure displays the simulated Fourier plane for two different values of the number of detector pixels, N P , across the exit-pupil diameter, D 2 . From Eq. (21), the autocorrelation term and the Dirac-delta term appear at the center of each plot; however, the strengths of these terms are negligible because we used a strong reference beam and subtracted off the average photoelectron count from each digital hologram. With that said, the turbulence-limited PSF and its complex conjugate are readily visible in each plot, and we isolate the correct turbulence-limited PSF with a window function. Note that the white circles represent the window function, and the green-dotted circles demarcate the turbulence-limited bucket diameter for a spherical wave D T s w for Scenario 5 in Table 1. Also note that the Fourier-plane side length, L F , and radius, ρ W , of the window function vary from subplot to subplot. Here, we express the size of ρ W and L F in terms of the approximate diffraction-limited bucket radius ρ I .
Fig. 4.
Fig. 4. This plot displays the percentage errors between the analytically and numerically calculated SNRs for 10 different SNRs as given by Eq. (39) for turbulence Scenarios 1 and 5 in Table 1. We calculated the SNRs by using the maximum allowed radius, ρ W , of the window function in the Fourier plane (i.e., ρ W was set to one fourth of the Fourier-plane side length L F ). To vary the SNRs, we adjusted the per-pixel mean number of photoelectrons from the signal beam. Here, the error bars depict the standard-deviation offsets for 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations.
Fig. 5.
Fig. 5. In this plot, we see the trade-space trends that arise from varying the number of FPA pixels, N P , across the exit-pupil diameter, D 2 , in (a) and the SNR in (b) for the turbulence scenarios in Table 1. All of the solid colored lines are averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations. The dashed lines are standard-deviation offsets from these averages. In both plots, the radius, ρ W , of the window function in the Fourier plane is always four times the approximate diffraction-limited bucket radius, ρ I .
Fig. 6.
Fig. 6. In this plot, we see the trade-space trends that arise from varying the radius ρ W of the window function in the Fourier plane in (a) and the SNR in (b) for the turbulence scenarios in Table 1. All of the solid colored lines are averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations. The dashed lines are standard-deviation offsets from these averages. It is important to note that the peaks in the curves in (a) are pushed out to higher values of ρ W / ρ I as the strength of turbulence increases ( ρ W / ρ I = 31.5 for Scenario 1 and ρ W / ρ I = 38.0 for Scenario 5). Again, ρ I is the approximate diffraction-limited bucket radius. In both plots, the number of FPA pixels, N P , across the diameter of the exit pupil, D 2 , is 256.
Fig. 7.
Fig. 7. Here, we show the wrapped-phase truth in (a) and the wrapped-phase estimates in (b)–(d) determined via a single independent realization of digital-holographic detection in the off-axis PPRG. The number of FPA pixels, N P , across the exit-pupil diameter, D 2 , is 256 for these results, and the turbulence strength corresponds to Scenario 5 in Table 1. Note that the black-dashed circle in (a) has a diameter equal to the Fried coherence diameter for a spherical wave r 0 s w .
Fig. 8.
Fig. 8. This figure displays the azimuthally averaged and normalized [with respect to the maximum of (a)] turbulence-limited PSFs with and without noise as well as the simulated-noise power and expected-noise power σ n 2 in the Fourier plane for Scenarios 1 and 5 [cf. Table 1] in (a) and (b), respectively. Note that the values of ρ / ρ I at the crossings of the blue and red curves for each turbulence scenario are roughly equal to the values of ρ W / ρ I at the peaks in performance for Fig. 6(a) ( ρ W / ρ I = 31.5 for Scenario 1 and ρ W / ρ I = 38.0 for Scenario 5). Also note that the y axes for both plots are normalized and displayed on a log scale to help visualize the high-spatial-frequency content of the turbulence-limited PSFs. Both plots show the averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise, that is, 400 realizations.
Fig. 9.
Fig. 9. This figure displays the differences between the window radii associated with maximum performance and the crossover radii where the turbulence-limited PSF dips below the noise floor in the Fourier plane for all of the turbulence scenarios in Table 1. As a function of SNR, all of the points waver around a mean value of 0.0498 with a standard deviation of 0.6449 in units of length normalized to the approximate diffraction-limited bucket radius ρ I (which is a small quantity compared with the radius ρ W of the window functions in the Fourier plane). The crossing points of the turbulence-limited PSFs with the noise floor were determined by taking the average of the first 10 points that were within 0.0005 pe 2 of the expected-noise power σ n 2 . The data show the averages of 20 independent realizations of turbulence and 20 independent realizations of detection noise.

Tables (1)

Tables Icon

Table 1. Turbulence Scenarios Used in the Trade-Space Analysis

Equations (42)

Equations on this page are rendered with MathJax. Learn more.

U S ( x 2 , y 2 ) = U S ( x 2 , y 2 ) cyl ( x 2 2 + y 2 2 D 2 ) ,
cyl ( x 2 + y 2 ) = { 1 0 x 2 + y 2 < 0.5 0.5 x 2 + y 2 = 0.5 0 x 2 + y 2 > 0.5
U R ( x 2 , y 2 ) = A R exp [ j 2 π x R x 2 M T λ z ] exp [ j 2 π y R y 2 M T λ z ] ,
M T = D 2 D 1 ;
I H ( x 2 , y 2 ) = | U S ( x 2 , y 2 ) + U R ( x 2 , y 2 ) | 2 = | U S ( x 2 , y 2 ) | 2 + | U R ( x 2 , y 2 ) | 2 + U S ( x 2 , y 2 ) U R * ( x 2 , y 2 ) + U R ( x 2 , y 2 ) U S * ( x 2 , y 2 ) ,
I ^ H ( n x s , m y s ) = [ I H ( x 2 , y 2 ) × 1 w x rect ( x 2 n x s w x ) 1 w y rect ( y 2 m y s w y ) ] d x 2 d y 2 ,
rect ( x ) = { 1 0 | x | < 0.5 0.5 | x | = 0.5 0 | x | < 0.5
m ¯ H ( n x s , m y s ) = η T h ν I ^ H ( n x s , m y s ) w x w y ,
d H ( x 2 , y 2 ) = m ¯ H ( x 2 , y 2 ) × 1 x s comb ( x 2 x s ) 1 y s comb ( y 2 y s ) × rect ( x 2 N x s ) rect ( y 2 M y s ) ,
m ¯ H ( x 2 , y 2 ) = η T h ν [ I H ( x 2 , y 2 ) × rect ( x 2 x 2 w x ) rect ( y 2 y 2 w y ) ] d x 2 y 2 = η T h ν I H ( x 2 , y 2 ) ** rect ( x 2 w x ) rect ( y 2 w y )
1 | w | comb ( x w ) = n = δ ( x n w )
δ ( x ) = 1 | w | lim w 0 p ( x w )
V ( x , y ) ** W ( x , y ) = V ( x , y ) W ( x x , y y ) d x d y ,
V ˜ ( ν x , ν y ) = V ( x , y ) e j 2 π ( x ν x + y ν y ) d x d y .
d ˜ H ( ν x , ν y ) = η T h ν α ˜ ( ν x , ν y ) I ˜ H ( ν x , ν y ) ** w x comb ( x s ν x ) w y comb ( y s ν y ) ** N x s sinc ( N x s ν x ) M y s sinc ( M y s ν y ) ,
α ˜ ( ν x , ν y ) = sinc ( w x ν x ) sinc ( w y ν y )
I ˜ H ( ν x , ν y ) = Γ ˜ ( ν x , ν y ) + | A R | 2 δ ( ν x ) δ ( ν y ) + A R * U ˜ S ( ν x + x R M T λ z , ν y + y R M T λ z ) + A R U ˜ S * ( ν x x R M T λ z , ν y y R M T λ z ) ,
Γ ˜ ( ν x , ν y ) = U ˜ S ( ν x , ν y ) ** U ˜ S * ( ν x , ν y ) .
d ˜ H ( x 0 M T λ z , y 0 M T λ z ) = η T h ν α ˜ ( x 0 M T λ z , y 0 M T λ z ) × { 1 ( M T λ z ) 2 Γ ˜ ( x 0 M T λ z , y 0 M T λ z ) + | A R | 2 δ ( x 0 M T λ z ) δ ( y 0 M T λ z ) + A R * U ˜ S ( x 0 + x R M T λ z , y 0 + y R M T λ z ) + A R U ˜ S * ( x 0 x R M T λ z , y 0 y R M T λ z ) } ** p 2 ( M T λ z ) 2 comb ( p x 0 M T λ z ) comb ( p y 0 M T λ z ) ** N P 2 p 2 ( M T λ z ) 2 sinc ( N P p x 0 M T λ z ) sinc ( N P p y 0 M T λ z ) .
L F = M T λ z p = D 2 D 1 λ z p = λ z D 1 N P = ρ I N P .
d ˜ H ( x 0 M T λ z , y 0 M T λ z ) = η T h ν α ˜ ( x 0 M T λ z , y 0 M T λ z ) × { 1 ( M T λ z ) 2 Γ ˜ ( x 0 M T λ z , y 0 M T λ z ) + | A R | 2 δ ( x 0 M T λ z ) δ ( y 0 M T λ z ) + A R * U ˜ S ( x 0 + x R M T λ z , y 0 + y R M T λ z ) + A R U ˜ S * ( x 0 x R M T λ z , y 0 y R M T λ z ) } ** 1 ρ I 2 N P 2 comb ( x 0 ρ I N P ) comb ( y 0 ρ I N P ) ** 1 ρ I 2 sinc ( x 0 ρ I ) sinc ( y 0 ρ I ) .
w ( x 0 , y 0 ) = cyl ( x 0 2 + y 0 2 2 ρ W ) ,
L F = M T r 0 p λ z r 0 = Q P λ z r 0 ,
D T = 2.44 λ z r 0 ,
Q P 2 λ z r 0 2.44 λ z r 0 ,
Q P 4.88 .
( x R M T λ z , y R M T λ z ) .
ψ ˜ ^ S ( x 0 M T λ z , y 0 M T λ z ) d ˜ H ( x 0 x R M T λ z , y 0 y R M T λ z ) w ( x 0 , y 0 ) η T h ν α ˜ ( x 0 x R M T λ z , y 0 y R M T λ z ) A R * A S = ψ ˜ S ( x 0 M T λ z , y 0 M T λ z ) w ( x 0 , y 0 ) ,
V ( x , y ) = V ˜ ( ν x , ν y ) e j 2 π ( x ν x + y ν y ) d ν x d ν y .
ψ ^ S ( M T x 1 , M T y 1 ) ψ S ( M T x 1 , M T y 1 ) ** π ρ W 2 ( λ z ) 2 somb ( 2 ρ W x 1 2 + y 1 2 λ z ) ,
somb ( x 2 + y 2 ) = 2 J 1 ( π x 2 + y 2 ) π x 2 + y 2
r 0 p w = 0.185 ( λ 2 Z C n 2 ) 3 / 5
r 0 s w = 0.33 ( λ 2 Z C n 2 ) 3 / 5 ,
σ χ p w 2 = 0.307 k 7 / 6 Z 11 / 6 C n 2
σ χ s w 2 = 0.124 k 7 / 6 Z 11 / 6 C n 2 ,
S F = | E { U S ( x 1 , y 1 ) U ^ S * ( x 1 , y 1 ) } | 2 E { | U S ( x 1 , y 1 ) | 2 } E { | U ^ S ( x 1 , y 1 ) | 2 } ,
σ n 2 = m ¯ S + m ¯ R + σ r 2 ,
SNR = E { | U ˜ S | 2 } Var { U ˜ N } ,
SNR A = m ¯ S m ¯ R γ σ n 2 = m ¯ S m ¯ R γ ( m ¯ S + m ¯ R + σ r 2 ) ,
γ = π ρ W 2 ρ I 2 N P 2
Var { U ˜ N } m ¯ R + σ r 2 .
SNR N = E { | U ˜ ^ S + N | 2 | U ˜ ^ N | 2 } Var { U ˜ ^ N } ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.