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

Multi-filter transport of intensity equation solver with equalized noise sensitivity

Open Access Open Access

Abstract

Phase retrieval based on the Transport of Intensity Equation (TIE) has shown to be a powerful tool to obtain the phase of complex fields. Recently, it has been proven that the performance of TIE techniques can be improved when using unequally spaced measurement planes. In this paper, an algorithm is presented that recovers accurately the phase of a complex objects from a set of intensity measurements obtained at unequal plane separations. This technique employs multiple band-pass filters in the frequency domain of the axial derivative and uses these specific frequency bands for the calculation of the final phase. This provides highest accuracy for TIE based phase recovery giving minimal phase error for a given set of measurement planes. Moreover, because each of these band-pass filters has a distinct sensitivity to noise, a new plane selection strategy is derived that equalizes the error contribution of all frequency bands. It is shown that this new separation strategy allows controlling the final error of the retrieved phase without using a priori information of the object. This is an advantage compared to previous optimum phase retrieval techniques. In order to show the stability and robustness of this new technique, we present the numerical simulations.

© 2015 Optical Society of America

1. Introduction

Quantitative phase imaging (QPI) has been used extensively for visualizing hidden features of biological and technical samples [1–4]. Moreover, QPI allows measurements on the sample without disturbing or destroying it. The advent of Interferometry and Holography provided the necessary tools to carry out QPI with high accuracy. These well-established techniques provide the integrated phase (averaged refractive index) of the light as it passes through the specimen. One implementation of these techniques is Digital Holographic Microscopy (DHM) [5–8] that is successfully applied in the areas of life sciences and technology giving remarkable results [7,8]. DHM is fast, accurate and with additional effort allows tracking in real time processes in life sciences and engineering. However, DHM and related Interferometry techniques have a number of weaknesses: they require a reference beam, high mechanical and environmental stability, and coherent illumination. Furthermore, the obtained phase needs to be unwrapped requiring an extra computational effort.

Single Beam Phase Retrieval Techniques (SBRTs) that seek to obtain the phase information from a sequence of through focus intensities [2,4,9] are an alternative solution to the phase problem. SBRTs based on iterative algorithms [1,9–11] recover the phase of the object by employing a backward-forward wave-propagation scheme of the captured intensities in a loop until certain numbers of iterations are accomplished or certain stagnation value is reached. However, iterative SBRTs may suffer from stagnation in local minima, giving a slow convergence and a solution that may not be guaranteed due to the non-convex nature of the problem [12]. Non-iterative SBRTs are referred as Deterministic Phase Retrieval Techniques (DPRTs) [13–15]. DPRTs based on the Transport of Intensity Equation (TIE) [2,3] have gained increased interest because they relate the transversal energy flux of the phase with axial energy flux of the intensity in the Fresnel region by means of a linear transformation [3,16]. Notably, TIE techniques give a unique solution [17], the retrieved phase does not need to be unwrapped [18] and it can be employed with partially coherent illumination [3,19,20]. Therefore, TIE based techniques have been applied in diverse areas ranging from optical microscopy to X-ray imaging [2,16,21]. However, major disadvantages of TIE methods are that they are strongly affected by Low Frequency Artifacts (LFAs) which are the effect of low frequency noise amplification [22]. The impact of these LFAs will depend on the plane separation between images [18]. Furthermore, TIE based algorithms have a poor efficiency for recovering high frequency components while the accuracy in the recovering of low frequency components is limited by the maximal plane separation [23]. Regularization Techniques (RTs) [21,24] have been proposed in order to suppress LFAs in the retrieved phase by TIE techniques. The aim of these RTs is to exclude low frequency components from the retrieved phase. Some RTs (for example the Tikhonov regularization [24]) are designed as High Pass Filters (HPFs) that can exclude low frequency components from the retrieved phase. Nevertheless, HPFs cannot distinguish between the low frequency components that belong to the LFAs and those that belong to the phase of the object [20]. Thus, RTs based on HPFs can therefore destroy important low frequency components corrupting the final phase reconstruction. More sophisticated RTs can be designed for suppressing LFAs. However, these techniques need a priori information of the object in order to distinguish between the low frequency components that belong to LFAs and those that belong to the phase of the object. In the particular case of the optical regime, it has been shown that the performance of TIE methods can be improved by choosing the proper plane separation strategy for capturing the intensities [15,22,25,26], and thus, the employment of RTs can be avoid. When selecting large plane separations, the effect of LFAs in the retrieved phase can be mitigated without employing RTs [18,26] but high frequency components of the phase cannot be recovered [18,27]. This problem can be overcome when capturing a large number of intensities obtained at equally spaced planes having small plane separations. This will mitigate LFAs while recovering high frequencies [22,23]. However, an efficient suppression of LFAs through this spacing technique does require a large amount of measurement planes with small plane separations. A different approach to this problem is an exponential plane selection strategy, which has been presented for the case of CTF based solvers [15]. That technique mitigates LFAs and recovers a wide frequency spectrum of the phase by employing a few captured intensities [10,15]. This strategy has been successfully implemented in the Gaussian Process regression TIE approach (GP-TIE) [23] giving high accuracy that outperforms current state of the art techniques. The exponential spacing strategy has been developed originally on the specific properties of CTF solvers [15], which did not consider the properties of TIE solvers. This is an important aspect, because the GP-TIE solver [23] seeks to find optimal estimation of the axial derivative. However, in [22], Martinez et al showed that the condition for an accurate estimation of the phase is different from the one for the axial derivative.

The goal of this paper is to introduce a new TIE technique that is able to recover the phase with minimal error when employing unequally spaced plane strategies. The principle of this technique is derived by exploiting the properties of the Classical TIE solver [18] with a set of three images. Initially, we study the frequency response of three plane TIE solver [3,18] with respect to LFAs, nonlinear error components, plane separation, and the Inverse Laplacian (IL). From this analysis, it is found that only a limited range of spatial frequencies can be retrieved accurately for a given plane separation. Thus, we show that the frequency components retrieved by the three plane TIE solver can be enlarged when combining the phase from a series of band pass filtered three plane TIE reconstructions at different defocused distances. It is shown that the noise sensitivity and accuracy of each TIE reconstruction depends on the size of the band pass filter. This method gives a robust and accurate solution for the phase retrieval problem over a wide frequency range. Moreover, we develop a new separation strategy that equalizes the noise sensitivity of all band pass filtered TIE reconstructions. This allows controlling the final error of the retrieved phase without a priori information of the phase distribution as it is the case for other optimum TIE techniques [21–23,25].

2. Frequency response of a three plane TIE solver

Consider a paraxial wave field [3,18] where the intensity variations along z are calculated from two images placed at ± z and are related to the transversal energy flux of the phase as

·[I0φ]=kI(r,z)I(r,z)2z,
where r is the transverse coordinate vector, φ the object phase, k = 2π, I(r, ± z) are the defocus intensities, I0 = exp[-2μ(r)] is the in-focus intensity, μ(r)>0 is the absorption distribution of the object, λ the wavelength, and the Right Hand Side (RHS) of Eq. (1) is the finite difference approximation of the axial intensity derivative. For convenience, the case of a pure phase object is considered for the derivation of this approach. Nevertheless, the case of absorbing objects is analyzed in Appendix A. Hence, if μ(r)<<1 then I0 can be considered as constant and Eq. (1) is solved as [18]
φ=FT1I^0[FT(A(r,z))4πλf2z],
where FT and FT−1 are the Fourier Transform and its inverse respectively, f 2 = u2 + v2 is the transversal Laplacian in the frequency domain, Î0 is the average of the in-focus intensity and A(z) = I(z)-I(-z). However, the inversion of the Laplacian (f −2) in Eq. (2) introduces two problems: firstly, the element at f = 0 is undetermined, and second, the IL amplifies significantly the noise near the zero frequency. In order to avoid both problems, RTs have been employed in TIE based solvers [24,28]. A well-known RT is the Tikhonov regularization where the original IL 1 / f 2 is modified as f 2/(f 4 + β) where β is known as the regularization parameter. The modified IL acts as HPF that can mitigate LFAs and β controls the lower cutoff frequency. However, this RT cannot distinguish between LFAs and frequencies of the phase object. Thus, the Tikhonov regularization may destroy important information of the phase, and often does so in experimental practice. More sophisticated RTs reduce LFAs [14,21,24] more efficiently, but their application is restricted to specific cases, e.g. piecewise constant objects [24], and may strongly corrupt the final phase solution [29]. When using visible radiation for phase retrieval techniques, the employment of RTs can be avoided by selecting the proper separation strategy. Recently, it has been shown that the exponential spacing strategy [15] is the most known effective solution for the LFAs suppression without employing RTs [15,23]. Moreover, the applicability of Eq. (2) is limited to the condition that the function
PTFTIE=πλf2z,
which is known as the Phase Transfer function of the TIE (PTFTIE), is small [18,21]. This means that neither the defocus distance z nor the frequency range f can be arbitrarily large. In consequence, only a specific range of frequencies can be recovered by the TIE solver accurately. When the inverse PTFTIE (IPTFTIE) is applied on the frequency spectrum of the function A, we have that the IPTFTIE damps the higher frequency components as a Low Pass Filter (LPF) i.e. larger plane separations have a stronger suppression of high frequency components [18,22]. The mathematical form of this LPF is 1/zf 2 and its corner frequency can be controlled by the defocus distance z. In order to know the range of frequencies that are retrieved reliably within this LPF, one compares Eq. (3) with the Phase Transfer Function (PTF)
PTF=sin(πλf2z).
which is the CTF for phase pure objects [23]. The validity of Eq. (4) is given by the condition |φ(x)- φ(x-a)|<<1, where a is a constant. This condition is known as the slowly varying phase condition, which is more general that the weak phase approximation [16]. Note that Eq. (3) is the first order Taylor approximation of Eq. (4) [15,18,23]. The discrepancy between these two transfer functions can be described by the ratio
PTFPTFTIE=1ε,
where ε is a measure for the maximum acceptable error. Thus, it suitable to define a radial cutoff frequency (COF) ζ that guarantees reliable reconstructions for the spatial frequency components within the LPF. Here, in accordance to [18], the COF is defined as
ζ(ε,z)=6επλz.
Additionally, the derivations in [22] showed that in presence of uncorrelated Additive White Gaussian Noise (UAWGN) the overall error in the retrieved phase for z in Eq. (2) can be described by the Root Mean Square Error (RMSE) as
RMSE2=K1z2c0+z4K2W|FT1[f2[z3I˜|z=0]]|2d2r,
where K1 = (kσW/π)2Δx2/32π2, K2 = (k/Wπ2)2, W = NpΔx is the detector size, Δx is the pixel size, Np is the number of pixels, σ is the standard deviation of the noise, c0≈6.0268 [22] and z3Ĩ is the FT of the third axial intensity derivative evaluated at the in-focus position. The first term in Eq. (7) is the error due to the UAWGN and second term describes the non-linear error due to the propagation [22]. Equation (7) allows visualizing the accuracy of the TIE solver for specific frequency components. However, for the computation of the error in Eq. (7) a given object knowledge is needed, i.e. the third axial derivate. Equation (7) can be simplified if a field within the validity of the TIE solver is considered [16], using the CTF model [18],
I˜(f,z)=δ(f)+2Mcos(πλf2z)2Φsin(πλf2z),
where δ(f) is the Dirac delta, Φ = FT(φ), and M = FT(μ). Thus, the third axial derivative can be calculated as
3z3I˜(f,z)|z=0=2(πλf2)3Φ.
Considering a sine grating φj = sin(2πxfj ), where fj is the sine frequency, as a phase object and employing the result of Eq. (9), it can be shown that Eq. (7) reduces to
RMSE2=C1[σ2(πλ)2ζ46ε+C2C1(6ε(πλ)2)2(fjζ)8],
where C1 = (c0/32)(kWΔx/π)2 and C2/C1 = (2/c0W3)(πλ3/24Δx)2. The expression in Eq. (10) implies that the non-linear error due to the propagation is proportional to fi. Moreover, if the ratio fi/ζ<1 then (fi/ζ)8 <<1. Hence, Eq. (10) can be written as
RMSE2=C1[σ2(πλ)2ζ46ε1].
Equation (11) shows that for all the frequencies below of the COF ζ the error in the retrieved phase will be given solely by the UAWGN term, and despite this, the overall error of the retrieved phase can be calculated without requiring information about the phase of the object.

Figure 1(a) shows the IPTFTIE for three defocus distances; the positions of the COFs for ε = 0.01 are denoted by the vertical dash lines. Figure 1(b) shows the RMSE of the TIE solver at a given spatial frequency that is simulated using a single frequency sine grating φj = A0sin(2πx/Tj) as a pure phase object, where Tj is the grating period and A0 = 0.5. It should be noted that for these simulation, the regularization parameter is β = 0, the zero frequency sample is excluded, and the DC term is removed from the phase error in order to estimate the RMSE. The simulations in Fig. 1(b) are carried out using Δx = 4μm, λ = 0.633μm, Np = 256, SNR = 40dB (σ = 0.01), and various grating periods Tj. The two intensities have a distance z to the either side of the in-focus plane. The selected separations for these simulations were z1 = 10μm (green line), z2 = 100μm (red line) and z3 = 1000μm (blue line). In Fig. 1(b), the solid lines represent the analytical RMSE (denoted as RMSET, using Eq. (7)), while the lines labeled with RMSES correspond to the RMSE obtained by simulation. These results were obtained by making an average of 25 simulations. The analytical result for Eq. (11) is shown in Fig. 1(b) with the horizontal dash lines and the positions of the COFs that are denoted by the vertical dash lines. Notably, the analytical predictions of Eq. (7) (solid line) and Eq. (11) follow accurately the simulation results (RMSES) for spatial frequencies that do not exceed ζ. Hence, larger defocus distances give a smaller RMSE but allow recovering reliably less frequency components.

 figure: Fig. 1

Fig. 1 a) IPTFTIE for three different defocus distances and b) RMSE frequency response of the retrieved phases. The vertical lines are placed in its corresponding COF. The lines labeled with the legend RMSES correspond to the RMSE of the retrieved phase through simulations, and the theoretical RMSE (RMSET) are denoted with the solid lines. Moreover, the analytical results of Eq. (11) and the position of the COFs are plotted with the horizontal and the vertical dashed lines respectively. The RMSES presented in this figure are the average of 25 simulations.

Download Full Size | PDF

3. Improving the 3 plane TIE solution: The Multi-Filter TIE algorithm

From Fig. 1(b), it can be seen that the frequency components of the phase below the COF can be recovered with the smallest error possible. Following this idea, the performance of the three plane TIE solver can be improved, if for each defocus distance we are able to select and extract the range of frequencies for those where the phase retrieval is done accurately. For instance, for a large defocus distance ± z3 (see Fig. 1 comparing z1 and z2), the retrieved phase with Eq. (2) will have the low frequency well resolved until the COF ζ(z3); this information can be extracted when suppressing higher frequencies with a LPF. By adding two more planes with a smaller separation distance ± z2, the frequencies in the radial range ζ(z3)<f ζ(z2) could be well-resolved. The set of recovered frequencies in this region can be extracted using a Band Pass Filter (BPF) with COFs (ζ(z3), ζ(z2)), and can be added to the other recovered frequencies so that the range of frequencies 0 < f ζ(z3) can be recovered accurately. The same process can be applied to the shortest defocus distance z1, and thus, the set of frequencies in the region (ζ(z2), fmax), where fmax = 1/(2Δx), could be extracted by a High Pass Filter (HPF) and added to the other data. Thus, the initial set of recovered frequencies (0, ζ(z3)) has been enlarged to the range (0, fmax). This methodology can be generalized to the case of several distances ± z1,…, ± zN with z1< z2<…< zN-1< zN and several COFs given as

0<ζ(zN)<ζ(zN1)<<ζ(z2)fmax,
Hence, the frequency domain can be divided into a single circular region (0 < ζ(|zN|)), several donut-shaped ranges of spatial frequencies (ζ(zi + 1)< f< ζ(zi)), and a high frequency region (ζ(z2)< f< fmax). Each of these regions in the frequency domain can be extracted using a BPF (denoted as F(i)) with a radial bandwidth of ζi < f ≤ ζi-1, as
φi=FT1[F(i)(A(zi))4πλzif2I0].
A similar procedure can be applied for 0 < f ≤ ζN and ζ2 < f ≤ fmax using a LPF and a HPF for F(1),(N) respectively. From Eq. (13), we can see that for the separations z1 to zN-1 the employment of the BPFs is filtering out the region where the IL amplifies the low frequency noise, and thus, the regularization parameter β is not needed any more. The low-frequency components correspond to the distance zN, a large value of zN can mitigate the effect of LFAs [18,22]. Hence, our aim is to find the proper value for zN that effectively suppresses LFAs in the retrieved phase without using RTs. There are no restriction in the design of the filters F(i) provided the condition that the sum of all filters equals one is fulfilled. Given these considerations, the object phase can be reconstructed by superimposing all filtered phases φi as follows
φ(x,y)=i=1Nφi(x,y).
This procedure is referred as Multi-Filter TIE (MF-TIE). Figure 2 shows the flow chart of the MF-TIE algorithm presented in this paper. The idea to employ BPFs in the TIE solver has been already applied [20,21]. However, these approaches have been made for the specific case of equally spaced separation strategy, which has been proven to be inefficient in high noise conditions [23,26,29]. Moreover the approaches developed in references [20,21] are unstable when high order derivatives are calculated [23], i.e. for larger number of planes. The MF-TIE overcomes this problem given that the backbone of this solver is the 3 plane TIE solver.

 figure: Fig. 2

Fig. 2 Flow chart of the MF-TIE based solver.

Download Full Size | PDF

3.1 Error response of the MF-TIE

Given that the filtered phases are not correlated, we can estimate the corresponding RMSE for each filtered phase. This RMSE will be denoted as RMSEi and can be calculated using similar considerations leading to Eq. (7), but in contrary to the case of Eq. (7) it is necessary to include the effect of the filter, as

RMSEi2=(kσπ)2132π2zi21W2W|FT1[F(i)/f2]|2d2r+zi4K2W|FT1[[F(i)×z3I˜/f2]]|2d2r.
Moreover, if the COF of the filter F(i) fulfills the conditions discussed in Section 2 (see Fig. 1(b)) then the second term in Eq. (15) can be neglected, and thus, the error can be estimated as
RMSEi2=(kσπ)2132π2zi2αi,
where αi = 1/W2∫∫|FT−1[F(i)(f)/ f 2] |2 d2r. The latter expression allows describing the error of the TIE solver solely by detector characteristics and noise. Thus, for each retrieved phase φi (corresponding to the BPF F(i)) the error is independent of the object phase and solely governed by noise. In contrary to other optimization schemes, it does not require particular a priori knowledge of the phase distribution as e.g. the value of the third axial derivative [3,18,21,22].

4. Plane selection strategy for the MF-TIE

As it was discussed in Section 2, the selection of the plane separation strategy is crucial for SBPRTs because this allows mitigate the effects of LFAs due to noise and determines the range of spatial frequencies to be recovered [15,23,26]. Recently reported, the exponential plane selection strategy [15,23] allows such phase reconstructions with small error. The defocus distances in that strategy are calculated as

zj=g0j1Δz,
where g0 is the geometric progression factor [15,30], j = 1,2,…,N and Δz is the smallest defocus distance. That plane separation strategy has been successfully implement in the GP-TIE approach [23] and was originally developed to match the needs of CTF based solvers (avoid singularities [15]), but it has not specifically developed to account for the properties of the TIE solver. Hence, a TIE specific unequal plane separation strategy needs to be developed for this purpose. The first step for modeling this separation strategy consists to find the distance zN which suppress effectively the LFAs. This can be done by writing Eq. (16) as
RMSEN2=(kσπ)2132π2zN2αN,
It should be noted that αN is the power of the IL between the frequencies (0,ζN]. The parameter αN can be evaluated numerically if the filter shape is known. However, for the specific case of an ideal stepwise circular LPF, and using Eq. (6) for re-expressing the corresponding COF in terms of the separation distance zN, αN yields:
αN=W2Δx2π[2Δf2πλzN(6ε)1/2],
where Δf = 1/(NpΔx). Substituting this result into Eq. (18) and solving for zN, yields:
zN=21(RMSEN2K1π)[b26ε+b226ε+8(RMSEN2K1π)],
where K1 = (kσW/π)2Δx2/32π2 and b2 = Δf 2λπ. A well-established optimization criterion when dealing with multiple independent random variables is to equalize the variance of all variables. When considering this principle for the MF-TIE solver, one wishes to obtain RMSE1 = … = RMSEN-1 = RMSEN. Thus, the initial constraint to be imposed is given by Eq. (18). This consideration allows calculating the defocus distance zN-1 by solving RMSEN = RMSEN-1. Again, this solution depends on the design of the BPFs, but when considering F(i) to be ideal stepwise circular BPFs, F(N-1) can be simplified as an donut shape filter, which leads
αN-1=W2Δx2π[ζN2ζN12],
and therefore
RMSEN2=K1πzN12Δf2[ζN2ζN12].
Re-expressing Eq. (22) in terms of the distances zN-1 and zN, results in
zN12+Γ[zN1zN]=0,
where Γ = π2λ K1/((RMSEN)2 ()1/2). The valid solution for Eq. (23) is given as
zN1=21(Γ+Γ2+4zNΓ).
This train of thoughts can be extended to the remaining planes, as
zi1=21(Γ+Γ2+4ziΓ).
The procedure in Eq. (25) is terminated when ζi> fmax, where ζ(zi) = 6ε/πλzi. Equation (25) allows finding the proper defocus distances necessary for ensuring a constant RMSE contribution for each BPF employed. Hence, the total error in the retrieved phase can be estimate as
RMSET=NRMSEN,
where N = (M-1)/2 is the number of defocus distances pairs generated with Eq. (25), and M is the number of planes. Notably, the minimization of RMSEN leads an effective minimization of the RMSET. With the separation strategy defined by Eq. (25), we can find the minimal and maximum separation distances according to the experimental conditions. The selection of the parameter ε will define how many are necessary to suppress efficiently the LFAs.

By using Eq. (26) the total error of the reconstructed phase can be estimated without specific object information. It should further be noted, that in contrary to other separation strategies (e.g. the exponential spacing [15,23]), the noise sensitive spacing strategy allows to include the information of an acceptable level of RMSE in the plane selection methodology. This important finding has not been reported by previous works [15,22,23]. The result given in Eq. (26) allows us to retrieve the phase of the object with a specific error for a fixed number of planes provided that a maximum value for ε is not exceeded. The maximum value for ε is object dependent and is related to the validity of TIE, i.e. cases where non-linear error components are negligible. This property is visualized by simulations. Consider a pure phase object in form of a Shepp-Logan phantom with a phase difference of π/4, π/8 and π/16 employing 11 intensities (N = 5) and different values of RMSET. These simulation were carried out for different levels of noise given as SNR = 30,40 and 50dB.

The results of these simulations are shown in Fig. 3. The solid line in this plot is the noise boundary limit given by Eq. (26) (no nonlinear error components). This line can be obtained by plotting the corresponding values of ε and RMSET for which Eq. (25) generates only five defocus distances. The red, green and black circles dots represent the actual RMSET of the retrieved phase (obtained by simulation) when employing the MF-TIE for the corresponding values of ε and different phase amplitudes (π/4, π/8 and π/16 respectively). When comparing the solid line and the circles, one can observe that the simulations results follow the noise boundary limit until certain value ε = εMAX is reached. In other words: the nonlinear error components can be neglected for ε < εMAX. The value of this εMAX depends of the noise and phase amplitude. The errors present for ε>εMAX result from the fact that at larger ε the conditions discussed in Sections 2 and 3 are not fulfilled, and thus, nonlinear error components are not suppressed. In consequence, the retrieved phase with the MF-TIE will be inaccurate. From Fig. 3, we can observe that the smaller is the phase amplitude of the object the larger is the range of ε where the error of the retrieved phase is similar to the noise boundary limit. In Fig. 3(a), εMAX≈0.035, 0.25, and 0.47 for the phase amplitudes of π/4, π/8 and π/16 respectively. Notably, for the case of weak phase objects the MF-TIE allows large values of ε without introduce non-linear errors. For the case of Fig. 3(b), we have that εMAX≈0.006, 0.09, and 0.17 for the phase amplitudes of π/4, π/8 and π/16 respectively. In these simulations, the value of εMAX has decreased at least in a factor of five, which is expected. The reason of these results comes from the fact that the noise amplitude has been decreased, and thus, the non-linear component error become more dominant (relative to noise) as ε increased. In Fig. 3(c) it is observed that εMAX≈0.002, 0.006, and 0.08 for π/4, π/8 and π/16 respectively. Similarly to Fig. 3(b), the value of εMAX are lower than in Fig. 3(b), because the noise in the system lower. Thus, the nonlinear error becomes more dominant in the RMSET at larger values of ε (see for example the red circles in Fig. 3).

 figure: Fig. 3

Fig. 3 RMSET vs ε using several phase difference amplitudes and different levels of noise. The red, black and green circles correspond to the phase difference amplitudes of π/4, π/8 and π/16 respectively. a) SNR = 30dB. b) SNR = 40dB c) SNR = 50dB. The solid line is the noise boundary limit given by Eq. (26). The circles are the actual RMSET when employing the MF-TIE. The triangle, pentagon, star and square markers indicate the number of planes necessary to retrieve the phase with a constant value of the RMSET (dash line).

Download Full Size | PDF

Furthermore, Fig. 3 shows four additional additional markers (triangle, pentagon, star and square) for which the RMSET has been fixed to RMSET = 0.021 (Fig. 3(a)), 0.011 (Fig. 3(b)), and 0.008rad (Fig. 3(c)) for the level of noise SNR = 30,40 and 50dB respectively. The fixed value for the RMSET is depicted by the dash-line in each picture. Table 1 shows the numerical conditions for which the markers were obtained in Fig. 3 by employing Eq. (20) and (25). Notably, for generating Table 1 no phase information about the object is necessary. Moreover, in Table 1 can be observed that the number of planes increases as ε decreases. This table also highlights that a high noise robustness can be achieved using only fewer measurement planes by increasing the value of ε. Similarly, the smaller the value for ε, the higher the ability to reconstruct stronger phase objects. Hence, MF-TIE together with the noise sensitivity separation strategy described here allows controlling the mount of error down to an acceptable level, i.e. RMSET of the retrieved the phase.

Tables Icon

Table 1. Defocus distances for various SNR, ε and RMSET

5. Numerical experiments and comparison to other TIE techniques

The performance and frequency response of the MF-TIE solver is investigated by simulation for the case of pure sine gratings having various periods and defocus distances. The accuracy in the reconstructed phase with the MF-TIE is compared to the results obtained by the GP-TIE [23].

Figure 4(a) shows the obtained simulation results for both the MF-TIE solver (solid line) and the GP-TIE (dash line) when using the exponential separation strategy. For this simulation we have employed 11 planes with exponential spacing given by zi = ± 6; ± 23; ± 89; ± 350; ± 1375μm, and various levels of noise SNR = 25,30, and 35dB. Figure 4(b) shows the effectiveness of MF-TIE and GP-TIE solvers under same simulation conditions as in Fig. 4(a), but using the plane selection strategy defined in Eq. (25). The plane separation for this simulation is given as zi = ± 6; ± 10; ± 24, ± 98, ± 1375μm. For a fair comparison, we carried out these simulations using z1 = ± 6μm (same maximum spatial frequency) and z4 = ± 1375μm (same low frequency artifacts) and ε = 0.01. Note that z2,3,4 are smaller than their GOMF counterpart. As it was the case with previous simulations, the zero frequency sample is excluded from the calculations and the DC term is removed from the phase error in order to estimate the RMSET. All simulation have been carried out for β = 0, because the unequal separation strategies mitigate LFAs without using regularization parameters [15,23]. When comparing Fig. 4(a) and Fig. 4(b), we see that despite the high accuracy that can be obtained by the GP-TIE algorithm, the MF-TIE solver gives, under same conditions, a lower RMSE (by a factor of two). Notably, the MF-TIE solver reported here, gives the more accurate result compared to the GP-TIE technique. The performance loss of the GP-TIE solver [23] originates from the fact that this approach only seeks an optimal estimation of the axial derivative; however, as Martinez et at shown in [22], this not imply a retrieved phase with minimal error. The MF-TIE algorithm suppresses LFAs through the BPFs but without affecting the retrieved frequencies of the object. A solver can adapt to this TIE specific behavior by either choosing appropriate measurement planes (as Martinez et at shown in [22]) or by adapting the corner frequencies of the filters (as it is the case here).

 figure: Fig. 4

Fig. 4 RMSE vs for various grating frequencies when using different plane selection strategies for the GP-TIE and MF-TIE for several levels of noise. a) Plane separation strategy defined in Eq. (17). b) Separation strategy defined in Eq. (25). The vertical lines are placed on its corresponding COF. The final results of these plots are the averages of 15 simulations.

Download Full Size | PDF

In the previous simulation, we have tested the response error of the GP-TIE and the MF-TIE for a wide range of frequencies showing an optimal performance for the MF-TIE by using the exponential or equal noise sensitivity plane selection strategy. For the following simulation, we want to compare the error response of the MF-TIE with two different TIE solvers: the classical TIE solver (CL-TIE) [18] with the optimized equally spaced plane separation presented in reference [22], and the GP-TIE solver with the exponential spacing [23]. For the exponential spacing without optimization, we have selected Δz = 10μm (in order to recover high frequency components for low noise conditions) and g0 = 5 (to suppress LFAs in high noise conditions). Moreover, we carry out an additional simulation to show that the exponential spacing can be optimized. The optimization regards the choice of the shortest and largest plane separation. This is not trivial when one considers the effect of noise and non-linear error components. Previous approaches [15,18,23], choose z1 according to the diffraction limit, and a value of zN that suppresses visible effects of LFAs in the reconstructed phase. In this work, we show that the exponential plane selection strategy can be optimized when using the values for z1 and zN that one would obtain for the plane selection strategy with equal noise sensitivity of this work, i.e. Equations (20) and (25). We select the distance z1 and zN from this strategy (Eq. (20)) in order to generate the exponential spacing, and thus, g0 = (zN /z1)1/(N-1) [15]. We consider the same phase object as in Fig. 3 with a maximum phase difference of π/8 and 11 planes. The noise level in this simulation is ranging from 20 to 50dB. It should be noted that for each level of noise the value of ε and the defocus distances will change (see Table 1). A further comparison is made with optimum plane selection strategies for equidistant measurement planes. Notably, the optimal separation dz for equally spaced planes depends on the level of noise. In order to estimate the corresponding plane separation, we applied Eq. (17) from reference [22].

The simulation results are shown in Fig. 5. As expected, the RMSET of the equidistant separation strategy is large for all levels of noise. Notably, the exponential spacing does not give a minimal RMSET for strong noise conditions (SNR<34dB). In fact, the performance of this technique is a slightly better than the optimized even separation strategy for low noise conditions (SNR>34dB). However, when the exponential spacing is optimized with the help of the noise sensitivity formula (Eq. (25)), it can be observed that the RMSET in the retrieved phase is improved in at least a factor of four in relation to the green markers. Thus, the exponential spacing and the GP-TIE can benefit from the spacing strategy developed in this contribution. Finally, Fig. 5 shows that the technique with minimal error is given by the MF-TIE solver combined with the plane selection strategy of this work (Eq. (25)). For this solver, it is observed that the RMSET stagnates for SNR>40dB; a similar effect can be seen in Fig. 3 due to the presence of nonlinear components errors.

 figure: Fig. 5

Fig. 5 RMSE of the retrieved phase for various levels of noise. Three different TIE solvers have been employed: the Classical TIE solver (CL-TIE) [27] with optimal equally spaced planes (triangles) [22], the GP-TIE solver [23] with exponential spacing (cross) and optimized exponential spacing (asterisk), and the MF-TIE with equal noise sensitivity strategy of Eq. (25) (circle). These simulations were carried out with eleven planes. Each marker is the average of 20 simulations.

Download Full Size | PDF

Figure 6 shows a further example of the retrieved phase for the case when the SNR = 40dB and eleven planes. The defocus distances were selected according to the values showed in Table 1. Figure 6(a), (b), and (c), shows the case of the classical TIE (CL-TIE) [18], GP-TIE [23], and the MF-TIE solver, respectively. Similar to previous simulations β = 0 in order to visualize the ability of the techniques to suppress LFAs. The retrieved phase with CL-TIE (Fig. 6a) is not affected by LFAs due to the large plane separation; however, all the sharp features of the phase have been lost. The retrieved phase in Fig. 6(a) has been obtained using the algorithm in [21,31]. As expected, GP-TIE (Fig. 6(b)) provides a sharp phase reconstruction but this result is affected by LFAs that are result of no considering the effect of the IL. Finally, the result obtained by the MF-TIE is not affected by LFAs. This is because the filtering process in the MF-TIE excludes all the noisy and redundant information from other measured planes, which is not the case for the GP-TIE approach. Moreover, the MF-TIE provides the best quantitative results having an RMSET that is four times lower than the RMSET of the GP-TIE and five times lower than the RMSET of CL-TIE. When comparing the corresponding RMSET of Table 1 one can see that these measurements planes have been chosen in order to obtain a RMSET = 0.01rad; when comparing this value with the error of the retrieved phase of Fig. 4(c) it can be seen that this anticipated results could be met.

 figure: Fig. 6

Fig. 6 Phase retrieval for three different solvers and employing the unequal phase separation defined by Eq. (25)

Download Full Size | PDF

5. Conclusions and discussion

In this work, a so called MF-TIE solver is reported that uses a series of low-pass, band-pass and high-pass filters in order to extract the particular object frequencies from a series of defocused intensities. The idea behind this approach is to deliberately exclude spatial frequencies from the measurement data if they do not correspond to the plane that provides minimal error at those frequencies. In this way, MF-TIE obtains phase reconstructions with higher noise robustness than the GP-TIE solver (that is currently the most accurate TIE technique but does not exclude less accurate data sets). The success of this method can be shown for the case of equidistant or non-equidistant measurement planes. Recent efforts in TIE have shown the advantageous use of non-equidistant plane selection strategies, i.e. exponential spacing [15,23]. However, exponential strategies have been originally developed for CTF based solvers and do not account for the properties of the TIE solver, i.e. the noise amplification of the IL. In particular, many TIE techniques as e.g. GP-TIE solver [23] seek to minimize the error in the axial derivative; however, in [22] Martinez et al shown that this not imply a minimal error in the retrieved phase. Hence, for the case of non-equidistant measurement planes there is room for improvement. Motivated by that, this paper reports a novel non-equidistant plane selection strategy that takes into account the properties of the TIE solver and thereby improves further the phase reconstruction. The idea behind this is based on the criterion to equalize the noise contribution to the RMSET of all measurement planes by appropriate selection of the measurement planes and the corner frequencies of the filters. Notably, the developed plane separation technique provides analytical expressions for the smallest and largest plane separation giving high LFA suppression with no need to employ RTs. The choice of the smallest and largest measurement planes for various levels of noise are not explicitly determined by exponential schemes [15,23]. Interestingly, as shown in Fig. 4 and 5, the exponential plane selection strategies may benefit from this information in terms of reconstruction accuracy. In conclusion, this works reports both a new TIE solver and a novel TIE plane selection methodology for high accuracy phase reconstructions. The so-called MF-TIE solver uses series of filter in order to extract the measurement data that gives higher accuracy, where optimal corner frequencies are chosen according to Eq. (6). The reported optimum non-equidistant plane selection strategy allows increasing further the accuracy without employing RTs. This strategy provides a minimum number of planes for a given level of noise and acceptable level of RMSET.

Appendix A

The MF-TIE algorithm of this work assumed the object to be a pure phase object, so that the in-focus intensity I0≈constant [32]. In practice, this may be a good approximation [28,33], but for absorbing objects the use of Eq. (2) is not adequate. For these cases, one has to introduce the potential ∇ψ=I0∇φ [34] into Eq. (1), which yields

ψ=kI(r,z)I(r,z)2z.
When solving Eq. (27), the solution of Eq. (1) can be found as [35,36]
φ=2{[I1ψ]},
where notably, the solution of Eq. (27) exists when ∇I0x∇φ = 0 [32,36], which is fulfilled for φ = constant, I0≈constant, or if the gradients of I0 and φ are parallel [32]. The last case occurs when the transmitted light is in function on the thickness or density of the sample [32,36]. Hence, the accuracy of the solution of Eq. (28) is given by the degree of correlation between the intensity and the phase distributions.

In order to retrieve the phase with the MF-TIE for the absorbing case, the solution presented by Eq. (27) and (28) must be adopted as well. Firstly, we apply the same COFs for the BPFs as for the case of no absorption for calculating the filtered phases ψi. In the next step, the potential function ψ is estimated using a superposition of filtered functions ψi similar to Eq. (14), as

ψ=i=1Nψi.
By substituting Eq. (29) in Eq. (28), the phase can be estimated as
φ=i=1Nφi=i=1N2{[I1ψi]}
For the absorbing case, the diffracted intensities for estimating ψi will contain information about the phase and absorption. In consequence, the BPFs will not be able to filter out the nonlinear errors from the retrieved phase. A practical solution to this problem is to employ the iterative TIE schemes proposed in references [35,36]. Such techniques allow removing the TIE solver errors from the retrieved phase caused by the absorption distribution. Hence, when applying the iterative technique to the retrieved phase by the MF-TIE, we will remove successfully the errors caused by absorption without the necessity of recalculating the COFs.

The applicability of this approach is tested by simulation for phase object of Fig. 6 but having an absorption distribution μ as it is shown in Fig 7(a). The maximum value for μ is set to μMAX=0.5. Simulations have been carried out using the 11 measurement planes of Table 1 for SNR=50dB. The reconstructed phase is shown in Figs. 7(b) and 7(c) for the MF-TIE and GP-TIE algorithm, respectively. Fig. 7(b) indicates that MF-TIE is less sensitive to absorbing samples than the GP-TIE algorithm. Given that the MF-TIE is based in the three plane TIE approach, and not by the weak absorption approximation (WAA) [23], the results showed in Fig. 7(b) are not strongly affected by the absorption of the object. Notably, the loss of accuracy in GP-TIE is because the fitting model only considers small variations due to the absorption of the object. Nevertheless, when having a closer look, the effect absorption does not remain unnoticed for the MF-TIE solver, and similar error may also occur for other TIE techniques [32,35–37]. Figure 8(a) shows the effect of the absorption in the RMSE, where the maximum absorption of Fig. 7(a) is varied from μ=0 (pure phase object) to μ=1 (strong absorption). Despite the resulting error, recently reported technique overcomes this problem within an iterative post-processing scheme [35,36]. The pseudo code of the iterative TIE approach is: (i) Retrieve the phase φ0 with the MF-TIE; (ii)Estimate the axial intensity derivative zIE from the experimental set of images with the GP Regression [23]; (iii)Initialize φi= φ0.; (iv)For i=1 to n do: (iv.a) Calculate numerically the axial derivative zIN with Eq. (29) of reference [23] when having I0 and φi as inputs; (iv.b) Calculate the residual ΔzI=∂zIN-∂zIE ; (iv.c)Use I0, and ΔzI as inputs to Eqs. (27) and (28) to obtain the phase Δφi; (iv.d) Correct initial phase estimate as φi+1= φi+ Δφi .

 figure: Fig. 7

Fig. 7 Phase retrieval in case of absorption for SNR = 50dB. a) Absorption distribution μ. b) Retrieved phase with MF-TIE. c) Retrieved phase with GP-TIE.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 RMSE of the retrieved phase for various values of μ. a) RMSE of the retrieved phase when the in-focus intensity is not constant. b) RMSE of the retrieved phase after employing the iterative TIE algorithm showed of reference [36]. These simulations were made with SNR = 60dB, n = 5. Each marker is the average of 25 simulations.

Download Full Size | PDF

Figure 8(b) shows the RMSE for the retrieved phase when applying the iterative scheme proposed by Zuo et al in [36]. In that simulations only five iterations have been employed (n=5) and each marker is the average of 25 simulations. Fig. 8(b) indicates that the previous iterative procedure reduces the RMSET for both the MF-TIE and GP-TIE solver, giving an increased accuracy by at least a factor of three. It is noticed that MF-TIE converges faster to the exact solution than the GP-TIE. The remaining error at larger values of μ may be reduced for both solvers by increasing the number of iterations; this however requires more computational effort.

Acknowledgments

The research leading to the described results are realized within program TEAM/2011-7/7 of Foundation for Polish Science, co-financed from European Funds of Regional Development. We would like to acknowledge the support of the statutory funds of Warsaw University of Technology. K. Falaggis also acknowledges the partial support of this work by the Outstanding Young Scientists Stipend Award 564/2014 of the Polish Ministry of Science and Higher Education. We also thank the authors of reference [23] for providing the Matlab implementation of the GP TIE algorithm.

References and links

1. J. R. Fienup, “Phase retrieval algorithms: a comparison,” Appl. Opt. 21(15), 2758–2769 (1982). [CrossRef]   [PubMed]  

2. K. A. Nugent, D. Paganin, and T. E. Gureyev, “A Phase odyssey,” Phys. Today 54(8), 27–32 (2001). [CrossRef]  

3. N. Streibl, “Phase imaging by the transport equation of intensity,” Opt. Commun. 49(1), 6–10 (1984). [CrossRef]  

4. K. A. Nugent, “The measurement of phase through the propagation of intensity: an introduction,” Contemp. Phys. 52(1), 55–69 (2011). [CrossRef]  

5. M. K. Kim, “Principles and techniques of digital holographic microscopy,” SPIE Rev. 1, 018005 (2010).

6. P. Langehanenberg, G. Von Bally, and B. Kemper, “Autofocusing in digital holographic microscopy,” 3D Res. 01, 1–11 (2011).

7. B. Kemper and G. von Bally, “Digital holographic microscopy for live cell applications and technical inspection,” Appl. Opt. 47(4), A52–A61 (2008). [CrossRef]   [PubMed]  

8. R. Porras-Aguilar, M. Kujawinska, and W. Zaperty, “Capture and display mismatch compensation for real-time digital holographic interferometry,” Appl. Opt. 53(13), 2870–2880 (2014). [CrossRef]   [PubMed]  

9. V. Y. Ivanov, V. P. Sivokon, and M. A. Vorontsov, “Phase retrieval from a set of intensity measurements: theory and experiment,” J. Opt. Soc. Am. A 9(9), 1515–1524 (1992). [CrossRef]  

10. K. Falaggis, T. Kozacki, and M. Kujawinska, “Accelerated single-beam wavefront reconstruction techniques based on relaxation and multiresolution strategies,” Opt. Lett. 38(10), 1660–1662 (2013). [CrossRef]   [PubMed]  

11. I. A. Shevkunov, N. S. Balbekin, and N. V. Petrov, “Comparison of digital holography and iterative phase retrieval methods for wavefront reconstruction,” Proc. SPIE 9271, 927128 (2014). [CrossRef]  

12. E. J. Candès, T. Strohmer, and V. Voroninski, “Phaselift: exact and stable signal recovery from magnitude measurements via convex programming,” Commun. Pure Appl. Math. 66(8), 1241–1274 (2013). [CrossRef]  

13. M. Reed Teague, “Deterministic phase retrieval: a Green’s function solution,” J. Opt. Soc. Am. 73(11), 1434 (1983). [CrossRef]  

14. M. Langer, P. Cloetens, and F. Peyrin, “Regularization of phase retrieval with phase-attenuation duality prior for 3-D holotomography,” IEEE Trans. Image Process. 19(9), 2428–2436 (2010). [CrossRef]   [PubMed]  

15. K. Falaggis, T. Kozacki, and M. Kujawinska, “Optimum plane selection criteria for single-beam phase retrieval techniques based on the contrast transfer function,” Opt. Lett. 39(1), 30–33 (2014). [CrossRef]   [PubMed]  

16. T. E. Gureyev, A. Pogany, D. M. Paganin, and S. W. Wilkins, “Linear algorithms for phase retrieval in the Fresnel region,” Opt. Commun. 231(1-6), 53–70 (2004). [CrossRef]  

17. F. Roddier, “Wavefront sensing and the irradiance transport equation,” Appl. Opt. 29(10), 1402–1403 (1990). [CrossRef]   [PubMed]  

18. D. Paganin, A. Barty, P. J. McMahon, and K. A. Nugent, “Quantitative phase-amplitude microscopy. III. The effects of noise,” J. Microsc. 214(1), 51–61 (2004). [CrossRef]   [PubMed]  

19. J. C. Petruccelli, L. Tian, and G. Barbastathis, “The transport of intensity equation for optical path length recovery using partially coherent illumination,” Opt. Express 21(12), 14430–14441 (2013). [CrossRef]   [PubMed]  

20. M. H. Jenkins, J. M. Long, and T. K. Gaylord, “Multifilter phase imaging with partially coherent light,” Appl. Opt. 53(16), D29–D39 (2014). [CrossRef]   [PubMed]  

21. C. Zuo, Q. Chen, Y. Yu, and A. Asundi, “Transport-of-intensity phase imaging using Savitzky-Golay differentiation filter--theory and applications,” Opt. Express 21(5), 5346–5362 (2013). [CrossRef]   [PubMed]  

22. J. Martinez-Carranza, K. Falaggis, and T. Kozacki, “Optimum plane selection for transport-of-intensity-equation-based solvers,” Appl. Opt. 53(30), 7050–7058 (2014). [CrossRef]   [PubMed]  

23. Z. Jingshan, R. A. Claus, J. Dauwels, L. Tian, and L. Waller, “Transport of Intensity phase imaging by intensity spectrum fitting of exponentially spaced defocus planes,” Opt. Express 22(9), 10661–10674 (2014). [CrossRef]   [PubMed]  

24. L. Tian, J. C. Petruccelli, and G. Barbastathis, “Nonlinear diffusion regularization for transport of intensity phase imaging,” Opt. Lett. 37(19), 4131–4133 (2012). [CrossRef]   [PubMed]  

25. J. Martinez-Carranza, K. Falaggis, and T. Kozacki, “Optimum measurement criteria for the axial derivative intensity used in transport of intensity-equation-based solvers,” Opt. Lett. 39(2), 182–185 (2014). [CrossRef]   [PubMed]  

26. C. T. Koch, “Towards full-resolution inline electron holography,” Micron 63, 69–75 (2014). [CrossRef]   [PubMed]  

27. L. Waller, L. Tian, and G. Barbastathis, “Transport of Intensity phase-amplitude imaging with higher order intensity derivatives,” Opt. Express 18(12), 12552–12561 (2010). [CrossRef]   [PubMed]  

28. C. Zuo, Q. Chen, W. Qu, and A. Asundi, “High-speed transport-of-intensity phase microscopy with an electrically tunable lens,” Opt. Express 21(20), 24060–24075 (2013). [CrossRef]   [PubMed]  

29. J. Martinez-Carranza, K. Falaggis, M. Jozwik, and T. Kozacki, “Comparison of phase retrieval techniques based on the transport of intensity equation using equally and unequally spaced plane separation criteria,” Proc. SPIE 9204, 92040G (2014). [CrossRef]  

30. K. Falaggis, D. P. Towers, and C. E. Towers, “Phase measurement through sinusoidal excitation with application to multi-wavelength interferometry,” J. Opt. 11, 54008A (2009).

31. R. Bie, X.-H. Yuan, M. Zhao, and L. Zhang, “Method for estimating the axial intensity derivative in the TIE with higher order intensity derivatives and noise suppression,” Opt. Express 20(7), 8186–8191 (2012). [CrossRef]   [PubMed]  

32. J. A. Schmalz, T. E. Gureyev, D. M. Paganin, and K. M. Pavlov, “Phase retrieval using radiation and matter-wave fields: Validity of Teague’s method for solution of the transport-of-intensity equation,” Phys. Rev. A 84(2), 023808 (2011). [CrossRef]  

33. C. Zuo, Q. Chen, W. Qu, and A. Asundi, “Noninterferometric single-shot quantitative phase microscopy,” Opt. Lett. 38(18), 3538–3541 (2013). [CrossRef]   [PubMed]  

34. D. Paganin and K. Nugent, “Noninterferometric phase imaging with partially coherent light,” Phys. Rev. Lett. 80(12), 2586–2589 (1998). [CrossRef]  

35. C. Zuo, Q. Chen, L. Huang, and A. Asundi, “Phase discrepancy analysis and compensation for fast Fourier transform based solution of the transport of intensity equation,” Opt. Express 22(14), 17172–17186 (2014). [CrossRef]   [PubMed]  

36. A. Shanker, L. Tian, M. Sczyrba, B. Connolly, A. Neureuther, and L. Waller, “Transport of intensity phase imaging in the presence of curl effects induced by strongly absorbing photomasks,” Appl. Opt. 53(34), J1–J6 (2014). [PubMed]  

37. A. Lubk, G. Guzzinati, F. Börrnert, and J. Verbeeck, “Transport of intensity phase retrieval of arbitrary wave fields including vortices,” Phys. Rev. Lett. 111(17), 173902 (2013). [CrossRef]   [PubMed]  

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 (8)

Fig. 1
Fig. 1 a) IPTFTIE for three different defocus distances and b) RMSE frequency response of the retrieved phases. The vertical lines are placed in its corresponding COF. The lines labeled with the legend RMSES correspond to the RMSE of the retrieved phase through simulations, and the theoretical RMSE (RMSET) are denoted with the solid lines. Moreover, the analytical results of Eq. (11) and the position of the COFs are plotted with the horizontal and the vertical dashed lines respectively. The RMSES presented in this figure are the average of 25 simulations.
Fig. 2
Fig. 2 Flow chart of the MF-TIE based solver.
Fig. 3
Fig. 3 RMSET vs ε using several phase difference amplitudes and different levels of noise. The red, black and green circles correspond to the phase difference amplitudes of π/4, π/8 and π/16 respectively. a) SNR = 30dB. b) SNR = 40dB c) SNR = 50dB. The solid line is the noise boundary limit given by Eq. (26). The circles are the actual RMSET when employing the MF-TIE. The triangle, pentagon, star and square markers indicate the number of planes necessary to retrieve the phase with a constant value of the RMSET (dash line).
Fig. 4
Fig. 4 RMSE vs for various grating frequencies when using different plane selection strategies for the GP-TIE and MF-TIE for several levels of noise. a) Plane separation strategy defined in Eq. (17). b) Separation strategy defined in Eq. (25). The vertical lines are placed on its corresponding COF. The final results of these plots are the averages of 15 simulations.
Fig. 5
Fig. 5 RMSE of the retrieved phase for various levels of noise. Three different TIE solvers have been employed: the Classical TIE solver (CL-TIE) [27] with optimal equally spaced planes (triangles) [22], the GP-TIE solver [23] with exponential spacing (cross) and optimized exponential spacing (asterisk), and the MF-TIE with equal noise sensitivity strategy of Eq. (25) (circle). These simulations were carried out with eleven planes. Each marker is the average of 20 simulations.
Fig. 6
Fig. 6 Phase retrieval for three different solvers and employing the unequal phase separation defined by Eq. (25)
Fig. 7
Fig. 7 Phase retrieval in case of absorption for SNR = 50dB. a) Absorption distribution μ. b) Retrieved phase with MF-TIE. c) Retrieved phase with GP-TIE.
Fig. 8
Fig. 8 RMSE of the retrieved phase for various values of μ. a) RMSE of the retrieved phase when the in-focus intensity is not constant. b) RMSE of the retrieved phase after employing the iterative TIE algorithm showed of reference [36]. These simulations were made with SNR = 60dB, n = 5. Each marker is the average of 25 simulations.

Tables (1)

Tables Icon

Table 1 Defocus distances for various SNR, ε and RMSET

Equations (30)

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

· [ I 0 φ ] = k I ( r , z ) I ( r , z ) 2 z ,
φ = FT 1 I ^ 0 [ FT ( A ( r , z ) ) 4 π λ f 2 z ] ,
PTF T I E = π λ f 2 z ,
PTF= sin ( π λ f 2 z ) .
PTF PTF T I E = 1 ε ,
ζ ( ε , z ) = 6 ε π λ z .
RMSE 2 = K 1 z 2 c 0 + z 4 K 2 W | F T 1 [ f 2 [ z 3 I ˜ | z = 0 ] ] | 2 d 2 r ,
I ˜ ( f , z ) = δ ( f ) + 2 M cos ( π λ f 2 z ) 2 Φ sin ( π λ f 2 z ) ,
3 z 3 I ˜ ( f , z ) | z = 0 = 2 ( π λ f 2 ) 3 Φ .
RMSE 2 = C 1 [ σ 2 ( π λ ) 2 ζ 4 6 ε + C 2 C 1 ( 6 ε ( π λ ) 2 ) 2 ( f j ζ ) 8 ] ,
RMSE 2 = C 1 [ σ 2 ( π λ ) 2 ζ 4 6 ε 1 ] .
0 < ζ ( z N ) < ζ ( z N 1 ) < < ζ ( z 2 ) f m a x ,
φ i = F T 1 [ F ( i ) ( A ( z i ) ) 4 π λ z i f 2 I 0 ] .
φ ( x , y ) = i = 1 N φ i ( x , y ) .
RMSE i 2 = ( k σ π ) 2 1 32 π 2 z i 2 1 W 2 W | FT 1 [ F ( i ) / f 2 ] | 2 d 2 r + z i 4 K 2 W | FT 1 [ [ F ( i ) × z 3 I ˜ / f 2 ] ] | 2 d 2 r .
RMSE i 2 = ( k σ π ) 2 1 32 π 2 z i 2 α i ,
z j = g 0 j 1 Δ z ,
RMSE N 2 = ( k σ π ) 2 1 32 π 2 z N 2 α N ,
α N = W 2 Δ x 2 π [ 2 Δ f 2 π λ z N ( 6 ε ) 1 / 2 ] ,
z N = 2 1 ( RMSE N 2 K 1 π ) [ b 2 6 ε + b 2 2 6 ε + 8 ( RMSE N 2 K 1 π ) ] ,
α N-1 = W 2 Δ x 2 π [ ζ N 2 ζ N 1 2 ] ,
RMSE N 2 = K 1 π z N 1 2 Δ f 2 [ ζ N 2 ζ N 1 2 ] .
z N 1 2 + Γ [ z N 1 z N ] = 0 ,
z N 1 = 2 1 ( Γ + Γ 2 + 4 z N Γ ) .
z i 1 = 2 1 ( Γ + Γ 2 + 4 z i Γ ) .
RMSE T = N RMSE N ,
ψ = k I ( r , z ) I ( r , z ) 2 z .
φ = 2 { [ I 1 ψ ] } ,
ψ = i = 1 N ψ i .
φ = i = 1 N φ i = i = 1 N 2 { [ I 1 ψ i ] }
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.