Open Access
8 April 2014 Polarization spatial heterodyne interferometer: model and calibration
Author Affiliations +
Abstract
Spatial heterodyne interferometry (SHI) is a technique based on Fourier transform spectroscopy. As such, many of the benefits, such as high spectral resolving power, can be realized. Furthermore, unlike a Fourier transform spectrometer, an SHI is able to minimize the number of required samples for a given resolving power and spectral range. The calibration and detailed modeling of a polarization spatial heterodyne interferometer (PSHI) are detailed. Unlike our original first-order ray tracing model, the new model is based on the Jones matrix formalism. Using this improved model, we explore the nonideal aspects of the PSHI, including interference effects caused by retardance errors in the polarization grating and quarter wave plate. To minimize the influence of these errors, a calibration procedure is described based on a linear operator theory. Finally, the Jones matrix model and calibration procedure are validated through a series of simulations and experiments.

1.

Introduction

Fourier transform spectrometers (FTSs) are well known for their high spectral resolving power.1,2 Typically, high spectral resolution measurements are based on Michelson interferometers. Although these offer measurement flexibility, high throughput, and spectral resolution, disadvantages include vibration sensitivity, temporal and mechanical scanning requirements, and alignment sensitivity. Fortunately, vibration sensitivity and the need for mechanical scanning can be overcome with common-path interferometers.3 These sensors map optical path difference (OPD) across a two-dimensional (2-D) focal plane array (FPA) or line array (LA) camera such that

Eq. (1)

OPD(x,y)ay+b,
where a is the slope and b is the offset. This enables multiple interferogram samples to be acquired simultaneously, enhancing the speed of the spectral measurement.

One issue with conventional FTS implementations, including common-path systems, is sampling inefficiency. For instance, a conventional nonaliased FTS must resolve all wave numbers to fully measure a spectrum.4 This limitation is best observed with an example. Consider a conventional FTS, designed to measure the atmospheric oxygen emission within the spectral band of 558±3nm (e.g., 17,825 to 18,018cm1) to infer atmospheric wind speed and temperature.58 Assuming that a complete interferogram is required (i.e., not the partial interferograms described in Refs. 56.7.8), then the required spectral resolution Δσ of the system is approximately 1.0cm1. If the spectrometer’s cutoff frequency were 540 nm (18,519cm1), then the FTS must measure all spectral points spanning 0cm1 to the Nyquist frequency. Therefore, while only 193 samples within the region of the emission line are actually of interest, approximately 18,519 samples, assuming a single-sided interferogram, are required to acquire the region of interest. Ultimately, this value would be slightly greater, since a small double-sided region is required for phase correction.1

Although sampling inefficiencies can be overcome in a conventional Michelson interferometer by aliasing the interferogram’s measurement,9 this is not a viable option with most FPA- or LA-based FTSs. Given a pixel width of dw, the interference’s spatial frequency ηI in units of cycles/pixel, generated by the OPD defined in Eq. (1), can be expressed as

Eq. (2)

ηI=aσdw,
where σ is the wave number. Since each pixel has a finite optical transfer function, the interference fringe’s visibility will generally decrease with increasing spatial frequency. Interference visibility is defined as

Eq. (3)

V=(ImaxImin)/(Imax+Imin),
where Imax and Imin are the maximum and minimum intensities within a local fringe. This tradeoff is illustrated in Fig. 1, which depicts V versus spatial frequency ηI. Note that this analysis assumes a pixel pitch that is equal to the pixel width.

Fig. 1

Fringe visibility versus the interferogram’s spatial frequency.

OE_53_4_044104_f001.png

Revisiting the previous oxygen emission example demonstrates the impracticality of the aliasing technique when using detector arrays. Assuming a linear array is used with a pixel pitch of dw means that the required slope of the OPD can be expressed as

Eq. (4)

a=1/(ΔσNsdw),
where Ns is the number of samples in the interferogram. Substituting Eq. (4) into Eq. (2) yields a normalized spatial frequency of

Eq. (5)

ηI=σ/(ΔσNs).

Given Ns=18,519 yields ηI=1.0 cycles/pixel, resulting in zero fringe visibility per Fig. 1. Ideally the spatial frequency must be decreased in order to increase the detector’s response. However, the issue is only worsened when using the aliasing technique. For instance, if aliasing is implemented on a 1000 pixels array, then ηI=17.9 cycles/pixel. Thus, while the spatial frequency is aliased, its visibility is too low to remain detectable.

Two techniques exist to increase the spectral resolution in these common-path instruments while minimizing the number of samples required: (1) modifying the offset b while maintaining a resolvable slope a or (2) heterodyning the interference to lower spatial frequencies.4,10,11 Since modifying b does little to negate the sampling limitation, we consider spatial heterodyning in the current approach. A spatial heterodyne interferometer (SHI) is able to conserve the high spectral resolution of conventional FTS systems.58 They do so by replacing the retro-reflecting mirrors of a Michelson interferometer with Littrow-configured blazed diffraction gratings.5,11 This enables any arbitrary wave number, σh, to generate a 0 cycle/m interference fringe, where σh is the heterodyne wave number (note that in the conventional nonaliased FTS, σh=0cm1).

The implementation of SHI has primarily leveraged the Michelson interferometer (MI) architecture.7,8 In these systems, the MIs vibration and thermal errors have been addressed by creating monolithic all-glass systems. Alternatives to this technique, which aim to create common-path vibration insensitive designs, are based on Sagnac interferometers.12,13 However, to minimize both size and vibration sensitivity, SHIs based on fiber-optic Mach–Zehnder interferometers14 and an SHI system based on birefringent prisms and polarization gratings (PGs)15,16 have also been introduced. With these implementations, the interferometer’s size, weight, and alignment complexity can be significantly reduced over reflective free-space Sagnac interferometers or uncommon-path monolithic Michelson interferometer designs.

In this paper, we focus on advancing the modeling and calibration procedures of a polarization spatial heterodyne interferometer (PSHI) that was originally presented in Ref. 15. Advantages of this architecture lie in its potential compactness, due to the use of monolithic birefringent crystals, and its vibration insensitivity, stemming from its common-path design.17,18 In Sec. 2, we discuss a Jones matrix model of the PSHI and investigate its nonideal properties. In Sec. 3, we present our experimental system while Sec. 4 contains the experimental measurements that are used to validate the model. Finally, in Sec. 5, we discuss the methods of calibration and experimentally demonstrate the calibration procedure, validating it against a separate spectral measurement.

2.

Theoretical Model

The goal of our theoretical model is to understand the nonideal characteristics that arise when performing spatial heterodyning with a PG.16 In the PSHI, interference cross talk was caused by the PG’s unwanted diffraction orders.15 These orders are presented in Fig. 2, superimposed on the schematic of the original PSHI system. Polarized light initially transmits through a Wollaston prism (WP), with an apex angle α, which causes the two incident polarization states to split at an angle θWP. These beams are then incident on a quarter wave plate (QWP). The QWP has a fast axis orientation of 45 deg with respect to the x, such that the orthogonal linear eigenpolarizations of the WP are converted into orthogonal circular polarization states. These states interact with the PG and are subsequently diffracted into either the m=0 or m=±1 diffraction orders. A linear analyzer (A) unifies the polarization state, enabling interference to be measured on the FPA.

Fig. 2

Schematic of the beams exiting the PSHI.

OE_53_4_044104_f002.png

Ideally, after the WP’s light transmits through the PG, only the two m=1 order beams, with angular separation θd (e.g., an angularly reduced component compared to θWP), should exist. However, the PG’s zero-order light leakage allows two unwanted beams to propagate with an angular separation of θ0, such that θ0=θWP. Furthermore, error in the QWP’s retardance will cause some light to be coupled into the m=+1 order, creating two additional unwanted beams propagating at an angle θu with respect to each other. Finally, the cross terms, denoted as θc1, θc2, and θc3, also create observable interference effects. Under the small angle approximation, the angle of diffraction from the PG can be expressed as

Eq. (6)

θm=θWP+mλΛ,
where Λ is the PG’s spatial period and λ is the free-space wavelength. Due to the linearity of Eq. (6), θc1=θc2 and θc3=2θc1.

The Jones matrix formalism19 was used to model the interference created by the beams exiting the PG. For a WP, the Jones matrix can be expressed as

Eq. (7)

JWP=(100eiUWP),
where UWP is the spatially dependent phase delay imparted to the polarization states by the WP. This term can be expressed as

Eq. (8)

UWP=4πσΔnCtan(α)y,
where α is the prism’s apex angle, σ=1/λ is the wave number, and ΔnC is the birefringence of the prism’s crystal such that

Eq. (9)

ΔnC=(ne,Cno,C),
where ne,C and no,C are the crystal’s extraordinary and ordinary refractive indices, respectively. Meanwhile, the Jones matrix for a general retarder is

Eq. (10)

JRet(θ,ϕ)=(ABBC),
in which

Eq. (11)

A=cos2(θ)+exp[iϕ]sin2(θ),

Eq. (12)

B=cos(θ)sin(θ)exp[iϕ]cos(θ)sin(θ),
and

Eq. (13)

C=exp[iϕ]cos2(θ)sin2(θ),
where θ is the fast axis orientation angle in the xy-plane and ϕ is the retardance. The Jones matrix for the PG can be expressed using ϕ=ϕPG with a fast axis angle θ=UPG such that

Eq. (14)

UPG=2πy/Λ,
where UPG is the PG’s spatially dependent phase term, ϕPG is the retardance of the liquid crystal layer, and Λ is the PG’s spatial period.20,21 The spatial frequencies of the WP and PG can be expressed as

Eq. (15)

ηWP=2σΔnCtan(α),
and

Eq. (16)

ηPG=1/Λ,
respectively. Meanwhile, the retardance of the PG’s polymerized liquid crystal layer can be expressed as

Eq. (17)

ϕPG=2πσdLCΔnLC,
where dLC is the liquid crystal layer’s thickness and ΔnLC is the liquid crystal’s birefringence such that

Eq. (18)

ΔnLC=(ne,LCno,LC),
where ne,LC and no,LC are the extraordinary and ordinary refractive indices of the liquid crystal. The refractive indices of the RMS03-001C (Merck) reactive mesogen, according to the vendor’s datasheet, are

Eq. (19)

ne,LC=1.629+18350/λ2,
and

Eq. (20)

no,LC=1.501+10010/λ2,
where the wavelength λ is in nm. It should be mentioned that the dispersion in ϕPG is the primary contributor to error addressed in the current paper.

Meanwhile, the QWP can be expressed using Eq. (10) with ϕ=ϕQWP and θ=θQWP, where θQWP and ϕQWP are the QWPs orientation and retardance, respectively. The Jones matrix for a linear polarizer, with a transmission axis oriented at 0 deg with respect to the x-axis, is

Eq. (21)

JLP=(1000).

Calculating the electric field incident onto the detector can be accomplished by

Eq. (22)

Eout=JLP·JPG·JQWP·JWP·Ein.

Using a 45-deg linearly polarized input makes

Eq. (23)

Ein=E(σ)[1/21/2]T,
where the superscript T represents the transpose operation and E(σ) represents the incident spectrum’s complex amplitude at a specific wave number. Setting θQWP=±45deg while maintaining generality with ϕPG and ϕQWP enables the intensity, detected by the FPA, to be calculated as

Eq. (24)

Iout(x,y,σ)=|Eout|2=16|E(σ)|2+I0(x,y,σ)+Ic1(x,y,σ)+Id(x,y,σ)+Ic3(x,y,σ)+Iu(x,y,σ),
where I0,Ic1,Id,Ic3, and Iu represent the interference between the beams corresponding to the angles θ0, θc1, θd, θc3, and θu, respectively, per Fig. 2. The intensity of these components can be represented as

Eq. (25)

I0(x,y,σ)=2|E(σ)|2[±cos(UWPϕPGϕQWP)cos(UWPϕPG+ϕQWP)±cos(UWP+ϕPGϕQWP)cos(UWP+ϕPG+ϕQWP)±2cos(UWP+ϕQWP)±2cos(UWPϕQWP)],

Eq. (26)

Ic1(x,y)=2|E(σ)|2[sin(UWPUPG+ϕPG+ϕQWP)sin(UWPUPGϕPGϕQWP)+sin(UWPUPG+ϕPGϕQWP)sin(UWPUPGϕPG+ϕQWP)],

Eq. (27)

Id(x,y)=|E(σ)|2[±cos(UWP2UPG+ϕPG+ϕQWP)cos(UWP2UPG+ϕPGϕQWP)cos(UWP2UPGϕPG+ϕQWP)cos(UWP2UPGϕPGϕQWP)2cos(UWP2UPG+ϕQWP)±2cos(UWP2UPGϕQWP)4sin(UWP2UPG)+2sin(UWP2UPGϕPG)+2sin(UWP2UPG+ϕPG)],

Eq. (28)

Ic3(x,y)=2|E(σ)|2[sin(UWP+UPG+ϕPG+ϕQWP)sin(UWP+UPGϕPGϕQWP)+sin(UWP+UPGϕPGϕQWP)sin(UWP+UPGϕPG+ϕQWP)],
and

Eq. (29)

Iu(x,y)=|E(σ)|2[±cos(UWP+2UPG+ϕPG+ϕQWP)cos(UWP+2UPG+ϕPGϕQWP)±cos(UWP+2UPGϕPG+ϕQWP)cos(UWP+2UPGϕPG+ϕQWP)2cos(UWP+2UPG+ϕQWP)±2cos(UWP+2UPGϕQWP)+4sin(UWP+2UPG)2sin(UWP+2UPG+ϕPG)2sin(UWP+2UPGϕPG)].

It should be mentioned that Eqs. (25)–(29) assume negligible ±2nd-order diffraction, which was measured at 0.21% in our experimental PG. Expanding the discussion to continuous spectral distributions means that Eq. (24) can be spectrally band-integrated such that

Eq. (30)

I(x,y)=T(x,y,σ)R(x,y,σ)Iout(x,y,σ)dσ,
where T(x,y,σ) is the transmission of the optics and R(x,y,σ) is the responsivity of the detector. Fourier transformation of Eq. (30) yields five frequency components, corresponding to the Fourier transformation of Eqs. (25)–(29). Since the calculation of these transforms is straightforward and providing their rigorous closed form representations offers limited utility, we have expressed only their proportionalities for clarity. The Fourier transformations of Eqs. (25)–(29) are

Eq. (31)

F0(x,y)|E(σ)|2*[A0(σ)δδ(σηWP)]dσ,

Eq. (32)

Fc1(x,y)|E(σ)|2*[Ac1(σ)δδ(σηWPηPG)]dσ,

Eq. (33)

Fd(x,y)|E(σ)|2*[Ad(σ)δδ(σηWP2ηPG)]dσ,

Eq. (34)

Fc3(x,y)|E(σ)|2*[Ac3(σ)δδ(σηWP+ηPG)]dσ,
and

Eq. (35)

Fu(x,y)|E(σ)|2*[Au(σ)δδ(σηWP+2ηPG)]dσ,
where A0, Ac1, Ad, Ac3, and Au are magnitude coefficients that are proportional to the superpositions of sinusoidal functions in Eqs. (25)–(29) and * represents a convolution. Meanwhile

Eq. (36)

δδ(σ/η)(1/2)[δ(σ+η)+δ(ση)],
where δ is the Dirac delta function. These Fourier transform proportionalities indicate that each frequency component carries with it the power spectrum |E(σ)|2, which is modified by the coefficients A0, Ac1, Ad, Ac3, and Au. In an ideal SHI, only Fd should exist and Ad would be unity for all σ. However, deviations in ϕPG and ϕQWP, away from their ideal values, yields nonzero values of A0, Ac1, Ac3, and Au. This creates undesired frequency components that coexist with the desired down-shifted spectrum.

2.1.

Simulated Results

Numerical simulations were performed, using Eq. (24), to investigate the systematic effects caused by retardance errors in the PG and QWP. Double-sided interferograms were created for monochromatic spectra at wavelengths λ0 of 460, 560, and 700 nm. A WP, with a wedge angle of α=6.2 deg and a PG with a period Λ=453μm, were used to maintain consistency with our previous experimental setup per Ref. 15. A fast Fourier transform (FFT) was applied to the simulated 1280 pixel element interferograms, along the y-dimension, such that

Eq. (37)

FF0(yj,σ0)=|FFT(H(y)I0(yj,σ0))|,

Eq. (38)

FFc1(yj,σ0)=|FFT(H(yj)Ic1(yj,σ0))|,

Eq. (39)

FFd(yj,σ0)=|FFT(H(yj)Id(yj,σ0))|,

Eq. (40)

FFc3(yj,σ0)=|FFT(H(yj)Ic3(yj,σ0))|,
and

Eq. (41)

FFu(yj,σ0)=|FFT(H(yj)Iu(yj,σ0))|,
where the constant term in Eq. (24) has been removed, the x-dimension has been suppressed because interference only occurs along y, σ0=1/λ0, E(σ0)=1 for all σ0, and the subscript j denotes y as a discrete quantity. A Hanning window, H(yj), was used to apodize the interferograms and is defined as

Eq. (42)

H(yj)=0.5(1+cos(πyj/w)),
where 2w is the Hanning window’s full-width.

Equations (37)–(41) were calculated for various values of ϕPG and ϕQWP. First, the WP’s spatial frequency power spectrum was calculated without the PG by setting ϕPG=ϕQWP=0deg for all σ0. This yielded the spatial frequency spectrum depicted in Fig. 3(a) in which FF0 is the only nonzero component. For an ideal PG and QWP, ϕPG=180deg and ϕQWP=90deg for all σ0. For these values, Figs. 3(b) and 3(c) depict the spectra that are heterodyned to both high (θQWP=+45deg) and low (θQWP=45deg) spatial frequencies, respectively. Meanwhile, the spatial frequency spectrum that is obtained with error in the PG’s retardance is depicted in Fig. 3(d), where ϕPG=135deg and ϕQWP=90deg. In this case, zero-order light leakage is manifested by the presence of the FF0 component. Conversely, the result of error in the QWP’s retardance is shown in Fig. 3(e), where ϕPG=180deg and ϕQWP=45deg. In this case, both the high and low spatial frequency components, corresponding to FFu and FFd, respectively, are present. Finally, simultaneous error in both the PG and QWP retardance values is shown in Fig. 3(f), where ϕPG=135deg and ϕQWP=45deg. Frequency components FF0, FFu, and FFd are present, as well as the cross-interference terms corresponding to FFc1 and FFc3. Notable is that the FFc1 component is multiplexed with the desired FFd component.

Fig. 3

Relative magnitude of the Fourier components versus spatial frequency. The three values of λ0 are shown alongside the channel magnitudes to indicate the approximate mapping of wavelength to spatial frequency within each channel. Simulation results for (a) Wollaston prism only with ϕPG=0deg, ϕQWP=0deg. (b) Ideal up-shifted spectra with ϕPG=180deg, ϕQWP=90deg, and θQWP=+45deg. (c) Ideal down-shifted spectra with ϕPG=180deg, ϕQWP=90deg, and θQWP=45deg. (d) Error in only the PGs retardance with ϕPG=135deg, ϕQWP=90deg, and θQWP=45deg. (e) Error in only the QWPs retardance with ϕPG=180deg, ϕQWP=45deg, and θQWP=45deg. (f) Error in both the PGs and QWPs retardance values with ϕPG=135deg, ϕQWP=45deg, and θQWP=45deg.

OE_53_4_044104_f003.png

Finally, a simulation was performed using a PG retardance spectrum that was calculated per Eq. (17) with dLC=2.05μm. A QWP, based on quartz crystal, was also simulated using

Eq. (43)

ϕQWP=2πσdQWPΔnquartz,
where dQWP=15.27μm was used for the quartz crystal’s thickness. The birefringence of quartz was simulated using

Eq. (44)

Δnquartz=Lσ3+Gσ2+Jσ+K,
where L=7.655E23, G=3.243E16, J=9.101E10, and K=8.137E3 and σ has units m1. The simulated retardance spectra for the PG and QWP are presented in Fig. 4(a) while the spatial frequency content is depicted in Fig. 4(b). Of particular importance is that the maximum magnitude of the FFd component corresponds to ϕPG=180deg and that the multiplexing between FFd and FFc1 is present due to simultaneous error in both ϕPG and ϕQWP.

Fig. 4

(a) Simulated retardance spectra for a PG, with a layer thickness d=2.05μm, and a quartz QWP with a thickness of 15.27 μm. (b) Simulated frequency content of each component.

OE_53_4_044104_f004.png

3.

Experimental Setup

A view of the system that was used to validate the theoretical model is depicted in Fig. 5. It consists of a diffuse source that illuminates a linear polarization generator (LPG). The LPG consisted of a linear polarizer with a transmission axis nominally oriented at θG=135deg, where the angle is measured with respect to the x-axis. Light from the LPG was then incident onto a quartz WP that has an apex angle α of 6.2 deg. The two linear eigenpolarizations of the WP are oriented at 0 and 90 deg. An achromatic quarter wave plate (AQWP) follows the WP with a nominal orientation of θQWP=±45deg such that the linear eigenpolarizations from the WP are converted into circular polarization states. A singlet, with a focal length f=100mm, was used to relay an image of the WP’s wedge onto a PG with a lateral magnification mg=1. A variable iris was used to stop the lens down to reduce aberrations. It should be mentioned that this low power singlet was incorporated, as opposed to a multiple-element lens, to minimize polarization aberrations between the AQWP and the PG. The PG was patterned with polymerized liquid crystal with a spatial period of Λ=453μm and a peak first-order diffraction efficiency at λ=610nm. Light from the PG was then transmitted through a linear polarization analyzer, which consisted of a linear polarizer with a transmission axis nominally oriented at θA=45deg. Finally, a 50 mm focal length c-mount objective lens imaged the interference pattern onto an 8-bit 1280×960 pixel element FPA.

Fig. 5

Experimental setup of the PSHI. The PG’s grating vector is indicated by the arrow, as are the WP’s fast axis orientations.

OE_53_4_044104_f005.png

The measured retardance spectrum of the AQWP is provided in Fig. 6(a). This measurement was taken by rotating the AQWP between crossed linear polarizers while measuring the transmitted spectrum using an Ocean Optics HR4000 spectrometer. The accuracy of this measurement is estimated at ±0.25%. Meanwhile, the PG’s retardance was calculated from the measured zero-order diffraction efficiency by fitting it to the theoretical closed-form expression

Eq. (45)

DE0(λ)=cos2(ϕPG(λ)/2),
where ϕPG(λ) is used as the fitting parameter.21 Using a least squares fitting procedure yielded the PG retardance depicted in Fig. 6(b). Here, ϕPG is a half wave at a wavelength of 610 nm, corresponding to the minimum zero-order diffraction efficiency.

Fig. 6

Measured (a) achromatic QWP retardance spectrum and (b) PG zero-order diffraction efficiency and retardance spectra.

OE_53_4_044104_f006.png

4.

Experimental Results

The experimental procedure for validating the model consisted of aligning the AQWP at θQWP=+45deg and θQWP=45deg to up- and down-shift the prism’s interference frequency, respectively. Light from a monochromator was used to illuminate the PSHI. The source’s wavelength, λs, was varied from 460 to 700 nm in 20 nm increments and at each of these wavelengths a 2-D interference pattern was recorded. Since the interference components amplitude modulate the spectrum of the light source used in the monochromator, the spectrum’s amplitude was removed by calibrating to a flat field at each wavelength.

In a procedure similar to Ref. 17, a flat field measurement was calculated at each monochromatic wavelength by acquiring two interferograms that are out-of-phase by 180 deg. The nominal interferogram, IG45, was acquired with the LPG at θG=45deg while the phase-shifted interferogram, IG135, was acquired at an LPG orientation of θG=135deg. These two interferograms can be expressed as

Eq. (46)

IG45(x,y,σs)=TR|E(σs)|2[16+I0+Ic1+Id+Ic3+Iu],
and

Eq. (47)

IG135(x,y,σs)=TR|E(σs)|2[16I0Ic1IdIc3Iu],
where T, R, I0, Ic1, Id, Ic3, and Iu are implicitly dependent on x, y, and σs for clarity and σs=1/λs. Averaging Eqs. (46) and (47) yields a flat field containing only the spectral magnitude, optical transmission, and responsivity quantities such that

Eq. (48)

IFlat(x,y,σs)=16TR|E(σs)|2.

A flat field measurement was calculated at both orientations of the QWP (θQWP=±45deg) in order to account for polarization dependencies in the optical transmission. Interferograms of monochromatic spectra were then acquired by the PSHI such that

Eq. (49)

IM(x,y,σs)=TRIout(x,y,σs).

Normalization of Eq. (49) by (48) produces an interferogram without responsivity, optical transmission, or spectral dependencies.

After all monochromator interferograms were normalized to the flat field, an average across the FPA’s x-dimension was taken in order to obtain a one-dimensional interferogram along y. Fourier spectra for each interferogram were then calculated per Eqs. (37)–(41). A summary of the experimental data is provided in Fig. 7 for both the (a) down-shifted (b) up-shifted cases. In each plot, the magnitude of the Fourier components for the experimental measurements (M) is plotted alongside the results from the theoretical model (T) for each channel. Due to the use of an achromatic QWP, minimal multiplexing is observed between FFd and FFc1. Hence, the most significant error is a result of zero-order light leakage through the PG, which can be reduced by incorporating an achromatic PG.22 It should be mentioned that this was not implemented in the current paper so that we could study the effects of zero-order light leakage.

Fig. 7

Comparison between the theoretical model (T) and measurements (M) for QWP orientations of (a) θQWP=+45deg and (b) θQWP=45deg. There is an excellent agreement between T and M for all wavelengths.

OE_53_4_044104_f007.png

The root-mean-square (RMS) error between the measured and theoretical Fourier components was calculated as

Eq. (50)

RMS=1001NsσsNs(FFTheo(σs)FFMeas(σs))2,
where Ns is the number of spectral measurements and FFTheo and FFMeas correspond to the theoretical and measured Fourier component magnitudes (e.g., FF0, FFd, FFc1, etc.). The RMS error was calculated for each frequency component in both the up- and down-shifted datasets, the results of which are summarized in Table 1. Note that for FFc3 in the down-shifted data, only four data points were used where overlap did not occur with FFd, per Fig. 7(a). The mean RMS error across all of the measured Fourier components is 1.02%, which is approximately equal to our experiment’s estimated accuracy of 0.9% accounting for quantization error, alignment error in θQWP, and error in the measured values of ϕPG and ϕQWP.

Table 1

Percent RMS error between theoretical and measured data for each Fourier component.

FF0FFc3FFuFFc1FFd
Up-shifted1.071.321.190.841.03
Down-shifted1.710.291.810.740.21

5.

Calibration Results

For most applications, the down-shifted spectra of Fig. 7(a) are of greatest utility. To account for linear mixing between the FFd and FFc1 components in this configuration, a linear system model was used to spectrally calibrate the PSHI.14,23 A matrix H was created such that

Eq. (51)

g=H·f,
where H is the system transfer matrix, f is the input spectrum, and g is the measured interferogram. The matrix H is configured with dimensions of OPD×λ, such that each column contains a monochromatic interferogram.

Using the experimental setup, depicted previously in Fig. 5, the H-matrix’s calibration performance was evaluated for calculating the output spectrum. This was conducted with the experimental setup depicted in Fig. 8, which depicts the inclusion of a 100-mm diameter integrating sphere (IS). An Ocean Optics HR4000 spectrometer, connected to the sphere via a 200-μm diameter fiber, was used to measure spectra for independent validation. The sphere was illuminated by either a white light tungsten halogen lamp or a monochromator. Optionally, the white light source could be filtered with removable gelatin filters. H-matrix characterization was accomplished using the monochromator. The monochromator was incremented from 420 to 720 nm in N=19 increments with a uniform wave number spacing of Δσ=551cm1. This spectral resolution corresponds closely to the spectral resolution of the WP.

Fig. 8

(a) Schematic of the experimental setup for calibrating the PSHI and (b) photo of the system on the benchtop.

OE_53_4_044104_f008.png

The measured H-matrix is depicted in Fig. 9. Notable is the presence of both the Id and I0 interference components that correspond to the high and low spatial frequency modulations versus y, respectively. As observed in the simulations, I0 has increasing contrast for values of ϕPG away from 180 deg per Fig. 6(b). Calculating the pseudo-inverse of H enables the input spectrum to be calculated using

Eq. (52)

f=W·g,
where W is the data reduction matrix, calculated from the pseudo-inverse functionality in MATLAB. Validation of the H-matrix calibration procedure was performed by illuminating the sphere with the white light tungsten halogen lamp. Spatially heterodyned interferograms were measured using the experimental setup, in addition to directly measured spectra that were acquired using the Ocean Optics HR4000 spectrometer. A white-light reference interferogram and spectrum were acquired first, followed by interferograms and spectra of four absorption gelatin filters (Roscolux brand, labeled F1 through F4) that were inserted between the lamp and integrating sphere. Spectra were then calculated, from the measured interferograms, by applying W, and the transmission of each filter was calculated by dividing each of the filter’s spectra by the white light reference spectrum. The transmission of each filter was also calculated from the HR4000 data for direct comparison. These results are depicted in Fig. 10 in which the ocean optics (OO) transmission spectra are plotted alongside the transmission spectra calculated from the PSHI.

Fig. 9

Measured H-matrix.

OE_53_4_044104_f009.png

Fig. 10

Measured transmission data comparison from the PSHI and the OO spectrometer for filters F1 to F4.

OE_53_4_044104_f010.png

The RMS error, averaged for all four transmission measurements, was calculated between the PSHI and OO data to be approximately 1.2% for these data.

6.

Conclusion

In this paper, we successfully demonstrated a theoretical model and experimental calibration procedure for a PSHI. The interferometer was based on a WP that was heterodyned using a PG. As demonstrated by the model and subsequent experiments, nonideal frequency components are caused by the PG’s zero-order light leakage and error in the QWP’s retardance. Fortunately, low zero-order diffraction efficiencies (<3%) in the PGs are achievable using broadband multilayer achromatic PG designs.22 Additionally, as demonstrated in the experiment, AQWPs can be leveraged to further reduce the undesirable frequency components. Finally, the H-matrix calibration procedure was implemented to calibrate the PSHI. This yielded an experimentally observed average RMS error of 1.2% when compared to an Ocean Optics spectrometer.

Acknowledgments

We gratefully acknowledge support from the National Science Foundation (NSF Grant No. ECCS-0955127).

References

1. 

P. GriffithsJ. A. D. Haseth, Fourier Transform Infrared Spectrometry, John Wiley & Sons(2007). Google Scholar

2. 

R. G. SellarG. D. BoremanL. E. Kirkland, “Comparison of signal collection abilities of different classes of imaging spectrometers,” Proc. SPIE, 4816 389 –396 (2002). http://dx.doi.org/10.1117/12.451649 PSISDG 0277-786X Google Scholar

3. 

Optical Shop Testing, 3rd ed.Wiley-Interscience, Hoboken, NJ (2007). Google Scholar

4. 

J. HarlanderR. J. ReynoldsF. L. Roesler, “Spatial heterodyne spectroscopy for the exploration of diffuse interstellar emission lines at far-ultraviolet wavelengths,” Astrophys. J., 396 (2), 730 –740 (1992). http://dx.doi.org/10.1086/171756 ASJOAB 0004-637X Google Scholar

5. 

G. G. Shepherdet al., “WINDII observations of the 558 nm emission in the lower thermosphere: the influence of dynamics on composition,” J. Atmos. Sol.-Terr. Phys., 59 (6), 655 –667 (1997). http://dx.doi.org/10.1016/S1364-6826(96)00142-3 JASPF3 1364-6826 Google Scholar

6. 

G. Shepherdet al., “Windii, the wind imaging interferometer on the upper-atmosphere research satellite,” J. Geophys. Res. Atmos., 98 (D6), 10725 –10750 (1993). http://dx.doi.org/10.1029/93JD00227 JGRDE3 0148-0227 Google Scholar

7. 

J. M. Harlanderet al., “Design and laboratory tests of a Doppler Asymmetric Spatial Heterodyne (DASH) interferometer for upper atmospheric wind and temperature observations,” Opt. Express, 18 (25), 26430 –26440 (2010). http://dx.doi.org/10.1364/OE.18.026430 OPEXFF 1094-4087 Google Scholar

8. 

C. R. EnglertD. D. BabcockJ. M. Harlander, “Doppler asymmetric spatial heterodyne spectroscopy (DASH): concept and experimental demonstration,” Appl. Opt., 46 (29), 7297 –7307 (2007). http://dx.doi.org/10.1364/AO.46.007297 APOPAI 0003-6935 Google Scholar

9. 

D. A. Nayloret al., “Astronomical spectroscopy using an aliased step-and-integrate Fourier transform spectrometer,” Proc. SPIE, 5498 685 –694 (2004). http://dx.doi.org/10.1117/12.552054 Google Scholar

10. 

D. Komisareket al., “High-performance nonscanning Fourier-transform spectrometer that uses a Wollaston prism array,” Appl. Opt., 43 (20), 3983 –3988 (2004). http://dx.doi.org/10.1364/AO.43.003983 APOPAI 0003-6935 Google Scholar

11. 

N. R. Gomeret al., “Raman spectroscopy using a spatial heterodyne spectrometer: proof of concept,” Appl. Spectrosc., 65 (8), 849 –857 (2011). http://dx.doi.org/10.1366/11-06298 APSPA4 0003-7028 Google Scholar

12. 

MeigsA. D., “Common path interferometer for spectral image generation,” U.S. Patent No. 6687007 B1 (2004).

13. 

M. W. Kudenovet al., “White light Sagnac interferometer for snapshot linear polarimetric imaging,” Opt. Express, 17 (25), 22520 –22534 (2009). http://dx.doi.org/10.1364/OE.17.022520 OPEXFF 1094-4087 Google Scholar

14. 

A. V. Velascoet al., “Optical fiber interferometer array for scanless Fourier-transform spectroscopy,” Opt. Lett., 38 (13), 2262 –2264 (2013). http://dx.doi.org/10.1364/OL.38.002262 OPLEDP 0146-9592 Google Scholar

15. 

M. W. Kudenovet al., “Spatial heterodyne interferometry with polarization gratings,” Opt. Lett., 37 (21), 4413 –4415 (2012). http://dx.doi.org/10.1364/OL.37.004413 OPLEDP 0146-9592 Google Scholar

16. 

M. W. Kudenovet al., “Compact spatial heterodyne interferometer using polarization gratings,” Proc. SPIE, 8873 88730Q (2013). http://dx.doi.org/10.1117/12.2024104 PSISDG 0277-786X Google Scholar

17. 

M. W. KudenovE. L. Dereniak, “Compact real-time birefringent imaging spectrometer,” Opt. Express, 20 (16), 17973 –17986 (2012). http://dx.doi.org/10.1364/OE.20.017973 OPEXFF 1094-4087 Google Scholar

18. 

M. J. PadgettA. R. Harvey, “A static Fourier transform spectrometer based on Wollaston prisms,” Rev. Sci. Instrum., 66 (4), 2807 –2811 (1995). http://dx.doi.org/10.1063/1.1145559 RSINAK 0034-6748 Google Scholar

19. 

R. C. Jones, “New calculus for the treatment of optical systems. VIII. Electromagnetic theory,” J. Opt. Soc. Am., 46 (2), 126 –131 (1956). http://dx.doi.org/10.1364/JOSA.46.000126 JOSAAH 0030-3941 Google Scholar

20. 

G. Cincotti, “Polarization gratings: design and applications,” IEEE J. Quantum Electron., 39 (12), 1645 –1652 (2003). http://dx.doi.org/10.1109/JQE.2003.819526 IEJQA7 0018-9197 Google Scholar

21. 

M. J. Escutiet al., “Simplified spectropolarimetry using reactive mesogen polarization gratings,” Proc. SPIE, 6302 630207 (2006). http://dx.doi.org/10.1117/12.681447 OPPHEL 1047-6938 Google Scholar

22. 

C. OhM. J. Escuti, “Achromatic diffraction from polarization gratings with high efficiency,” Opt. Lett., 33 (20), 2287 –2289 (2008). http://dx.doi.org/10.1364/OL.33.002287 OPLEDP 0146-9592 Google Scholar

23. 

D. H. Goldstein, Polarized Light, CRC Press, Boca Raton, Florida (2011). Google Scholar

Biography

Michael W. Kudenov received his BS degree in electrical and computer engineering from the University of Alaska Fairbanks, Fairbanks, AK, in 2005 and his PhD degree in optical sciences from the University of Arizona, Tucson, AZ, in 2009. He is an assistant professor of ECE at North Carolina State University in Raleigh, NC. His lab researches compact high-speed hyperspectral, polarimetric, and interferometric sensors and sensing systems within multidisciplinary applications spanning remote sensing, defense, process monitoring, and biological imaging. He is a member of SPIE.

Matthew N. Miskiewicz received BS degrees in electrical engineering and computer engineering from North Carolina State University (NCSU) in 2006. He has since been a member of the Opto-Electronics and Lightwave Engineering Group, NCSU, working toward a PhD degree. His research focuses on FDTD methods, polarization holography, computer generated holograms, beam-combining, beam-shaping, beam-steering, and novel direct-write systems.

Michael J. Escuti received his PhD degree in electrical engineering from Brown University, Providence, RI, in 2002. Currently, he is an associate professor of electrical and computer engineering at North Carolina State University (NCSU) at Raleigh, where he pursues interdisciplinary research in photonics, opto-electronics, flat-panel displays, diffractive optics, remote sensing, and beyond. He is a named inventor on 17 issued and 12 pending patents, has published more than 93 refereed journal and conference publications, presented 24 invited talks, and has coauthored one book chapter. He is a member of SPIE.

James F. Coward is cofounder and president/CEO of SA Photonics, where he leads development of photonic sensing and communication systems. He has been the system architect/principal engineer on hardware fielded in many market segments, including mission critical sensor hardware flying on over 20 satellites, critical cockpit display hardware on every F18 E/F in the world, mission critical hardware in the terrestrial telecom, and semiconductor inspection industries.

© 2014 Society of Photo-Optical Instrumentation Engineers (SPIE) 0091-3286/2014/$25.00 © 2014 SPIE
Michael W. Kudenov, Matthew N. Miskiewicz, Michael J. Escuti, and James F. Coward "Polarization spatial heterodyne interferometer: model and calibration," Optical Engineering 53(4), 044104 (8 April 2014). https://doi.org/10.1117/1.OE.53.4.044104
Published: 8 April 2014
Lens.org Logo
CITATIONS
Cited by 5 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Calibration

Polarization

Heterodyning

Interferometers

Spatial frequencies

Fourier transforms

Spectral resolution


CHORUS Article. This article was made freely available starting 08 April 2015

Back to Top