Open Access
8 March 2019 Optical phantoms for biomedical polarimetry: a review
Joseph Chue-Sang, Mariacarla Gonzalez, Angie Pierre, Megan Laughrey, Ilyas Saytashev, Tatiana Novikova, Jessica C. Ramella-Roman
Author Affiliations +
Abstract
Calibration, quantification, and standardization of the polarimetric instrumentation, as well as interpretation and understanding of the obtained data, require the development and use of well-calibrated phantoms and standards. We reviewed the status of tissue phantoms for a variety of applications in polarimetry; more than 500 papers are considered. We divided the phantoms into five groups according to their origin (biological/nonbiological) and fundamental polarimetric properties of retardation, depolarization, and diattenuation. We found that, while biological media are generally depolarizing, retarding, and diattenuating, only one of all the phantoms reviewed incorporated all these properties, and few considered at least combined retardation and depolarization. Samples derived from biological tissue, such as tendon and muscle, remain extremely popular to quickly ascertain a polarimetric system, but do not provide quantifiable results aside from relative direction of their principal optical axis. Microspheres suspensions are the most utilized phantoms for depolarization, and combined with theoretical models can offer true quantification of depolarization or degree of polarization. There is a real paucity of birefringent phantoms despite the retardance being one of the most interesting parameters measurable with polarization techniques. Therefore, future work should be directed at generating truly reliable and repeatable phantoms for this metric determination. Diattenuating phantoms are rare and application-specific. Given that diattenuation is considered to be low in most biological tissues, the lack of such phantoms is seen as less problematic. The heterogeneity of the phantoms reviewed points to a critical need for standardization in this field. Ultimately, all research groups involved in polarimetric studies and instruments development would benefit from sharing a limited set of standardized polarimetric phantoms, as is done earlier in the round robin investigations in ellipsometry.

1.

Introduction

The use of polarized light in clinical and preclinical applications is expanding,1 and several recent reviews by Tuchin,2 Ghosh and Vitkin,3 Qi and Elson,4 de Boer et al.,5 and Baumann6 have illustrated the fast progress of this approach in the medical field.

As polarimetric techniques reach the clinical and commercial stage, there is a need to validate them with replicative systems that could serve as biological proxies and mimic the characteristic trends of typical biological observations. Over the past several decades, a variety of such systems—commonly referred to as phantoms—have been implemented for the use of general optical imaging and sensing; Pogue and Patterson7 illustrated these tools in an exhaustive review. Here, we focus uniquely on phantoms used for polarimetry in biomedicine; these phantoms were not included in previous reviews and are relevant for scientists and engineers working on polarimetric applications.

Three dominant mechanisms influence polarized light as it travels through a biological media: depolarization, retardation, and diattenuation.

Scattering is a primary contributor to the process of depolarization. Loss of polarization is mainly due to the disarrayed changes of amplitude and phase of the scattered electromagnetic field reaching a detector.8

Scattering is generally very high in biological media due to the high density and large variety of sub- and extracellular components (such as organelles, nuclei, collagen fiber bundles, cell membrane, to name a few). Different polarization states of incident radiation—linear, circular, or elliptical—depolarize at different rates. As for the mathematical representation of depolarization, its theoretical premise is generally supported by the Mueller matrix of an intrinsic (or diagonal) depolarizer [Eq. (1a)], satisfying the covariance conditions [Eq. (1b)]:9

Eq. (1a)

MΔ=d0(10000a0000b0000c),0<d0<1,|a|,|b|,|c|1,

Eq. (1b)

abc1,a+b+c1,ab+c1,a+bc1.

It follows from Eq. (1a) that 1|a| and 1|b| represent the linear depolarization power (horizontal-vertical and ±45  deg frameworks). Similarly, 1|c| specifies the power of circular depolarization.

From this, the total depolarization power Δ can be calculated as

Eq. (2)

Δ=1|a|+|b|+|c|3=1|tr(MΔ)1|3,0Δ1.

In birefringent media, light experiences changes in propagation speeds for its different polarization components, which leads to phase differences (also called retardation) between those components. Linear retardation is the phase shift between two orthogonal linear polarization states (e.g., 0 deg and 90 deg, or +45  deg and 45  deg). Circular retardation (also referred to as optical rotation) is the difference in phase between the right and the left circular polarized components of light, which happens due to circular birefringence (optical activity). The Mueller matrix of a linear retarder [see Eq. (3)] depends on its phase difference parameter δ and on the azimuth θ of its fast axis:

Eq. (3)

R=(10000cos2(2θ)+sin2(2θ)cosδsin2θcos2θ(1cosδ)sin2θsinδ0sin2θcos2θ(1cosδ)sin2(2θ)+cos2(2θ)cosδcos2θsinδ0sin2θsinδcos2θsinδcosδ)

The retardation (δ) property of a uniaxial medium is frequently expressed through its birefringence and can be written as shown in Eq. (4), where ne and no are extraordinary and ordinary refractive indices of a birefringent material, respectively, d is the distance traveled by light (wavelength λ0) through the birefringent medium, Δn=n(θ)no, where n(θ) is the refractive index seen by the photon propagating in the direction u(ux,uy,uz). Angle θ is the angle between the direction u and extraordinary axis of birefringent material defined as e(cosη,sinη,0):

Eq. (4)

δ=2πdΔnλ,

Eq. (5)

n(θ)=none(ne2cos2θ+no2sin2θ)0.5,

Eq. (6)

θ=cos1[uxcosη+uysinη(ux2+uy2+uz2)0.5].

Birefringence itself can be divided into “intrinsic birefringence” and “form birefringence.”2 Typically, biological tissues rich in extracellular matrix (ECM) fibers, for example, skin, cornea, sclera, tendon, uterine cervix, and cardiac tissue, exhibit retardation.

Mueller matrix polarimetry24 and polarization-sensitive optical coherence tomography (PS-OCT)10 are techniques capable of quantifying many of the aforementioned parameters of interest. Calculation of the Mueller matrix requires the modulation of both light source and detector into a minimum of four different polarization states for a total of 16 measurements. Once the Mueller matrix is determined, it can be decomposed11 as a sequence of elementary polarization components: a diattenuator, a retarder, and a depolarizer. PS-OCT is an extension of OCT, a technique based on low-coherence interferometry that can provide high-resolution cross-sectional imaging of biological tissue, and it too can be used to quantify birefringence, diattenuation, and depolarization index, a parameter related to depolarization.

Diattenuation, also called dichroism, is generally considered to have the smallest impact on polarized light propagating in biological media. Diattenuation arises from polarization-selective attenuation of the electrical field. Related to diattenuation is the property of optical activity, also known as circular birefringence, which is characterized by the rotation of the polarization plane of linearly polarized light about the axis of propagation.2 This property is prevalent for chiral molecules such as glucose, proteins, and nucleic acids.3

The use of polarimetry in monitoring biological tissue often focuses on quantification of the tissue preferential azimuth (i.e., the orientation of optical axis of uniaxial birefringent medium) related to the arrangement of a collagenous ECM or other cellular assembly. Skeletal muscle and cardiac tissue are both strongly depolarizing and birefringent due to cellular components and layered structure.

Collagen, animal cornea, retina, and optic nerves have all been shown to have large birefringence and preferential alignment through PS-OCT and polarized light microscopy.1214 Several studies using PS-OCT imaging on articular cartilage, which is rich in oriented collagen fibers, have shown changes in collagen retardation in depth.1518 Nerves have also been shown to yield retardation with polarization-sensitive spectroscopy.19 Since birefringence is the most common source of retardation and signal for this modality, in general most retardance phantoms can be used as PS-OCT phantoms.

Microtubules made from extracted elements of the porcine brain and axonemes prepared from sea urchin have been examined using polarized light microscopy, where fibers can be visualized.20 The ECM of the cervix is composed of about 70% collagen and, therefore, has shown to have a significant retardation. Chue-Sang et al.21 used Mueller matrix polarimetry to calculate retardance, depolarization, and collagen fiber azimuth of ex vivo porcine cervix samples (seen in Fig. 1). Pierangelo et al. used wide-field multiwavelength Mueller matrix polarimeters to image cervical neoplasia and colon cancer.2230

Fig. 1

Mueller matrix-derived parameters (Lu–Chipman decomposition11) of an ex vivo porcine cervix: (a) CCD image, (b) depolarization power, (c) scalar retardation, and (d) azimuth of optical axis. Darkened area in the center of the image is the cervical OS. Overlaid lines are calculated from the mean of subsections of the azimuth depicted by the false color.21 Mueller matrix-derived parameters (Lu–Chipman decomposition) of healthy human cervical specimen: (e) CCD image, (f) depolarization power, (g) scalar retardation, and (h) azimuth of optical axis.22

JBO_24_3_030901_f001.png

Vitkin et al. used Mueller polarimetry to determine the local structural disorders of the bladder31 and myocardium.32 Enhancement of superficial structure by eliminating deep penetrating scattered photons is also a common use of polarimetry in medicine. Groner et al.33 used cross-polarization to highlight superficial vascular contrast in intravital microscopy, applying this technique, among others, to study brain perfusion and pancreatic and hepatic microcirculation.3338

Polarized light imaging has been used extensively to enhance surface contrast for dermatologic applications.39 Demarcation of margins of skin cancers, not visible to the naked eye, has been conducted by several researchers, starting with setups focusing on linear depolarization to other systems,4043 utilizing full Stokes vector polarimetry and out-of-plane approaches.4446 The skin stratum corneum has been shown to be highly scattering hence producing strong depolarization.47,48 Changes in retardation have been associated with the presence of collagen in the dermis. For this reason, scars have a strong response to polarized light as collagen in wounds recombines in the direction of local forces.46

2.

Optical Phantoms

We have categorized all phantoms by their dominant polarization property—namely, depolarization, retardation, diattenuation, or optical activity. We have also introduced a separate table for biological tissues used as phantoms. Many phantoms exhibit more than one property; hence, they may appear in more than one table, these repeated phantoms are identified by an asterisk (*). The retardation phantoms table includes an induced retardation column. This column is included to differentiate phantoms which are inherently birefringent due to their structure from phantoms that are mechanically stressed, strained, or otherwise manipulated in order to change their birefringence. Many of the phantoms cited in this review have been used by the same investigators in multiple journals, for simplicity, we have not cited all the articles using the same phantoms and limited the review to the ones that were substantially different to each other.

2.1.

Biological Phantoms

The construction of polarimetric phantoms is a complex process; hence, biological samples are commonly used in polarization-sensitive optical modalities (Table 1). Collagen-rich tissues, for example, tendons or rat tails, are the most commonly used in polarimetry. As most biological tissues, collagen scatters (and, consequently, depolarizes); more importantly, collagen introduces a phase shift between orthogonal polarization states of incident polarized light2 due to its strong birefringence. Since many healthy collagen-rich tissues behave as uniaxial birefringent media, the azimuth of optical axis of linear retardation related to collagen alignment can often be measured.21,22,27,51,59,77,78

Table 1

Biological tissues used as polarization phantoms.

Tissue typePreparationPolarization propertyTransmission/reflectanceRef.
Axonemes (sea urchin)Extraction from sea urchin sperm and purification stepsRetardationR20
Bladder (porcine)Excised, freshDepolarization, retardation, diattenuationR49
Brain (porcine)Phosphate-buffered saline solution (0.02 M)DepolarizationR50
Cartilage (animal)Excised, freshDepolarization, retardationR1518
Cartilage (porcine)Excised, freshRetardation, depolarization, diattenuationT18
Cervix (porcine)Fixed in 4% paraformaldehyde and embedded in paraffinDepolarization, retardationR51
Eye (cornea)Excised, freshRetardationR13 and 14
Eye (optic nerve)CryosectionedRetardationR12
Eye (retina)Excised, freshRetardationR14
Fibroblast (rat)SuspensionDepolarizationR52 and 53
Heart (myocardium)Excised, fixedDepolarization, retardationR32, 54, 55, and 56
Heart (porcine myocardium)Phosphate-buffered saline solution (0.02 M)DepolarizationR50
Heart (rat myocardium)10% formalin and cut into 1 mm slicesRetardation, diattenuation, depolarizationR57
Heart (valve leaflet)Excised, freshDepolarization, retardationR58 and 59
Heart (porcine valve)Excised, freshRetardationR60 and 58
Heart (porcine aorta)Excised, freshRetardationR61
Heart (bovine right ventricle)Cut into 2  cm×2  cm×1  cm sectionsRetardation, diattenuationR62
Heart (swine right ventricle)Excised, freshRetardationR63
Heart (rabbit right ventricular wall)3.7% formaldehyde for 1 day and 20% sucrose solution for an additional 2 daysRetardationR64
Kidney cortexPhosphate-buffered saline solution (0.02 M)DepolarizationR50
LiverPhosphate-buffered saline solution (0.02 M)DepolarizationR40 and 50
Melanin granulesSuspensionDepolarization, retardationR65
MicrotubulesExtraction from porcine brain and purification stepsRetardationR20
Nerve (lobster leg)Excised, freshDepolarization, retardationR19
Skeletal muscleExcised, freshDepolarization, retardationR32, 40, 50, 5456, 64, and 6668
SkinIn vivoDepolarization, retardationR47, 48, and 69
Skin (calf)Excised, freshRetardationT70
Skin equivalent modelFixed and cut into few μm slicesDepolarization, retardationT71
Tail (rat)Frozen and thawedDepolarization, retardationR72
TendonExcised, freshDepolarization, retardationR45, 50, 60, and 7376
Yeast cellsSuspensionDepolarizationR53

Chicken or cow tendons have been used by many groups45,50,60,7376 to validate polarization-based optical instruments. Azimuth angle is calculated45,50,58,60,7276,7983 as well as an increase in scalar retardance due to birefringence. Similar to tendon, murine tails also contain collagen fibers which are strongly aligned. Since the azimuth of the collagen fibers preferential orientation can be directly observed, a typical validation test for polarimeters includes positioning a tendon or rat tail at predetermined angles and then measuring samples at different and well-known angular positions.21,59,72

While muscle tissue can be used for the same purposes as collagen-based phantoms, the interpretation of the results is less straightforward due to the increased cellularity of these tissues.61 Studies of myocardium muscle32,40,50,5456,64,6668 have been conducted by several investigators showing loss of retardation and local order for infarcted tissue. For this reason, samples of myocardium have been used to validate different polarimetric systems. Ghosh et al.57 used Mueller matrix decomposition to calculate depolarization, diattenuation, and retardance of fixed rat myocardial tissue.

Heart valve leaflets are another highly collagenous and anisotropic tissue that have been used as a depolarization and retardation phantom.60 As in previous example, the azimuth of collagen fibers’ preferential orientation can be detected and used for instrument characterization. Changes in depolarization can also be observed by treating the sample with collagenase.58,59

Artificial skin models grown from epidermal keratinocytes forming a multilayered epidermis on top of collagen I hydrogel with dermal fibroblasts have also been used to mimic the interaction of polarized light with the skin.71 Unstained cuts of fixed skin equivalents of varying thickness (range: 5 to 30  μm) were measured in transmission with Mueller microscopy and the values of retardation and depolarization parameters were extracted using logarithmic decomposition84 of the measured Mueller matrices. The measurements confirmed parabolic dependence of depolarization and linear dependence of retardation on thickness, as follows from differential Mueller matrix formalism.

2.2.

Depolarizing Phantoms

Several authors have studied the effect of particle size, density, and index of refraction on the polarization of scattered light.85,86 As suggested by the results of these studies, the main scatterers in biological tissues are nuclei, organelles, and bulk tissue structures that limit the photon penetration depth and depolarize light traveling through these media.53 The cell nuclei and organelles are frequently modeled as spherical scattering particles87 of refractive index varying between 1.33 and 1.47. The components of ECM, such as collagen and elastin, have been represented by spherical88 or cylindrical25 structures.

Work by MacKintosh et al.89 showed that circular polarization was maintained for longer depths as compared to linearly polarized light in Mie scattering regime (scatterer size light wavelength in the medium). In one of the relevant studies, Monte Carlo simulations supported this finding by showing that mean penetration depth was 2 mean free paths (MFP) for linearly and 10 MFP for circularly polarized light in Mie scattering regime.86

Suspensions of microspheres and other small particles are commonly used to create phantoms with scattering properties (Table 2). The amount of scattering can be adjusted depending on the size and concentration of the microspheres based on the Mie scattering theory. On a smaller scale, nanoparticles have also been widely used to create scattering phantoms in Rayleigh scattering regime. These particles can also be embedded in solid host media, such as gels or polymers, to ensure scattering properties of those materials. In addition, India ink, hemoglobin, and dyes are commonly added to influence the absorbing characteristics.

Table 2

Depolarizing phantoms. *Phantoms that were also tested for other polarization properties in corresponding reference paper.

Depolarizing agentEmbedding materialTissue mimickingPhantom thicknessTransmission/reflectanceRef.
GNPs (50 nm)IntralipidContrast agentSemi-infiniteR90
Intralipid*Water, India inkBladder wallSemi-infiniteR49
IntralipidWaterTurbid biological mediaSemi-infiniteR53, 69, and 91
Intralipid or polystyrene microspheresWater, Naphthol greenPorcine liver1  μm, 1.4  μmR92
Kapton tape (stacked)*Layered against a rigid baseTheoretical standardSemi-infiniteR93
Mylar (biaxially oriented polyethylene terephthalate)*Laid against a plexiglass baseTheoretical standardSemi-infiniteR93
Polystyrene microspheresWaterTurbid biological mediaSemi-infiniteR24, 40, 53, and 9497
Polystyrene microspheresIntralipidTurbid biological mediaSemi-infiniteR53 and 2
Polystyrene microspheresPolyacrylamide, sucroseTurbid biological media1  cm3T3
Polystyrene microspheres (0.5  μm) and fiber glass*PolyacrylamideAnisotropic sample1×2×4  cm3T98
Polystyrene microspheres and silk fibers*WaterAnisotropic sample2.1 cmR88 and 99
Quartz plate (wedged)*NoneN/A3 mmT100
Melanin granules*WaterRetina/retinal pigment epitheliumSemi-infiniteR65
Silicon phantom (extruded)Air between layersAnisotropic sample2 mmR51
Silicon (amorphous)*NoneTheoretical polarization standardSemi-infiniteR93
Silicon (poly-)*NoneTheoretical standardSemi-infiniteR93
Silicon gratingSilicon waferTheoretical standardSemi-infiniteR101
TiO2 nanoparticles (530 nm)PVC-based transparent materialBiopsy samples1 mmT23
TiO2WaxSkin2 and 5 mmR44
ZnO nanoparticles (340 nm)PVCP stock solutionHuman skin0.2 –to 2 mmT102 and 103

Several studies, such as Antonelli, Rakovic et al., and Cote and Vitkin,75,94,95 have used aqueous polystyrene microsphere suspensions as backscattering polarization phantoms. In order to measure the change in scattering (i.e., depolarization power) calculated for different suspensions, microsphere diameter was varied.24,40,53,9496 This class of phantoms has also been shown to depolarize linear polarization less with smaller-diameter microspheres as compared to circular polarization, while, with an increase of the microsphere diameter, circular polarization has been reported to be better preserved as compared to linear polarization.96

While purely aqueous monodispersed suspensions of microspheres are most commonly used in scattering experiments, intralipid has also been used to create depolarizing phantoms.2,53 Intralipid is commonly used as a nutrition supplement and is an emulsion of fatty micelles; therefore, scattering is due to multidispersed spherical structures. Aqueous intralipid suspensions with different dilution factors starting at 1:500 to 1:1 have been used to test depolarization with reflectance polarimetry.53,69,91 An example of such experiment can be seen in Fig. 2, where loss of elliptical polarization is measured as a function of depth in an intralipid suspension as reported by Sridhar and Da Silva.69 While intralipid suspension exhibits monotonic dependence of depolarization on light wavelength, the use of gold nanoparticles (GNPs) suspended in intralipid creates more complicated depolarization behavior.90

Fig. 2

Image from Ref. 69. (a) Ruler placed obliquely in a tank containing Intralipid® solution, (b) elliptical channel image at 45 deg after subtraction method 1, (c) elliptical channel image at 45 deg after subtraction method 2. (b) and (c) have a common colorbar represented at the right edge of the figure. Yellow-dotted line represents the Intralipid®–air interface. Each graduation on the ruler (i.e., 1 mm) corresponds to 0.35 mm in actual depth. Wavelength: 633 nm. Text is from Ref. 69.

JBO_24_3_030901_f002.png

Titanium dioxide (TiO2) is another material commonly used to produce scattering in optical phantoms. TiO2 particles have been used in solid host media, such as polydimethylsiloxane or polyurethane, where, before the curing process, these particles are mixed into the polymer. Adjusting the concentration of TiO2 particles makes it possible to change the amount of depolarization.7,23 Zinc oxide (ZnO) particles are also commonly mixed into polymers.102,103 Melanin suspensions of rising concentrations can be used to test depolarization with PS-OCT and model the same phenomenon in the retinal pigmented epithelium. As demonstrated by Baumann et al.,65 the change in depolarization based on melanin concentration has a linear relationship with degree of polarization uniformity (DOPU).

2.3.

Retarding Phantoms

Polymer-based materials are a common source of retardation. Due to their molecular structure or preparation process, many polymers possess intrinsic birefringence (i.e., behave as uniaxial crystals).104 Others can be induced to become birefringent by applying mechanical stress to the material.3,105 Many of these polymers are transparent; hence, scattering particles such as microspheres can be added to better simulate biological media. Electrospun polymer fibers, fabricated by charging droplets of polymer at high voltages which creates an interconnected network of small fibers,106 were used by Goth et al.60 to determine the degree of anisotropy of the overall structure. The anisotropic biological elements in the ECM (particularly collagen and elastin) have been simulated with several materials, including silk88,99 and glass fibers.98,107 An example of fibrous phantom is shown in Fig. 3. Here, the phantom is composed of polystyrene microspheres and well-aligned glass fibers embedded in polyacrylamide (glass fibers have a 10-μm diameter and 1.547 refractive index).

Fig. 3

Image from Ref. 107. (a) Cylinder model, (b) sphere-cylinder model, and (c) sphere-cylinder birefringence model.

JBO_24_3_030901_f003.png

Phantoms for PS-OCT require a strong backscattering to generate a high image contrast and have ideally well-defined layers with homogeneous yet different values of birefringence (Table 3). Accordingly, Liu et al. have used a phantom consisting of a long birefringent polymer band laid over four smaller bands of differing birefringence. The optical axes of bottom four bands were oriented at 45 deg with the optical axis of top layer allowing for a depth-dependent change in retardation.108 An example of this retarding phantom is shown in Fig. 4.

Table 3

Retardation phantoms. The “induced retardation” column is for differentiating between phantoms which inherently exhibit their birefringence due to their structure and phantoms that are mechanically stressed, strained, or otherwise manipulated in order to change their birefringence. *Phantoms that were also tested for other polarization properties in corresponding reference paper.

Retardation materialEmbedded materialInduced retardationTissue mimickingPhantom thicknessTransmission/reflectanceRef.
Birefringent filmIntralipid, India inkStructureECMSemi-infiniteR49
Electrospun fibers (0.6 to 1.0  μm)NoneStructureHeart valve leafletSemi-infiniteR60
Human hairNoneStructureHuman hairN/AR15
Kapton tape (stacked)Layered against a rigid baseStructure (layers)Theoretical standardSemi-infiniteR93
Mylar (biaxially oriented polyethylene terephthalate)Laid against a plexiglass baseStructureTheoretical standardSemi-infiniteR93
Plastic cap*NoneStructureTheoretical standardSemi-infiniteR76
PolycarbonateNoneLongitudinal stretch (heating and cooling)Turbid biological tissue250  μmR108
Polyacrylamide polymer (elastic)None4 mm stretchTurbid biological tissue4 mmR109
Polyacrylamide gelsPolystyrene microspheres, 1 M sucroseStretchingTurbid biological tissue1×1×4  cm3T105
Polyacrylamide*Sucrose, polystyrene microspheresStretchingTurbid biological tissue1×1×1  cm3T3
Polyacrylamide*Polystyrene microspheres and well-aligned fiber glassStretching (1 to 5 mm), birefringence = 0 to 105Turbid biological tissue1×2×4  cm3T98 and 107
Polyethylene (low density)NoneBending (up to 2.5 MPa)Turbid biological tissue1 mmR110
Polystyrene sphereNoneStructureInfarcted myocardium8 mm diameterT111
Polystyrene microspheresWaterStructureTurbid biological mediaSemi-infiniteR75, 97, 112
PolyurethaneParticle filled polypropyleneLongitudinal stretchTheoretical standard1 mmR113
Silicon (extruded)Air between layersStructureTheoretical standard2 mmR51
Silicon (amorphous)NoneStructureTheoretical standardSemi-infiniteR93
Silicon (poly-)NoneStructureTheoretical standardSemi-infiniteR93
Silk fibers*WaterStructureAnisotropic sampleSemi-infiniteR88 and 99

Fig. 4

Image from Ref. 108. (a)–(c) Intensity, birefringence, and DOP images of the slab and (d and e) cylindrical phantoms. (a) Representative cross-sectional images of the birefringence phantom for galvanometer-scanning system. (b) and (c) En-face images at different depths as indicated by the dashed red lines in (a). Horizontal and vertical scale bars for (a)–(c): 2 mm and 250  μm, respectively. (d) Representative images obtained from one rotational scan with the catheter. Scale bar: 1 mm. (e) Longitudinal sections obtained from a pull-back data set, with its corresponding location indicated by the dashed red line in (d). Radial and horizontal scale bars: 250  μm and 1 mm, respectively. (Text from Ref. 108.)

JBO_24_3_030901_f004.png

Ghosh et al. induced changes in retardation by stretching a polyacrylamide phantom. Moreover, changing birefringence, and mixing polystyrene microspheres and sucrose into the polymer, produced phantoms that could be used to characterize retardance, depolarization, and diattenuation.3,98,109 Extruded silicon, silicon wafers with gratings, and other types of silicon (poly and amorphous), as well as different tapes (e.g., Kapton and Mylar) normally used in solar panels, have been used to create phantoms containing different combinations of diattenuation, depolarization, and retardation properties.51,93,101 Figure 5 shows an example of an experimental setup used to induce birefringence in a polymer through mechanical strain by Wood et al.114

Fig. 5

Image from Ref. 114. Apparatus to create birefringent phantoms.

JBO_24_3_030901_f005.png

In order to account for different geometries and extract geometry-independent metrics of anisotropy, retardance measurements have been taken using an 8-mm-diameter polystyrene sphere of known anisotropy axis azimuth.111 Fan et al.76 imaged a plastic cap to determine its retardation with PS-OCT.

2.4.

Diattenuating Phantoms

The asymmetry of a molecule can result in selective transmission of an incident state of polarized light. Swami et al.115 measured diattenuation as a parameter to identify the general shape of GNPs (Table 4). Differently shaped GNPs displayed different spectroscopic diattenuation values. Chen et al.116 and Lung et al.117 used a quarter-wave plate and a polarizer to test the performance of an analytical model for low diattenuating optical components as they were rotated from 0 deg to 150 deg with a step of 30 deg. Moreover, these authors also used a polymer polarizer baked at 150°C as a sample with both diattenuating and birefringent properties. Chenault and Chipman118 used a rotating sample polarimeter to find linear diattenuation and retardance of the sample calculated from intensity modulation.

Table 4

Diattenuation phantoms. *Phantoms that were also tested for other polarization properties in corresponding reference paper.

Diattenuation agentSolvent/preparationTissue mimickingPhantom thicknessTransmission/reflectanceRef.
GNPs (nonspherical shapes)CTAB-coated GNPsTheoretical standardSemi-infiniteT115
Kapton tape (stacked)*Layered against a rigid baseTheoretical standardSemi-infiniteR93
Mylar (biaxially oriented polyethylene terephthalate)*Laid against a plexiglass baseTheoretical standardSemi-infiniteR93
PolarizerNoneTheoretical standard21.59 mmT116 and 117
Polarizer (baked)150°C for 80 minTheoretical standardN/AT116 and 117
Polarizer (rotating)NoneTheoretical standardN/AT118
Quarter-wave plateNoneTheoretical standardN/AT116 and 117
Silicon (amorphous)*NoneTheoretical standardSemi-infiniteR93
Silicon (poly-)*NoneTheoretical standardSemi-infiniteR93

2.5.

Circular Retardation Phantoms

The effect of circular birefringence is frequently associated with the presence of chiral molecules,119 such as glucose. The aggregation of the presence chiral molecules in media causes the rotation of polarization plane of linearly polarized light as it travels through that volume. Manhas et al.,97 Ortega-Quijano et al.,120 and Ossikovski et al.112 added glucose to a polystyrene microsphere mixture in order to induce chirality and provide optical activity properties to the phantom, (Table 5).

Table 5

Optical activity phantoms.

Optical activity agentSolvent/preparationTissue mimickingPhantom thicknessTransmission/reflectanceRef.
Glucose (l-lysine)Distilled water, β-alanine, intralipid suspension, trypan blue dyeTurbid biological mediaSemi-infiniteR121
Glucose (d-)Water, 2  μm polystyrene microspheres, lipofundin, blood plasma, SiO2 nanoparticlesTurbid biological media40 mmR/T70, 97, 112, 120, 122, and 123
GlucoseWaterEye aqueous humorSemi-infiniteR124126
GlucoseWaterEye aqueous humor1×1  cm2T127
HoneyNoneTurbid biological mediaSemi-infiniteR75
SucrosePolyacrylamide, polystyrene microspheresTurbid biological tissue1×1×1  cm3T3
l-(+) –arabinose, M. racemicWater and polystyrene microspheresTurbid biological tissue1×1×1  cm3Side T119

Malik et al.124 developed several ocular models to investigate the feasibility of measuring glucose in the eye aqueous humor with polarization-based techniques (Fig. 6). The model shown also accounts for the cornea birefringence utilizing a PMMA-based phantoms overlaying a chamber mimicking the aqueous humor. A similar approach was used by Rawer et al.125

Fig. 6

Optical phantom from Ref. 124. The custom-built ocular model. Glucose concentration in the anterior section is varied through the two infusion tubes.

JBO_24_3_030901_f006.png

Other intralipid suspension liquid phantoms can be made with absorbers, such as dye, and optically active molecules, such as glucose and l-lysine, to test optical activity in samples.18,121 Antonelli75 used honey to calculate the optical activity of the sample. Pham et al.70 and Chang et al.18 studied the concentration of glucose by measuring the optical rotation angle of circular birefringence (optical activity) in human blood plasma and porcine cartilage samples.

3.

Conclusions

Optical phantoms that can be used for the calibration and benchmarking of polarimetric techniques and for mimicking the optical response of tissues have been used by several investigators.

It is to be noted that polarimetric optical phantoms are often unique to each research group and, aside from tests conducted on depolarization with microspheres suspensions, no standardization has been attempted. To our knowledge, only one company offers birefringent phantoms for polarized microscopy (NBS 1963A Birefringent Resolution Target by Thorlabs). As the biomedical applications of polarimetric techniques move toward quantification of directionality and retardation, more standardized phantoms are necessary. The PS-OCT phantoms proposed by Liu et al.108 are a good example of such approach. The measurements of PS-OCT’s two core parameters, namely, retardation and azimuth of optical axis can be easily reproduced, and different instruments can be benchmarked using such standardized phantoms. These mixed properties phantoms, particularly ones that include both depolarization and retardation, are needed for many applications. Phantoms that have birefringence of form rather than just intrinsic birefringence are also needed to simulate fibrous tissues, such as the cervix, cardiac tissue, or muscle. Nevertheless, the task of creating general use phantoms is complicated by the heterogeneity of tissues, the complexity of polarized light–tissue interaction, and the strong wavelength dependence of polarization-based techniques.

For these reasons, the use of biological tissue as measurement standards is very common in polarimetric applications, but unless these samples are well-known or measured with an alternative modality (e.g., PS-OCT or second harmonic generation), the scientific rigor of these experiments remains limited.

As new fabrication modalities, such as 3-D printing and lithography, are becoming available to researchers worldwide, we believe that a collaborative effort in the development of a standardized optical phantom for polarimetry could truly benefit the scientific community.

Disclosures

The authors have no relevant financial interests in this article and no potential conflicts of interest to disclose.

Acknowledgments

We gratefully acknowledge the support of the Herbert Wertheim Foundation and NSF ERC Grant PATHS-UP EEC-1648451 and STROBE: A National Science Foundation Science and Technology Center under Grant No. DMR 1548924.

References

1. 

J. Vanhellemont et al., “Round robin investigation of silicon oxide on silicon reference materials for ellipsometry,” Appl. Surf. Sci., 63 45 –51 (1993). https://doi.org/10.1016/0169-4332(93)90062-G Google Scholar

2. 

V. V. Tuchin, “Polarized light interaction with tissues,” J. Biomed. Opt., 21 (7), 071114 (2016). https://doi.org/10.1117/1.JBO.21.7.071114 JBOPFO 1083-3668 Google Scholar

3. 

N. Ghosh and I. A. Vitkin, “Tissue polarimetry: concepts, challenges, applications, and outlook,” J. Biomed. Opt., 16 (11), 110801 (2011). https://doi.org/10.1117/1.3652896 JBOPFO 1083-3668 Google Scholar

4. 

J. Qi and D. S. Elson, “Mueller polarimetric imaging for surgical and diagnostic applications: a review,” J. Biophotonics, 10 950 –982 (2017). https://doi.org/10.1002/jbio.v10.8 Google Scholar

5. 

J. F. de Boer, C. K. Hitzenberger and Y. Yasuno, “Polarization sensitive optical coherence tomography: a review [Invited],” Biomed. Opt. Express, 8 1838 –1873 (2017). https://doi.org/10.1364/BOE.8.001838 BOEICL 2156-7085 Google Scholar

6. 

B. Baumann, “Polarization sensitive optical coherence tomography: a review of technology and applications,” Appl. Sci.-Basel, 7 474 (2017). https://doi.org/10.3390/app7050474 Google Scholar

7. 

B. W. Pogue and M. S. Patterson, “Review of tissue simulating phantoms for optical spectroscopy, imaging and dosimetry,” J. Biomed. Opt., 11 (4), 041102 (2006). https://doi.org/10.1117/1.2335429 JBOPFO 1083-3668 Google Scholar

8. 

N. Ghosh, M. F. G. Wood and I. A. Vitkin, “Polarimetry in turbid, birefringent, optically active media: a Monte Carlo study of Mueller matrix decomposition in the backscattering geometry,” J. Appl. Phys., 105 102023 (2009). https://doi.org/10.1063/1.3116129 JAPIAU 0021-8979 Google Scholar

9. 

J. J. Gil and R. Ossikovski, “Polarized light and the Mueller matrix approach,” Series in Optics and Optoelectronics, CRC Press, Boca Raton, Florida (2016). Google Scholar

10. 

M. R. Hee et al., “Polarization-sensitive low-coherence reflectometer for birefringence characterization and ranging,” J. Opt. Soc. Am. B, 9 903 –908 (1992). Google Scholar

11. 

S. Y. Lu and R. A. Chipman, “Interpretation of Mueller matrices based on polar decomposition,” J. Opt. Soc. Am. A, 13 1106 –1113 (1996). https://doi.org/10.1364/JOSAA.13.001106 JOAOD6 0740-3232 Google Scholar

12. 

N.-J. Jan, K. Lathrop and I. A. Sigal, “Collagen architecture of the posterior pole: high-resolution wide field of view visualization and analysis using polarized light microscopy,” Invest. Ophth. Vis. Sci., 58 735 –744 (2017). https://doi.org/10.1167/iovs.16-20772 IOVSDA 0146-0404 Google Scholar

13. 

Y. Lim et al., “Birefringence measurement of cornea and anterior segment by office-based polarization-sensitive optical coherence tomography,” Biomed. Opt. Express, 2 2392 –2402 (2011). https://doi.org/10.1364/BOE.2.002392 BOEICL 2156-7085 Google Scholar

14. 

M. Wojtkowski, “High-speed optical coherence tomography: basics and applications,” Appl. Opt., 49 D30 –D61 (2010). https://doi.org/10.1364/AO.49.000D30 APOPAI 0003-6935 Google Scholar

15. 

W. Kaminsky et al., “Simultaneous false‐colour imaging of birefringence, extinction and transmittance at camera speed,” J. Microsc., 228 153 –164 (2007). https://doi.org/10.1111/jmi.2007.228.issue-2 JMICAR 0022-2720 Google Scholar

16. 

N. Ugryumova et al., “Novel optical imaging technique to determine the 3-D orientation of collagen fibers in cartilage: variable-incidence angle polarization-sensitive optical coherence tomography,” Osteoarthritis Cartilage, 17 33 –42 (2009). https://doi.org/10.1016/j.joca.2008.05.005 Google Scholar

17. 

D. K. Kasaragod et al., “Experimental validation of an extended Jones matrix calculus model to study the 3-D structural orientation of the collagen fibers in articular cartilage using polarization-sensitive optical coherence tomography,” Biomed. Opt. Express, 3 378 –387 (2012). https://doi.org/10.1364/BOE.3.000378 BOEICL 2156-7085 Google Scholar

18. 

C.-M. Chang et al., “Optical characterization of porcine articular cartilage using a polarimetry technique with differential Mueller matrix formulism,” Appl. Opt., 57 2121 –2127 (2018). https://doi.org/10.1364/AO.57.002121 APOPAI 0003-6935 Google Scholar

19. 

R. Lu et al., “A polarization-sensitive light field imager for multi-channel angular spectroscopy of light scattering in biological tissues,” Quant. Imaging Med. Surg., 5 1 –8 (2015). https://doi.org/10.3978/j.issn.2223-4292.2014.11.01 Google Scholar

20. 

R. Oldenbourg, E. D. Salmon and P. T. Tran, “Birefringence of single and bundled microtubules,” Biophys. J., 74 645 –654 (1998). https://doi.org/10.1016/S0006-3495(98)77824-5 BIOJAU 0006-3495 Google Scholar

21. 

J. Chue-Sang et al., “Use of Mueller matrix polarimetry and optical coherence tomography in the characterization of cervical collagen anisotropy,” J. Biomed. Opt., 22 (8), 086010 (2017). https://doi.org/10.1117/1.JBO.22.8.086010 JBOPFO 1083-3668 Google Scholar

22. 

A. Pierangelo et al., “Polarimetric imaging of uterine cervix: a case study,” Opt. Express, 21 14120 –14130 (2013). https://doi.org/10.1364/OE.21.014120 OPEXFF 1094-4087 Google Scholar

23. 

N. Agarwal et al., “Spatial evolution of depolarization in homogeneous turbid media within the differential Mueller matrix formalism,” Opt. Lett., 40 5634 –5637 (2015). https://doi.org/10.1364/OL.40.005634 OPLEDP 0146-9592 Google Scholar

24. 

M. R. Antonelli et al., “Mueller matrix imaging of human colon tissue for cancer diagnostics: how Monte Carlo modeling can help in the interpretation of experimental data,” Opt. Express, 18 10200 –10208 (2010). https://doi.org/10.1364/OE.18.010200 OPEXFF 1094-4087 Google Scholar

25. 

M. R. Antonelli et al., “Impact of model parameters on Monte Carlo simulations of backscattering Mueller matrix images of colon tissue,” Biomed. Opt. Express, 2 1836 –1851 (2011). https://doi.org/10.1364/BOE.2.001836 BOEICL 2156-7085 Google Scholar

26. 

T. Novikova et al., “Special section guest editorial: Antonello de Martino (1954–2014): in memoriam,” J. Biomed. Opt., 21 (7), 071101 (2016). https://doi.org/10.1117/1.JBO.21.7.071101 JBOPFO 1083-3668 Google Scholar

27. 

A. Pierangelo et al., “Ex-vivo characterization of human colon cancer by Mueller polarimetric imaging,” Opt. Express, 19 1582 –1593 (2011). https://doi.org/10.1364/OE.19.001582 OPEXFF 1094-4087 Google Scholar

28. 

A. Pierangelo et al., “Ex vivo photometric and polarimetric multilayer characterization of human healthy colon by multispectral Mueller imaging,” J. Biomed. Opt., 17 (6), 066009 (2012). https://doi.org/10.1117/1.JBO.17.6.066009 JBOPFO 1083-3668 Google Scholar

29. 

A. Pierangelo et al., “Multispectral Mueller polarimetric imaging detecting residual cancer and cancer regression after neoadjuvant treatment for colorectal carcinomas,” J. Biomed. Opt., 18 (4), 046014 (2013). https://doi.org/10.1117/1.JBO.18.4.046014 JBOPFO 1083-3668 Google Scholar

30. 

J. Rehbinder et al., “Ex vivo Mueller polarimetric imaging of the uterine cervix: a first statistical evaluation,” J. Biomed. Opt., 21 (7), 071113 (2016). https://doi.org/10.1117/1.JBO.21.7.071113 JBOPFO 1083-3668 Google Scholar

31. 

S. Alali et al., “Assessment of local structural disorders of the bladder wall in partial bladder outlet obstruction using polarized light imaging,” Biomed. Opt. Express, 5 621 –629 (2014). https://doi.org/10.1364/BOE.5.000621 BOEICL 2156-7085 Google Scholar

32. 

M. F. Wood et al., “Polarization birefringence measurements for characterizing the myocardium, including healthy, infarcted, and stem-cell-regenerated tissues,” J. Biomed. Opt., 15 (4), 047009 (2010). https://doi.org/10.1117/1.3469844 JBOPFO 1083-3668 Google Scholar

33. 

W. Groner et al., “Orthogonal polarization spectral imaging: a new method for study of the microcirculation,” Nat. Med., 5 1209 –1212 (1999). https://doi.org/10.1038/13529 1078-8956 Google Scholar

34. 

O. Genzel-Boroviczeny et al., “Orthogonal polarization spectral imaging (OPS): a novel method to measure the microcirculation in term and preterm infants transcutaneously,” Pediatr. Res., 51 386 –391 (2002). https://doi.org/10.1203/00006450-200203000-00019 Google Scholar

35. 

S. Langer et al., “Orthogonal polarization spectral imaging versus intravital fluorescent microscopy for microvascular studies in wounds,” Ann. Plast. Surg., 48 646 –653 (2002). https://doi.org/10.1097/00000637-200206000-00014 APCSD4 Google Scholar

36. 

S. Langer et al., “Orthogonal polarization spectral imaging as a tool for the assessment of hepatic microcirculation: a validation study,” Transplantation, 71 1249 –1256 (2001). https://doi.org/10.1097/00007890-200105150-00012 TRPLAU 0041-1337 Google Scholar

37. 

E. Uhl et al., “Intraoperative detection of early microvasospasm in patients with subarachnoid hemorrhage by using orthogonal polarization spectral imaging,” Neurosurgery, 52 1307 –1317 (2003). https://doi.org/10.1227/01.NEU.0000065154.04824.9E NEQUEB Google Scholar

38. 

E. von Dobschuetz et al., “Noninvasive in vivo assessment of the pancreatic microcirculation: orthogonal polarization spectral imaging,” Pancreas, 26 139 –143 (2003). https://doi.org/10.1097/00006676-200303000-00009 PANCE4 0885-3177 Google Scholar

39. 

R. R. Anderson, “Polarized light examination and photography of the skin,” Arch. Dermatol., 127 1000 –1005 (1991). https://doi.org/10.1001/archderm.1991.01680060074007 Google Scholar

40. 

S. L. Jacques, J. R. Roman and K. Lee, “Imaging superficial tissues with polarized light,” Lasers Surg. Med., 26 119 –129 (2000). https://doi.org/10.1002/(SICI)1096-9101(2000)26:2<119::AID-LSM3>3.0.CO;2-Y LSMEDI 0196-8092 Google Scholar

41. 

S. L. Jacques, J. C. Ramella-Roman and K. Lee, “Imaging skin pathology with polarized light,” J. Biomed. Opt., 7 (3), 329 –340 (2002). https://doi.org/10.1117/1.1484498 JBOPFO 1083-3668 Google Scholar

42. 

S. P. Morgan et al., “Rotating orthogonal polarization imaging,” Opt. Lett., 33 1503 –1505 (2008). https://doi.org/10.1364/OL.33.001503 OPLEDP 0146-9592 Google Scholar

43. 

J. C. Ramella-Roman et al., “Design, testing, and clinical studies of a handheld polarized light camera,” J. Biomed. Opt., 9 (6), 1305 –1310 (2004). https://doi.org/10.1117/1.1781667 JBOPFO 1083-3668 Google Scholar

44. 

P. Ghassemi et al., “Out-of-plane Stokes imaging polarimeter for early skin cancer diagnosis,” J. Biomed. Opt., 17 (7), 076014 (2012). https://doi.org/10.1117/1.JBO.17.7.076014 JBOPFO 1083-3668 Google Scholar

45. 

P. Ghassemi et al., “A new approach for optical assessment of directional anisotropy in turbid media,” J. Biophotonics, 9 100 –108 (2016). https://doi.org/10.1002/jbio.201400124 Google Scholar

46. 

P. Ghassemi et al., “A polarized multispectral imaging system for quantitative assessment of hypertrophic scars,” Biomed. Opt. Express, 5 3337 –3354 (2014). https://doi.org/10.1364/BOE.5.003337 BOEICL 2156-7085 Google Scholar

47. 

M. Bonesi et al., “High-speed polarization sensitive optical coherence tomography scan engine based on Fourier domain mode locked laser,” Biomed. Opt. Express, 3 2987 –3000 (2012). https://doi.org/10.1364/BOE.3.002987 BOEICL 2156-7085 Google Scholar

48. 

S. Sakai et al., “In vivo evaluation of human skin anisotropy by polarization-sensitive optical coherence tomography,” Biomed. Opt. Express, 2 2623 –2631 (2011). https://doi.org/10.1364/BOE.2.002623 BOEICL 2156-7085 Google Scholar

49. 

J. Qi and D. S. Elson, “A high definition Mueller polarimetric endoscope for tissue characterization,” Sci. Rep., 6 25953 (2016). https://doi.org/10.1038/srep25953 SRCEC3 2045-2322 Google Scholar

50. 

S. Alali et al., “Quantitative correlation between light depolarization and transport albedo of various porcine tissues,” J. Biomed. Opt., 17 (4), 045004 (2012). https://doi.org/10.1117/1.JBO.17.4.045004 JBOPFO 1083-3668 Google Scholar

51. 

K. A. Montejo et al., 10043 1004303 –1004310 (2017). Google Scholar

52. 

A. H. Hielscher and S. Bartel, “Diffuse backscattering Mueller matrix analysis for tissue diagnostics with polarized light,” Proc. SPIE, 3917 43 –53 (2000). https://doi.org/10.1117/12.382748 Google Scholar

53. 

A. H. Hielscher, J. R. Mourant and I. J. Bigio, “Influence of particle size and concentration on the diffuse backscattering of polarized light from tissue phantoms and biological cell suspensions,” Appl. Opt., 36 125 –135 (1997). https://doi.org/10.1364/AO.36.000125 APOPAI 0003-6935 Google Scholar

54. 

M. Sun et al., “Characterizing the microstructures of biological tissues using Mueller matrix and transformed polarization parameters,” Biomed. Opt. Express, 5 4223 –4234 (2014). https://doi.org/10.1364/BOE.5.004223 BOEICL 2156-7085 Google Scholar

55. 

M. A. Wallenburg et al., “Comparison of optical polarimetry and diffusion tensor MR imaging for assessing myocardial anisotropy,” J. Innovative Opt. Health Sci., 3 109 –121 (2010). https://doi.org/10.1142/S1793545810000976 Google Scholar

56. 

M. F. G. Wood et al., “Effects of formalin fixation on tissue optical polarization properties,” Phys. Med. Biol., 56 N115 (2011). https://doi.org/10.1088/0031-9155/56/8/N01 PHMBA7 0031-9155 Google Scholar

57. 

N. Ghosh et al., “Mueller matrix decomposition for polarized light assessment of biological tissues,” J. Biophotonics, 2 145 –156 (2009). https://doi.org/10.1002/jbio.v2:3 Google Scholar

58. 

B. Yang et al., “Polarized light spatial frequency domain imaging for non-destructive quantification of soft tissue fibrous structures,” Biomed. Opt. Express, 6 1520 –1533 (2015). https://doi.org/10.1364/BOE.6.001520 BOEICL 2156-7085 Google Scholar

59. 

J. Chue-Sang et al., “Use of combined polarization-sensitive optical coherence tomography and Mueller matrix imaging for the polarimetric characterization of excised biological tissue,” J. Biomed. Opt., 21 (7), 071109 (2016). https://doi.org/10.1117/1.JBO.21.7.071109 JBOPFO 1083-3668 Google Scholar

60. 

W. Goth et al., 9710 971018 –971019 (2016). Google Scholar

61. 

S. Sugita and T. Matsumoto, “Quantitative measurement of the distribution and alignment of collagen fibers in unfixed aortic tissues,” J. Biomech., 46 1403 –1407 (2013). https://doi.org/10.1016/j.jbiomech.2013.02.003 JBMCB5 0021-9290 Google Scholar

62. 

C. Fan and G. Yao, “Imaging myocardial fiber orientation using polarization sensitive optical coherence tomography,” Biomed. Opt. Express, 4 460 –465 (2013). https://doi.org/10.1364/BOE.4.000460 BOEICL 2156-7085 Google Scholar

63. 

Y. Gan and C. P. Fleming, “Extracting three-dimensional orientation and tractography of myofibers using optical coherence tomography,” Biomed. Opt. Express, 4 2150 –2165 (2013). https://doi.org/10.1364/BOE.4.002150 BOEICL 2156-7085 Google Scholar

64. 

C. P. Fleming et al., “Quantification of cardiac fiber orientation using optical coherence tomography,” J. Biomed. Opt., 13 (3), 030505 (2008). https://doi.org/10.1117/1.2937470 JBOPFO 1083-3668 Google Scholar

65. 

B. Baumann et al., “Polarization sensitive optical coherence tomography of melanin provides intrinsic contrast based on depolarization,” Biomed. Opt. Express, 3 1670 –1683 (2012). https://doi.org/10.1364/BOE.3.001670 BOEICL 2156-7085 Google Scholar

66. 

C. He et al., “Characterizing microstructures of cancerous tissues using multispectral transformed Mueller matrix polarization parameters,” Biomed. Opt. Express, 6 2934 –2945 (2015). https://doi.org/10.1364/BOE.6.002934 BOEICL 2156-7085 Google Scholar

67. 

M. Yamanari, S. Makita and Y. Yasuno, “Polarization-sensitive swept-source optical coherence tomography with continuous source polarization modulation,” Opt. Express, 16 5892 –5906 (2008). https://doi.org/10.1364/OE.16.005892 OPEXFF 1094-4087 Google Scholar

68. 

X. Li, J. C. Ranasinghesagara and G. Yao, “Polarization-sensitive reflectance imaging in skeletal muscle,” Opt. Express, 16 9927 –9935 (2008). https://doi.org/10.1364/OE.16.009927 OPEXFF 1094-4087 Google Scholar

69. 

S. Sridhar and A. Da Silva, “Enhanced contrast and depth resolution in polarization imaging using elliptically polarized light,” J. Biomed. Opt., 21 (7), 071107 (2016). https://doi.org/10.1117/1.JBO.21.7.071107 JBOPFO 1083-3668 Google Scholar

70. 

H. T. T. Pham et al., “Optical parameters of human blood plasma, collagen, and calfskin based on the Stokes-Mueller technique,” Appl. Opt., 57 4353 –4359 (2018). https://doi.org/10.1364/AO.57.004353 APOPAI 0003-6935 Google Scholar

71. 

H. R. Lee et al., SPIE Photonics Europe, 10677 8 2018). Google Scholar

72. 

P. J. Wu and J. T. Walsh, “Stokes polarimetry imaging of rat-tail tissue in a turbid medium using incident circularly polarized light,” Lasers Surg. Med., 37 396 –406 (2005). https://doi.org/10.1002/lsm.20242 LSMEDI 0196-8092 Google Scholar

73. 

M. Todorović et al., “Determination of local polarization properties of biological samples in the presence of diattenuation by use of Mueller optical coherence tomography,” Opt. Lett., 29 2402 –2404 (2004). https://doi.org/10.1364/OL.29.002402 OPLEDP 0146-9592 Google Scholar

74. 

L. X. Chin et al., “En face parametric imaging of tissue birefringence using polarization-sensitive optical coherence tomography,” J. Biomed. Opt., 18 (6), 066005 (2013). https://doi.org/10.1117/1.JBO.18.6.066005 JBOPFO 1083-3668 Google Scholar

75. 

M. R. Antonelli, “Biomedical applications of polarimetric imaging contrast. Initial studies for scattering media and human tissues,” France (2011). Google Scholar

76. 

C. Fan and G. Yao, “Mapping local retardance in birefringent samples using polarization sensitive optical coherence tomography,” Opt. Lett., 37 1415 –1417 (2012). https://doi.org/10.1364/OL.37.001415 OPLEDP 0146-9592 Google Scholar

77. 

S. Bancelin et al., “Determination of collagen fiber orientation in histological slides using Mueller microscopy and validation by second harmonic generation imaging,” Opt. Express, 22 22561 –22574 (2014). https://doi.org/10.1364/OE.22.022561 OPEXFF 1094-4087 Google Scholar

78. 

J. Chue-Sang et al., “Use of Mueller matrix colposcopy in the characterization of cervical collagen anisotropy,” J. Biomed. Opt., 23 (12), 121605 (2018). https://doi.org/10.1117/1.JBO.23.12.121605 JBOPFO 1083-3668 Google Scholar

79. 

E. M. Spiesz, W. Kaminsky and P. K. Zysset, “A quantitative collagen fibers orientation assessment using birefringence measurements: calibration and application to human osteons,” J. Struct. Biol., 176 302 –306 (2011). https://doi.org/10.1016/j.jsb.2011.09.009 JSBIEM 1047-8477 Google Scholar

80. 

S. Jiao and L. V. Wang, “Two-dimensional depth-resolved Mueller matrix of biological tissue measured with double-beam polarization-sensitive optical coherence tomography,” Opt. Lett., 27 101 –103 (2002). https://doi.org/10.1364/OL.27.000101 OPLEDP 0146-9592 Google Scholar

81. 

S. Jiao et al., “Fiber-based polarization-sensitive Mueller matrix optical coherence tomography with continuous source polarization modulation,” Appl. Opt., 44 5463 –5467 (2005). https://doi.org/10.1364/AO.44.005463 APOPAI 0003-6935 Google Scholar

82. 

Z. Lu, D. Kasaragod and S. J. Matcher, “Conical scan polarization-sensitive optical coherence tomography,” Biomed. Opt. Express, 5 752 –762 (2014). https://doi.org/10.1364/BOE.5.000752 BOEICL 2156-7085 Google Scholar

83. 

Z. Lu, D. K. Kasaragod and S. J. Matcher, “Optic axis determination by fibre-based polarization-sensitive swept-source optical coherence tomography,” Phys. Med. Biol., 56 1105 –1122 (2011). https://doi.org/10.1088/0031-9155/56/4/014 PHMBA7 0031-9155 Google Scholar

84. 

R. Ossikovski, “Differential matrix formalism for depolarizing anisotropic media,” Opt. Lett., 36 2330 –2332 (2011). https://doi.org/10.1364/OL.36.002330 OPLEDP 0146-9592 Google Scholar

85. 

J. M. Schmitt, A. H. Gandjbakhche and R. F. Bonner, “Use of polarized-light to discriminate short-path photons in a multiply scattering medium,” Appl. Opt., 31 6535 –6546 (1992). https://doi.org/10.1364/AO.31.006535 APOPAI 0003-6935 Google Scholar

86. 

S. Rehn et al., “Depth probing of diffuse tissues controlled with elliptically polarized light,” J. Biomed. Opt., 18 (1), 016007 (2013). https://doi.org/10.1117/1.JBO.18.1.016007 JBOPFO 1083-3668 Google Scholar

87. 

J. C. Ramella-Roman, S. A. Prahl and S. L. Jacques, “Three Monte Carlo programs of polarized light transport into scattering media: part II,” Opt. Express, 13 10392 –10405 (2005). https://doi.org/10.1364/OPEX.13.010392 OPEXFF 1094-4087 Google Scholar

88. 

T. Yun et al., “Monte Carlo simulation of polarized photon scattering in anisotropic media,” Opt. Express, 17 16590 –16602 (2009). https://doi.org/10.1364/OE.17.016590 OPEXFF 1094-4087 Google Scholar

89. 

F. C. Mackintosh et al., “Polarization memory of multiply scattered-light: reply,” Phys. Rev. B, 45 8165 –8165 (1992). https://doi.org/10.1103/PhysRevB.45.8165 Google Scholar

90. 

M. K. Swami et al., “Effect of gold nanoparticles on depolarization characteristics of intralipid tissue phantom,” Opt. Lett., 38 2855 –2857 (2013). https://doi.org/10.1364/OL.38.002855 OPLEDP 0146-9592 Google Scholar

91. 

B. Kunnen et al., “Application of circularly polarized light for non-invasive diagnosis of cancerous tissues and turbid tissue-like scattering media,” J. Biophoton., 8 317 –323 (2015). https://doi.org/10.1002/jbio.v8.4 Google Scholar

92. 

M. Ahmad et al., “Do different turbid media with matched bulk optical properties also exhibit similar polarization properties?,” Biomed. Opt. Express, 2 3248 –3258 (2011). https://doi.org/10.1364/BOE.2.003248 BOEICL 2156-7085 Google Scholar

93. 

J. C. Petermann, (2012). Google Scholar

94. 

M. J. Rakovic et al., “Light backscattering polarization patterns from turbid media: theory and experiment,” Appl. Opt., 38 3399 –3408 (1999). https://doi.org/10.1364/AO.38.003399 APOPAI 0003-6935 Google Scholar

95. 

D. Cote and I. Vitkin, “Robust concentration determination of optically active molecules in turbid media with validated three-dimensional polarization sensitive Monte Carlo calculations,” Opt. Express, 13 148 –163 (2005). https://doi.org/10.1364/OPEX.13.000148 OPEXFF 1094-4087 Google Scholar

96. 

N. Ghosh et al., “Depolarization of light in tissue phantoms: effect of collection geometry,” Opt. Commun., 222 93 –100 (2003). https://doi.org/10.1016/S0030-4018(03)01567-0 OPCOB8 0030-4018 Google Scholar

97. 

S. Manhas et al., “Mueller matrix approach for determination of optical rotation in chiral turbid media in backscattering geometry,” Opt. Express, 14 190 –202 (2006). https://doi.org/10.1364/OPEX.14.000190 OPEXFF 1094-4087 Google Scholar

98. 

Y. Guo et al., “A study on forward scattering Mueller matrix decomposition in anisotropic medium,” Opt. Express, 21 18361 –18370 (2013). https://doi.org/10.1364/OE.21.018361 OPEXFF 1094-4087 Google Scholar

99. 

H. He et al., “Two-dimensional backscattering Mueller matrix of sphere-cylinder scattering medium,” Opt. Lett., 35 2323 –2325 (2010). https://doi.org/10.1364/OL.35.002323 OPLEDP 0146-9592 Google Scholar

100. 

B. Laude-Boulesteix et al., “Mueller polarimetric imaging system with liquid crystals,” Appl. Opt., 43 2824 –2832 (2004). https://doi.org/10.1364/AO.43.002824 APOPAI 0003-6935 Google Scholar

101. 

E. Garcia-Caurel et al., “Application of spectroscopic ellipsometry and Mueller ellipsometry to optical characterization,” Appl. Spectrosc., 67 1 –21 (2013). https://doi.org/10.1366/12-06883 APSPA4 0003-7028 Google Scholar

102. 

M. S. Wróbel et al., 20 10 (2015). Google Scholar

103. 

M. S. Wróbel et al., “Multi-layered tissue head phantoms for noninvasive optical diagnostics,” J. Innovative Opt. Health Sci., 8 1541005 (2015). https://doi.org/10.1142/S1793545815410059 Google Scholar

104. 

A. A. Hamza et al., “Determination of the intrinsic birefringence of polymeric fibres,” Polymer Testing, 23 203 –208 (2004). https://doi.org/10.1016/S0142-9418(03)00081-3 POTEDZ 0142-9418 Google Scholar

105. 

N. Ghosh, M. F. Wood and I. A. Vitkin, “Mueller matrix decomposition for extraction of individual polarization parameters from complex turbid media exhibiting multiple scattering, optical activity, and linear birefringence,” J. Biomed. Opt., 13 (4), 044036 (2008). https://doi.org/10.1117/1.2960934 JBOPFO 1083-3668 Google Scholar

106. 

P. W. Gibson, H. L. Schreuder‐Gibson and D. Rivin, “Electrospun fiber mats: transport properties,” AIChE J., 45 190 –195 (1999). https://doi.org/10.1002/aic.690450116 AICEAC 0001-1541 Google Scholar

107. 

Y. Wang et al., “Study on the validity of 3 × 3 Mueller matrix decomposition,” J. Biomed. Opt., 20 (6), 065003 (2015). https://doi.org/10.1117/1.JBO.20.6.065003 JBOPFO 1083-3668 Google Scholar

108. 

X. Liu et al., “Tissue-like phantoms for quantitative birefringence imaging,” Biomed. Opt. Express, 8 4454 –4465 (2017). https://doi.org/10.1364/BOE.8.004454 BOEICL 2156-7085 Google Scholar

109. 

S. Alali, A. Gribble and I. A. Vitkin, “Rapid wide-field Mueller matrix polarimetry imaging based on four photoelastic modulators with no moving parts,” Opt. Lett., 41 1038 –1041 (2016). https://doi.org/10.1364/OL.41.001038 OPLEDP 0146-9592 Google Scholar

110. 

B. Heise et al., “Spatially resolved stress measurements in materials with polarisation-sensitive optical coherence tomography: image acquisition and processing aspects,” Strain, 46 61 –68 (2010). https://doi.org/10.1111/str.2009.46.issue-1 Google Scholar

111. 

M. A. Wallenburg et al., “Polarimetry-based method to extract geometry-independent metrics of tissue anisotropy,” Opt. Lett., 35 (15), 2570 –2572 (2010). https://doi.org/10.1364/OL.35.002570 OPLEDP 0146-9592 Google Scholar

112. 

R. Ossikovski, “Analysis of depolarizing Mueller matrices through a symmetric decomposition,” J. Opt. Soc. Am. A, 26 1109 –1118 (2009). https://doi.org/10.1364/JOSAA.26.001109 JOAOD6 0740-3232 Google Scholar

113. 

D. Stifter et al., “Spectral domain polarization sensitive optical coherence tomography at 1.55 μm: novel developments and applications for dynamic studies in materials science,” Proc. SPIE, 7889 78890Z (2011). https://doi.org/10.1117/12.873542 Google Scholar

114. 

M. F. Wood, X. Guo and I. A. Vitkin, “Polarized light propagation in multiply scattering media exhibiting both linear birefringence and optical activity: Monte Carlo model and experimental methodology,” J. Biomed. Opt., 12 (1), 014029 (2007). https://doi.org/10.1117/1.2434980 JBOPFO 1083-3668 Google Scholar

115. 

M. K. Swami et al., “Spectral Mueller matrix measurements for characterization of depolarization from non-spherical gold nanoparticles,” Opt. Commun., 308 136 –141 (2013). https://doi.org/10.1016/j.optcom.2013.06.016 OPCOB8 0030-4018 Google Scholar

116. 

P.-C. Chen et al., “Measurement of linear birefringence and diattenuation properties of optical samples using polarimeter and Stokes parameters,” Opt. Express, 17 15860 –15884 (2009). https://doi.org/10.1364/OE.17.015860 OPEXFF 1094-4087 Google Scholar

117. 

L. Yu Lung et al., “Using polarimeter and Stokes parameters for measuring linear birefringence and diattenuation properties of optical samples,” in EPJ Web of Conf., 32001 (2010). https://doi.org/10.1051/epjconf/20100632001 Google Scholar

118. 

D. B. Chenault and R. A. Chipman, “Measurements of linear diattenuation and linear retardance spectra with a rotating sample spectropolarimeter,” Appl. Opt., 32 3513 –3519 (1993). https://doi.org/10.1364/AO.32.003513 APOPAI 0003-6935 Google Scholar

119. 

K. C. Hadley and I. A. Vitkin, “Optical rotation and linear and circular depolarization rates in diffusively scattered light from chiral, racemic, and achiral turbid media,” J. Biomed. Opt., 7 (3), 291 –299 (2002). https://doi.org/10.1117/1.1483880 JBOPFO 1083-3668 Google Scholar

120. 

N. Ortega-Quijano et al., “Comparative study of optical activity in chiral biological media by polar decomposition and differential Mueller matrices analysis,” Proc. SPIE, 7906 790612 (2011). https://doi.org/10.1117/12.875080 Google Scholar

121. 

M. Mehrubeoglu et al., “Effect of molecular concentrations in tissue-simulating phantoms on images obtained using diffuse reflectance polarimetry,” Opt. Express, 3 286 –298 (1998). https://doi.org/10.1364/OE.3.000286 OPEXFF 1094-4087 Google Scholar

122. 

T. L. Chen et al., “Noninvasive measurement of glucose concentration on human fingertip by optical coherence tomography,” J. Biomed. Opt., 23 (4), 047001 (2018). https://doi.org/10.1117/1.JBO.23.4.047001 JBOPFO 1083-3668 Google Scholar

123. 

Q. H. Phan and Y. L. Lo, “Stokes-Mueller matrix polarimetry system for glucose sensing,” Opt. Laser Eng., 92 120 –128 (2017). https://doi.org/10.1016/j.optlaseng.2016.08.017 Google Scholar

124. 

B. H. Malik, C. W. Pirnstill and G. L. Cote, “Dual-wavelength polarimetric glucose sensing in the presence of birefringence and motion artifact using anterior chamber of the eye phantoms,” J. Biomed. Opt., 18 (01), 017007 (2013). https://doi.org/10.1117/1.JBO.18.1.017007 JBOPFO 1083-3668 Google Scholar

125. 

R. Rawer, W. Stork and C. F. Kreiner, “Non-invasive polarimetric measurement of glucose concentration in the anterior chamber of the eye,” Graefes Arch. Clin. Exp. Ophthalmol., 242 1017 –1023 (2004). https://doi.org/10.1007/s00417-004-1031-7 Google Scholar

126. 

R. R. Ansari, S. Bockle and L. Rovati, “New optical scheme for a polarimetric-based glucose sensor,” J. Biomed. Opt., 9 (1), 103 –115 (2004). https://doi.org/10.1117/1.1626664 JBOPFO 1083-3668 Google Scholar

127. 

Z. F. Yu, C. W. Pirnstill and G. L. Cote, “Dual-modulation, dual-wavelength, optical polarimetry system for glucose monitoring,” J. Biomed. Opt., 21 (8), 087001 (2016). https://doi.org/10.1117/1.JBO.21.8.087001 JBOPFO 1083-3668 Google Scholar

Biographies of the authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Joseph Chue-Sang, Mariacarla Gonzalez, Angie Pierre, Megan Laughrey, Ilyas Saytashev, Tatiana Novikova, and Jessica C. Ramella-Roman "Optical phantoms for biomedical polarimetry: a review," Journal of Biomedical Optics 24(3), 030901 (8 March 2019). https://doi.org/10.1117/1.JBO.24.3.030901
Received: 11 September 2018; Accepted: 29 January 2019; Published: 8 March 2019
Lens.org Logo
CITATIONS
Cited by 35 scholarly publications and 1 patent.
Advertisement
Advertisement
KEYWORDS
Polarimetry

Tissues

Polarization

Birefringence

Biomedical optics

Optical phantoms

Collagen

Back to Top