Open Access
1 November 2005 Optical coherence tomography speckle reduction by a partially spatially coherent source
Author Affiliations +
Abstract
Speckle in optical coherence tomography (OCT) images originates in the high spatial coherence of incident light that enables interference of light backscattered from spatially heterogenous tissue specimens. We report results of a numerical simulation and an experiment to test speckle reduction using a partially spatially coherent source. A Gaussian-Schell model for a partially spatially coherent source is used in the OCT simulation. For the experiment, such a source was generated by a spatially coherent boardband light source and a multimode fiber. The advantage of using a multimode fiber in combination with a broadband source is the large number of photons per coherence volume. To illustrate speckle reduction with a partially spatially coherent source, we record low-coherence interferograms of a scattering surface using single-mode and multimode source fibers. Interferograms recorded using a single-mode source fiber are indicative of those observed using conventional OCT. Speckle in OCT images recorded using a multimode source fiber is substantially reduced.

1.

Introduction

Optical coherence tomography (OCT) uses the short temporal coherence properties of broadband light to extract structural information from heterogeneous samples such as tissue. Two approaches are recognized to generate 2-D and 3-D OCT images: single-point raster scanning1, 2, 3, 4 and parallel/wide-field imaging.5, 6, 7, 8 Speckle noise is a major obstacle in measuring specific structural properties (e.g., birefringence9, 10, 11) of tissue specimens and observing specific features that approach the diffraction limit.12 Although speckle in OCT images is still being investigated, a common view is that the effect originates in distorted wavefronts of light returning from tissue that result in random appearing constructive and destructive features in the interferogram.

Figure 1 shows OCT images of in vivo primate retina and porcine mesenteric artery that present high-contrast speckle in regions of strong backscatter in two different biological tissue samples. The appearance of speckle in the retinal nerve fiber layer (RNFL) of Fig. 1(a) does not vary with depth, but rather present relatively constant over the depth. The media and adventitia area of Fig. 1(b) also display a gradual distribution of the speckle feature, indicating that the speckle is not dependent on depth because multiple backscattering and random phase variation of the returning light dominate the statistical properties of the speckle.

Fig. 1

Examples of OCT images with presence of speckle: (a) PS-OCT intensity image of a primate retina, inferior to the optic nerve head, where N is nasal while T is temporal and RNFL is the retinal nerve fiber layer. (b) OCT intensity image of a mesenteric artery.

064034_1_025506jbo1.jpg

Speckle reduction methods have been addressed by several investigators and include postsignal processing,13, 14, 15, 16, 17, 18 angular compounding,19, 20, 21, 22 and use of spatially incoherent illumination.23 Postsignal processing techniques have been reported, including a zero-amplitude procedure,18 deconvolution, and rotating kernel transformation. The zero-amplitude procedure (ZAP), which operates in the complex number domain, shows speckle reduction in OCT images but blurs boundaries between tissue structures. Other complex-number-domain processing methods applied to OCT include iterative point deconvolution (i.e., CLEAN in Ref. 16), and constrained iterative deconvolution.17 Deconvolution techniques require some prior knowledge of the point spread function of the imaging optics,24 as well as the optical properties of the imaged sample. The computation time of the rotating kernel transformation (RKT) technique, applied by Rogowsk and Brezinski14 increases substantially with kernel size. Angular compounding methods are based on acquiring/averaging multiple beams each with a different incidence/reflecting angle from the sample but have the disadvantages of either time-restricted implementation or high complexity.

Use of spatially incoherent illumination with a thermal light source has neither time constraints nor high complexity, but suffers from reduced SNR due to the low photon number per spatial mode.23 Since the average number of photons within a coherence volume for thermal light sources is of the order of unity, these sources are nondegenerate.25 Use of quasihomogeneous (partially spatially coherent) illumination has long been recognized to improve lateral resolution by decreasing speckle in recorded images.25, 26, 27 Although partially spatially coherent illumination can be easily implemented using a multimode optical fiber,28 its application in OCT imaging has not been reported, possibly due to problems associated with modal dispersion and alignment constraints.

We present a mathematical analysis and an experimental result showing a simple and rapid method to reduce speckle in OCT imaging using a partially spatially coherent light source generated by a multimode fiber. Although the coherence length of this source is reduced, the photon number per spatial mode is much larger than unity. The paper is organized as follows. First, we derive an interferometric imaging equation for an OCT system utilizing a Gaussian-Schell model light source. A numerical simulation and experimental setup are described in Sec. 2. Experimental results are presented in Sec. 3 and discussion in Sec. 4. Summary and conclusions follow in Sec. 5.

2.

Theory and Methods

Figure 2 shows a two-beam imaging Michelson interferometer. Broadband and partially spatially coherent light is emitted from a source and illuminates the sample and a reference mirror, which are imaged onto a CCD camera by lens L3. Spatial coordinates at the sample, reference, and CCD camera are denoted by (ξ,ζ) , (ξ,ζ) , and (x,z) respectively, where ξ , ξ , and x are in a plane perpendicular to light propagation, while ζ , ζ , and z are along a line parallel to light propagation. The amplitude of light As at the CCD camera plane produced from the sample reflection can be written as a coherent integral29 of amplitudes reflected from the sample modulated by point-spread function [h()] and the local reflectivity [r(ξ)] ,

Eq. 1

As(x,ν)=Ai(ξ,ν)r(ξ)exp(2πjντ)h(xξ,ν)dξ,
where subscripts s indicates sample. The factor exp(j2πντ) is included to account for the optical path length difference between reference and sample paths and will become important when sample and reference beams are combined. In Eq. 1 we assume amplitude of light returning from the sample is the simple product of the incident amplitude [Ai(ξ,ν)] and local reflectivity [r(ξ)] . In thick heterogeneous media such as tissue, the amplitude of light returning from the sample is an integral of incident amplitude and tissue refractive index [n(ξ)] . Similarly the amplitude field Ar at the CCD camera plane imaged from the reference mirror can be written as

Eq. 2

Ar(x,ν)=Ai(ξ,ν)rmirh(xξ,ν)dξ,
where subscripts r indicates the reference, and rmir is a constant reflectivity of the mirror. When sample and reference fields ( As and Ar ) interfere at the image plane, intensity at position x consists of two dc intensity components and a fringe signal. The fringe signal is a complex correlation between amplitudes returned from sample and reference paths. The detected signal can be written as

Eq. 3

Id(x,ν,τ)=As(x,ν,τ)+Ar(x,ν)2=Is(x,ν)+Ir(x,ν)+2ReAs(x,ν,τ)Ar*(x,ν),
where Is(x,ν) and Ir(x,ν) are the dc intensities from sample and reference paths, respectively, represented by

4.

Eq. 4a

Is(x,ν)=Ai*(ξ,ν,τ)Ai(ξ,ν,τ) r*(ξ)r(ξ)h*(xξ,ν)h(xξ,ν)dξdξ,

Eq. 4b

Ir(x,ν)=Ai(ξ,ν)Ai*(ξ,ν)rmir2h(xξ,ν)h*(xξ,ν)dξdξ.
Expressions for Is(x,ν) and Ir(x,ν) are the well-known imaging equations for partially spatial coherent light.25

Fig. 2

Imaging Michelson interferometer with wide-field illumination.

064034_1_025506jbo2.jpg

The complex fringe signal optical frequency (ν) is given by,

Eq. 5

If(x,ν,τ)=As(x,ν,τ)Ar*(x,ν)=rmirAi(ξ,ν)Ai*(ξ,ν)r*(ξ) exp(j2πντ)h(xξ,ν)h*(xξ,ν)dξdξ.
Inclusion of the complex sample reflectivity [r(ξ)] distinguishes the interference imaging equation [Eq. 5] from reference and sample intensity integrals [Eq. 4].

By summing over optical frequencies (ν) the expression for the complex fringe signal If(x,τ) becomes

Eq. 6

If(x,τ)=As(x,ν,τ)Ar*(x,ν)dν.
We model Eq. 6 with a Gaussian-Schell model source. The Gaussian-Schell model expression of the complex fringe can be written as

Eq. 7

If(x,τ)=rmirW(ξ,ξ,ν)r(ξ) exp(j2πντ)h(xξ)h*(xξ)dξdξdν,
where we have neglected the optical frequency (ν) dependence of the point-spread function [h()] and local reflectivity [r(ξ)] . Here W(ξ,ξ,ν) is the cross-spectral density function, and has the coherent-mode representation30

Eq. 8

W(ξ1,ξ2,ω)=nαn(ω)φn*(ξ1,ω)φn(ξ2,ω),
where the asterisk denotes the complex conjugate. The functions φn(ξ,ω) are the eigenfunctions (modes), and the coefficients αn(ω) are the eigenvalues of the integral equation. From the physical viewpoint, Eq. 8 indicates that a partially coherent source can be represented as a superposition of mutually uncorrelated modes each of which is spatially completely coherent in the space-frequency domain.31 For the simulation study we model the cross-spectral density function [W(ξ,ξ,ν)] as a Gaussian-Schell model source, having the form

Eq. 9

W(ξ,ξ,ν)=[S(ξ,ν)]12[S(ξ,ν)]12g(ξξ,ν),
with

Eq. 10

S(ξ,ν)=A2(ν)exp[ξ22σs2(ν)],

Eq. 11

g(ξξ,ν)=exp[(ξξ)22σg2(ν)],
representing the spectral density and the spectral degree of coherence of light with positive values σs and σg . The source power spectral density is given by A2(ν) . Values of σs and σg , which generally depend on optical frequency (ν) , determine the effective widths of the beam and spatial coherence, across the beam spot at the detection plane imaged from the sample and the reference mirror, respectively. The limit σgσs represents a globally incoherent (the so-called quasihomogeneous) source, while the opposite limit σgσs corresponds to a completely coherent source. The point spread function with a square aperture with width D is

Eq. 12

h(xξ)=sinc[Dczν0(xξ)],
where z is the distance between the sample and lens, c is speed of light, and ν0 is center optical frequency of the source power spectral density.

If the power spectral density A2(ν) is given as

Eq. 13

A2(ν)=exp[(νν0)22Δν2],
where Δν is the FWHM spectral width of the source, we can evaluate Eq. 7. Substituting Eqs. 9, 10, 11, 12 into the complex fringe signal [Eq. 7] and integrating over ν , we find

Eq. 14

If(x,τ)=πexp(iπτν0)exp(π2τ22Δν2)sinc[Dczν0(xξ)]sinc[Dczν0(xξ)]r(ξ) [12Δν2+A(ξ,ξ)]exp[A(ξ,ξ)π2τ2]dξdξ,
where σs and σg are dependent on optical frequency (ν) and estimated as σs(ν0)(ν0ν) and σg(ν0)(ν0ν) , respectively, and A(ξ,ξ) is
A(ξ,ξ)=σg(ν0)2(ξ2+ξ2)+2σs(ν0)2(ξξ)24σs(ν0)2σg(ν0)2ν02.
We note that the complex fringe signal If(x,τ) in Eq. 14 contains the OCT complex fringe signal with carrier frequency [exp(iπτν0)] with envelope [exp(π2τ22Δν2)] , but also a multiplicative term, which is the integral over spatial coordinates [ξ,ξ] . The integral over ξ and ζ accumulates in regions where the point spread functions are nonzero and is proportional to r(ξ) . Because the phase of r(ξ) can fluctuate with sample position (ξ) , the multiplicative term introduces speckle.

Because the exact form of the complex reflectivity of the sample [r(ξ)] is not generally known, we evaluate the statistics of If(x,τ) . We assume that the local complex reflectivity [r(ξ)] in a sample has a fixed amplitude distribution [r(ξ)] with randomly varying but uniformly distributed phase19 (ϕk) . The complex fringe signal at position x in Eq. 14 can be viewed as a sum of all field amplitudes reflected back within a coherence patch. Each field amplitude reflected from the sample contributes as a complex-valued phasor at position x in the observation plane. The complex fringe signal with speckle can be analyzed as a random phasor sum:

Eq. 15

If=aexp(iθ)=1Nk=1Nαkexp(iϕk).
The amplitude (αk) represents the comprehensive amplitude terms in Eq. 14 and exp(iϕk) is the randomly varying phase of the complex reflectivity of the sample. The probability density function of the amplitude of the random phasor sum follows a Rayleigh function and has mean and variance:

16.

a¯=π2σ,
σa2=(2π2)σ2,
where σ2 is the second moment α2¯ divided by two. The mean and variance of the amplitude are dependent only on and proportional to the second moment α2¯ . Although the second moment of the complex fringe signal can be calculated analytically by finding the second derivative of a characteristic function after integrating over ξ in Eq. 14, this study includes a numerical evaluation of the second moment of the complex fringe signal [Eq. 14] with variable coherent size (σg) .

2.1.

Numerical Simulation

Because speckle in OCT arises from coherent interference of backscattered light from the sample, we investigate speckle reduction by considering a Gaussian-Schell model source with different spatial coherence lengths (σg) used in conjunction with a 1-D rough surface. The complex fringe signal [Eq. 14] is computed along a line using software written by the author. We assume pupil of lens L3 has a diameter of 1cm and is placed 10cm away from the sample or reference mirror giving a numerical aperture (NA)=0.1 . The width of the point spread function (PSF) is calculated to be 1.2λ0NA=8.5μm assuming λ0=850nm . Grid spacing for ξ and ξ is 0.34μm or 25 times smaller than the PSF width. Beam spot diameter (σs) at the sample and the reference mirror is set at σs=100μm and several spatial coherent lengths (σg) of the Gaussian-Schell model source are chosen to observe the speckle reduction effect. The sample reflectivity is set at unity between 40 and 60μm and zero elsewhere. Phases vary from π to π with uniformly distributed random probability. The image plane is positioned at 10cm from lens L3 and complex fringe signal [If(x)] space is computed over [x=0,100μm] at discrete points separated by 1μm .

2.2.

Experimental Setup

A depth-resolving OCT system based on a two-beam Michelson interferometer was constructed to investigate speckle reduction using a partially spatial coherent source. Partially spatial coherent light was generated from a coherent light source by using a long segment of high-NA multimode optical fiber. For coherent illumination, we used a single mode fiber with a cuffoff wavelength λc=800nm . Both single- and multi-mode optical fibers are placed in the source path of a Michelson interferometer to investigate effect of source coherence length on speckle reduction. First, we used a 100-m -long, 0.48-NA multimode glass optical fiber with a 200-μm core diameter, producing about 63,000 coherent spatial modes.32 Second, we used a 30-m -long single mode optical fiber (Rifocs Corp.) similar to that used in a conventional OCT setup. In both cases, a mode-locked Ti:sapphire laser with an output centered at 850nm and spectral width Δλ50nm is coupled into the fibers and used as source light. When coupled into a single mode fiber, light gives a TEM00 spatial mode. In the first setup (Fig. 3 ), light is coupled into a multimode fiber to generate many spatial modes. The 100-m -long multimode fiber is mounted on a spool to reduce any variations in mechanical stress that might introduce variable mode coupling. Following the multimode fiber, lens L2 directs light into both sample and reference paths through a broadband beamsplitter (BS). In the second setup using a single-mode fiber, the multimode fiber is removed from the source path of the interferometer. Reflected light from both sample and reference paths are imaged onto a CCD camera (FASTCAM Super 10K ) using a focusing lens (L3). A voice coil with a retroreflector is used in the reference path to provide variable delay (τ) between sample and reference paths. The voice coil is driven by a function generator creating a sinusoidal wave at 0.1Hz and 10Vp-p . Scan range is 1.6mm , yielding a fringe carrier frequency fc=752Hz . Frame rate of the CCD camera is set at 3000framess sampling the fringe at four points per cycle or twice the Nyquist limit. Integration time of the camera is 50μs . Output images from the CCD camera are recorded in the digital video (DV) format to avoid any loss in spatial resolution. Incoherent demodulation is used to determine the envelope of the complex fringe signal [Eq. 15] from recorded 3-D data cubes. Fringe amplitude is determined at each pixel in time (Fig. 4 ). Other than incoherent demodulation no additional image enhancement is used.

Fig. 3

Optical setup with multimode fiber in source path of Michelson interferometer: L1 (f=10mm) , L2 (f=150mm) , and L3 (f=100mm) ; lenses; BS; beamsplitter; ND, neutral density filter.

064034_1_025506jbo3.jpg

Fig. 4

Flow diagram for incoherent demodulation: X(t) ; detected signal; BPF, bandpass filter; Hilbert, Hilbert transform; ABS, absolute value; A(t) , envelope of the detected signal.

064034_1_025506jbo4.jpg

A mill-finished aluminum plate is used as a scattering surface. The rough surface of the aluminum plate produces speckles caused by distorted wavefronts in reflected light. The aluminum plate was tilted slightly to produce a uniform phase gradient and produce linear fringes in the CCD. For comparison, the same physical location covering a 2-mm beam diameter on the aluminum plate is imaged with both single and multimode fiber systems. To maximize fringe visibility and to avoid saturation of CCD photocells, light intensity incident on the CCD camera was adjusted by using a neutral density filter in the detection path of the interferometer.

3.

Results

Effect of speckle reduction in OCT imaging was simulated using a Gaussian-Schell model source and Eq. 14 for If(x,τ) . For this simulation, the illuminating beam spot [σs] was fixed at 100μm . Simulation results shown in Fig. 5 and Table 1 are determined for a depth-resolving OCT system with several different spatial coherent lengths (σg) corresponding to a Gaussian-Schell model source. In Fig. 5, the dotted lines represent the envelope of the complex fringe signal without speckle, while solid lines are with speckle. The speckle-free case was simulated by removing the random phase term in Eq. 15. Coherent illumination (σg=σs=100μm) produces the greatest speckle [Fig. 5(a)]. For incoherent illumination, spatial coherence length was set much smaller (σg=0.01μm) than diameter (σs) of the illuminating beam [Fig. 5(d)]. Two additional spatial coherence lengths (σg=10,1μm) in Figs. 5(b) and 5(c), respectively, were investigated to demonstrate speckle reduction with decreasing coherence length. To assess speckle reduction quantitatively, speckle root-mean-square (rms) values were calculated by subtracting computed complex fringe signal with speckle from the speckle-free case. Speckle rms variation with spatial coherence length (σg) is shown in Table 1. Entries in Table 1 were computed by averaging 30 simulations for each spatial coherence length (σg) . Note the logarithmic scale to display σg .

Fig. 5

Simulated normalized fringe intensity for different spatial coherence lengths of Gaussian-Schell model source: σg , coherent length of a Gaussian-Schell model source; solid lines, complex fringe signal with speckle; and dotted lines, envelope of complex fringe signal without speckle. Beam spot size at target (σs) was set at 100μm .

064034_1_025506jbo5.jpg

Table 1

Speckle rms variation according to spatial coherence length (σg) .

σg (μm) Speckle rms
1000.188
100.174
10.081
0.10.031

To observe difference in sample illumination through single or multimode fibers, light from the reference path was blocked and the aluminum plate was placed in the sample path. Figures 7(a) and 7(b) show intensity images of the sample when illuminated by single-mode and multimode fibers, respectively. Due to a small mode field diameter (5μm) at the single-mode fiber tip, Airy rings in the illuminated area are observed in Fig. 7(a), whereas relatively flat illumination is evident in the multimode case [Fig. 7(b)].

Fig. 7

CCD camera images of an aluminum plate through (a) single-mode and (b) multimode source fiber. Reference path of interferometer is blocked.

064034_1_025506jbo7.jpg

After restoring the reference path, an interference pattern was measured by the CCD camera corresponding to a 1.6-mm pathlength delay between sample and reference paths. During measurement of the interference pattern, no significant motion artifacts in the fringes, caused by environmental changes, were observed. Measured interference patterns were demodulated using an incoherent demodulation algorithm ( ).

Figures 8(a) and 8(b) show OCT cross-sectional images of the sample at a fixed height, with the normalized complex fringe signal represented along the vertical axis, recorded with single-mode and multimode fiber systems, respectively. When using the multimode fiber for illumination, speckle is significantly reduced [Fig. 8(b)] compared to that recorded using the single-mode fiber [Fig. 8(a)].

Fig. 8

Cross-section Parallel OCT image recoded from a scattering surface using (a) single-mode and (b) multimode source fiber. Fringe intensities are normalized and correspond to a single depth scan.

064034_1_025506jbo8.jpg

Demodulated interference patterns at a fixed depth recorded with single- and multi-mode fiber systems are shown in (a) and (b) respectively. For comparison, both images were recorded at the same position on the aluminum plate, so surface micro-irregularities of the surface are identical. The image in (a) was recorded after rotating the aluminum plate 20degrees counterclockwise. The image recorded using the multimode fiber, (b), shows continuous linear fringes corresponding to tilt of the aluminum plate, whereas the image recorded using single mode fiber for illumination, (a), displays disrupted and fragmented bright grey bands over the surface due to speckle.

4.

Discussion

Light emitted from a Gaussian-Schell model source considered here has a variable degree of spatial coherence from incoherent to fully coherent. When partially spatial coherent light is launched into the interferometer, the beam may be regarded as an assembly of mutually uncorrelated finite-sized phase cells. In OCT, light input into the interferometer is divided into two beams, which illuminate the sample and the reference mirror. While mirror reflectivity in the reference path is considered constant, the sample has a complex reflectivity function and backscatters incident light. When light reflected from sample and reference paths recombine at the detection plane, cells of the beams overlap and partially interfere. Interference can occur only for light originating from the same phase cell in sample and reference paths. The speckled appearance of the images in is the effect of interference between light backscattered from different particles within a coherence area. With a decrease in the coherence area, speckle reduction is expected, as indicated in Table 1, but as Fig. 5(d) illustrates speckle still exists despite a very small spatial coherence length (σg=0.01μm) corresponding to incoherent illumination.

An interesting point when looking at the complex fringe signal [Eq. 14] as a random phasor sum is that speckle increases with the second moment of the complex fringe signal, which follows the second moment (α2¯) of amplitude (αk) , not the random phase distribution (ϕk) . The result of spatial coherence length dependence on the second moment of the complex fringe signal is shown at Fig. 6. The second moment α2¯ of the intensity fluctuation increases as the degree of coherence increases and approaches unity for a spatially coherent beam. The second moment never reaches zero even at the smallest spatial coherence length.

Fig. 6

Second moment of normalized complex fringe signal versus spatial coherence length (σg) .

064034_1_025506jbo6.jpg

Producing incoherent sources has an inherent trade-off in view of radiation efficiency and the degeneracy parameter.25 The radiant emittance of a partially spatial coherent source is expressed30 by

Eν(ρ)=Cg(ν)S(ρ,ν),
where the radiation efficiency (Cg) is a proportionality factor between radiant emittance [Eν(ρ)] and spectral density [S(ρ,ν)] , and ρ is a position vector of a radiation source point. The radiation efficiency increases monotonically with σg , from zero for a completely incoherent source to unity for a completely coherent source.30 The degeneracy parameter indicates average number of identical photons (same polarization sate) that are contained in a coherence volume. For example, use of spatially incoherent illumination with a thermal light source suffers from a low degeneracy number.23 The advantage of using a multimode fiber to produce partially spatially coherent light is high number of photons per coherence volume. The source and multimode fiber used in experiments reported here provided 0.3×109 photons per coherence volume substantially higher than that provided by a thermal source. Spatial coherence length at the tip of the multimode fiber was calculated33, 34 to be 22μm by
σg={4σ02(πDNA2λ)21}12,
where D is a core diameter of multimode fiber.

Another advantage of this method to modify coherence properties of the light source is relative simplicity over other methods including use of a ground glass plate or liquid crystals.30

In Eq. 1 we assumed amplitude of light returning from the sample is the product of the incident amplitude [Ai(ξ,ν)] and the local reflectivity [r(ξ)] , which is reasonable for the scattering object studied. In tissue, however, the amplitude of light returning from the sample will be a complex function of Ai(ξ,ν) due to multiple photon paths in the tissue and spatial variation of the refractive index of the tissue.

Speckle reduction using partially spatial coherent sources can be also understood as an incoherent mode summation. When broadband light is coupled into a multimode fiber, a number of orthogonal spatial modes are generated which are spatially and temporally coherent. Each spatial mode propagates through the multimode fiber at a distinct group velocity producing modal dispersion. After propagating a sufficient length along the fiber, spatial modes become temporally decorrelated due to differences in group velocity. Detected interference is an incoherent sum of interferograms formed between spatial modes. Because each mode forms a statistically independent speckle field, mode summation has the effect of reducing speckle by averaging. The appearance of unbroken grooves in the multimode fiber case [Fig. 9(b) ] is consistent with the speckle reduction effect resulting from superposition of many interferograms each representing mutually incoherent spatial modes.

Fig. 9

Low-coherence interferograms recorded from a scattering surface using (a) single-mode and (b) multimode source fibers.

064034_1_025506jbo9.jpg

As sample-reference path length delay increases, the delay may compensate the mode separation introduced by the multimode fiber, and the decorrelated modes become correlated again. To avoid this intermodal interference, the multimode fiber should be sufficiently long so that the modal separation is much greater than maximum sample-reference delay time in scanning.

To verify no intermodal interference was present in our system, signal to noise ratios for single mode and multimode cases were measured and compared across the full scanning range (1.6mm) of the voice coil stage. A mirror was placed in the sample path with a photoreceiver and a 100-μm pinhole in the detection path replaced the CCD camera. We express noise sources in terms of the photocurrent variance σi2 . Primary noise sources include receiver (σre2) , and shot noise (σsh2) . Intensity noise was omitted because of the small contribution from a mode locked laser.35 Receiver noise was calculated from the manufacturer’s specification. Shot noise is given by,

σsh2=2qIdcB,
where Idc represents the mean detector photocurrent, and B is the electronic detection bandwidth, and q is the charge on an electron.

Total photocurrent variance is given by

σi2=σre2+σsh2.
Measured and calculated shot noises are compared in Table 2 . The shot noise is calculated from the mean detector photocurrent Idc . Measured shot noise is comparable with calculated values.

Table 2

Noise comparison between multimode and single-mode source fiber setups.

σre σsh (multimode) σsh (singlemode)
Calculated noise 9.25μV 10.12nV 27.2nV
Measured noise 9.8μV 11.73nV 28nV

We define SNR as peak-to-peak current of the coherence function divided by the total photocurrent standard variance (σi) . SNRs were 80 and 79dB over 1.6mm of reference and sample path delays for single-mode and multimode fiber cases, respectively. The Small difference in SNR between coherent and partially spatial coherent cases indicates degradation of SNR due to the multimode source fiber was insignificant and supports the hypothesis that mode separation between adjacent modes is longer than the scanning range (1.6mm) after 100m of travel through the multimode fiber. Mode cross-coupling in the fiber was below the measurement limit over the entire scan (1.6mm) .

5.

Conclusion

We proposed and demonstrated a speckle reduction method using a partially spatial coherent source in a Michelson interferometer in a parallel detecting OCT system. Speckle in interferograms recorded using a partially spatial coherent source is substantially reduced compared to the fully coherent case. No degradation in SNR is observed, and no measurable mode cross-coupling is observed in the multimode fiber. A partially spatial coherent source is also preferable for en face imaging not only for reducing speckle but also for eliminating Airy rings in the image caused by the small field diameter of most single-mode fibers. Our simulations model a partially spatially coherent source with a Gaussian-Schell model source to indicate speckle reduction in OCT interferograms. As the physical viewpoint, our experiments explain that superposition of coherent modes has the effect of reducing speckle by averaging because each mode forms a statistically independent speckle field. The results of both simulations and experiments indicate that broadband light sources with reduced spatial coherence that provide a large number of photons per coherence volume may be effectively utilized to reduce speckle in OCT interferograms.

Acknowledgments

This research has been supported by National Institutes of Health (NIH) (R01 EB002021) and the Texas ATP.

References

1. 

D. Huang, E. A. Swanson, C. P. Lin, J. S. Schuman, W. G. Stinson, W. Chang, M. R. Hee, T. Flotte, K. Gregory, C. A. Puliafito, and J. G. Fujimoto, “Optical coherence tomography,” Science, 254 (5035), 1178 –1181 (1991). 0036-8075 Google Scholar

2. 

J. G. Fujimoto, “Optical coherence tomography for ultrahigh resolution in vivo imaging,” Nat. Biotechnol., 21 (11), 1361 –1367 (2003). https://doi.org/10.1038/nbt892 1087-0156 Google Scholar

3. 

M. E. Brezinski, G. J. Tearney, S. A. Boppart, B. E. Bouma, E. A. Swanson, J. F. Southern, and J. G. Fujimoto, “High speed catheter based OCT imaging of coronary microstructure,” Circulation, 94 (8), 1494 –1494 (1996). 0009-7322 Google Scholar

4. 

W. Drexler, U. Morgner, F. X. Kartner, C. Pitris, S. A. Boppart, X. D. Li, E. P. Ippen, and J. G. Fujimoto, “In vivo ultrahigh-resolution optical coherence tomography,” Opt. Lett., 24 (17), 1221 –1223 (1999). 0146-9592 Google Scholar

5. 

L. Vabre, A. Dubois, and A. C. Boccara, “Thermal-light full-field optical coherence tomography,” Opt. Lett., 27 (7), 530 –532 (2002). 0146-9592 Google Scholar

6. 

M. Ducros, M. Laubscher, B. Karamata, S. Bourquin, T. Lasser, and R. P. Salathe, “Parallel optical coherence tomography in scattering samples using a two-dimensional smart-pixel detector array,” Opt. Commun., 202 (1–3), 29 –35 (2002). https://doi.org/10.1016/S0030-4018(02)01073-8 0030-4018 Google Scholar

7. 

S. Bourquin, P. Seitz, and R. P. Salathe, “Two-dimensional smart detector array for interferometric applications,” Electron. Lett., 37 (15), 975 –976 (2001). https://doi.org/10.1049/el:20010669 0013-5194 Google Scholar

8. 

S. Bourquin, V. Monterosso, P. Seitz, and R. P. Salathe, “Video-rate optical low-coherence reflectometry based on a linear smart detector array,” Opt. Lett., 25 (2), 102 –104 (2000). 0146-9592 Google Scholar

9. 

Z. P. Chen, T. E. Milner, D. Dave, and J. S. Nelson, “Optical Doppler tomographic imaging of fluid flow velocity in highly scattering media,” Opt. Lett., 22 (1), 64 –66 (1997). 0146-9592 Google Scholar

10. 

D. P. Dave, T. Akkin, and T. E. Milner, “Polarization-maintaining fiber-based optical low-coherence reflectometer for characterization and ranging of birefringence,” Opt. Lett., 28 (19), 1775 –1777 (2003). 0146-9592 Google Scholar

11. 

J. F. de Boer and T. E. Milner, “Review of polarization sensitive optical coherence tomography and Stokes vector determination,” J. Biomed. Opt., 7 (3), 359 –371 (2002). https://doi.org/10.1117/1.1483879 1083-3668 Google Scholar

12. 

D. T. Miller, J. Qu, R. S. Jonnal, and K. Thorn, “Coherence garing and adaptive optics in the eye,” Proc. SPIE, 4956 65 –72 (2003). 0277-786X Google Scholar

13. 

A. I. Kholodnykh, I. Y. Petrova, K. V. Larin, M. Motamedi, and R. O. Esenaliev, “Precision of measurement of tissue optical properties with optical coherence tomography,” Appl. Opt., 42 (16), 3027 –3037 (2003). 0003-6935 Google Scholar

14. 

J. Rogowska and M. E. Brezinski, “Evaluation of the adaptive speckle suppression filter for coronary optical coherence tomography imaging,” IEEE Trans. Med. Imaging, 19 (12), 1261 –1266 (2000). https://doi.org/10.1109/42.897820 0278-0062 Google Scholar

15. 

J. Rogowska and M. E. Brezinski, “Image processing techniques for noise removal, enhancement and segmentation of cartilage OCT images,” Phys. Med. Biol., 47 (4), 641 –655 (2002). https://doi.org/10.1088/0031-9155/47/4/307 0031-9155 Google Scholar

16. 

J. M. Schmitt, “Restoration of optical coherence images of living tissue using the CLEAN algorithm,” J. Biomed. Opt., 3 (1), 66 –75 (1998). https://doi.org/10.1117/1.429863 1083-3668 Google Scholar

17. 

M. D. Kulkarni, C. W. Thomas, and J. A. Izatt, “Image enhancement in optical coherence tomography using deconvolution,” Electron. Lett., 33 (16), 1365 –1367 (1997). https://doi.org/10.1049/el:19970913 0013-5194 Google Scholar

18. 

K. M. Yung, S. L. Lee, and J. M. Schmitt, “Phase-domain processing of optical coherence tomography images,” J. Biomed. Opt., 4 (1), 125 –136 (1999). https://doi.org/10.1117/1.429942 1083-3668 Google Scholar

19. 

J. M. Schmitt, S. H. Xiang, and K. M. Yung, “Speckle in optical coherence tomography,” J. Biomed. Opt., 4 (1), 95 –105 (1999). https://doi.org/10.1117/1.429925 1083-3668 Google Scholar

20. 

N. Iftimia, B. E. Bouma, and G. J. Tearney, “Speckle reduction in optical coherence tomography by path length encoded angular compounding,” J. Biomed. Opt., 8 (2), 260 –263 (2003). https://doi.org/10.1117/1.1559060 1083-3668 Google Scholar

21. 

J. M. Schmitt, “Array detection for speckle reduction in optical coherence microscopy,” Phys. Med. Biol., 42 (7), 1427 –1439 (1997). https://doi.org/10.1088/0031-9155/42/7/015 0031-9155 Google Scholar

22. 

M. Bashkansky and J. Reintjes, “Statistics and reduction of speckle in optical coherence tomography,” Opt. Lett., 25 (8), 545 –547 (2000). 0146-9592 Google Scholar

23. 

B. Karamata, P. Lambelet, M. Laubscher, R. P. Salathe, and T. Lasser, “Spatially incoherent illumination as a mechanism for cross-talk suppression in wide-field optical coherence tomography,” Opt. Lett., 29 (7), 736 –738 (2004). https://doi.org/10.1364/OL.29.000736 0146-9592 Google Scholar

24. 

L. Buttner and J. Czarske, “Multi-mode fibre laser Doppler anemometer (LDA) with high spatial resolution for the investigation of boundary layers,” Exp. Fluids, 36 (1), 214 –216 (2004). 0723-4864 Google Scholar

25. 

L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, 664 –666 Cambridge University Press, Cambridge, UK (1995). Google Scholar

26. 

H. Ambar, Y. Aoki, N. Takai, and T. Asakura, “Fringe contrast improvement in speckle photography by means of speckle reduction using vibrating optical fiber,” Optik (Jena), 74 (2), 60 –64 (1986). 0030-4026 Google Scholar

27. 

H. Ambar, Y. Aoki, N. Takai, and T. Asakura, “Relationship of speckle size to the effectiveness of speckle reduction in laser microscope images using rotating optical fiber,” Optik (Jena), 74 (1), 22 –26 (1986). 0030-4026 Google Scholar

28. 

N. Takai and T. Asakura, “Statistical properties of laser speckles produced under illumination from a multimode optical fiber,” J. Opt. Soc. Am. A, 2 (8), 1282 –1290 (1985). 0740-3232 Google Scholar

29. 

J. W. Goodman, Introduction to Fourier Optics, 1996). Google Scholar

30. 

L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, 298 –300 Cambridge University Press, Cambridge, UK (1995). Google Scholar

31. 

T. Shirai, A. Dogariu, and E. Wolf, “Mode analysis of spreading of partially coherent beams propagating through atmospheric turbulence,” J. Opt. Soc. Am. A, 20 (6), 1094 –1102 (2003). 0740-3232 Google Scholar

32. 

M. Young, Optics and Lasers Including Fibers and Optical Waveguides, 5th ed.Springer, New York (2000). Google Scholar

33. 

M. Santarsiero, F. Gori, R. Borghi, G. Cincotti, and P. Vahimaa, “Spreading properties of beams radiated by partially coherent Schell-model sources,” J. Opt. Soc. Am. A, 16 (1), 106 –112 (1999). 0740-3232 Google Scholar

34. 

L. Buettner and J. Czarske, “A multimode-fibre laser-Doppler anemometer for highly spatially resolved velocity measurements using low-coherence light,” Meas. Sci. Technol., 12 (11), 1891 –1903 (2001). https://doi.org/10.1088/0957-0233/12/11/319 0957-0233 Google Scholar

35. 

P. Dufour, G. Rousseau, and N. McCarthy, “Optical noise reduction in a femtosecond Ti:sapphire laser,” Proc. SPIE, 4833 894 –899 (2002). 0277-786X Google Scholar
©(2005) Society of Photo-Optical Instrumentation Engineers (SPIE)
Jeehyun Kim, Donald T. Miller, Eunha K. Kim, Sanghoon Oh, Jung Hwan Oh, and Thomas E. Milner "Optical coherence tomography speckle reduction by a partially spatially coherent source," Journal of Biomedical Optics 10(6), 064034 (1 November 2005). https://doi.org/10.1117/1.2138031
Published: 1 November 2005
Lens.org Logo
CITATIONS
Cited by 89 scholarly publications and 11 patents.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Speckle

Optical coherence tomography

Multimode fibers

Optical fibers

Spatial coherence

CCD cameras

Light

Back to Top