Sunteți pe pagina 1din 10

1400

J. Opt. Soc. Am. A / Vol. 23, No. 6 / June 2006

Jesacher et al.

Spiral interferogram analysis


Alexander Jesacher, Severin Frhapter, Stefan Bernet, and Monika Ritsch-Marte
Division for Biomedical Physics, Innsbruck Medical University, Mllerstrasse 44, A-6020 Innsbruck, Austria Received July 27, 2005; revised November 25, 2005; accepted December 3, 2005; posted January 10, 2006 (Doc. ID 63660) Interference microscopy using spatial Fourier ltering with a vortex phase element leads to interference fringes that are spirals rather than closed rings. Depressions and elevations in the optical thickness of the sample can be distinguished immediately by the sense of rotation of the spirals. This property allows an unambiguous reconstruction of the objects phase prole from one single interferogram. We investigate the theoretical background of spiral interferometry and suggest various demodulation techniques based on the processing of one single interferogram or multiple interferograms. 2006 Optical Society of America OCIS codes: 070.6110, 100.2650, 180.3170.

1. INTRODUCTION
The properties of a spiral phase element as spatial lter in imaging applications has been discussed in several recent publications.17 Placing the lter in a Fourier plane of the optical system leads to strong edge contrast enhancement for amplitude and phase objects similar to that of the Nomarski, or differential interference contrast, technique.8 In a recently published paper9 we considered the application of a phase vortex to interferometry. We reported the observation of spiral-shaped interference patterns that appear when thick phase samples are examined. A special feature of such spiral interference patterns is that they contain sufcient topographic information to allow the reconstruction of the objects phase prole. Conventional methods10 typically require three different interference patterns in order to eliminate the ambiguity between a step up or down from a given contour line with respect to its neighbor. The present paper investigates the theoretical background of spiral interferometry and suggests methods for the demodulation of the coiled fringe patterns. On one hand, it is demonstrated how the overall topographic information of a sample phase prole can be evaluated from a single interferogram. This might have applications in high-speed interferometry with a single laser pulse, or in the recording of an interferometric video of a rapidly varying surface in which each video frame can be postprocessed separately. On the other handif imaging speed is not of primary importancethe setup allows highaccuracy object reconstruction from a selectable number of interferograms, which in our approach can easily be produced using a straightforward nonmechanical phasestepping method.

by another, identical lens. As indicated in Fig. 1, the phase vortex is slightly modied, i.e., its center is replaced by a circular area of constant phase shift. The reason for this modication is our intention to use the zeroth order spot of the object light eld as reference wave for producing so-called self-referenced interferograms (see Ref. 9 for details of the setup) that are captured in the plane x , y . Since the zeroth order light focuses in the center of the spiral lter, it is now phase-shifted by a constant value (determined by the phase of the central circular area) and becomes a plane reference wave after the second lens, i.e., in the image plane. In our experiments we use the LC-R 3000 reective-liquid-crystal spatial light modulator of HOLOEYE Photonics for generating the lter functions. Its small pixel size of 10 m allows us to match the diameter of the central disk area to the size of the zeroth order spot. The following mathematical considerations are concerned at rst only with the effect of pure spiral phase ltering, i.e., they neglect the mentioned modication, assuming the Fourier lter to be a perfect phase spiral. The effect of the superposition with a phase-controlled plane wave, given by the zeroth Fourier order, is considered later. According to the Fourier convolution theorem, the result of the vortex lter process can be derived by a convolution of the original object function Ein x , y with the Fourier transform of the lter:

Eout x ,y = Ein * KV x ,y =
x= y=

Ein x,y KV x x,y y dxdy. 1

2. VORTEX FILTERING
Figure 1 shows the generic experimental setup, which is a so-called 4-f system. The object of interest in the x , y plane is illuminated with sufciently coherent light. A convex lens performs a Fourier transform of the eld distribution Ein x , y . After inuencing the image by a vortex phase lter, a second Fourier transform is accomplished
1084-7529/06/061400-10/$15.00

One has to use the mirrored lter function exp i which is identical to exp i as a result of the inverting property of the 4-f system. The Fourier transform of exp i is also called the convolution kernel KV. In polar coordinates it has the form
2006 Optical Society of America

Jesacher et al.

Vol. 23, No. 6 / June 2006 / J. Opt. Soc. Am. A

1401

Fig. 1. Schematic vortex lter setup. The spiral lter is modied such that its central area is assigned a constant phase value. The gray tones of the lter correspond to respective phase shifts. Note that the coordinate system x , y is mirrored compared with the x , y system.

Fig. 2. Comparison of lter kernels K and KV for f = 0.1 m and max = 0.01 m. (a) and (b) show the absolute values, (c) and (d) the phase of the kernels. Note that the axis scaling for (c) and (d) has been changed with respect to (a) and (b) for better visualization; (e) and (f) show the real parts of the cross section dened by = / 2.

1402

J. Opt. Soc. Am. A / Vol. 23, No. 6 / June 2006

Jesacher et al.

2 KV r, = f i exp i

max

2 J1 f r d . 2 Eout P =

1 N

R2

exp i
P=0

Ein rP,

rPdrPd

P,

=0

rP=R1

J1 is the rst-order Bessel function of the rst kind, the light wavelength, and the focal length of the two lenses. The explicit analytical form of KV r , is derived in Appendix A. It is related to the eld distribution of the * LaguerreGaussian mode TEM01, which is also known as optical vortex or doughnut mode.11 In the limit max , KV simplies to i f exp i / 2 r2 .12 For comparison, Eq. (3) describes the convolution kernel of the setup shown in Fig. 1 without a vortex lter, hence representing a simple two-lens imaging system: 2 K r, = f
max

5 where P = Px , Py . Here rP , P denes a polar coordinate system with its origin in the center of the kernel (see Fig. 3) and Ein the input light eld expressed in this local coordinate system. The convolution integral has to be evalu ated for every point P of the input light eld Ein. For an investigation of the basic effects of such a con volution procedure, we expand Ein xP , yP = Ein xP , yP exp i in xP , yP in a Taylor series to rst order: Ein xP,yP Ein 0 exp i in 0 + exp i in 0 + i Ein 0 gPh 0 rP . gAm 0 rP 6

2 J0 f r d =

max

2 J1 f
maxr

=0

. 3

K r , represents the point-spread function (PSF) of a circular aperture with radius max. Comparing the convolution kernels of Eq. (2) and (3) (see also Fig. 2), one can identify the main differences to be the different orders of the Bessel functions and the vortex phase factor exp i , which causes the vortex kernel KV to be dependent. This anisotropy is responsible for the spiral interference patterns, as will be shown in the following. A descriptive way to achieve the convolution of two functions KV x , y and Ein x , y at a certain point P = Px , Py is to mirror KV at the origin, then shift it to point P, and nally integrate over the product of the shifted kernel with Ein. Consider, for simplicity, an approximated lter kernel 1 KV r, = N 0 exp i for R1 r R2

Here gAm = Ein and gPh = in describe the amplitude and phase gradient, respectively, of Ein evaluated at point P =kernel center , which is the same as the origin 0 in the local coordinate system xP , yP . Together with Eq. (5), the output light eld can, after integration (see Appendix A) nally be written Eout P exp i
in

P gAm P exp i
Ph

Am

P 7

+ iEin P gPh P exp i

P .

elsewhere

2 2 where N is a scaling factor dened as N = R2 R1 and R1, R2 are the inner and the outer radius of a coil aperture, respectively. This assumption allows the result of the convolution at point P to be derived as

Here, Am P and Ph P are the polar angles of the corresponding gradients (see Fig. 3). Relation (7) allows a qualitative examination of the vortex ltering properties: It is apparent that it consists of two terms that describe the effects of amplitude and phase variations of the input object on the lter result. The two terms are proportional to the absolute values of the gradients gAm and gPh respectively, which explains the observed strong isotropic amplication of amplitude and phase edges. The factors exp i Ph and exp i Am can be interpreted as the manifestation of gradient-dependent geometric phases in the following sense: Ph and Am were originally geometric angles that indicated the spatial direction of the respective gradients. These factors now appear in the ltered image no longer as directions, but as additional phase off-

Fig. 3. Graphical scheme to explain the convolution process for the case of a pure amplitude object. The result at a certain location is derived by shifting the (mirrored) kernel to this point and integrating over the product of the shifted kernel with Ein x , y .

Jesacher et al.

Vol. 23, No. 6 / June 2006 / J. Opt. Soc. Am. A

1403

Fig. 4. (a) Gaussian-shaped elevation as phase object. (b) After the lter process; a phase factor proportional to the geometrical direction of the local phase gradient has been added.

sets of the image wave at the appropriate positions. Similar to other manifestations of geometric phases,13 the phase offset does not depend on the magnitude of the amplitude or phase gradient, but only on its geometric characteristics, one of which in this case is the direction of the eld gradient. A consequence of the factors exp i Ph and exp i Am is that the edge amplication becomes anisotropic when interfering Eout with a plane reference wave, because they produce a phase difference of between a rising and falling edge of equal orientation. The resulting shadow effect gives the impression of a sidewise illumination in accordance with our earlier observations.7 Because of the factor i in relation (7), this pseudo-illumination shows a 90 rotation between pure amplitude and phase samples. In the following discussion we assume the object to be a pure phase sample (with a thickness in the range of the illumination wavelength or larger). Thus we can neglect the rst term in relation (7). In this case, the anisotropy in edge enhancement that emerges when superposing the ltered wavefront with an external plane reference wave enters the interference fringes: The shape of the fringes depends not only on the phase distribution in of the sample, but on the direction of the local phase gradient. Two different points within the sample showing identical values for in but unequal values for Ph lead to different interference results. In this manner it becomes clear that a closed isoline of equal phase in that surrounds a local extremum and would lead to a closed fringe of equal brightness in classical interferometry will cause an open spiral-shaped fringe in spiral interferometry, because the local phase gradients along this isoline must necessarily cover direction angles ranging from zero to 2 . The rotational direction of such a spiral fringe allows us to distinguish between local maxima and minima. In the context of interferometry, the weighting of the resulting amplitude by gPh is unwanted because of its inuence on the fringe positions. However, this effect, in our experience, is of little importance, except very close to local extrema and saddle points; moreover it can be completely eliminated by using multiple spiral images for object reconstruction (see Section 3). To get a better understanding of how the vortex lter alters phase proles, Fig. 4 shows the result of a numerical simulation. It considers a Gaussian-shaped phase sample as the object of interest (left image). As is apparent in the gure, the

vortex-ltered wavefront (right image) looks somehow similar to the original, but with an imprinted phase spiral.

3. DEMODULATION OF SPIRAL INTERFEROGRAMS


The analysis of conventional interference patterns is usually based on three interferograms, each showing the same object but taken at different values of the phase of the reference wave. The images therefore show slightly shifted fringe patterns, such that together they contain the whole topographic information of the examined sample. Mathematical combination of the three single images yields the objects surface structure modulo 2 , which can nally be unwrapped using one of several phase unwrapping algorithms. Spiral interferograms, in contrast, allow feasible object reconstruction based on only a single spiral-ltered image. To the best of our knowledge this efcient phase surface reconstruction at a glance is a novelty in interferometry. Technically, the reconstruction of objects from spiral fringes can be done in analogy to methods developed for traditional interferometry,10 since the interference spirals can be transformed into normal closed fringe patterns by a numerical reverse spiral transform (similar to the method indicated below in Fig. 8), which can then be processed in the usual way with high accuracy. However, for resolving the ambiguity between elevations and depressions in the object topography, the required complementary information can be extracted from an additional direct processing of the spiral fringe pattern. In the following we present two alternative one-image demodulation methods of spiral fringes that deliver complete quantitative information about the topography of a specimen and that can be used to provide the missing topographic information for supplementing the more precise demodulation techniques of traditional interferometry, as proposed, e.g., by Larkin.12 A. Single-Image Demodulation The demodulation techniques presented here are based on the assumption that the ltered wavefront is of the form Eout exp i in P + Ph P , with constant eld amplitude. Consequently, the fringe positions reect the local values for in + Ph, modulo 2 , which implies that

1404

J. Opt. Soc. Am. A / Vol. 23, No. 6 / June 2006

Jesacher et al.

mod in + Ph , 2 is constant at positions of maximum fringe intensity. The value of this constant can be selected in the experiment by adjusting the phase of the zeroth Fourier component (center of the spiral phase plate).9 Since gPh is always perpendicular to the tangent of the local spiral fringe, characterized by the angle tan, we obtain the relation mod in + tan , 2 = C. Setting the arbitrary constant C to zero, we may write
in = tan

other, continuously assigning height information h to each point, the height being determined by the angle of the spirals tangents. In the following, two variants of single-image demodulation methods are presented, which differ in the way they represent interference fringes by single lines. 1. Contour Line Demodulation A quite simple way to obtain the surface prole is based on processing contour lines. Fig. 5 demonstrates the process considering a practical example. Figure 5(b) shows the spiral interferogram of a deformation in a transparent glue strip. Phase modulations of such a shape emerge due to internal stresses where the rigidity of the lm is decreased by local heating. The opposite coiling directions of the spirals in the interferogram indicate that the deformation consists of an elevation adjacent to a depression. In a rst step one single contour line [Fig. 5(c)] must be constructed that is a closed line and connects points of equal intensity in Fig. 5(b). This can be done using standard image processing software. In our case, the software

up to multiples of 2 .

Scaled to length units, this yields n h=


tan

up to multiples of

n ,

where n is the difference between the refractive indices of object and surrounding medium and the illumination wavelength. According to Eq. (9), a basic way to obtain the topographic information of a pure phase sample from the interferogram is to process the spiral from one end to the

Fig. 5. Demodulation using contour lines. (a) Classical closed-fringe interferogram of a deformation in a plastic lm, (b) corresponding spiral interferogram, (c) and (d) single contour line raw and after preprocessing, respectively. The reconstructed three-dimensional shape is shown in (e) and (f).

Jesacher et al.

Vol. 23, No. 6 / June 2006 / J. Opt. Soc. Am. A

1405

Fig. 6. Demodulation based on center lines. (a) Skeleton of the spiral fringe pattern, which roughly consists of connected intensity maxima, (b) and (c) reconstructed three-dimensional shapes.

represents contour lines by an oriented array of L pairs of corresponding x , y vectors (L being the length of the line), which makes their handling quite simple, because the height assignment according to Eq. (9) can be carried out pointwise following the contour line array from its beginning to its end. Unfortunately, this process generates a systematic error, since after assigning the height information, i.e., after adding a third height dimension, the resulting curve always shows a discontinuity: Starting and ending points of the numerical processingwhich are direct neighbors in the two-dimensional contour lineare now separated in height by a step of n . This is a consequence of the number of clockwise and counterclockwise revolutions, which necessarily differ by 1 in a closed two-dimensional curve. However, this error is tolerable in many cases where an overview of the topography of an extended phase object is required with a minimum of computational expense. In addition, the error can be corrected by further processing, e.g., by cutting certain parts out of the line, as demonstrated in the example of Fig. 5. Figures 5(d)5(f) show the contour plot after smoothing the lines, the reconstructed shell, and the surface, respectively. As will be explained in Subsection 3.A.2, an alternative way of singleimage demodulation can be devised that avoids this specic systematic error. 2. Center Line Demodulation In the center line demodulation method, the spirals are represented by curves that follow the intensity maxima. They are constructed from the spiral interferogram by applying an algorithm that continuously removes pixels

from the spiral boundaries until a skeleton remains [Fig. 6(a)]. As a result, one obtains a connected line following the maxima of the spiral fringes that is further processed similar to the contour line method described in Subsection 3.A.1, i.e., the local tangential direction is calculated and transformed into height information. In contrast to the contour line method, the maximum intensity line avoids the articial phase jump. However, the calculation of the maximum intensity positions cuts a line into two parts at possible branching points of the fringes. After calculation of the height characteristics for each branch according to Eq. (9), the two parts are nally reconnected. This procedure ensures a quite accurate reproduction of the prole [see Fig. 6(b)]. In a nal step, interpolation can be used to construct a continuous surface [Fig. 6(c)]. In order to compare the still-remaining reconstruction errors of the contour line method and the center line method, Fig. 7 shows the result of a numerical simulation that assumes an input sample consisting of two Gaussian-shaped phase deformations. The topographical reconstruction based on contour lines can be taken from Fig. 7(b), which also shows the deviation from the original object. The standard deviation in this example is 1.18 rad. Generally, the emerging error is inuenced by many factors, such as specic object shape, illumination wavelength, and the interpolation method used. In the present example, the surface data between the contour lines have been acquired using cubic spline interpolation. As in the contour line case, Fig. 7(c) illustrates the error analysis of the center line method by means of a simulation. Here, the standard deviation was determined to be 0.87 rad. The center line method thus provides a more accurate

1406

J. Opt. Soc. Am. A / Vol. 23, No. 6 / June 2006

Jesacher et al.

reconstruction of the object phase prole from a single interferogram compared with the contour line method described before. This advantage, however, is at the expense of more computational effort, since a contouring algorithm is usually available in software packages, while a fringe tracking algorithm may not be. B. Multi-Image Demodulation If highest surface reconstruction accuracy is desired, a method based on processing numerous interferograms can be devised that is based on processing more than one image, each showing an interferogram captured at a different value for the phase of the reference wave. The benets include more accurate results and applicability to objects of nonuniform transmission. As will be shown in the following, nearly exact results are obtainable by including three (or more) different interferograms. The intensity of a general interference pattern can be described as I x ,y = Aobj x ,y exp i
obj

the zeroth order of the object itself, which ideally forms a plane wavefront of uniform intensity in the image plane x , y . The reference phase ref is thereby adjustable by adding appropriate phase shifts j to the center region of the lter hologram. In the case of an off-axis hologram, the center region is replaced by a blazed grating of adjustable phase offset. After acquisition of three interferograms at different reference phase values j, the reference phase information is added to the intensity patterns by multiplication with exp i j . The interferogram I1 captured at the reference phase value 1 thus yields a complex image of the I form Ic1 = I1 exp i 1 . The arithmetic mean c of the three complex images Icj yields x ,y = Ic 1
n

n j=1

Icj = ArefAobj x ,y exp i

obj

x ,y

, 11

x ,y

+ Aref exp i

ref

. 10

Here the reference wave parameters Aref and ref are assumed to be constants. In our case9 the reference wave is

provided that the reference phase values j are evenly distributed within the interval 0 , 2 , that is, n exp i j = 0. c x , y now includes the complete object I j=1 topography. The factor Aref can be determined from the three interferograms; see Appendix A.

Fig. 7. Quantitative error analysis of the single-image-based demodulation methods. The input object (a) of the simulation consists of two Gaussian-shaped phase deformations, (b) and (c) show the results of the contour line and center line reconstruction methods, respectively, together with their difference from the original phase sample. Note that the axis scalings of the error plots are different from the scalings of the phase object graphs.

Jesacher et al.

Vol. 23, No. 6 / June 2006 / J. Opt. Soc. Am. A

1407

8, consists of an inverse spiral lter process, which means numerical spatial ltering with the function exp i in order to get the true object prole, followed by image unwrapping. It compensates for all errors caused by the spiral lter, i.e., the inuence of object phase variations on the amplitude of the light eld and vice versa [see Eq. (7)]. An alternative way could again make use of contour lines: One gray level in the vortex ltered image can be chosen to achieve the height deconvolution analog to the single-image techniques. The complete topography can nally be restored by either repeating this process for every gray level or by matching the remaining parts of the image to the contour line whose height characteristics have formerly been acquired. Although this procedure would not correct the errors caused by the spiral lter, it would avoid the unwrapping process, which is in many cases a delicate task. Generally speaking, to achieve better resolution and noise suppression, the multi-image method can incorporate an arbitrary number of interferograms whose reference phase values are evenly distributed. Obtaining interferograms with different phases of the reference wave is done straightforwardly by shifting the phase of the central circular area of the spiral phase plate, resulting in revolving interference spirals. For the case of a nonholographic spiral phase element designed as an on-axis phase plate (as in Fig. 1), phase shifting could be performed merely by rotating the spiral phase element by the desired phase shifting angle. Compared with an alternative method proposed in Ref. 14 where self-referenced phase-stepping was performed in a phase-contrast microscope setup, the spiral method has the signicant advantage that the fringe contrast does not change during stepping of the phase of the zeroth Fourier order, thus allowing one to process a large number of interferograms with smoothly changing fringe positions (and equal fringe contrast) to obtain the highest precision.

4. DISCUSSION
We have investigated a Fourier ltering technique that gives an unambiguous impression of the topography of an examined object at a glance. Using the phase singularity of a vortex lter leads to open interference fringes in the form of spirals with opposite senses of rotation for optical elevations and depressions. It has been shown theoretically and experimentally that the presented method promises major advantages compared with established techniques in interferometry, in particular, interference microscopy. Additionally, various demodulation procedures were suggested based on single- and multi-interferogram demodulation. Single-image methods are potentially interesting for the examination of very fast processes, for instance, for uid dynamics imaging, because they allow three-dimensional topography reconstruction based on a movie recording the interferograms developing in time. The multi-image technique that we have introduced is based on established demodulation methods and has the advantage of higher accuracy compared with single-image demodulation.

Fig. 8. Principle of multi-image demodulation. The mean value of three complexied images is proportional to the spiral ltered object eld, but without its zeroth order. After restoration of the missing eld amplitude, the spiral back transformation is accomplished. Finally, the original phase distribution is restored by using a standard phase unwrapping algorithm.

However, strictly speaking, Eq. (11) contains the complete object only for interferograms created by a separate external reference wave. In the case of self-referenced interferometry, we must keep in mind that the objects zeroth order itself represents the reference beam. According to Eq. (10), this means that Aobj exp i obj describes the object without its zeroth order. Consequently, Aref has to be added for a complete image reconstruction. So far, the demodulation does not exist in the original sample, but in its vortex ltered image (see Fig. 8). At this point, different ways of further numerical image processing are possible. The rst one, which is explained in Fig.

1408

J. Opt. Soc. Am. A / Vol. 23, No. 6 / June 2006

Jesacher et al.

APPENDIX A
1. Derivation of the Spiral Kernel The convolution kernel basically is a Fourier transform of the lter function exp i 15: 1 KV x,y = f exp i
Aperture

Eout P

Ein P N exp i + cos +i


P

R2

exp i
P=0

rPdrPd

rP=R1

in

P gAm P N

exp i
P 2 rP

2 rP

Am

P drPd

Ein P gPh P N
P

2 exp i f x +y d d , A1

exp i
P.

cos

Ph

P drPd

A7

where is the light wavelength, f is the focal length of the lens, and the aperture is a circularly symmetric disk. The structure of the vortex lter suggests the introduction of polar coordinates x = r cos , y = r sin , = cos , = sin , A2 which lead (after a simplication using a trigonometric sum formula) to exp i KV r, = f 2 exp i f r cos d d . A3
max 2

Because of the integration over exp i P , the rst term equals zero. The remaining terms can be simplied using the exponential forms of the cosine functions: exp i Eout P
in

P gAm P

2N exp i2
P

exp i

Am

2 rPdrPd

rP

+ exp i exp i

Am

P
P

2 rPdrPd rP

=0

=0

+i

Ein P gPh P 2N exp i2

exp i

Ph

2 rPdrPd

Equation (A3) can nally be expressed as 2 KV r, = f i exp i


=0 max

rP

2 J1 f r d , A4

+ exp i

Ph

P
P

2 rPdrPd rP

A8

using Eq. (A5), which gives an integral representation of Bessel functions of the rst kind16: in Jn z = 2
2

The integrals containing exp i2 P again yield zero. Integration of the remaining terms nally leads to the form
3 3 1 R2 R1 2 2 3 R2 R1

Eout P exp iz cos exp in d . A5

exp i

in

P gAm P exp i
Ph

Am

P A9

=0

+ iEin P gPh P exp i

In its integrated form, the kernel has the form1


max

2 Here, the scaling factor N has been replaced by R2 2 R1 .

2 J1 2 f r r H0

2 f r 3. Multiple-Image Demodulation The intensity distribution of a general interferogram can be written as A6 I = Aobj exp i
obj

KV r,

= i exp i 2 J0 f r

2r H1

+ Aref exp i

ref obj

where H0 and H1 are Struve functions of zero and rst order, respectively. 2. Derivation of the Vortex Filter Result Inserting the rst-order approximation of relation (6) into Eq. (5) yields

2 2 = Aobj + Aref + AobjAref exp i

ref

+ exp i

obj

ref

A10 conse-

and the related complex image Ic = I exp i quently as

ref

Jesacher et al.
2 2 Ic = Aobj + Aref exp i

Vol. 23, No. 6 / June 2006 / J. Opt. Soc. Am. A

1409

ref obj

+ AobjAref exp i exp i2


ref

obj

REFERENCES
A11
1. 2. S. N. Khonina, V. V. Kotlyar, M. V. Shinkaryev, V. A. Soifer, and G. V. Uspleniev, The phase rotor lter, J. Mod. Opt. 39, 11471154 (1992). Z. Jaroszewicz and A. Kolodziejczyk, Zone plates performing generalized Hankel transforms and their metrological applications, Opt. Commun. 102, 391396 (1993). G. A. Swartzlander, Peering into darkness with a vortex spatial lter, Opt. Lett. 26, 497499 (2001). K. Crabtree, J. A. Davis, and I. Moreno, Optical processing with vortex-producing lenses, Appl. Opt. 43, 13601367 (2004). J. A. Davis, D. E. McNamara, D. M. Cottrell, and J. Campos, Image processing with the radial Hilbert transform: theory and experiments, Opt. Lett. 25, 99101 (2000). S. Frhapter, A. Jesacher, S. Bernet, and M. Ritsch-Marte, Spiral phase contrast imaging in microscopy, Opt. Express 13, 689694 (2005). A. Jesacher, S. Frhapter, S. Bernet, and M. Ritsch-Marte, Shadow effects in spiral phase contrast microscopy, Phys. Rev. Lett. 94, 233902 (2005). M. R. Arnison, K. G. Larkin, C. J. R. Sheppard, N. I. Smith, and C. J. Cogswell, Linear phase imaging using differential interference contrast microscopy, J. Microsc. (Oxford) 214, 712 (2004). S. Frhapter, A. Jesacher, S. Bernet, and M. Ritsch-Marte, Spiral interferometry, Opt. Lett. 30, 19531955 (2005). D. W. Robinson and G. T. Reid, eds., Interferogram Analysis: Digital Fringe Pattern Measurement Techniques (IOP, 1993). S. Sundbeck, I. Gruzberg, and D. G. Grier, Structure and scaling of helical modes of light, Opt. Lett. 30, 477479 (2005). K. G. Larkin, D. J. Bone, and M. A. Oldeld, Natural demodulation of two-dimensional fringe patterns. I. General background of the spiral phase quadrature transform, J. Opt. Soc. Am. A 18, 18621870 (2001). J. Anandan, J. Christian, and K. Wanelik, Resource Letter GPP-1: Geometric Phases in Physics, Am. J. Phys. 65, 180185 (1997). A. Y. M. Ng, C. W. See, and M. G. Somekh, Quantitative optical microscope with enhanced resolution using a pixelated liquid crystal spatial light modulator, J. Microsc. (Oxford) 214, 334340 (2004). G. O. Reynolds, J. B. Develis, and B. J. Thompson, The New Physical Optics Notebook: Tutorials in Fourier Optics (SPIE, 1989). M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions (Dover, 1970).

+ AobjAref exp i

The mean value c of three images Ic that differ only by I their reference phase is = Ic 1 3
2 2 Aobj + Aref exp i

+ exp i 1

+ exp i

3.
3

4.
obj

+ AobjAref exp i + exp i2


2

+ AobjAref exp i 3
3

obj

exp i2

5.

+ exp i2

A12
6. 7. 8.

If the phase values n are now equally distributed within n the interval 0 , 2 , i.e., j=1 exp i j = 0, the terms in square brackets give zero, and the expression of Eq. (11) results. Together with the mean value of the three interferograms, = I 1 3
n 2 2 In = Aobj + Aref ,

A13

9. 10. 11. 12.

it can nally be shown that the reference amplitude is given by I Aref = /2 /2 2 2 I Ic


1/2

A14

Note the ambiguity of Aref due to the sign of the inner square root.

13. 14.

ACKNOWLEDGMENTS
The authors thank the Austrian Science Foundation (project P18051-N02) as well as the Austrian Academy of sciences for support (A. J.). Corresponding author S. Bernets e-mail address is stefan.bernet@uibk.ac.at.

15. 16.

S-ar putea să vă placă și