11
Hu et al. Light: Science & Applications (2020)9:119 Ofcial journal of the CIOMP 2047-7538 https://doi.org/10.1038/s41377-020-00362-z www.nature.com/lsa ARTICLE Open Access Ef cient full-path optical calculation of scalar and vector diffraction using the Bluestein method Yanlei Hu 1,2 , Zhongyu Wang 1 , Xuewen Wang 3 , Shengyun Ji 1 , Chenchu Zhang 4 , Jiawen Li 1 , Wulin Zhu 1 , Dong Wu 1 and Jiaru Chu 1 Abstract Efcient calculation of the light diffraction in free space is of great signicance for tracing electromagnetic eld propagation and predicting the performance of optical systems such as microscopy, photolithography, and manipulation. However, existing calculation methods suffer from low computational efciency and poor exibility. Here, we present a fast and exible calculation method for computing scalar and vector diffraction in the corresponding optical regimes using the Bluestein method. The computation time can be substantially reduced to the sub-second level, which is 10 5 faster than that achieved by the direct integration approach (~hours level) and 10 2 faster than that achieved by the fast Fourier transform method (~minutes level). The high efciency facilitates the ultrafast evaluation of light propagation in diverse optical systems. Furthermore, the region of interest and the sampling numbers can be arbitrarily chosen, endowing the proposed method with superior exibility. Based on these results, full-path calculation of a complex optical system is readily demonstrated and veried by experimental results, laying a foundation for real-time light eld analysis for realistic optical implementation such as imaging, laser processing, and optical manipulation. Introduction Diffraction is a classic optical phenomenon accounting for the propagation of light waves. The efcient calcula- tion of light diffraction is of signicant value toward the real-time prediction of light elds in microscopy 1 , laser fabrication 25 , and optical manipulation 6,7 . The diffrac- tion of electromagnetic (EM) waves can be cataloged into scalar diffraction and vector diffraction according to the validation of different approximation conditions. Scalar diffraction considers only the scalar amplitude of one transverse component of either the electric or the mag- netic eld with certain simplications and approxima- tions 8 . Scalar diffraction can yield sufciently accurate results if the diffracting aperture and observing distance are both far larger than a wavelength, which is most valid for optical systems with a low numerical aperture (NA). For high-NA optical systems, polarization effects play a paramount role near the focal spot, and thus, vector dif- fraction must be adopted for light eld tracing 911 . Although mathematical expressions for optical diffrac- tions have been presented authoritatively for ages, fun- damental breakthroughs have rarely been achieved in diffraction computations. The direct integration method was rst used to calculate both scalar and vector diffrac- tion 1214 . However, the point-by-point calculation fashion renders the computation extremely tedious and inef- cient. Fast Fourier transform (FFT)-based algorithms have been developed to perform fast calculations of light dif- fraction 1519 . However, these methods can generate only the light eld distribution within a xed region of interest (ROI) and sampling numbers (i.e., resolution) determined by the intrinsic characteristic of the Fourier transform (FT), lacking exibility in computing the desired local distribution with variable sampling intervals. Therefore, © The Author(s) 2020 Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the articles Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the articles Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/. Correspondence: Jiawen Li ([email protected]) or Dong Wu ([email protected]) 1 CAS Key Laboratory of Mechanical Behavior and Design of Materials, Key Laboratory of Precision Scienti c Instrumentation of Anhui Higher Education Institutes, Department of Precision Machinery and Precision Instrumentation, University of Science and Technology of China, Hefei 230026, China 2 Department of Mechanical Engineering and Department of Civil and Environmental Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA Full list of author information is available at the end of the article 1234567890():,; 1234567890():,; 1234567890():,; 1234567890():,;

Efficient full-path optical calculation of scalar and ...mane.ustc.edu.cn/uploadfile/2020/0813/20200813090006681.pdfthe versatile computation of optical diffraction in an efficient

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • Hu et al. Light: Science & Applications (2020) 9:119 Official journal of the CIOMP 2047-7538https://doi.org/10.1038/s41377-020-00362-z www.nature.com/lsa

    ART ICLE Open Ac ce s s

    Efficient full-path optical calculation of scalar andvector diffraction using the Bluestein methodYanlei Hu1,2, Zhongyu Wang1, Xuewen Wang3, Shengyun Ji1, Chenchu Zhang4, Jiawen Li1, Wulin Zhu1, Dong Wu1 andJiaru Chu1

    AbstractEfficient calculation of the light diffraction in free space is of great significance for tracing electromagnetic fieldpropagation and predicting the performance of optical systems such as microscopy, photolithography, and manipulation.However, existing calculation methods suffer from low computational efficiency and poor flexibility. Here, we present afast and flexible calculation method for computing scalar and vector diffraction in the corresponding optical regimesusing the Bluestein method. The computation time can be substantially reduced to the sub-second level, which is 105

    faster than that achieved by the direct integration approach (~hours level) and 102 faster than that achieved by the fastFourier transform method (~minutes level). The high efficiency facilitates the ultrafast evaluation of light propagation indiverse optical systems. Furthermore, the region of interest and the sampling numbers can be arbitrarily chosen,endowing the proposed method with superior flexibility. Based on these results, full-path calculation of a complex opticalsystem is readily demonstrated and verified by experimental results, laying a foundation for real-time light field analysis forrealistic optical implementation such as imaging, laser processing, and optical manipulation.

    IntroductionDiffraction is a classic optical phenomenon accounting

    for the propagation of light waves. The efficient calcula-tion of light diffraction is of significant value toward thereal-time prediction of light fields in microscopy1, laserfabrication2–5, and optical manipulation6,7. The diffrac-tion of electromagnetic (EM) waves can be cataloged intoscalar diffraction and vector diffraction according to thevalidation of different approximation conditions. Scalardiffraction considers only the scalar amplitude of onetransverse component of either the electric or the mag-netic field with certain simplifications and approxima-tions8. Scalar diffraction can yield sufficiently accurate

    results if the diffracting aperture and observing distanceare both far larger than a wavelength, which is most validfor optical systems with a low numerical aperture (NA).For high-NA optical systems, polarization effects play aparamount role near the focal spot, and thus, vector dif-fraction must be adopted for light field tracing9–11.Although mathematical expressions for optical diffrac-tions have been presented authoritatively for ages, fun-damental breakthroughs have rarely been achieved indiffraction computations. The direct integration methodwas first used to calculate both scalar and vector diffrac-tion12–14. However, the point-by-point calculation fashionrenders the computation extremely tedious and ineffi-cient. Fast Fourier transform (FFT)-based algorithms havebeen developed to perform fast calculations of light dif-fraction15–19. However, these methods can generate onlythe light field distribution within a fixed region of interest(ROI) and sampling numbers (i.e., resolution) determinedby the intrinsic characteristic of the Fourier transform(FT), lacking flexibility in computing the desired localdistribution with variable sampling intervals. Therefore,

    © The Author(s) 2020OpenAccessThis article is licensedunder aCreativeCommonsAttribution 4.0 International License,whichpermits use, sharing, adaptation, distribution and reproductionin any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if

    changesweremade. The images or other third partymaterial in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to thematerial. Ifmaterial is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtainpermission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

    Correspondence: Jiawen Li ([email protected]) or Dong Wu ([email protected])1CAS Key Laboratory of Mechanical Behavior and Design of Materials, KeyLaboratory of Precision Scientific Instrumentation of Anhui Higher EducationInstitutes, Department of Precision Machinery and Precision Instrumentation,University of Science and Technology of China, Hefei 230026, China2Department of Mechanical Engineering and Department of Civil andEnvironmental Engineering, Massachusetts Institute of Technology, Cambridge,MA 02139, USAFull list of author information is available at the end of the article

    1234

    5678

    90():,;

    1234

    5678

    90():,;

    1234567890():,;

    1234

    5678

    90():,;

    www.nature.com/lsahttp://creativecommons.org/licenses/by/4.0/mailto:[email protected]:[email protected]

  • the versatile computation of optical diffraction in anefficient and flexible fashion is highly demanded for wideapplications.In addition, scalar and vector diffractions are separately

    analyzed in conventional studies because different integralformulas are needed for each case. However, in mostpractical apparatuses, scalar and vector diffractions co-exist for different parts of the optical system. For example,in typical systems for optical microscopy, fabrication andmanipulation, a monochromatic beam propagates over along distance by passing optical elements such as focusinglenses, expanders, and collimators before entering anobjective lens with a high NA. For the preceding partwhere the paraxial condition is valid, scalar diffraction issatisfactory for the light propagation evaluation. For thepart behind the high-NA objective that meets the Debyeapproximation, vector diffraction is required for theaccurate evaluation of the light propagation by taking intoaccount each polarization component and non-paraxialpropagation of light as well as apodization function ofoptical systems. Therefore, a facile and efficient methodwith the capacity for light propagation calculation alongthe entire optical path, which is termed full-path calcu-lation, is highly desired for the comprehensive analysis ofnumerous realistic application scenarios.Here, we propose an efficient full-path calculation

    method by exploring the mathematical similarities inscalar and vector diffraction. The scalar and vector dif-fraction are both expressed using the highly flexibleBluestein method. A fast light field evaluation over theentire optical path is achieved with arbitrarily definedROIs and sampling numbers. This paper is organized asfollows: first, the integral formulas for scalar and vectordiffraction are revisited and deduced in FT forms. Second,the Bluestein method is utilized and reformed to com-pletely supplant the FT in a more flexible fashion. Basedon this, optical diffractions are evaluated with designatedROIs and sampling numbers. Third, representativeexamples are given for both scalar and vector diffractionto demonstrate the improvement in efficiency and flex-ibility. Finally, full-path light tracing of a laser holographicsystem is presented with unprecedented computationspeed and agrees well with the experimental results,showcasing the superior ability of the Bluestein-baseddiffraction calculation. The proposed method holds greatpromise in the universal applications of optical micro-scopy, fabrication, and manipulation.

    ResultsScalar and vector diffraction integral in the form of aFourier transformFor scalar diffraction, as shown in Fig. 1a, the electric

    field at a point (x, y, z) in the Cartesian coordinates can beobtained based on the Huygens–Fresnel principle

    and expressed by the Rayleigh–Sommerfeld diffractionintegral20:

    E x; y; zð Þ ¼ � iλ

    Z ZΩE0 u; v; 0ð Þ ´ exp ikrð Þr ´ cos θ dudv

    ð1Þ

    where r ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffix� uð Þ2þ y� vð Þ2þz2

    qis the distance

    between the source point and the observation point ofinterest. k= 2π/λ is the wavenumber. In the condition ofthe Fresnel approximation with a Fresnel number F ≥ 1,the third term and higher orders in the Taylor expression

    of r can be ignored, that is, r � z þ x�uð Þ2þ y�vð Þ22z . In thedenominator of Eq. (1), r can be further approximatedwith only the first term (r ≈ z). Moreover, the paraxialapproximation ensures cosθ ≈ 1. In this way, the complexelectric field can be described by the Fresnel diffractionintegral:

    E x; y; zð Þ ¼ exp ikzð Þiλz

    Z ZΩE0 u; v; 0ð Þ

    ´ expik2z

    x� uð Þ2þ y� vð Þ2� �� �

    dudv

    ð2Þwhich can be further rewritten as:

    E x; y; zð Þ ¼exp ikzð Þ ´ exp ik x2þy22z

    � �iλz

    Z ZΩE0 u; v; 0ð Þ

    ´ expiπλz

    u2 þ v2

    � �

    ´ exp � 2iπλz

    xuþ yvð Þ� �

    dudv

    ð3Þ

    Here, we define:

    F0 ¼exp ikzð Þ ´ exp ik x2þy22z

    � �iλz

    ð4Þ

    F ¼ exp iπλz

    u2 þ v2

    � �

    ð5Þ

    Therefore, the integral Eq. (3) can be expressed in termsof the two-dimensional FT:

    E ¼ F0 ´F E0 ´ Fð Þ ð6Þ

    here F represents the two-dimensional FT. Moreover, aswith the other type of scalar diffraction, Fraunhoferdiffraction in the far field can be expressed byE ¼ F0 ´F E0ð Þ, which can be regarded as a special caseof Fresnel diffraction passing through a converging lens.

    Hu et al. Light: Science & Applications (2020) 9:119 Page 2 of 11

  • Therefore, scalar diffraction can be computed across thexy-plane using an FT-based approach.Scalar diffraction can be used to effectively compute the

    complex amplitude distribution of many optical systemswith a few approximations, as described above. However,it is known that the polarization components are changeddue to large refractivity after passing through a high-NAnon-paraxial system, and scalar diffraction is incapable ofachieving proper results. The vectorial Debye diffractionintegral, established by Richards and Wolf21, has to beadopted to analyze the complex EM field of eachpolarization component (Supplementary InformationSection 1). The optical layout is shown in Fig. 1b.Due to the refraction of the non-paraxial tight focusing

    system, the electric field components (polarization com-ponents e!s and e!p) on the entrance pupil Pe are trans-formed into a spherical reference surface Pr ( e

    !s, e!

    th, ande!r). The transformation can be expressed in Cartesiancoordinates as20:

    E!

    r ¼ A0ffiffiffiffiffiffiffiffiffifficos θ

    p´M ´ E

    !i ð7Þ

    M is the transform matrix of the polarization from theentrance surface to the converging spherical surface.A0

    ffiffiffiffiffiffiffiffiffifficos θ

    pis the apodization factor accounting for the

    energy conservation. The propagation of the electric fieldfrom the reference surface Pr to the imaging point p (x, y, z)

    near the focus is expressed by the Debye integral:

    E!¼ � iC

    λ

    Z ZΣE!

    r ´ exp i kzz � kxx� kyy � �

    dΣ ð8Þ

    The definition of k!

    rcan be found in SupplementaryInformation Section 1. By performing the integration overthe planar surface Pe instead of the surface Pr (Supple-mentary Information Section 2):

    E!¼ � iC

    λ

    Z ZΩ

    E!

    r ´ exp ikzzð Þ=cos θh i

    ´ exp �i kxxþ kyy � �

    dkxdky ð9Þ

    which can be rewritten in the form of an FT:

    E!

    x; y; zð Þ ¼ � iCλF E

    !r ´ exp ikzzð Þ=cos θ

    h i

    ¼ � iCλF M ´ E

    !i ´ exp ikzzð Þ=

    ffiffiffiffiffiffiffiffiffifficos θ

    ph i

    ð10Þ

    In brief, both scalar diffraction and vector diffractioncan be expressed by the FT. FFT algorithms in moderncomputer systems allow for fast and accurate calculations.The similarity between these two diffractions is obvious

    X

    Y

    Z

    U

    V

    q (u,v,0 )

    p (x,y,z )

    O

    X

    Y

    Z

    O�

    ep

    es es �

    er

    E i

    kikr

    Er

    Pi Pr

    a

    b

    Scalar diffraction

    Vector diffraction

    Ω Σ

    p (x,y,z )P

    r

    Fig. 1 Illustrative diagrams of scalar and vector diffraction. a Geometry for scalar diffraction calculation. b Geometry for vector diffraction calculation

    Hu et al. Light: Science & Applications (2020) 9:119 Page 3 of 11

  • from a mathematical point of view: the vector diffractionintegral is equivalent to the scalar Fraunhofer diffractionin the case of a low-NA optical system where 1/cosθ ≈ 1.Although the FFT-based optical calculation is much

    faster than the direct integration method, it results ininevitable drawbacks: the resultant output field has a fixedtransverse dimension and unchangeable sampling num-bers determined by the dimension and sampling size ofthe input aperture for a given distance. The dimension ofthe output field is:

    Dm ¼ λdps ð11Þwhere d is the distance between the input aperture andoutput plane. ps is the sampling size of the inputaperture. The sampling numbers of the output planeare rigidly equivalent to those of the input aperture.The restriction is brought about by the intrinsiccharacteristic of the FT and greatly limits the flexibilityin light propagation calculations. For example, theinput aperture must be enormously expanded with theaid of the zero-padding approach when a small portionof the output plane is required with high resolution,which inevitably leads to a large increment of thecomputation time.

    Bluestein method to compute Fourier transform witharbitrary ROI and sampling resolutionRegarding mathematics, to achieve the required band-

    width and resolution in the frequency domain, theappropriate zero-padding operation is needed to extendthe dimension of the original input sequence15. For mostapplications in laser manipulation and lithography, only asmall fraction of the output field with high resolution isneeded to obtain sufficient details, resulting in largeamounts of zero-padding. This results in a severe waste ofresources, as most of the results are discarded. Theoperation of the zero-padding inevitably increases thecomputation time and the demand for memory usage.Moreover, the resultant output region remainsunchangeable, greatly limiting its potential in practicalapplications. Here, the Bluestein method is adopted toevaluate the scalar and vector diffraction calculations. TheBluestein method is an elegant method conceived by L.Bluestein22 and further generalized by L. Rabiner et al.23

    that is capable of performing more general FTs at arbi-trary frequencies as well as boosting the resolution overthe full spectrum. The Bluestein method offers us aspectral zoom operation with high resolution and arbi-trary bandwidth. This advantage is enabled by computingthe z-transform along spiral contours in the z-plane for aninput sequence (Supplementary Information Section 3and Fig. S1). The computational complexity is O[(M+N)log2(M+N)], manifesting an FFT algorithm. The method

    is based on the z-transform along a spiral contour in thez-plane defined by A and W:

    X m½ � ¼XN�1n¼0

    x n½ � ´ z�n

    z¼A ´W�m

    ¼XN�1n¼0

    x n½ � ´A�n ´Wmn

    ð12Þhere m ¼ 0; � � � ;M � 1½ �.M is the length of the transform.N is the length of input sequence. A specifies the complexstarting point of the z-plane spiral contour of interest, andW specifies the complex scalar describing the complexratio between points along the contour. Note that the caseof A= 1, W= exp(−i2π/N), and M=N corresponds tothe discrete Fourier transform (DFT), which computes thez-transform along the unit circle with a finite duration.More generally, the method can be used to calculate theDFT between an arbitrary starting point f1 and endingpoint f2 (i.e., the tuneable frequency bandwidth relative tothe total frequency range fs) with arbitrary samplingnumbers M.The practical implementations of the Bluestein

    method for enhanced DFT computation deserve addi-tional comments. First, a 2D FT is needed for the com-putation of both scalar and vector diffraction. TheBluestein method should be adopted in both the columnand the row dimensions to fulfill this requirement. Sec-ond, the Bluestein method internalizes padding of theinput array with zeros at the tail. However, symmetriczero-padding around the input array is needed for thesimulation of realistic optical systems. Third, an addi-tional operation is needed to shift the zero-frequencycomponent to the center of the array before and after theDFT to eliminate the high-frequency oscillation in thephase information. To address these issues, the defini-tion of parameters A and W should be rearranged, andphase shifting factor Pshift should be included at the endof the calculation (see Supplementary Information Sec-tion 3 and Figs. S2–S4).By performing these adjustments, the Bluestein

    method can be developed as a fast approach for lightdiffraction calculation with superior flexibility: it allowsfor the selection of arbitrary segments in the imagingplane with arbitrary resolution, providing competitiveefficiency and flexibility over direct integration and theFFT methods.

    Fast numerical implementation of the Bluestein method inscalar Fresnel diffractionFigure 2 illustrates the scalar calculation with a para-

    digm of the converging spherical wave propagation, whichis generated by a plane wave passing through a convexlens. The phase profile of the lens is shown in Fig. 2a,which is equivalent to the phase plate after being wrappedbetween 0 and 2π (Fig. 2b). The optical configuration is

    Hu et al. Light: Science & Applications (2020) 9:119 Page 4 of 11

  • sketched in Fig. 2c, with the parameters λ= 800 nm,f= 600mm, and D= 8.64 mm. Figure 2d, e shows theoptical field distribution in the focal plane in terms of theintensity and phase. Figure 2f, g shows the cross-sectionalintensity and phase distributions in the light propagationdirection. The corresponding line plots of the intensityand phase are given in Fig. 2h–k. A comparison betweenthe Bluestein method and traditional direct integrationand FFT methods is also made, from which we can seeexcellent agreements. It is revealed that the Bluestein

    method can calculate the scalar light diffraction with highaccuracy.The Bluestein method has a superior advantage in the

    computation time cost over the direct integration andFFT methods. Due to the tedious point-by-point calcu-lation method, the direct integration method is associatedwith two cycling loops, and the computation timeincreases drastically with the calculation points of thetarget plane (with a computational complexity of O (M2 ×N2)). For the case of the FFT method, a zero-padding

    ~13.7 h

    68 s

    0.67 s

    xy plane

    a120

    80

    Pha

    se p

    rofil

    e (r

    ad)

    Inte

    nsity

    (a.

    u.)

    Com

    puta

    tion

    time

    (s)

    Com

    puta

    tion

    time

    (s)

    Com

    puta

    tion

    time

    (s)

    Inte

    nsity

    (a.

    u.)

    Pha

    se (

    rad)

    Pha

    se (

    rad)

    40

    0

    0.1

    0.75

    1

    0.5

    0.25

    0.75

    1

    0.5

    0.25

    1.5

    0 0

    3

    –3–1.5

    0

    –0.1

    1.02

    1.04

    2

    –2

    –4

    0

    0.8

    0.6

    0.4

    0.2

    0.0

    1

    0

    –1

    –2

    0.8

    0.6

    0.4

    0.2

    0.0

    60

    4.96 × 104

    80

    60

    40

    20

    0

    3

    2

    10 100 1000

    1

    40

    20

    0

    0.1

    0

    –0.1

    0.1

    0

    –0.1

    0.1

    0

    –0.1–0.1

    –0.10

    10 100

    x (points)

    x (points)

    1000

    Dire

    ctFF

    T

    Blue

    stein

    –0.05 0.00 0.05 0.10

    0 0.1

    DirectFFTBiuestein

    Direct

    FFT

    Biuestein

    DirectFFTBiuestein

    DirectFFTBiuestein

    DirectFFTBiuestein

    –0.1 0 0.58 0.6 0.62 0.58 0.6 0.620.1

    –4 –2 0

    x (mm)

    x (mm)

    –0.10 –0.05 0.00 0.05 0.10 0.58 0.59 0.60 0.61 0.62

    x (mm) z (m)

    0.58 0.59 0.60 0.61 0.62

    z (m)

    2 4

    b c

    f = 600 mm

    D =

    8.6

    4 m

    m

    Intensity-XY Intensity-YZPhase-XY Phase-YZd e f g

    h i j k

    l m~85 d

    .8 h2

    53 s

    2.8 h

    2 h

    83 s

    Volumetric 3D

    yz plane

    n

    Dire

    ctFF

    T

    Blue

    stein

    7.4 × 106

    1.2 × 104

    9.0 × 103

    6.0 × 103

    3.0 × 103

    0.0

    Com

    puta

    tion

    time

    (s)

    Fig. 2 Scalar calculation of the converging spherical wave. a Phase profiles of the convex lens (gray line) and the corresponding phase plate (redline). b 3D rendered diagram of the phase plate. c Illustration of the optical setup. d Intensity and e phase distributions in the focal plane (z=600mm). f Intensity and g phase distributions in the longitudinal direction. h–k Line plots corresponding to (d–g), calculated using three differentmethods. l Dependence of the computation time on the number of sampling points in one dimension. An incident light field with sampling pointsof 1080 × 1080 and an interval of 8 μm (i.e., width of 8.64 mm) is fixed for each calculation (the same hereinafter unless otherwise specified).m Comparison of the computation time for the light field in the xy-plane using different methods. Here, the target region with a width of 0.2 mm isfixed with sampling points of 1080 × 1080. n Comparison of the computation time for the light field in volumetric three dimensions and the cross-sectional yz-plane using different methods. Here, 150 sliced layers are calculated

    Hu et al. Light: Science & Applications (2020) 9:119 Page 5 of 11

  • operation is needed to fulfill the requirement for the pre-set target sampling numbers, resulting in a rapid increasein computation time with the sampling points. As shownin Fig. 2l, with the increase in the sampling points alongone coordinate axis, the Bluestein method exhibits itsobvious superiority compared with the other two meth-ods. This advantage makes the method particularlyapplicable to scenarios where large sampling points areneeded, such as high-resolution microscopy. For the casein Fig. 2d, e, where the sampling points in the entrancepupil and output field are the same (M=N= 1080) andthe ROI is 0.2 × 0.2 mm, the computational cost is ~13.7 hfor the direct integration method, making it unsuitable forpractical applications. For the FFT method, the compu-tational cost is improved to 68 s, as shown in Fig. 2m. Incomparison, the computation time is only 0.67 s using ourproposed Bluestein method, which is 105 and 102 timesless than those of the direct integration method and FFTmethod, respectively. The three-dimensional volumetriclight field (Supplementary Information Fig. S5) can bereconstructed using cross-sectional light fields by calcu-lating the lateral planes layer by layer. As depicted inFig. 2n, the computation time for the direct method isexcessively long to obtain the volumetric light field(~85 days). It takes 2 h to calculate the cross-sectionallight field in the longitudinal yz-plane. By using the FFTmethod, the computational cost is the same (2.8 h) forboth the volumetric and cross-sectional light fieldsbecause the ROI cannot be tuned due to the intrinsiccharacteristic of the FT. Owing to the fast computationproperty of the Bluestein method, calculation of the 3Doptical field can be accomplished in

  • 22.78 min is needed for the sliced yz-plane. An acceptabletime (2.88 s) is needed for the FFT method to calculate thexy-plane. However, an impractical 280.4 s is needed toobtain the light distribution in the volumetric threedimensions and the two-dimensional yz-plane. In con-trast, only 0.2 s is consumed by the Bluestein method forcalculation in the xy-plane. Moreover, only 9.34 and12.19 s is needed to achieve the 2D cross-sectional and 3Dvolumetric light fields. Note that the computation timeincreases much more quickly with the sampling numbersof the ROI for the direct method and FFT method thanfor the Bluestein method, e.g., more than 10 days areneeded for the direct method and 126.5 s is needed for theFFT method to acquire transverse light distributions inthe xy-plane when the number of sampling pointsincreases to ~106 (1080 × 1080), while only 1.78 s is nee-ded for the Bluestein method, which is five orders ofmagnitude less than that needed for the direct methodand 102 times less than that for the FFT method.

    Full-path optical calculation with superior flexibility andefficiencyAs discussed above, both the scalar and vector diffrac-

    tion can be efficiently calculated by the Bluestein method.Based on this, the full-path optical calculation and tracingcan be performed with high flexibility and efficiency.Figure 5a illustrates a representative optical layout forlaser holographic processing and holographic manipula-tion. This setup can be further adopted for two-photonscanning confocal microscopy. Here, a phase-only spatiallight modulator (SLM, Holoeye Pluto NIR-II, resolution:1920 × 1080) is used to modulate the wavefront of thelaser by loading a predesigned CGH. A combination of ahalf-wave plate and polarization beam splitter is utilizedto attenuate the laser power. A 4f configuration consistingof Lens 1 (f= 600mm) and Lens 2 (f= 200mm) is placedbetween the SLM and aplanatic objective (100×, NA: 1.4).It is a typical optical system involving both scalar dif-fraction and vector diffraction during light propagation.

    FT lensCGH ReconstructuionDigital CGHa

    b

    c d e

    g h

    L = 24 mm L = 8 mm L = 2 mm

    Inte

    nsity

    0.8

    0.6

    0.4

    0.2

    3

    2

    1

    0

    –1

    –2

    –3

    3

    2

    1

    0

    –1

    –2

    –3

    3

    2

    1

    0

    –1

    –2

    –3

    3

    2

    1

    0

    –1

    –2

    –3

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    Pha

    se

    L = 16 mm

    f

    i j

    Fig. 3 Scalar calculation of the CGH-modulated light wave. a Digital CGH derived from the GSW algorithm. b Optical setup for holographicreconstruction. Here the focal length of the FT lens is 600 mm. The dimension of the CGH is 8.64 mm (1080 × 1080 pixels), and the incidentwavelength is 800 nm. c–f Calculated intensity distributions with changing ROIs. g–j Calculated phase distributions with changing ROIs. Here,sampling points are fixed at 1080 × 1080

    Hu et al. Light: Science & Applications (2020) 9:119 Page 7 of 11

  • First, we simulate the multi-foci optical system, whichcan be used for holographic tweezers, laser parallel pro-cessing and data recording. Figure 5b is the correspondingCGH for the generation of a 9 × 9 multi-foci array. Alinearly polarized femtosecond laser (800 nm, emittedfrom Chameleon Vision-S, Coherent) is modulated by theCGH. After the FT of Lens 1, a multi-foci array is gen-erated (Fig. 5c). At the back of the objective, the phase andintensity distributions are retrieved as shown in Fig. 5d, e.The phase profile closely resembles the CGH, validatingthe accuracy of the Bluestein-enabled calculation of scalardiffraction. The light beam is slightly smaller than the sizeof the entrance pupil of the objective, ensuring that thephase-modulated beam can be fully transformed by theobjective. In the focal plane of the objective, a diffraction-

    limited 9 × 9 multi-foci array is generated (Fig. 5f). Thefull-path calculation can be accomplished with high effi-ciency in

  • consistent with the experimental result (Fig. 5k). By takingadvantage of the high flexibility of the Bluestein method, amagnified image of an arbitrary ROI can be calculatedwith arbitrary resolution and good accuracy in compar-ison with the experimental result, as shown in Fig. 5l, k.Another example of the usage of the Bluestein method forvector diffraction is shown in Supplementary InformationSection 6 and Fig. S8. In brief, full-path light tracing of theentire optical system can be accomplished by the Blue-stein method with high efficiency and flexibility, unfolding

    its capacities in the real-time prediction and evaluation ofoptical performance in advanced microscopy, lasermanipulation, and photolithography.

    DiscussionThe proposed Bluestein-based method provides a fun-

    damental improvement in optical diffraction calculations.The advantages of the method lie in the following threeaspects. First, the computation method for light diffrac-tion is superfast, allowing for the real-time prediction of

    F

    Laser

    HWP

    PBS

    Mirror

    SLM

    Lens1

    Lens2

    Objective

    Attenuator

    λ

    f1f1

    f2f2

    a

    b c d

    f h i

    j l

    1 mm 1 mm 0.5 mm

    10 μm 100 nm

    S

    P

    EF

    S P E

    F

    F

    k

    2 μm2 μm5 μm5 μm

    F

    Simulation

    Simulation

    Experiment

    Experiment Simulation Experiment

    e

    g

    m

    Phase3

    2

    1

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    1

    0.8

    1.0

    SimulationExperiment

    0.5

    Inte

    nsity

    (a.

    u.)

    0.0–24 –12

    x (μm)0 12 24

    0.6

    0.4

    0.2

    1

    0.8

    0.6

    0.4

    0.2

    0

    –1

    –2

    –3

    3

    2

    1

    0

    –1

    –2

    –3

    Intensity

    Polarization

    10 μm

    0.5 mm

    E

    F F

    F

    Fig. 5 Full-path calculation of a representative optical system. a Sketch of the optical system. S: the plane on the panel of the SLM. P: the focalplane of Lens 1. E: the entrance pupil of the objective. F: the focal plane of the objective. (b) CGH displayed on the SLM for the generation of a 9 × 9foci array. c The foci array on the focal plane of Lens 1 (P-plane). d Phase distribution and e intensity distribution on the entrance pupil of theobjective (E-plane). f Simulated and gmeasured multi-foci array generated on the focal plane of the objective (F-plane). h Enlarged intensity profile ofa single focal spot in the array. The arrows indicate the polarization directions. i Longitudinal intensity profile and corresponding line plot of the fociarray. j Simulated and k measured intensity distribution on the F-plane when the CGH for the generation of the pattern “E” is encoded on the SLM.l, m Enlarged intensity profiles of the pattern corresponding to (j) and (k) with the same sampling points as in (i)

    Hu et al. Light: Science & Applications (2020) 9:119 Page 9 of 11

  • light field propagation for diverse implementations.Second, the method has great flexibility, without loss ofaccuracy and efficiency. The desired ROI can be freelychosen, and the sampling numbers can be arbitrarilytuned. Third, the method shows good universality. Itsuits all diffraction conditions, such as phase modula-tion, amplitude filtering, polarization conversion, andfocusing transform. In particular, this method facilitatesthe simulation and optimal design of metasurfaces34–36,as exemplified in Supplementary Information Section 7and Fig. S9. Both scalar and vector diffraction can becomputed using this method, making this method pro-mising for full-path propagation evaluation in broadapplications of optical microscopy, lithography, andoptical manipulation.In addition, the applicability of this method needs to be

    explicated for realistic implementations. First, someapproximation conditions are assumed for both vectorand scalar diffraction. For vector diffraction, the lens isassumed to obey Abbe’s sine condition. For scalar dif-fraction, the Fresnel approximation should be valid. It isworth noting that Fraunhofer diffraction can also beimplemented using the Bluestein method with slightmodification. Second, it is important to take stringentprecautions against aliasing effects. When the diffractiondistance of scalar diffraction is too small or the focal shiftof vector diffraction is too long, obvious aliasing is likelyto occur because the sampling condition no longersatisfies the Nyquist sampling condition.In summary, an efficient calculation method is devel-

    oped to evaluate light diffraction with high flexibility andefficiency. First, a set of mathematical preliminaries isgiven to express the scalar and vector diffraction integralsin the form of an FT and then unified using the Bluesteinmethod. Examples for both scalar and vector diffractionare demonstrated to reveal that the computational effi-ciency and flexibility are greatly improved. Calculation ofthe light field is realized at the sub-second time levelcompared with several minutes using the FFT method orhours using the direct integration method. Full-path lighttracing is finally demonstrated using the Bluesteinmethod. This method holds great potential not only in thefast prediction of numerous optical systems but also in therealm of signal processing for acoustic and othercommunication waves.

    Materials and methodsComputational environmentAll the calculations are performed on a personal laptop

    with an Intel processor I5 2.50 GHz and 8 GB of memory,running the Windows 10 Professional operating system.The code is written, compiled and run in the MATLABR2019a software. All the comparison studies on efficiencyare performed in the same computational environment.

    Laser holographic systemThe laser holographic system consists of a Ti:sapphire

    femtosecond laser oscillator (Chameleon Vision-S,Coherent) with a central wavelength of 800 nm, a repe-tition rate of 80MHz, and a pulse width of 75 fs. A phase-only reflective liquid crystal SLM (Pluto NIR-2, Holoeye)is utilized for the phase modulation, which features a1920 × 1080 resolution and a 8 μm pixel pitch. In theexperiment, only the central portion of the SLM with1080 × 1080 pixels is used to modulate the wavefront,while the other pixels are set to zero. A phase hologrampattern with 256 different shades of gray is loaded ontothe SLM, corresponding to the phase modulation depthfrom 0 to 2π. A CCD camera is used to capture the lightfield distribution.

    AcknowledgementsThis work was supported by the National Natural Science Foundation of China(Nos. 51875544, 91963127, 51675503, 61805230, 51805509), USTC ResearchFunds of the Double First-Class Initiative (Grant No. YD2090002005), YouthInnovation Promotion Association of the Chinese Academy of Sciences(2017495), and National Key R&D Program of China (2018YFB1105400). TheUSTC Center for Micro and Nanoscale Research and Fabrication is alsoacknowledged.

    Author details1CAS Key Laboratory of Mechanical Behavior and Design of Materials, KeyLaboratory of Precision Scientific Instrumentation of Anhui Higher EducationInstitutes, Department of Precision Machinery and Precision Instrumentation,University of Science and Technology of China, Hefei 230026, China.2Department of Mechanical Engineering and Department of Civil andEnvironmental Engineering, Massachusetts Institute of Technology,Cambridge, MA 02139, USA. 3State Key Laboratory of Advanced Technologyfor Materials Synthesis and Processing, International School of MaterialsScience and Engineering, Wuhan University of Technology, Wuhan 430070,China. 4Institute of Industry and Equipment Technology, Hefei University ofTechnology, Hefei 230009, China

    Author contributionsY.H. conceived the idea and developed the theory. Y.H., Z.W., and X.W.performed the simulations. S.J., C.Z. W.Z., and D.W. performed the experimentalmeasurements. Y.H., X.W., J.L., D.W., and J.C. analyzed the data. Y.H., C.Z., J.L., D.W., and J.C. wrote the paper. J.L. and D.W. supervised the project. All authorsdiscussed the results and commented on the paper.

    Conflict of interestThe authors declare that they have no conflict of interest.

    Supplementary information is available for this paper at https://doi.org/10.1038/s41377-020-00362-z.

    Received: 12 May 2020 Revised: 23 June 2020 Accepted: 28 June 2020

    References1. Willig, K. I., Harke, B., Medda, R. & Hell, S. W. STED microscopy with continuous

    wave beams. Nat. Methods 4, 915–918 (2007).2. Kawata, S. et al. Finer features for functional microdevices. Nature 412,

    697–698 (2001).3. Andrew, T. L., Tsai, H. Y. & Menon, R. Confining light to deep sub-

    wavelength dimensions to enable optical nanopatterning. Science 324,917–921 (2009).

    4. Scott, T. F. et al. Two-color single-photon photoinitiation and photoinhibitionfor subdiffraction photolithography. Science 324, 913–917 (2009).

    Hu et al. Light: Science & Applications (2020) 9:119 Page 10 of 11

    https://doi.org/10.1038/s41377-020-00362-zhttps://doi.org/10.1038/s41377-020-00362-z

  • 5. Li, L. J. et al. Achieving λ/20 resolution by one-color initiation and deactivationof polymerization. Science 324, 910–913 (2009).

    6. Grier, D. G. A revolution in optical manipulation. Nature 424, 810–816 (2003).7. Block, S. M. Making light work with optical tweezers. Nature 360, 493–495

    (1992).8. Mezouari, S. & Harvey, A. R. Validity of fresnel and fraunhofer approximations in

    scalar diffraction. J. Opt. A 5, S86–S91 (2003).9. Sheppard, C. J. R. & Matthews, H. J. Imaging in high-aperture optical systems. J.

    Opt. Soc. Am. A 4, 1354–1360 (1987).10. Lee, K. G. et al. Vector field microscopic imaging of light. Nat. Photonics 1,

    53–56 (2007).11. Dorn, R., Quabis, S. & Leuchs, G. Sharper focus for a radially polarized light

    beam. Phys. Rev. Lett. 91, 233901 (2003).12. Wolf, E. Electromagnetic diffraction in optical systems. I. An integral repre-

    sentation of the image field. Proc. R. Soc. A 253, 349–357 (1959).13. Hao, X., Kuang, C. F., Wang, T. T. & Liu, X. Effects of polarization on the de-

    excitation dark focal spot in STED microscopy. J. Opt. 12, 115707 (2010).14. Urbach, H. P. & Pereira, S. F. Field in focus with a maximum longitudinal

    electric component. Phys. Rev. Lett. 100, 123904 (2008).15. Leutenegger, M. et al. Fast focus field calculations. Opt. Express 14,

    11277–11291 (2006).16. Lin, J. et al. Fast vectorial calculation of the volumetric focused field dis-

    tribution by using a three-dimensional Fourier transform. Opt. Express 20,1060–1069 (2012).

    17. Lin, J. et al. Direct calculation of a three-dimensional diffracted field. Opt. Letters36, 1341–1343 (2011).

    18. Boruah, B. R. & Neil, M. A. A. Focal field computation of an arbitrarily polarizedbeam using fast Fourier transforms. Opt. Commun. 282, 4660–4667 (2009).

    19. Nie, Z. Q. et al. Three-dimensional super-resolution longitudinal magnetizationspot arrays. Light Sci. Appl. 6, e17032 (2017).

    20. Born, M. & Wolf, E. Principles of Optics 7th edn (Cambridge University Press,Cambridge, 1999).

    21. Richards, B. & Wolf, E. Electromagnetic diffraction in optical systems, II. Struc-ture of the image field in an aplanatic system. Proc. R. Soc. A 253, 358–379(1959).

    22. Bluestein, L. A linear filtering approach to the computation of discrete Fouriertransform. IEEE Trans. Audio Electroacoust. 18, 451–455 (1970).

    23. Rabiner, L., Schafer, R. & Rader, C. The chirp z-transform algorithm. IEEE Trans.Audio Electroacoust. 17, 86–92 (1969).

    24. Kim, H. et al. In situ single-atom array synthesis using dynamic holographicoptical tweezers. Nat. Commun. 7, 13317 (2016).

    25. Li, X. P. et al. Athermally photoreduced graphene oxides for three-dimensionalholographic images. Nat. Commun. 6, 6984 (2015).

    26. Hu, Y. L. et al. High-efficiency fabrication of aspheric microlens arrays byholographic femtosecond laser-induced photopolymerization. Appl. Phys. Lett.103, 141112 (2013).

    27. Ni, J. C. et al. Three-dimensional chiral microstructures fabricated by structuredoptical vortices in isotropic material. Light Sci. Appl. 6, e17011 (2017).

    28. Zalevsky, Z., Mendlovic, D. & Dorsch, R. G. Gerchberg–Saxton algorithmapplied in the fractional Fourier or the Fresnel domain. Opt. Lett. 21, 842–844(1996).

    29. Hu, Y. L. et al. Fast bits recording in photoisomeric polymers by phase-modulated femtosecond laser. IEEE Photonics Technol. Lett. 26, 1154–1156(2014).

    30. Wang, H. F. et al. Creation of a needle of longitudinally polarized light invacuum using binary optics. Nat. Photonics 2, 501–505 (2008).

    31. Wang, S. C. et al. Ultralong pure longitudinal magnetization needle inducedby annular vortex binary optics. Opt. Lett. 39, 5022–5025 (2014).

    32. Hell, S. W. & Wichmann, J. Breaking the diffraction resolution limit by stimu-lated emission: stimulated-emission-depletion fluorescence microscopy. Opt.Lett. 19, 780–782 (1994).

    33. Gan, Z. S. et al. Three-dimensional deep sub-diffraction optical beam litho-graphy with 9 nm feature size. Nat. Commun. 4, 2061 (2013).

    34. Yu, N. F. et al. Light propagation with phase discontinuities: generalized lawsof reflection and refraction. Science 334, 333–337 (2011).

    35. Khorasaninejad, M. et al. Metalenses at visible wavelengths: diffraction-limitedfocusing and subwavelength resolution imaging. Science 352, 1190–1194(2016).

    36. Yu, N. F. & Capasso, F. Flat optics with designer metasurfaces. Nat. Mater. 13,139–150 (2014).

    Hu et al. Light: Science & Applications (2020) 9:119 Page 11 of 11

    Efficient full-path optical calculation of scalar and vector diffraction using the Bluestein methodIntroductionResultsScalar and vector diffraction integral in the form of a Fourier transformBluestein method to compute Fourier transform with arbitrary ROI and sampling resolutionFast numerical implementation of the Bluestein method in scalar Fresnel diffractionFast numerical calculation of the vectorial Debye diffractionFull-path optical calculation with superior flexibility and efficiency

    DiscussionMaterials and methodsComputational environmentLaser holographic system

    Acknowledgements