3,578
Views
21
CrossRef citations to date
0
Altmetric
Review Articles

Review on vortex beams with low spatial coherence

, &
Article: 1626766 | Received 21 Apr 2019, Accepted 28 May 2019, Published online: 30 Jun 2019

ABSTRACT

Vortex beams with helical phase, carrying phase singularity and orbital angular momentum, have attract great attention in the past decades due to their wide applications in optical communications, optical manipulation, super-resolution imaging and so on. Vortex beams with low spatial coherence, i.e. partially coherent vortex beams, carrying correlation singularity, display some unique properties during propagation, e.g. self-shaping, self-splitting and self-reconstruction. Partially coherent vortex beams exhibit some advantages over coherent vortex beams in some applications, such as remote sensing, laser radar and free-space optical communications. This review summarizes research progress on partially coherent vortex beams, including theoretical models, propagation properties, generation and topological charge determination.

Graphical Abstract

1 Introduction

Since laser beam was invented in the 1960s, it became one of the most powerful instruments and has been used widely in various fields [Citation1]. Beams with prescribed amplitude, coherence, phase and polarization have been studied in depth. Among them, coherence and phase are two extremely important properties that have been investigated extensively in the past decades [Citation2Citation5]. Coherence is regarded as a consequence of the correlation between two or more points in the fluctuating electric field. Laser beam with low spatial coherence named partially coherent beam, which is characterized by the cross-spectral density (CSD) function, has advantages over coherent beam in some applications. For example, decreasing the spatial coherence of a light beam can increase the signal–noise ratio and reduce the bit-error rate in free-space optical communications [Citation6Citation11]. Partially coherent beam can be used to overcome speckle effectively in laser nuclear fusion [Citation12], reduce noise in photograph [Citation13], and realize classical ghost imaging [Citation14]. In addition, partially coherent beam also displays advantages in particle trapping [Citation15], atom cooling [Citation16], second-harmonic generation [Citation17,Citation18], optical scattering [Citation19,Citation20] and laser scanning [Citation21]. Before a class of novel partially coherent beam with nonconventional correlation function was introduced by Gori and co-workers [Citation22Citation24], most of the literatures were focused on conventional partially coherent beam (i.e. Gaussian Schell-model beam), whose intensity and degree of coherence satisfy Gaussian distributions [Citation14Citation16] Some extraordinary properties of the nonconventional partially coherent beams have been found, e.g. self-shaping [Citation25Citation28], self-splitting [Citation29], self-focusing [Citation30,Citation31] and self-reconstruction [Citation32], which are expected be useful in many applications. A review on generation and propagation of partially coherent beams with nonconventional correlation functions can be found in [Citation33].

Figure 1. Focused intensity of optical coherence vortex lattices in the focal plane for different values of the initial coherence width [Citation124].

Figure 1. Focused intensity of optical coherence vortex lattices in the focal plane for different values of the initial coherence width [Citation124].

The phase of light can be characterized by its wave front on propagation. In general, there are three types of wave fronts, i.e. planar, spherical and helical wave fronts [Citation34]. Vortex beam with helical phase carries phase singularity and orbital angular momentum (OAM) [Citation34,Citation35]. Phase singularity demonstrates wave front dislocation, which means the point in the field has a zero amplitude with indefinite phase. Phase singularity was first introduced by Nye and Berry in 1974 [Citation35], since then, much work on phase singularity has been introduced, and a new branch of optics named singular optics was formed [Citation36,Citation37]. On the other hand, in 1992, Allen and coworkers discovered that light beam with helical phase [i.e. expilφ] carries an OAM of l per photon with l being the topological charge,φ being the azimuthal angle and being Planck’s constant divided by 2π [Citation38]. Due to their extraordinary properties, vortex beams have been applied for particle trapping and manipulations [Citation39Citation44], optical measurement [Citation45,Citation46], optical communications [Citation47Citation50], super-resolution imaging [Citation51Citation53] and so on. Various vortex beams have been introduced, e.g. Bessel beam [Citation54Citation56], Gaussian-like beam [Citation57], anomalous vortex beam [Citation58] and perfect vortex beam [Citation59,Citation60]. Besides scalar vortex beams, vector vortex beams, e.g. radial-polarized vortex beam [Citation61] and circularly polarized vortex beam [Citation62], were also introduced. Different methods, e.g. spiral phase plate (SPP) [Citation63Citation65], computational hologram [Citation66Citation68], mode converter [Citation69Citation71], long-period fiber gratings [Citation72] and plasmonic metasurface [Citation73] have been introduced to generate vortex beam. Many methods for detecting the topological charge of a vortex beam have been developed [Citation74Citation99], e.g. Shack Hartmann wave front sensor [Citation74Citation77], interference [Citation78Citation80], diffraction [Citation81Citation90], scattering [Citation91], Fourier patterns of the intensity [Citation92], mode transformation [Citation93], OAM density [Citation94] and rotational Doppler effect [Citation95].

Vortex beam with low spatial coherence named partially coherent vortex beam was first proposed by Gori and coworkers in 1998 [Citation100], and such beam displays some unique properties [Citation101Citation115], e.g. correlation singularity. Correlation singularity is defined as the point in the field whose CSDor degree of coherence equals to zero while the corresponding phase is indefinite. In general, the number of the correlation singularities depends on the topological charge of the partially coherent vortex beam. Partially coherent vortex beam possesses the extraordinary properties of both partially coherent beam and vortex beam, e.g. lower beam scintillation and beam wander during propagation in random media [Citation116Citation122], better self-reconstruction ability [Citation32], and is expected to be useful in free-space optical communications, optical imaging and information transfer. Most of the literatures have been focused on the conventional partially coherent vortex beams whose correlation functions are of Gaussian distributions [Citation101Citation122]. In recent years, partially coherent vortex beams with nonconventional correlation functions have been introduced [Citation123,Citation124]. In this review, we introduce research progress on partially coherent vortex beams, including theoretical models, propagation properties, generation and topological charge determination.

2 Theoretical models for partially coherent vortex beams

The statistical properties of a scalar partially coherent beam can be characterized by the mutual coherence function in the spatial-time domain or the CSD in the space-frequency domain [Citation3,Citation4]. The CSD of a partially coherent vortex beam in the source plane can be expressed in the following general form

(1) Wr1,r2=E(r1)E(r2)=Ar1Ar2expilφ1φ2gr1r2.(1)

where Er and Ar are the electric field and the amplitude, respectively, the angular bracket and the asterisk denote ensemble average and complex conjugate, respectively. r1 and r2 are two arbitrary vector coordinates in the source plane, φ=arctan2y/xdenotes the angular coordinate, gr1r2 is the correlation function between two arbitrary points. By varying the amplitude distribution Ar, various partially coherent vortex beams have been proposed. For a Collet-Wolf source [Citation3,Citation4], the correlation function between two arbitrary points satisfies Gaussian distribution, i.e.

(2) gr1r2=expr1r222δ02,(2)

where δ0 denotes the initial coherence width. When δ0, Equation (1) reduces to a coherent vortex beam, and when δ00, Equation (1) reduces to an incoherent vortex beam. Gaussian Schell-model (GSM) vortex beam is a classical partially coherent vortex beam, whose CSD function in the source plane is written as [Citation112]

(3) WGSMVr1,r2,φ1,φ2=expr12+r224σ02r12+r222r1r2cosφ1φ22δ02+ilφ1φ2,(3)

where r and φ are the radial and angle coordinates, respectively. σ0 denotes the transverse beam width. For δ0, a GSM vortex beam reduces to a coherent Gaussian vortex beam. Furthermore, partially coherent vortex beams with more complicated amplitude were proposed in [Citation118Citation122], e.g. partially coherent Laguerre-Gaussian (LG) beam, partially coherent Bessel–Gaussian beam, partially coherent Airy vortex beam, and some extraordinary propagation properties were found, e.g. non-diffraction, self-acceleration and self-reconstruction. Partially coherent LG beam is a classical partially coherent vortex beam with complicated amplitude whose CSD function in the source plane reads as [Citation113]:

(4) WLGr1,r2,φ1,φ2=2r1ω0l2r2ω0lLpl2r12ω02Lpl2r22ω02expr12+r22ω02×expr12+r222r1r2cosφ1φ22δ02expilφ1φ2,(4)

where ω0=2σ0, Lpl is the Laguerre polynomial with mode orders p and l. For p=l=0, Equation (4) reduces to a classical GSM beam. For p=0 and l0, Equation (4) reduces to a fundamental GSM vortex beam.

One can modulate not only the amplitude Ar of a partially coherent vortex beam but also its correlation function gr1r2. The correlation function of a conventional partially coherent vortex beam is of Gaussian distribution. Partially coherent vortex beams with nonconventional correlation functions were proposed and generated [Citation123,Citation124], and some interesting properties were found. Laguerre-Gaussian correlated Schell-model (LGCSM) vortex beam is a typical nonconventional partially coherent vortex beam, whose correlation function is of LGdistribution, and its CSD function in the source is expressed as [Citation123]:

(6) WLGCSMVr1,r2=expr12+r224σ02r1r222δ02Lp0r1r222δ02expilθ1θ2,(6)

where Lp0 denotes the Laguerre polynomial. Equation (6) reduces to a GSM vortex beam whenp=0. More recently, partially coherent vortex beam with periodical coherence properties was introduced [Citation124] and such beam displays interesting propagation properties, i.e. a Gaussian beam spot evolves into multiple beam spots (i.e. intensity lattices) in the focal plane (i.e. in the far field), which are useful for particle trapping and information transfer.

A vector partially coherent vortex beam has both x- and y- components and can be characterized by the CSD matrix in the space-frequency domain [Citation3,Citation4]. The degree of polarization and state of polarization of vector partially coherent beam may vary on propagation in free space [Citation4], which is much different from a vector coherent beam [Citation4]. An electromagnetic Gaussian Schell-model (EGSM) vortex beam is a typical vector partially coherent vortex beam with uniform state of polarization, and the CSD matrix of an EGSM vortex beam in the source plane is expressed as [Citation125]:

(7) WˆEGSMVr1,r2=Wxxr1,r2Wxyr1,r2Wyxr1,r2Wyyr1,r2,(7)

and its elements are given as

(8) Wαβr1,r2=AαAβBαβexpr124σα2r224σβ2r1r222δαβ2expilφ1φ2,(8)

where Ax and Ay are the amplitudes of x and y components of the electric field, respectively. Bxx=Byy=1, Bxy=Bxyexpiϕxy is the complex correlation coefficient between x and y components of the electric field with ϕxy being the phase difference between the two components. σi is the r.m.s width of the intensity distribution along i direction, δxx, δyy and δxy=δyx are the r.m.s widths of the autocorrelation functions of the x component of the electric field, of the y component of the electric field and of the mutual correlation function of x and y components, respectively. The nine real parameters Ax, Ay, σx, σy, δxx, δyy, δxy, Bxy, ϕxy are shown to satisfy several intrinsic constraints and obey some simplifying assumptions [Citation4]. Vector partially coherent vector beam with uniform state of polarization named partially coherent radially polarized beam was proposed and generated in [Citation126].

Partially coherent beam can carry not only vortex phase and but also twist phase. Twist phase only exists in a partially coherent beam [Citation127]. In the past decades, the vortex phase and vortex beam have been studied separately, and both phase will induce OAM. More recently, a new partially coherent beam named twisted Laguerre-Gaussian Schell-model (TLGSM) beam was introduced, which carries both vortex phase and twist phase, and its CSD function is given by [Citation128]

(9) WTLGSMr1,r2=r1lr2lexpr12+r224σ02expr1r222δ02×expilφ1φ2expikμ0x1y2+ikμ0x2y1,(9)

The last term in Equation (9) represents the twist phase. μ0 is a real-valued twist factor with dimension of an inverse length. k=2π/λ with λ being the wavelength of light. According to [Citation127], the twist factor must satisfy the inequality μ02k2δ041. The vortex phase’s and twist phase’s contributions to the OAM are interrelated, which can greatly increase the amount of OAM [Citation128].

3 Propagation properties of partially coherent vortex beams

Propagation of partially coherent beams in free space can be treated by the generalized Huygens-Fresnel diffraction integral [Citation3], and the generalized Collins formula was used to treat the propagation of partially coherent beams through a paraxial ABCD optical system [Citation129]. An efficient tensor method was proposed to treat the propagation of complicated partially coherent beam recently [Citation130]. Based on above methods, the propagation properties of partially coherent vortex beams and correlation vortices have been studied in detail [Citation101Citation122,Citation131Citation136]. Both numerical results and experimental results have shown that a partially coherent vortex beam has advantage over a coherent vortex beam or a GSM beam for reducing turbulence-induced degradation and scintillation [Citation118Citation122,Citation137], which is expected to be useful for free-space optical communications. Recently, Zeng et al. introduced a new kind of partially coherent vortex named partially coherent fractional vortex beam, which has fractional topological charge and its intensity pattern has a controllable opening gap and can be manipulated through varying its coherence width [Citation138].

It is known that it is time-consuming to obtain analytical solution of a partially coherent beam propagating in turbulent atmosphere. In recent years, a simulation method using multiple phase screens has been introduced to simulate the propagation of coherent beam through turbulent atmosphere. However, this method cannot be applied directly to treat the propagation of partially coherent beams, since they were characterized by the CSD function instead of the electric field. A feasible method is the coherent mode decomposition, we can expand a partially coherent beam as the incoherent superposition of multiple coherent modes [Citation139]. By simulating the propagation of every single mode through turbulence, then superposing them incoherently, one can obtain the statistical properties of a partially coherent beam in turbulent atmosphere. Another method resorts to a random phase screen which satisfies certain statistical properties to simulate partially coherent beams [Citation140]. A pure phase screen can be easily generated by the spatial light modulator (SLM) to simulate the propagation of a conventional-correlated partially coherent beam in turbulent atmosphere, e.g. GSM beam and partially coherent flat-topped beam [Citation141,Citation142]. However, for a partially coherent beam with non-conventional correlation function, a complex phase screen is needed in this case [Citation143]. Hence, one can simulate the propagation of a partially coherent beam through turbulent atmosphere by multiple random phase screens (see ) [Citation144].

Figure 2. Simulation of the propagation of light beam through turbulent atmosphere by multiple random phase screens [Citation144].

Figure 2. Simulation of the propagation of light beam through turbulent atmosphere by multiple random phase screens [Citation144].

Some interesting propagation properties of partially coherent vortex beams have been found in [Citation112,Citation114,Citation145,Citation146], e.g. self-shaping, self-splitting and self-reconstruction. Unlike the hollow profile of a coherent vortex beam, the intensity distribution of a partially coherent vortex beam varies during propagation, e.g. dark hollow, flat-topped and Gaussian beam profiles can be formed at different propagation distances. When a partially coherent vortex beam with conventional correlation function (e.g. partially coherent LG0l beam) is focused, one can obtain dark hollow or flat-topped or Gaussian beam spot in the focal plane by choosing suitable initial coherence width (see ) [Citation112]. The topological charge of partially coherent vortex beam plays a role of anti-degradation, i.e. the dark hollow beam profile evolves into a Gaussian beam profile more slowly as the topological charge increases. For a partially coherent vortex beam with periodical coherence properties (i.e. optical coherence vortex lattices) focused by a thin lens, one can obtain different intensity lattices in the focal plane through varying initial coherence width (see )[Citation124], and each beamlet of the intensity lattices in the focal plane can display dark hollow or flat-topped or Gaussian beam spot. The intensity lattices are useful for simultaneously trapping multiple particles. In addition, for a vector partially coherent vortex beam with uniform state of polarization (i.e. EGSM vortex beam), one can control its focused intensity distribution by varying its initial degree of polarization, topological charge and coherence widths [Citation125].

Figure 3. Focused intensity distribution of a partially coherent LG0l beam for different values of the topological charge and initial coherence width [Citation145].

Figure 3. Focused intensity distribution of a partially coherent LG0l beam for different values of the topological charge and initial coherence width [Citation145].

In principle, the beam spot of a partially coherent vortex beam rotates on propagation, which is caused by the topological charge, while it is hard to observe the beam spot rotation directly when the intensity distribution has a circular symmetry. Some special partially coherent vortex beam with non-circular symmetry, e.g. partially coherent fractional vortex beam, displays the phenomenon of beam spot rotation on propagation clearly. The intensity distribution of a partially coherent fractional vortex beam has a gap, and the gap rotates clockwise or anti-clockwise depending on the sign of the topological charge [Citation138]. One also can observe the rotation of the beam spot when a circular partially coherent vortex beam passes through an anisotropic optical system [Citation131], e.g. cylindrical lens. For a partially coherent radially polarized vortex beam, the beam spot of x or y component also demonstrates rotation on propagation [Citation126].

Another interesting and useful property of partially coherent vortex beam is its self-reconstruction ability. In general, the self-reconstruction is considered as the property of coherent diffraction-free beams, e.g. Bessel beam, Bessel-Gaussian beam and Airy beam [Citation147Citation149], and optical self-reconstruction has been applied for microscopic particle manipulation [Citation150] and human tissue microscopy [Citation151]. Usually diffraction-free beams reconstitute their spatial shapes upon interaction with the obstacle. In [Citation32], Wang et al. found that partially coherent beam can self-reconstruct its intensity profile and state of polarization upon scattering from an opaque obstacle provided the beam coherence area is reduced well below the obstacle area.

For an obstructed partially coherent vortex beam, the distribution of the degree of coherence also reconstructs upon propagation as shown in , and the self-reconstruction ability of the intensity and degree of coherence increases as the initial coherence width decreases. The reconstructed intensity does not reveal any information about the topological charge, while the reconstructed degree of coherence contain the information of the topological charge, i.e. the ring dislocation number of the degree of coherence is related to the magnitude of the topological charge.

Figure 4. Normalized intensity and modulus of the degree of coherence of a focused partially coherent LGpl beam with p= 1 and l= 1 obstructed by a sector-shaped opaque obstacle with center angle α =90° in the focal plane for different values of initial coherence width [Citation146].

Figure 4. Normalized intensity and modulus of the degree of coherence of a focused partially coherent LGpl beam with p= 1 and l= 1 obstructed by a sector-shaped opaque obstacle with center angle α =90° in the focal plane for different values of initial coherence width [Citation146].

3. Generation of partially coherent vortex beams

In general, generation of a partially coherent vortex beam mainly include two steps: the first step is to produce a partially coherent beam, and the second step is to load the vortex phase into the generated partially coherent beam. Different methods have been introduced for generating partially coherent beams: one is to increase the coherence width of incoherent beam by using the filter plate (e.g. an aperture filter), which has been widely used before the invention of laser beam. The energy loss of this method is large and the initial coherence width of the generated beam is not controllable. It is worth pointing out that an incoherent beam can also evolve into a partially coherent beam after long-distance propagation. The second method is to decrease the coherence width of a coherent beam with the help of random media (e.g. rotating ground-glass disk). By using this method, one can precisely control the initial coherence width of the generated partially coherent beam by manipulating the beam spot size on the surface of the rotating-ground glass disk [Citation26], however, the energy loss of this method is also inevitable. The third method is to superpose a sequence of coherent modes because partially coherent beam can be decomposed as a superposition of multiple coherent modes [Citation152,Citation153], and one can increase the power of the generated partially coherent beam with this method. On the other hand, various optical elements and methods have been proposed to generate vortex phase, e.g. SPP, diffractive optical elements, computer-generated hologram spiral fiber, uniaxial crystal and nanostructures.

For a scalar conventional partially coherent vortex beam, e.g. a typical GSM vortex beam, the corresponding experimental setup was shown in [Citation112]. A linearly polarized He-Ne laser beam is focused onto a rotating ground glass disk (RGGD) by thin lens (L1), the transmitted incoherent light then is collimated and filtered by the thin lens (L2) and Gaussian amplitude filter (GAF), respectively, producing a GSM beam. By adding vortex phase expilφ into the GSM beam with help of a SPP, one can generate a GSM vortex beam. The neutral density filter (NDF) is used to modulate the amplitude of the beam. One can modulate the coherence width of the generated beam by changing the focused beam spot size on the RGGD, here the focused beam spot size depends on the distance between thin lens L1 and RGGD. Based on above experimental setup, if we replace the SPP with a SLM, both the amplitude and the phase of the generated partially coherent beam can be modulated, and then more complicated partially coherent vortex beam with conventional-correlated function can be obtained, e.g. partially coherent LG beam and Bessel Gaussian Schell-model (BGSM) beam [Citation154,Citation155].

Figure 5. Experimental setup for generating a GSM vortex beam. NDF, neutral density filter; RGGD, rotating ground-glass disk; GAF, Gaussian amplitude filter; SPP, spiral phase plate; L1, L2, thin lenses [Citation112].

Figure 5. Experimental setup for generating a GSM vortex beam. NDF, neutral density filter; RGGD, rotating ground-glass disk; GAF, Gaussian amplitude filter; SPP, spiral phase plate; L1, L2, thin lenses [Citation112].

For a nonconventional correlated partially coherent vortex beam, e.g. optical coherent vortex lattices, the corresponding experimental setup is shown in . A linearly polarized laser beam is expanded by a beam expander and reflected to an amplitude mask by a mirror, then the shaped coherent beam illuminates the RGGD. The transmitted beam becomes a partially coherent beam with nonconventional correlation function after passing through thin lens L1 and GAF. Here the amplitude mask can be controlled by the SLM to generate various beam spots, e.g. doughnut spot, flat-topped beam spot, spot array. When the amplitude mask is a Gaussian beam spot array, optical coherence lattices can be formed in experiment. By adding the vortex phase into the optical coherence lattices with the help of the SPP, one can generate optical coherent vortex lattices. shows the experimental results of the normalized-focused intensity distribution of optical coherence vortex lattices for different initial coherence widths, and we indeed obtain intensity lattices in the focal plane, and the beam profile of each beamlet is controlled through varying the initial coherence width.

Figure 6. Experimental setup for generating nonconventional correlated partially coherent vortex beams [Citation124].

Figure 6. Experimental setup for generating nonconventional correlated partially coherent vortex beams [Citation124].

Figure 7. Experimental results of the normalized-focused intensity distribution of optical coherence vortex lattices for different initial coherence width [Citation124].

Figure 7. Experimental results of the normalized-focused intensity distribution of optical coherence vortex lattices for different initial coherence width [Citation124].

We can generate vector partially coherent vortex beam using similar method. Two types of vector partially coherent vortex beams with uniform state of polarization (i.e. EGSM vortex beam) and nonuniform state of polarization (i.e. partially coherent radially polarized vortex beam) have been proposed and generated in [Citation125,Citation126]. The state of polarization and degree of polarization of vector partially coherent vortex beams varies on propagation in free space, and the vortex phase plays a role of resisting coherence-induced degradation and depolarization.

In above introduction, a RGGD has been used to generate various partially coherent vortex beams with the help of a SPP or spatial light modulator. Recently, a digital method using a SLM has been proposed to generate partially coherent vortex beams [Citation156], and the corresponding experimental setup is shown in . A series of complex random phase screens are loaded dynamically by the SLM to generate partially coherent vortex beams. However, this method depends on the performance of the liquid crystal (e.g. refresh rate) of the SLM, and could not be used for generating a partially coherent vortex beam with high energy density. Mode-decomposition method may be used to generate partially coherent vortex beam with high power in future.

Figure 8. Experimental setup for digital generation of partially coherent vortex beams. BE, beam expander; SLM, spatial light modulator; L1, L2, thin lens; D, aperture [Citation156].

Figure 8. Experimental setup for digital generation of partially coherent vortex beams. BE, beam expander; SLM, spatial light modulator; L1, L2, thin lens; D, aperture [Citation156].

4. Determination the topological charge of partially coherent vortex beams

The photon of partially coherent vortex beam carries OAM, which is closely related to the topological charge and can be used for information encoding and decoding [Citation49]. Generally, to encode the information, we need to know (or determine) the magnitude and sign of the topological charge. One can determine the topological charge of a coherent vortex beam through measuring the intensity distribution after diffraction or interference [Citation77Citation95]. It is known that the intensity distribution of a partially coherent vortex beam gradually evolve from doughnut to Gaussian distribution with the decrease of the initial coherence width, and the traditional intensity measurement methods for determining the topological charge become invalid. shows the influence of the initial coherence width on the interference pattern of a GSM vortex beam. One sees that one cannot infer any information about the topological charge when the initial coherence width is small.

Figure 9. Numerical results of the interference pattern of a GSM vortex beam for different values of the initial coherence width with topological charge l= 2.

Figure 9. Numerical results of the interference pattern of a GSM vortex beam for different values of the initial coherence width with topological charge l= 2.

On the other hand, for a partially coherent vortex beam, the phase singularity gradually disappears on propagation, while the correlation singularity (i.e. ring dislocations) in the correlation function appears [Citation103], and the number of ring dislocations of a partially coherent LG0l beam in the focal plane or in the far field equals to the magnitude of the topological charge (see ) [Citation145,Citation155]. For a partially coherent LGpl beam, the number of the ring dislocations equals to 2p+|l| [Citation113], and one can determine p and l through measuring the double-correlation function [Citation157,Citation158].

Figure 10. Distribution of the modulus of the degree of coherence of a partially coherent LG0l beam with different values of the topological charge l in the focal plane for different state of coherence [Citation145].

Figure 10. Distribution of the modulus of the degree of coherence of a partially coherent LG0l beam with different values of the topological charge l in the focal plane for different state of coherence [Citation145].

In addition to the magnitude, the sign of the topological charge also plays an important role for information encoding and transfer. Above mentioned literatures are limited to the measurement of the magnitude of the topological charge. Actually, the sign of the topological charge also can be determined by the correlation function [Citation131]. shows the theoretical simulation (a-e) of the logarithm of the correlation function of a partially coherent LG0l beam for different topological charges at certain propagation distance after passing through a couple of perpendicular cylindrical lenses. The anisotropic correlation function pattern rotates anti-clockwise or clockwise corresponding to the positive or negative sign of the topological charge, respectively. In addition, one finds that the number of the bright fringes equals to 2l+1. More recently, Lu et al. proposed a phase-analysis method for measuring the correlation singularities through introducing a movable perturbation at a certain point in an illumination window of a finite size. Using the proposed method, the correlation singularities of a partially coherent vortex beam in the focal plane were measured. From the results, the magnitude and sign of the topological charge can be determined simultaneously from the phase distribution of the correlation singularities [Citation159].

Figure 11. (a–e) Theoretical simulations of the logarithm of the correlation function of a partially coherent LG0l beam for different topological charges at certain propagation distance after passing through a couple of perpendicular cylindrical lenses [Citation131].

Figure 11. (a–e) Theoretical simulations of the logarithm of the correlation function of a partially coherent LG0l beam for different topological charges at certain propagation distance after passing through a couple of perpendicular cylindrical lenses [Citation131].

For practical optical communications with partially coherent vortex beams, more complicated problems will be concerned. How can we determine the topological charge of a multiplexed partially coherent vortex beam? Recently, Chen and Li proposed a new method by using the lensless ghost imaging system to discriminate incoherent LG0l modes [Citation160], which provides an effective way to determine the topological charge of a multiplexed partially coherent vortex beam and can be used for optical communications.

4. Summary

As a summary, we have given a brief review on partially coherent vortex beams including theoretical models, propagation properties, generation and topological charge determination. Different from coherent vortex beams, partially coherent vortex beams exhibit some unique and extraordinary propagation properties, e.g. correlation singularity, self-shaping, self-splitting and self-reconstruction, which are useful for particle manipulations, material process and super-resolution imaging. One can determine the topological charge of a partially coherent vortex beam by measuring the correlation singularity. Partially coherent vortex beams have advantage over coherent vortex beams for reducing turbulence-induced degradation and scintillation [Citation118Citation122,Citation137], and are expected to be useful for free-space optical communications, imaging and information transfer. We believe this field will grow further and expand rapidly, and more and more interesting results, and potential applications will be revealed.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work was supported by the National Natural Science Foundation of China [Grant Nos. 11525418, 91750201, 11804198] and the Natural Science Foundation of Shandong Province [Grant No. ZR2019BA030].

References

  • Maiman TH. Stimulated optical radiation in ruby. Nature. 1960;187:507–528.
  • Born M, Wolf E. Principles of optics: electromagnetic theory of propagation, interference and diffraction of light. Cambridge: Cambridge University Press; 1999.
  • Mandel L, Wolf E. Optical coherence and quantum optics. Cambridge: Cambridge University Press; 1995.
  • Wolf E. Introduction to the theory of coherence and polarization of light. Cambridge: Cambridge University Press; 2007.
  • Cai Y, Chen Y, Yu J, et al. Generation of partially coherent beams. Prog Opt. 2017;62:157–223.
  • Ricklin JC, Davidson FM. Atmospheric turbulence effects on a partially coherent Gaussian beam: implications for free-space laser communication. J Opt Soc Am A. 2002;19:1794–1802.
  • Wang F, Cai Y, Eyyuboğlu HT, et al. Twist phase-induced reduction in scintillation of a partially coherent beam in turbulent atmosphere. Opt Lett. 2012;37:184–186.
  • Wang F, Liu X, Liu L, et al. Experimental study of the scintillation index of a radially polarized beam with controllable spatial coherence. Appl Phys Lett. 2013;103:091102.
  • Cai Y, Chen Y, Eyyuboğlu HT, et al. Scintillation index of elliptical Gaussian beam in turbulent atmosphere. Opt Lett. 2007;32:2405–2407.
  • Aksenov VP, Kolosov VV, Pogutsa CE. The influence of the vortex phase on the random wandering of a Laguerre-Gaussian beam propagating in a turbulent atmosphere: a numerical experiment. J Opt. 2013;15:044007.
  • Liu X, Wang F, Wei C, et al. Experimental study of turbulence-induced beam wander and deformation of a partially coherent beam. Opt Lett. 2014;39:3336–3339.
  • Kato Y, Mima K, Miyanaga N, et al. Random phasing of high-power lasers for uniform target acceleration and plasma-instability suppression. Phys Rev Lett. 1984;53:1057.
  • Beléndez A, Carretero L, Fimia A. The use of partially coherent light to reduce the efficiency of silver halide noise gratings. Opt Commun. 1993;98:236–240.
  • Cai Y, Zhu SY. Ghost imaging with incoherent and partially coherent light radiation. Phys Rev E. 2005;71:056607.
  • Zhao C, Cai Y, Lu X, et al. Radiation force of coherent and partially coherent flat-topped beams on a Rayleigh particle. Opt Express. 2009;17:1753–1765.
  • Zhang JF, Wang ZY, Cheng B, et al. Atom cooling by partially spatially coherent lasers. Phys Rev A. 2013;88:023416.
  • Zubairy MS, Mciver JK. Second-harmonic generation by a partially coherent beam. Phys Rev A. 1987;36:202–206.
  • Cai Y, Peschel U. Second-harmonic generation by an astigmatic partially coherent beam. Opt Express. 2007;15:15480–15492.
  • Van DT, Fischer DG, Visser TD, et al. Effects of spatial coherence on the angular distribution of radiant intensity generated by scattering on a sphere. Phys Rev Lett. 2010;104:173902.
  • Ding C, Cai Y, Korotkova O, et al. Scattering-induced changes in the temporal coherence width and the pulse duration of a partially coherent plane-wave pulse. Opt Lett. 2011;36:517–519.
  • Kermisch D. Partially coherent image processing by laser scanning. J Opt Soc Am. 1975;65:887–891.
  • Gori F, Guattari G, Padovani C. Modal expansion for J0-correlated Schell-model sources. Opt Commun. 1987;64:311–316.
  • Gori F, Santarsiero M. Devising genuine spatial correlation functions. Opt Lett. 2007;32:3531–3533.
  • Gori F, Ramirezsanchez V, Santarsiero M, et al. On genuine cross-spectral density matrices. J Opt A. 2009;11:85706–85707.
  • Sahin S, Korotkova O. Light sources generating far fields with tunable flat profiles. Opt Lett. 2012;37:2970–2972.
  • Wang F, Liu X, Yuan Y, et al. Experimental generation of partially coherent beams with different complex degrees of coherence. Opt Lett. 2013;38:1814–1816.
  • Mei Z, Korotkova O. Random sources generating ring-shaped beams. Opt Lett. 2013;38:91–93.
  • Chen Y, Liu L, Wang F, et al. Elliptical Laguerre-Gaussian correlated Schell-model beam. Opt Express. 2014;22:13975–13987.
  • Liang C, Wang F, Liu X, et al. Experimental generation of cosine-Gaussian-correlated Schell-model beams with rectangular symmetry. Opt Lett. 2014;39:769–772.
  • Tong Z, Korotkova O. Electromagnetic nonuniformly correlated beams. J Opt Soc Am A. 2012;29:2154–2158.
  • Lajunen H, Saastamoinen T. Non-uniformly correlated partially coherent pulses. Opt Express. 2013;21:190–195.
  • Wang F, Chen Y, Liu X, et al. Self-reconstruction of partially coherent light beams scattered by opaque obstacles. Opt Express. 2016;24:23735–23746.
  • Cai Y, Chen Y, Wang F. Generation and propagation of partially coherent beams with nonconventional correlation functions: a review [Invited]. J Opt Soc Am A. 2014;31:2083–2096.
  • Yao AM, Padgett MJ. Orbital angular momentum: origins, behavior and applications. Adv Opt Photon. 2011;3:161–204.
  • Nye J, Berry M. Dislocations in wave trains. Prog Roy Soc A. 1974;336:165–190.
  • Soskin M, Vasnetsov M. Singular optics. Prog Opt. 2001;42:219–276.
  • Gj G. Singular optics. Boca Raton: CRC Press; 2017.
  • Allen L, Beijersbergen MW, Spreeuw RJC, et al. Orbital angular-momentum of light and the transformation of Laguerre-Gaussian laser modes. Phys Rev A. 1992;45:8185–8189.
  • Allen L, Padgett M, Babiker M. The orbital angular momentum of light. Prog Opt. 1999;39:291–372.
  • Grier DG. A revolution in optical manipulation. Nature. 2003;424:810–816.
  • O’Neil AT, Padgett MJ. Axial and lateral trapping efficiency of Laguerre-Gaussian modes in inverted optical tweezers. Opt Commun. 2001;193:45–50.
  • Ng J, Lin Z, Chan C. Theory of optical trapping by an optical vortex beam. Phys Rev Lett. 2010;104:103601.
  • Wang X, Rui G, Gong L, et al. Manipulation of resonant metallic nanoparticle using 4Pi focusing system. Opt Express. 2016;24:24143–24152.
  • Chen J, Wan C, Kong LJ, et al. Tightly focused optical field with controllable photonic spin orientation. Opt Express. 2017;25:19517–19528.
  • Gu Y, Gbur G. Measurement of atmospheric turbulence strength by vortex beam. Opt Commun. 2010;283:1209–1212.
  • Li X, Tai Y, Zhang L, et al. Characterization of dynamic random process using optical vortex metrology. Appl Phys B. 2014;116:901–909.
  • Lavery MPJ, Speirits FC, Barnett SM, et al. Detection of a spinning object using light’s orbital angular momentum. Science. 2013;341:537–540.
  • Nagali E, Sciarrino F, De Martini F, et al. Quantum information transfer from spin to orbital angular momentum of photons. Phys Rev Lett. 2009;103:013601.
  • Wang J, Yang J, Fazal IM, et al. Terabit free-space data transmission employing orbital angular momentum multiplexing. Nat Photon. 2012;6:488–496.
  • Paterson C. Atmospheric turbulence and orbital angular momentum of single photons for optical communication. Phys Rev Lett. 2005;94:153901.
  • Thidé B, Then H, Sjöholm J, et al. Utilization of photon orbital angular momentum in the low-frequency radio domain. Phys Rev Lett. 2007;99:087701.
  • Tamburini F, Anzolin G, Umbriaco G, et al. Overcoming the Rayleigh criterion limit with optical vortices. Phys Rev Lett. 2006;97:163903.
  • Yu W, Ji Z, Dong D, et al. Super-resolution deep imaging with hollow Bessel beam STED microscopy. Laser Photon Rev. 2016;10:147–152.
  • Zhu K, Zhou G, Li X, et al. Propagation of Bessel-Gaussian beams with optical vortices in turbulent atmosphere. Opt Express. 2008;16:21315–21320.
  • Schwarz U, Sogomonian S, Maier M. Propagation dynamics of phase dislocations embedded in a Bessel light beam. Opt Commun. 2002;208:255–262.
  • Orlov S, Regelskis K, Smilgevičius V, et al. Propagation of Bessel beams carrying optical vortices. Opt Commun. 2002;209:155–165.
  • Flossmann F, Schwarz U, Maier M. Propagation dynamics of optical vortices in Laguerre–gaussian beams. Opt Commun. 2005;250:218–230.
  • Yang Y, Dong Y, Zhao C, et al. Generation and propagation of an anomalous vortex beam. Opt Lett. 2013;38:5418–5421.
  • Vaity P, Rusch L. Perfect vortex beam: fourier transformation of a Bessel beam. Opt Lett. 2015;40:597–600.
  • Li P, Zhang Y, Liu S, et al. Generation of perfect vectorial vortex beams. Opt Lett. 2016;41:2205–2208.
  • Chen W, Zhan Q. Realization of an evanescent Bessel beam via surface plasmon interference excited by a radially polarized beam. Opt Lett. 2009;34:722–724.
  • Zhan Q. Properties of circularly polarized vortex beams. Opt Lett. 2006;31:867–869.
  • Kristensen M, Beijersbergen MW, Woerdman JP. Angular momentum and spin-orbit coupling for microwave photons. Opt Commun. 1994;104:229–233.
  • Beijersbergen MW, Coerwinkel RPC, Kristensen M, et al. Helical-wavefront laser beams produced with a spiral phaseplate. Opt Commun. 1994;112:321–327.
  • Turnbull GA, Robertson DA, Smith GM, et al. The generation of free-space Laguerre-Gaussian modes at millimetre-wave frequencies by use of a spiral phaseplate. Opt Commun. 1996;127:183–188.
  • Hariharan P. Optical holography. Cambridge: Cambridge University Press; 1996.
  • Leith EN, Upatnieks J. Reconstructed wavefronts and communication theory. J Opt Soc Am A. 1962;52:1123–1130.
  • Leith EN, Upatnieks J. Wavefront reconstruction with continuous-tone objects. J Opt Soc Am A. 1963;53:1377–1381.
  • Beijersbergen MW, Allen L, van der Veen HELO, et al. Astigmatic laser mode converters and transfer of orbital angular momentum. Opt Commun. 1996;96:183–188.
  • Padgett M, Arlt J, Simpson N. An experiment to observe the intensity and phase structure of Laguerre-Gaussian laser modes. Am J Phys. 1996;64:77–82.
  • Li S, Mo Q, Hu X, et al. Controllable all-fiber orbital angular momentum mode converter. Opt Lett. 2015;40.
  • Du C, Wang J, Mo Q. Controllable all-fiber orbital angular momentum mode converter. Opt Lett. 2015;40:4376–4379.
  • Karimi E, Schulz SA, De Leon I, et al. Generating optical orbital angular momentum at visible wavelengths using a plasmonic metasurface. Light Sci Appl. 2014;3:e167.
  • Platt BC, Shack R. History and principle of Shack-Hartmann wavefront sensing. J Refract Srug. 2001;17:S573–S577.
  • Leach J, Keen S, Padgett MJ, et al. Direct measurement of the skew angle of the Poynting vector in a helically phased beam. Opt Express. 2006;14:11919–11924.
  • Chen M, Roux FS, Olivier JC. Detection of phase singularities with a Shack-Hartmann wavefront sensor. J Opt Soc Am A. 2007;24:1994–2002.
  • Luo J, Huang H, Matsui Y, et al. High-order optical vortex position detection using a Shack-Hartmann wavefront sensor. Opt Express. 2015;23:8706–8719.
  • Berkhout GC, Beijersbergen MW. Method for probing the orbital angular momentum of optical vortices in electromagnetic waves from astronomical objects. Phys Rev Lett. 2008;101:100801.
  • Sztul H, Alfano R. Double-slit interference with Laguerre-Gaussian beams. Opt Lett. 2006;31:999–1001.
  • Harris M, Hill CA, Tapster PR, et al. Laser modes with helical wave fronts. Phys Rev A. 1994;49:3119.
  • Hickmann J, Fonseca E, Soares W, et al. Unveiling a truncated optical lattice associated with a triangular aperture using light’s orbital angular momentum. Phys Rev Lett. 2010;105:053904.
  • De Araujo LE, Anderson ME. Measuring vortex charge with a triangular aperture. Opt Lett. 2011;36:787–789.
  • Berkhou GCG, Lavery MPJ, Courtial J, et al. Efficient sorting of orbital angular momentum states of light. Phys Rev Lett. 2010;105:153601.
  • Schulze C, Dudley A, Flamm D, et al. Measurement of the orbital angular momentum density of light by modal decomposition. New J Phys. 2013;15:073025.
  • Araujo LEE, Anderson ME. Measuring vortex charge with a triangular aperture. Opt Lett. 2011;36:787–789.
  • Liu R, Long J, Wang F, et al. Characterizing the phase profile of a vortex beam with angular-double-slit interference. J Opt. 2013;15:125712.
  • Fu D, Chen D, Liu R, et al. Probing the topological charge of a vortex beam with dynamic angular double slits. Opt Lett. 2015;40:788–791.
  • Guo CS, Yue SJ, Wei GX. Measuring the orbital angular momentum of optical vortices using a multipinhole plate. Appl Phys Lett. 2009;94:231104.
  • Denisenko V, Shvedov V, Desyatnikov AS, et al. Determination of topological charges of polychromatic optical vortices. Opt Express. 2009;17:23374–23379.
  • Kotlyar VV, Kovalev AA, Porfirev AP. Astigmatic transforms of an optical vortex for measurement of its topological charge. Appl Opt. 2017;56:4095–4104.
  • Vinu RV, Singh RK. Determining helicity and topological structure of coherent vortex beam from laser speckle. Appl Phys Lett. 2016;109:111108.
  • Prabhakar S, Kumar A, Banerji J, et al. Revealing the order of a vortex through its intensity record. Opt Lett. 2011;36:4398–4400.
  • Zhao P, Li S, Feng X, et al. Measuring the complex orbital angular momentum spectrum of light with a mode-matching method. Opt Lett. 2017;42:1080–1083.
  • Dudley A, Litvin IA, Forbes A. Quantitative measurement of the orbital angular momentum density of light. Appl Opt. 2012;51:823–833.
  • Zhou HL, Fu DZ, Dong JJ, et al. Orbital angular momentum complex spectrum analyzer for vortex light based on the rotational Doppler effect. Light Sci Appl. 2017;6:e16251.
  • Golub MA, Prokhorov AM, Sisakyan IN, et al. Synthesis of spatial filter for investigation of the transverse mode composition of coherent radiation. Sov J Quantum Electron. 1982;12:1208–1209.
  • Golub MA, Karpeev SV, Krivoshlykov SG, et al. Experimental investigation of spatial filters separating transverse modes of optical fields. Sov J Quantum Electron. 1983;13:1123–1124.
  • Golub MA, Karpeev SV, Krivoshlykov SG, et al. Spatial filter investigation of the distribution of power between transverse modes in a fiber waveguide. Sov J Quantum Electron. 1984;14:1255–1256.
  • Yang Y, Dong Y, Zhao C, et al. Autocorrelation properties of fully coherent beam with and without orbital angular momentum. Opt Express. 2014;2:2925–2932.
  • Gori F, Santarsiero M, Borghi R, et al. Partially coherent sources with helicoidal modes. J Mod Opt. 1998;45:539–554.
  • Ponomarenko SA. A class of partially coherent beams carrying optical vortices. J Opt Soc Am A. 2001;18:150–156.
  • Boggatyryova VG, Felde VC, Polyanskii PV, et al. Partially coherent vortex beams with a separable phase. Opt Lett. 2003;28:878–880.
  • Palacios D, Maleev I, Marathay A, et al. Spatial correlation singularity of a vortex field. Phys Rev Lett. 2004;92:143905.
  • Gbur G, Visser TD, Wolf E. ‘Hidden’ singularities in partially coherent wavefields. J Opt A Pure Appl Opt. 2004;6:S239–S242.
  • Visser TD, Gbur G, Wolf E. Effect of the state of coherence on the three-dimensional spectral intensity distribution near focus. Opt Commun. 2002;213:13–19.
  • Bouchal Z, Perina J. Non-diffracting beams with controlled spatial coherence. Opt Acta Int J Opt. 2002;49:1673–1689.
  • Gbur G, Visser TD. Coherence vortices in partially coherent beams. Opt Commun. 2003;222:117–125.
  • Ponomarenko SA. A class of partially coherent beams carrying optical vortices. J Opt Soc Am A. 2001;18:150–156.
  • Maleev ID, Palacios DM, Marathay AS, et al. Spatial correlation vortices in partially coherent light: theory. J Opt Soc Am B. 2004;21:1895–1900.
  • Jeng CC, Shih MF, Motzek K, et al. Partially incoherent optical vortices in self-focusing nonlinear media. Phys Rev Lett. 2004;92:043904.
  • Schouten HF, Visser TD, Dijk TV. Evolution of singularities in a partially coherent vortex beam. J Opt Soc Am A. 2009;26:741–744.
  • Wang F, Zhu S, Cai Y. Experimental study of the focusing properties of a Gaussian Schell-model vortex beam. Opt Lett. 2011;36:3281–3283.
  • Yang Y, Chen M, Mazilu M, et al. Effect of the radial and azimuthal mode indices of a partially coherent vortex field upon a spatial correlation singularity. New J Phys. 2013;15:113053.
  • Zhang Z, Fan H, Xu HF, et al. Three-dimensional focus shaping of partially coherent circularly polarized vortex beams using a binary optic. J Opt. 2015;17:065611.
  • Ostrovsky AS, Garcíagarcía J, Rickenstorffparrao C, et al. Partially coherent diffraction-free vortex beams with a Bessel-mode structure. Opt Lett. 2017;42:5182–5185.
  • Singh RK, Sharma AM, Senthilkumaran P. Vortex array embedded in a partially coherent beam. Opt Lett. 2015;40:2751–2754.
  • Stahl C, Gbur G. Partially coherent vortex beams of arbitrary order. J Opt Soc Am A. 2017;34:1793–1799.
  • Liu D, Wang Y, Yin H. Evolution properties of partially coherent flat-topped vortex hollow beam in oceanic turbulence. Appl Opt. 2015;54:10510.
  • Qin Z, Tao R, Zhou P, et al. Propagation of partially coherent Bessel–gaussian beams carrying optical vortices in non-Kolmogorov turbulence. Opt Laser Technol. 2014;56:182–188.
  • Cheng M, Guo L, Li J, et al. Propagation of an optical vortex carried by a partially coherent Laguerre-Gaussian beam in turbulent ocean. Appl Opt. 2016;55:4642–4648.
  • Zhang Y, Ma D, Zhou Z, et al. Research on partially coherent flat-topped vortex hollow beam propagation in turbulent atmosphere. Appl Opt. 2017;56:2922–2926.
  • Liu D, Yin H, Wang G, et al. Propagation of partially coherent Lorentz-Gauss vortex beam through oceanic turbulence. Appl Opt. 2017;56:8785–8792.
  • Chen Y, Wang F, Zhao C, et al. Experimental demonstration of a Laguerre-Gaussian correlated Schell-model vortex beam. Opt Express. 2014;22:5826–5838.
  • Liu X, Liu L, Peng X, et al. Partially coherent vortex beam with periodical coherence properties. J Quant Spectrosc Ra. 2019;222–223:138–144.
  • Liu X, Wang F, Liu L, et al. Generation and propagation of an electromagnetic Gaussian Schell-model vortex beam. J Opt Soc Am A. 2015;32:2058–2065.
  • Guo L, Chen Y, Liu X, et al. Vortex phase-induced changes of the statistical properties of a partially coherent radially polarized beam. Opt Express. 2016;24:13714–13728.
  • Simon R, Mukunda N. Twisted Gaussian Schell-model beams. J Opt Soc Am A. 1993;10:95–109.
  • Peng X, Liu L, Wang F, et al. Twisted Laguerre-Gaussian Schell-model beam and its orbital angular moment. Opt Express. 2018;26:33956–33969.
  • Lin Q, Cai Y. Tensor ABCD law for partially coherent twisted anisotropic Gaussian-Schell model beams. Opt Lett. 2002;27:216–218.
  • Chen J, Zhang E, Peng X, et al. Efficient tensor approach for simulating paraxial propagation of arbitrary partially coherent beams. Opt Express. 2017;25:24780–24789.
  • Chen J, Liu X, Yu J, et al. Simultaneous determination of the sign and the magnitude of the topological charge of a partially coherent vortex beam. Appl Phys B. 2016;122:1–12.
  • Maleev ID, Palacios DM, Marathay AS, et al. Spatial correlation vortices in partially coherent light: theory. J Opt Soc Am B. 2004;21:1895–1900.
  • Maleev ID, SwartzlandSwartzlander, Jr. GA. Propagation of spatial correlation vortices. J Opt Soc Am B. 2008;25:915–922.
  • Swartzlander GA, Schmit J. Temporal correlation vortices and topological dispersion. Phys Rev Lett. 2004;93:093901.
  • van Dijk T, Visser TD. Evolution of singularities in a partially coherent vortex beam. J Opt Soc Am A. 2009;26:741–744.
  • Gu Y, Gbur G. Topological reactions of optical correlation vortices. Opt Commmun. 2009;282:709–716.
  • Liu X, Shen Y, Liu L, et al. Experimental demonstration of vortex phase-induced reduction in scintillation of a partially coherent beam. Opt Lett. 2013;38:5323–5326.
  • Zeng J, Liu X, Wang F, et al. Partially coherent fractional vortex beam. Opt Express. 2018;26:26830–26844.
  • Wolf E. New theory of partial coherence in the space-frequency domain. J Opt Soc Am. 1980;70:1622.
  • Xiao X, Voelz D. Wave optics simulation approach for partial spatially coherent beams. Opt Express. 2006;14:6986–6992.
  • Qian X, Zhu W, Rao R. Numerical investigation on propagation effects of pseudo-partially coherent Gaussian Schell-model beams in atmospheric turbulence. Opt Express. 2009;17:3782–3791.
  • Zhang Y, Ma D, Zhou Z, et al. The research on partially coherent flat-topped vortex hollow beams propagation in turbulent atmosphere. Appl Opt. 2017;56:2922–2926.
  • Hyde MWH Iv, Basu S, Xiao X, et al. Producing any desired far-field mean irradiance pattern using a partially-coherent Schell-model source. J Opt. 2015;17:055607.
  • Wang F, Yu J, Liu X, et al. Research progress in propagation of partially coherent beams in turbulent atmosphere. Acta Phys Sin. 2018;67:184203.
  • Zhao C, Wang F, Dong Y, et al. Effect of spatial coherence on determining the topological charge of a vortex beam. Appl Phys Lett. 2012;101:261104.
  • Liu X, Peng X, Liu L, et al. Self-reconstruction of the degree of coherence of a partially coherent vortex beam obstructed by an opaque obstacle. Appl Phys Lett. 2017;110:181104.
  • Bouchal Z, Wagner J, Chlup M. Self-reconstruction of a distorted nondiffracting beam. Opt Commun. 1998;151:207–211.
  • Wu G, Wang F, Cai Y. Generation and self-healing of a radially polarized Bessel-Gaussian beam. Phys Rev A. 2014;89:043807.
  • Broky J, Siviloglou GA, Dogariu A, et al. Self-healing properties of optical Airy beams. Opt Express. 2008;16:12880–12885.
  • Garces-Chavez V, McGloin D, Melville H, et al. Simultaneous micromanipulation in several planes using a self-reconstructing light beam. Nature. 2002;419:145–147.
  • Fahrbach FO, Simon P, Rohrbach A. Microscopy with self-reconstructing beam. Nat Photon. 2010;4:780–786.
  • Chen X, Li J, Rafsanjani SMH, et al. Synthesis of Im-Bessel correlated beams via coherent modes. Opt Lett. 2018;43:3590–3593.
  • Zhu X, Wang F, Zhao C, et al. Experimental realization of dark and antidark diffraction-free beams. Opt Lett. 2019.
  • Zhao C, Dong Y, Wang Y, et al. Experimental generation of a partially coherent Laguerre-Gaussian beam. Appl Phys B. 2012;109:345–349.
  • Liu R, Wang F, Chen D, et al. Measuring mode indices of a partially coherent vortex beam with Hanbury Brown and Twiss type experiment. App Phys Lett. 2016;108:051107.
  • Perez-Garcia B, Yepiz A, Hernandez-Aranda RI, et al. Digital generation of partially coherent vortex beams. Opt Lett. 2016;41:3471.
  • Yang Y, Liu Y. Measuring azimuthal and radial mode indices of a partially coherent vortex field. J Opt. 2016;18:015604.
  • Liu X, Wu T, Liu L, et al. Experimental determination of the azimuthal and radial mode orders of a partially coherent LGpl beam. Chin Opt Lett. 2017;15:030002.
  • Lu X, Zhao C, Shao Y, et al. Phase detection of coherence singularities and determination of the topological charge of a partially coherent vortex beam. Appl Phys Lett. 2019.
  • Chen J, Li Y. Discrimination of incoherent vortex states of light. Opt Lett. 2018;43:5595–5598.