• Photonics Research
  • Vol. 9, Issue 5, 856 (2021)
Di Lin1、*, Joel Carpenter2, Yutong Feng1, Yongmin Jung1, Shaif-ul Alam1, and David J. Richardson1
Author Affiliations
  • 1Optoelectronics Research Centre, University of Southampton, Southampton SO17 1BJ, UK
  • 2School of Information Technology and Electrical Engineering, The University of Queensland, Brisbane, QLD 4072, Australia
  • show less
    DOI: 10.1364/PRJ.412342 Cite this Article Set citation alerts
    Di Lin, Joel Carpenter, Yutong Feng, Yongmin Jung, Shaif-ul Alam, David J. Richardson. High-power, electronically controlled source of user-defined vortex and vector light beams based on a few-mode fiber amplifier[J]. Photonics Research, 2021, 9(5): 856 Copy Citation Text show less

    Abstract

    Optical angular momentum (OAM)-based structured light beams provide an additional degree of freedom for practical applications ranging from optical communication to laser-based material processing. Many techniques exist for generating such beams within laser sources and these primarily rely upon the use of specially designed optical components that limit laser power scaling and ready tunability of the topological charge and polarization of the output OAM beams. Here we show that some of these limitations can be overcome by employing a computer controlled reflective phase-only spatial light modulator to adaptively tailor the input (and subsequent output) beam wavefront and polarization in a few-mode fiber amplifier. In this way modal-coupling-induced beam distortion within the fiber amplifier can be mitigated and we are able to generate at will any desired supported spatial mode guided in the fiber, including conventional linearly polarized (LP) modes, scalar OAM modes, and cylindrical vector modes, at average powers >10 W and with a peak power of >11 kW. Our results pave the way to the realization of practical high-power structured laser sources with tunable chirality and polarization.

    1. INTRODUCTION

    Structured laser beams have recently become a topic of significant interest due to the fact that they provide additional degrees of freedom in terms of optical phase, polarization, and amplitude and have subsequently found many novel applications in, among others, the optical manipulation of particles [1,2], optical free space and fiber-based communications [3,4], laser-based material processing [5,6], particle acceleration [7], and microscopy [8,9]. Foremost among the family of structured laser beams are scalar vortex beams and cylindrical vector beams. Scalar vortex beams are characterized by an annular intensity profile associated with an azimuthally varying phase structure of exp(ilφ) (where l is the integer of topological charge) so that they carry optical angular momentum (OAM). They are typically generated using dynamic phase approaches by imposing a helical phase front on an incoming Gaussian beam using a phase plate [10], computer-generated hologram [11], or phase-only spatial light modulator (SLM) [12]. On the other hand, cylindrical vector beams are associated with l-dependent azimuthally varying polarization distributions (such as radial and azimuthal polarizations), which can be interpreted as combinations of OAM and spin states that result in a plane wavefront that does not carry OAM [13]. Such beams are created with geometric phase approaches by employing a q-plate [14], s-wave plate [15], or meta-surfaces [16] to transform an incident beam with homogenous polarization (e.g., linear polarization) into a beam with the desired vector polarization state. While these passive beam conversion approaches are relatively easy to implement, they tend to lack flexibility in terms of the generation of different structured beams due to the static properties of the optical elements, have limited power handling capability, and the conversion efficiency is ultimately limited by the intensity overlap of the incident Gaussian beam with the desired structured beam (which usually induces a high loss when a high-order structured beam is desired [17]).

    Over the past decade, remarkable advances have been made in the direct generation of such structured light beams from gain media including bulk solid-state lasers/amplifiers [1820], fiber lasers/amplifiers [21,22], organic lasers [23], and gas lasers [24], offering advantages of higher modal purity and much higher power scalability. However, such “structured light” laser systems can typically only generate a specific fixed spatial mode. The selective and flexible generation of such beams (i.e., tuning of the polarization and topological charge of the OAM states on demand), ideally from a compact laser source at high output powers with high modal purity and with simultaneous control of the temporal characteristics, is of great interest in direct laser writing with ultrafast lasers which brings flexibility to the laser-induced morphology of features [25], but remains a great challenge. A few methods that provide some flexibility in switching the laser output intensity profiles, or tuning the chirality of the OAM states, have been implemented in bulk solid-state lasers [2628] and semiconductor-based microlasers [2932] but generally the performance has been limited by damage to the special optical components used due to the high intracavity powers involved. One promising way to overcome this limitation is to exploit the master-oscillator power amplifier (MOPA) approach that amplifies a reconfigurable structured light seed beam through an appropriate gain medium with a high gain while preserving a high mode purity. We have recently demonstrated a proof-of-concept demonstration that flexibly generates various structured light beams from a compact multicore fiber amplifier by coherently combining the individual Gaussian-shaped output beamlets with appropriate complex amplitude [33]. However, the unavoidable relatively higher loss associated with the non-collinear beam combination (typically 50%60% in theory) could ultimately hinder its use in the high-power regime and provides the incentive to develop more efficient approaches.

    The vector nature of the diversity of electric field distributions of the eigenmodes guided in a cylindrically symmetric isotropic large-core, few-mode optical fiber makes this an attractive compact platform for the dynamic generation of different structured spatial modes at high powers given the high efficiency and high-power scalability of traditional Gaussian-shaped beams in such gain fibers. However, the vector modes with the same radial and azimuthal index are either strictly degenerate with exactly the same effective refractive indices (i.e., EHl1,meven and EHl1,modd), or nearly degenerate with a small difference in the effective refractive indices (Δn105107, i.e., EHl1,meven and HEl+1,meven). Thus, even the slightest perturbations to the cylindrical symmetry of the fiber (i.e., bends, twists, and refractive index inhomogeneities) can cause strong coupling between such modes, resulting in the formation of the so-called scalar LP modes that are usually observed at the output of fibers/fiber devices [34,35]. Although, various mode excitation/selection techniques have been demonstrated in fiber lasers/amplifiers, these have generally only allowed the generation of a single, non-tailorable lowest-order vortex or vector mode for a given laser cavity design [22,36,37]. In such systems, mode-coupling-induced distortion of the vortex or vector beam can be reduced, to some extent at least, by deliberate mechanical manipulation of the fiber, e.g.,  by applying extra pressure, bends, or twists at certain points to ensure a doughnut-shaped intensity profile at the fiber output. These mitigation measures though are wholly empirical, typically lack repeatability, and become increasingly ineffective at higher power levels and for fibers supporting a larger number of guided modes. This ultimately hinders the utilization of such exotic laser sources for end-users who are not laser experts. More advanced fibers with annular index profiles that reduce mode coupling and support the stable propagation and amplification of vortex and vector modes over long distance have been proposed and demonstrated [35,38,39]. However, such fibers rely on critical design and rigorous control of the refractive index profile and thickness of the annular core required and are characterized by a small effective mode area. Thus, fiber fabrication is challenging, and the tight mode confinement is not compatible with high-power laser operation.

    In this work, we overcome the aforementioned limitations through the use of an adaptive input wavefront and polarization shaping technique which allows for precise control of the amplification of structured spatial modes in a commercially available few-mode large mode area (FM-LMA) fiber. A reflective phase-only SLM is employed to excite and correct the wavefront and polarization of the input beam to a final FM-LMA fiber amplifier stage in a picosecond pulsed MOPA system to obtain the desired structured spatial modes at the MOPA output. We experimentally demonstrate the computer-controlled generation of a variety of spatial modes supported by the FM-LMA fiber, obtaining very good modal purity (>90%), an average power of >10  W, and a corresponding peak power of 11 kW. This includes the arbitrary generation of conventional scalar LP modes, linearly polarized OAM modes, and vector eigenmodes.

    2. METHODS

    A. Principles

    Light propagation through passive fibers remains deterministic even in the case of strong modal coupling. Indeed, a number of techniques based on manipulation of the incident wavefront (including digital phase conjugation [40], transmission matrix [41], and adaptive wavefront shaping [42]) have been demonstrated with great success to selectively generate a desired electric field distribution at the output of a passive multimode fiber (MMF) by using a phase-only SLM [4345]. Recently the iterative optimization-based adaptive wavefront shaping technique has been adapted to shape the light in FM-LMA fiber amplifiers and MMF amplifiers in which light transmission becomes non-linear due to the gain competition among the transverse modes induced by non-uniform gain saturation. For instance, by employing a photonic lantern as the beam shaping element, the amplification of selected scalar LP modes in an LMA fiber amplifier has been demonstrated [46,47], and amplification in an MMF amplifier resulting in a tightly confined, single-spot (or multiple spots) output beam has been obtained using a deformable mirror to tailor the incident beam wavefront [48]. However, the deformable mirror has a very limited number of actuators, which means it cannot readily be used to generate more sophisticated structured beams with spatially varying phase and/or polarization distributions. To date, no technique has been demonstrated with the capability of allowing the user to generate a specific selected complex guided spatial mode in a fiber laser/amplifier at high output power.

    Spatial beam shaping with SLMs is widely used as one of the most versatile techniques for generating structured light beams in free space with high fidelity; however, there are issues with this approach. First, shaping efficiency can be an issue particularly due to diffraction loss when there is a limited intensity overlap between the incident beam and the target beam. Second, liquid crystals devices have a relatively low damage threshold (typically 2  Wcm2) and this limits the power level of the beams that can be reflected from these devices. Here, we propose a “digital fiber amplifier” [26] to overcome these issues. One of the central ideas behind our approach is to place the beam shaping SLM prior to an FM-LMA fiber amplifier, where the incident power levels are modest, allowing a pre-shaped beam to be amplified to a target beam with much higher output powers than what could be applied directly to the SLM if it was used to shape the beam at the amplifier output, while also avoiding any shaping losses at this critical point in the system. The SLM provides a reprogrammable means of converting a pre-amplified Gaussian-shaped beam into the required field which must be excited at the input of the fiber to generate the desired output. Any beam distortion and depolarization induced by mode coupling during the amplification are pre-compensated by adaptively shaping the incident wavefront and polarization state through an iterative optimization process. The shaped incident wavefront can be encoded as a superposition of a basis set of orthogonally linearly polarized modes generated by the SLM with an appropriate complex amplitude: Ein=l=Nl=N(al|H,l+bl|V,l),where |H,l=eilφ|H and |V,l=eilφ|V represent the basis modes (horizontal polarization |H and vertical polarization |V) with a helical wavefront eilφ (and with an intensity profile matched to those of the fiber eigenmodes), and al=al0eiθl1 and bl=bl0eiθl2 represent the complex amplitude of each mode (where al0 and bl0 represent the normalized amplitude of the horizontal and vertical components, respectively; and θl1(2) represents the relative phase of each mode). Any spatial mode can be expressed as an appropriate superposition of these basis modes.

    At the output, an analyzer SLM (used in the reverse direction to the beam shaping SLM) can be used as a correlation filter to detect the portion of the field in the desired state. It is digitally configured with a transmission function T(r)=Ψtar*(r) [where Ψtar*(r) represents the complex conjugate of the target mode in terms of both amplitude and phase] so that the on-axis intensity of the correlation signal in the far field is proportional to the power of the output beam within that target mode [49]. The on-axis signal power can be measured and used as the merit function to provide iterative feedback to adaptively shape the input wavefront and polarization. The optimization process is executed by the optimization package based on the conjugate gradient algorithm from the ALGLIB numerical library which solves unconstrained problems [50]. The coefficients of the amplitude and phase of each basis mode in each polarization state result in a total of 24 parameters to be optimized. Therefore, the output beam can be electronically controlled simply by displaying different (complex) phase masks on the SLM.

    B. Experimental Setup

    Experimental setup of the digital fiber amplifier. AMP, amplifier; PM, polarization maintaining; SMF, single-mode fiber; FMF, few-mode fiber; PBS, polarization beam splitter; DM, dichroic mirror; BS, beam splitter; PC, polarization controller; M, mirror; λ/2, half-wave plate; ISO, isolator.

    Figure 1.Experimental setup of the digital fiber amplifier. AMP, amplifier; PM, polarization maintaining; SMF, single-mode fiber; FMF, few-mode fiber; PBS, polarization beam splitter; DM, dichroic mirror; BS, beam splitter; PC, polarization controller; M, mirror; λ/2, half-wave plate; ISO, isolator.

    The re-shaped beam is first coupled to a 10  m length of passive single-clad fiber with a refractive index profile matched to that of the commercially available step-index FM-LMA Yb-doped fiber (Liekki Yb1200-30/250DC) used in the final stage amplifier. The passive single-clad fiber is used as an aperture that effectively blocks any forward propagated unwanted light that otherwise might be amplified, and also works with the free-space isolator to block any backward propagated amplified spontaneous emissions (ASEs) to protect the SLM. The resulting signal is passed through a home-made free-space polarization-independent isolator which ensures a reduced polarization-dependent loss and is injected into the final stage amplifier fiber which has a core diameter of 30 μm with an NA of 0.075±0.005 and a cladding diameter of 250 μm with an NA of 0.46. Our calculations show that the fiber can theoretically guide 20 spatial modes (accounting for the two polarizations) if kept straight. In practice, however, as the fiber was coiled with a diameter of 25  cm, only 12 spatial modes can be effectively guided (up to the LP02 mode) due to the high bend loss sensitivity of the higher-order modes. The fiber has a peak nominal cladding absorption of 14  dB/m at 976 nm, and 1.5  m length of fiber was used in the experiment. Both ends of the fiber were spliced to 1  mm long silica end caps with a diameter of 250 μm to suppress any potential parasitic lasing. The input end facet was perpendicularly cleaved but the output end facet was angle polished with an angle of 8° to further reduce the backreflection and unwanted optical feedback.

    At the fiber amplifier output, a fraction of the output beam is characterized by a digital spatial mode correlation filter based on a second SLM, and the resulting correlation signal is used in an iterative optimization process to adjust the wavefront and polarization of the beam launched into the fiber as needed to generate the target beam at the output. Note that we used a basis set of linearly polarized modes with helical wavefronts (and with intensity profiles matched to those of the fiber) to generate and analyze the spatial beam profile at the SLMs. Any scalar or vector fiber mode can thus be described in terms of a superposition of modes in this basis.

    3. RESULTS

    A. Generation of Scalar LP Modes

    Scalar LP modes. (a) Examples of input and output beam profiles without input wavefront shaping. (b), (c) Modal decomposition of (b) the unshaped LP01 mode and (c) the optimized LP01 mode. (d) Measured intensity profiles of the different orders of output LP mode after input wavefront shaping. (e) Measured output intensity profiles of the rotated LP11 mode.

    Figure 2.Scalar LP modes. (a) Examples of input and output beam profiles without input wavefront shaping. (b), (c) Modal decomposition of (b) the unshaped LP01 mode and (c) the optimized LP01 mode. (d) Measured intensity profiles of the different orders of output LP mode after input wavefront shaping. (e) Measured output intensity profiles of the rotated LP11 mode.

    We then tested the feasibility of our beam shaping concept with a targeted vertically polarized fundamental LP01 mode. In this case, only the section of the second SLM allocated to controlling the vertical polarization component of the beam was coded with a phase mask corresponding to the conjugated LP01 mode such that the correlation signal can be detected by the SMF attached to the power meter. This is then used as the merit function in the iterative optimization process. The initial phase of each mode was preset to 0 and the amplitude of each mode was randomly preset but with a higher value for the target LP01 mode. The preset parameters are not critical in determining whether we can find a good solution or not, but a good choice can reduce the optimization time which typically takes about 1015  min. Figure 2(c) shows the measured intensity profile of the optimized LP01 mode, and the modal decomposition result shows that the PER was increased to 12  dB with an improved modal purity of 90.1%.

    Next, we chose to shape the input signal wavefront to obtain other selected vertically polarized scalar LP modes. Such modes can be represented as the in-phase and out-of-phase coherent superposition of two equally weighted vertically polarized OAM modes with opposite chirality (LPa(b)=|V,l±|V,l). The appropriate correlation filter was encoded to display the conjugate phase of the desired LP mode at the second SLM, and the iterative optimization procedure was again executed to obtain the target LP modes by controlling the input wavefront. We successfully generated the individual LP modes supported by the fiber, as shown in Fig. 2(d). In addition, the orientation (α) of the LP modes which can be interpreted in terms of a superposition of the two linearly polarized OAM modes with equal amplitude but different relative phase β [LP(α)=|V,l+exp(iβ)|V,l, where α=β/2l] can be completely controlled. One example of the measured intensity profiles of the rotated LP11 output beams is shown in Fig. 2(e), indicating that the complex amplitudes of the four vector modes within the LP11 group are well controlled. Note that all generated LP modes had an output power of 10  W with a PER of >12  dB.

    B. Generation of Vector Fiber Eigenmodes

    We further investigated the generation of degenerate fiber eigenmodes in the LP11 and LP21 groups supported by the fiber. Mathematically, such vector modes can be represented as an appropriate linear combination of the four OAM modes basis in the same LP group [the opposite chirality and orthogonal linear polarizations, i.e., TM01=(|H,l+|H,li|V,l+i|V,l), where l=1]. Thus, the conjugate phase of the desired vector mode can be encoded on the second SLM to construct an appropriate correlation filter. In order to get an unambiguous correlation signal, the PC in Fig. 1 needs to be properly adjusted to ensure that the polarization of a horizontally polarized beam launched into the SMF is rotated to 45° with respect to the transmission axis of the fiberized polarization beam splitter (PBS). This ensures that the correlation signal of the desired vector mode is solely transmitted to the horizontal polarization output port of the PBS and is measured by the power meter as the merit function during the iterative optimization process. Note that there is a fixed slight path length difference (100  μm) between the horizontal and vertical polarizations through the correlation filter system so that the relative phase between them should be calibrated before implementing the iterative optimization process. Afterwards, the conjugate amplitude and phase of the desired vector mode were encoded in the second SLM, and by executing the adaptive wavefront shaping procedure, various fiber eigenmodes were successfully generated with high quality.

    Cylindrical vector modes. (a) Theoretical (first column) and experimentally measured (second column) far-field intensity distributions of doughnut-shaped output vector modes and the intensity profiles of the corresponding mode passing through a linear polarizer at different orientations (white arrows) as a verification of the vector polarization state. (b) Average output power of the TM01 and EH11e modes versus the launched pump power. (c) The measured temporal pulse shape having a pulse duration (FWHM) of ∼150 ps. (d), (e) Examples of vector modal decomposition analysis for (d) the TM01 mode and (e) the EH11e mode, showing that >91% of power is contained within the desired vector.

    Figure 3.Cylindrical vector modes. (a) Theoretical (first column) and experimentally measured (second column) far-field intensity distributions of doughnut-shaped output vector modes and the intensity profiles of the corresponding mode passing through a linear polarizer at different orientations (white arrows) as a verification of the vector polarization state. (b) Average output power of the TM01 and EH11e modes versus the launched pump power. (c) The measured temporal pulse shape having a pulse duration (FWHM) of 150  ps. (d), (e) Examples of vector modal decomposition analysis for (d) the TM01 mode and (e) the EH11e mode, showing that >91% of power is contained within the desired vector.

    C. Generation of Linearly Polarized OAM Modes

    OAM modes. (a)–(d) The measured intensity profiles of the linearly polarized OAM output beams with topological charge of (a) l=−1, (b) l=1, (c) l=2, and (d) l=−2, respectively. The insets show the orthogonal components of the horizontal and vertical polarization states, indicating that the linear polarization dominates the output beams. The different spiral interference patterns indicate the output beams have the desired topological charges of the OAM state. (e), (f) Examples of modal decompositions results show that >90% of the power is contained in the desired OAM mode of (e) |V,1⟩ and (f) |H,2⟩.

    Figure 4.OAM modes. (a)–(d) The measured intensity profiles of the linearly polarized OAM output beams with topological charge of (a) l=1, (b) l=1, (c) l=2, and (d) l=2, respectively. The insets show the orthogonal components of the horizontal and vertical polarization states, indicating that the linear polarization dominates the output beams. The different spiral interference patterns indicate the output beams have the desired topological charges of the OAM state. (e), (f) Examples of modal decompositions results show that >90% of the power is contained in the desired OAM mode of (e) |V,1 and (f) |H,2.

    D. Mode Stability

    Modal stability measurement. (a) Measured modal purity of the EH11e mode at a fixed output power of ∼11 W as a function of time, indicating that a stable vector mode was generated and preserved over 60 min. (b)–(d) Mode evolution versus output power: the measured intensity profile of the optimized EH11e mode at (b) 11 W; the intensity profile of the output beam when the output power was decreased to (c) 7 W and (d) 2 W without changing the input wavefront. (e) The intensity profile of the output beam at 2 W was restored to a doughnut shape after the input wavefront was re-shaped.

    Figure 5.Modal stability measurement. (a) Measured modal purity of the EH11e mode at a fixed output power of 11  W as a function of time, indicating that a stable vector mode was generated and preserved over 60 min. (b)–(d) Mode evolution versus output power: the measured intensity profile of the optimized EH11e mode at (b) 11 W; the intensity profile of the output beam when the output power was decreased to (c) 7 W and (d) 2 W without changing the input wavefront. (e) The intensity profile of the output beam at 2 W was restored to a doughnut shape after the input wavefront was re-shaped.

    4. CONCLUSION

    We have demonstrated a novel digital fiber amplifier capable of generating optical vortex beams of user-specified topological charge and vector polarization states on demand at high power levels under electronic control. Our approach is based on adaptive shaping of the beam wavefront and polarization by means of a phase-only SLM placed at the input to a few-mode fiber amplifier. Using this approach, we successfully avoided optical damage to the SLM (by reducing the incident powers to below 250 mW) and were able to generate various spatial modes supported by the fiber including conventional scalar LP modes, scalar OAM modes, and cylindrical vector modes. The amplifier produced 150 ps pulses at a repetition rate of 5.9 MHz with an average power >10  W corresponding to a peak power of above 11  kW. The net gain of the final digital fiber amplifier stage reached >16  dB. The output power might be further increased by a factor of 35 if we increase the incident power to 1.5  W at which point we reach the theoretical damage threshold of the SLM. As the SLM is expensive, and any contamination on the liquid display surface might reduce the damage threshold, we only operated the SLM at a power well below its damage threshold. Recent advances in SLM technology have resulted in devices with a higher damage threshold and a faster refresh rate than the device used herein. This fact, in conjunction with the use of more advanced optimization algorithms capable of faster convergence rate, should allow the construction of higher-power systems with faster spatial mode switching. The demonstrated versatility in mode generation and power-handling capabilities of our approach opens a new pathway for the dynamic generation of high-power structured light beams from simple and compact laser sources. This capability may be advantageous in a wide range of applications such as laser material processing, free-space optical communication, and high-resolution imaging.

    APPENDIX A: VECTOR MODE DECOMPOSITION

    To analyze the modal purity of the generated vector modes, a vector mode decomposition technique is implemented. First, the output beam is decomposed into the orthogonal linearly polarized OAM mode basis through an inner product measurement as follows: cε,L=?Ψε,L|Eout?,where Ψε,L represents the horizontal OAM basis mode (|H,±L?) for ε=1, and vertical polarization OAM basis mode (|V,±L?) for ε=?1, respectively (L3 in practice). The cε,L2 are the corresponding modal weights which are proportional to the correlation signal of the on-axis intensity coupled to the SMF. The relative modal weight of each LP group can be derived as the sum of the modal weights of the OAM mode basis with the same order of L and orthogonal polarizations as ρL=cε,±L2l=0l=3cε,±|l|2.

    The next step is to measure the relative modal weight of each degenerate vector mode within the same LP group, which is also achieved by an inner product measurement. Consider the case where the output is the pure HEL+1even mode, then the inner product with the conjugate phase of each degenerate vector mode has the following form: {?HEL+1,meven|HEL+1,meven?=2(11)|0?+(1?1)|2L?+(1?1)|?2L??HEL+1,modd|HEL+1,meven?=i(?11)|2L?+i(1?1)|?2L??EHL?1,meven|HEL+1,meven?=2(1?1)|0?+(11)|2L?+(11)|?2L??EHL?1,modd|HEL+1,meven?=?i(11)|2L?+i(11)|?2L?.

    The SMF located on the optical axis in the far field can only detect the non-OAM component (|0?) of the electric field within the resulting correlation signal. It can be seen that the inner product of the conjugate phase of the HEL+1even mode contains an on-axis intensity component with a linear polarization aligned to 45° with respect to the horizontal direction. The inner product of the conjugate phase of the EHL?1even mode contains the on-axis intensity component as well, but with orthogonal linear polarization (?45°). To differentiate between these two signals, a PC is adjusted to rotate the linear polarization of the correlation signal of the HEL+1even phase mask to the horizontal direction, which is aligned to the first output port of the PBS. Correspondingly, the linear polarization of the correlation signal of the EHL?1even phase mask is rotated to the vertical direction, which is aligned to the second output port of the PBS. By sequentially displaying the conjugate phase of each vector mode, the measured power (Pj) of the first output port of the PBS can be considered as the weight of each vector mode. Then the relative weight (ηj) of each vector mode can be expressed as ηj=PjPj,where j represents the jth vector mode in the LP group (each LP group contains four degenerate vector modes). As a result, the overall relative weight of each vector mode (γLj) can be expressed as γLj=ηjρL.

    References

    [1] D. G. Grier. A revolution in optical manipulation. Nature, 424, 810-816(2003).

    [2] M. Padgett, R. Bowman. Tweezers with a twist. Nat. Photonics, 5, 343-348(2011).

    [3] J. Wang, J. Y. Yang, I. M. Fazal, N. Ahmed, Y. Yan, H. Huang, Y. X. Ren, Y. Yue, S. Dolinar, M. Tur, A. E. Willner. Terabit free-space data transmission employing orbital angular momentum multiplexing. Nat. Photonics, 6, 488-496(2012).

    [4] N. Bozinovic, Y. Yue, Y. X. Ren, M. Tur, P. Kristensen, H. Huang, A. E. Willner, S. Ramachandran. Terabit-scale orbital angular momentum mode division multiplexing in fibers. Science, 340, 1545-1548(2013).

    [5] J. Hamazaki, R. Morita, K. Chujo, Y. Kobayashi, S. Tanda, T. Omatsu. Optical-vortex laser ablation. Opt. Express, 18, 2144-2151(2010).

    [6] R. Weber, A. Michalowski, M. Abdou-Ahmed, V. Onuseit, V. Rominger, M. Kraus, T. Graf. Effects of radial and tangential polarization in laser material processing. Phys. Procedia, 12, 21-30(2011).

    [7] L. J. Wong. Direct acceleration of an electron in infinite vacuum by a pulsed radially-polarized laser beam. Opt. Express, 18, 25035-22501(2010).

    [8] X. A. Hao, C. F. Kuang, T. T. Wang, X. Liu. Effects of polarization on the de-excitation dark focal spot in STED microscopy. J. Opt., 12, 115707(2010).

    [9] S. Furhapter, A. Jesacher, S. Bernet, M. Ritsch-Marte. Spiral phase contrast imaging in microscopy. Opt. Express, 13, 689-694(2005).

    [10] M. W. Beijersbergen, R. P. C. Coerwinkel, M. Kristensen, J. P. Woerdman. Helical-wavefront laser beams produced with a spiral phaseplate. Opt. Commun., 112, 321-327(1994).

    [11] N. R. Heckenberg, R. McDuff, C. P. Smith, A. G. White. Generation of optical phase singularities by computer-generated holograms. Opt. Lett., 17, 221-223(1992).

    [12] G. Gibson, J. Courtial, M. Padgett, M. Vasnetsov, V. Pas’ko, S. Barnett, S. Franke-Arnold. Free-space information transfer using light beams carrying orbital angular momentum. Opt. Express, 12, 5448-5456(2004).

    [13] G. Milione, H. I. Sztul, D. A. Nolan, R. R. Alfano. Higher-order Poincare sphere, Stokes parameters, and the angular momentum of light. Phys. Rev. Lett., 107, 053601(2011).

    [14] L. Marrucci, C. Manzo, D. Paparo. Optical spin-to-orbital angular momentum conversion in inhomogeneous anisotropic media. Phys. Rev. Lett., 96, 163905(2006).

    [15] M. Beresna, M. Gecevicius, P. G. Kazansky, T. Gertus. Radially polarized optical vortex converter created by femtosecond laser nanostructuring of glass. Appl. Phys. Lett., 98, 201101(2011).

    [16] N. F. Yu, P. Genevet, M. A. Kats, F. Aieta, J. P. Tetienne, F. Capasso, Z. Gaburro. Light propagation with phase discontinuities: generalized laws of reflection and refraction. Science, 334, 333-337(2011).

    [17] A. Longman, R. Fedosejevs. Mode conversion efficiency to Laguerre-Gaussian OAM modes using spiral phase optics. Opt. Express, 25, 17382-17392(2017).

    [18] D. Lin, J. M. O. Daniel, W. A. Clarkson. Controlling the handedness of directly excited Laguerre-Gaussian modes in a solid-state laser. Opt. Lett., 39, 3903-3906(2014).

    [19] A. Loescher, J. P. Negel, T. Graf, M. A. Ahmed. Radially polarized emission with 635  W of average power and 2.1  mJ of pulse energy generated by an ultrafast thin-disk multipass amplifier. Opt. Lett., 40, 5758-5761(2015).

    [20] H. Sroor, I. Litvin, D. Naidoo, A. Forbes. Amplification of higher order Poincare sphere beams through Nd:YLF and Nd:YAG crystals. Appl. Phys. B, 125, 49(2019).

    [21] D. Lin, J. M. O. Daniel, M. Gecevicius, M. Beresna, P. G. Kazansky, W. A. Clarkson. Cladding-pumped ytterbium-doped fiber laser with radially polarized output. Opt. Lett., 39, 5359-5361(2014).

    [22] D. Lin, N. Baktash, S. U. Alam, D. J. Richardson. 106  W, picosecond Yb-doped fiber MOPA system with a radially polarized output beam. Opt. Lett., 43, 4957-4960(2018).

    [23] D. Stellinga, M. E. Pietrzyk, J. M. E. Glackin, Y. Wang, A. K. Bansal, G. A. Turnbull, K. Dholakia, I. D. W. Samuel, T. F. Krauss. An organic vortex laser. ACS Nano, 12, 2389-2394(2018).

    [24] M. A. Ahmed, J. Schulz, A. Voss, O. Parriaux, J. C. Pommier, T. Graf. Radially polarized 3  kW beam from a CO2 laser with an intracavity resonant grating mirror. Opt. Lett., 32, 1824-1826(2007).

    [25] P. S. Salter, M. J. Booth. Adaptive optics in laser processing. Light Sci. Appl., 8, 110(2019).

    [26] S. Ngcobo, I. Litvin, L. Burger, A. Forbes. A digital laser for on-demand laser modes. Nat. Commun., 4, 2289(2013).

    [27] D. Naidoo, F. S. Roux, A. Dudley, I. Litvin, B. Piccirillo, L. Marrucci, A. Forbes. Controlled generation of higher-order Poincare sphere beams from a laser. Nat. Photonics, 10, 327-333(2016).

    [28] H. Sroor, Y. W. Huang, B. Sephton, D. Naidoo, A. Valles, V. Ginis, C. W. Qiu, A. Ambrosio, F. Capasso, A. Forbes. High-purity orbital angular momentum states from a visible metasurface laser. Nat. Photonics, 14, 498-503(2020).

    [29] N. C. Zambon, P. St-Jean, M. Milicevic, A. Lemaitre, A. Harouri, L. Le Gratiet, O. Bleu, D. D. Solnyshkov, G. Malpuech, I. Sagnes, S. Ravets, A. Amo, J. Bloch. Optically controlling the emission chirality of microlasers. Nat. Photonics, 13, 283-289(2019).

    [30] M. J. Strain, X. L. Cai, J. W. Wang, J. B. Zhu, D. B. Phillips, L. F. Chen, M. Lopez-Garcia, J. L. O’Brien, M. G. Thompson, M. Sorel, S. Y. Yu. Fast electrical switching of orbital angular momentum modes using ultra-compact integrated vortex emitters. Nat. Commun., 5, 4856(2014).

    [31] X. L. Cai, J. W. Wang, M. J. Strain, B. Johnson-Morris, J. B. Zhu, M. Sorel, J. L. O’Brien, M. G. Thompson, S. T. Yu. Integrated compact optical vortex beam emitters. Science, 338, 363-366(2012).

    [32] P. Miao, Z. F. Zhang, J. B. Sun, W. Walasik, S. Longhi, N. M. Litchinitser, L. Feng. Orbital angular momentum microlaser. Science, 353, 464-467(2016).

    [33] D. Lin, J. Carpenter, Y. T. Feng, S. Jain, Y. M. Jung, Y. J. Feng, M. N. Zervas, D. J. Richardson. Reconfigurable structured light generation in a multicore fibre amplifier. Nat. Commun., 11, 3986(2020).

    [34] B. Ndagano, R. Brüning, M. McLaren, M. Duparré, A. Forbes. Fiber propagation of vector modes. Opt. Express, 23, 17330-17336(2015).

    [35] S. Ramachandran, P. Kristensen, M. F. Yan. Generation and propagation of radially polarized beams in optical fibers. Opt. Lett., 34, 2525-2527(2009).

    [36] R. S. Chen, J. H. Wang, X. Q. Zhang, A. T. Wang, H. Ming, F. Li, D. Chung, Q. W. Zhan. High efficiency all-fiber cylindrical vector beam laser using a long-period fiber grating. Opt. Lett., 43, 755-758(2018).

    [37] T. Wang, F. Shi, Y. P. Huang, J. X. Wen, Z. Q. Luo, F. F. Pang, T. Y. Wang, X. L. Zeng. High-order mode direct oscillation of few-mode fiber laser for high-quality cylindrical vector beams. Opt. Express, 26, 11850-11858(2018).

    [38] Y. M. Jung, Q. Y. Kang, R. Sidharthan, D. Ho, S. Yoo, P. Gregg, S. Ramachandran, S. U. Alam, D. J. Richardson. Optical orbital angular momentum amplifier based on an air-hole erbium-doped fiber. J. Lightwave Technol., 35, 430-436(2017).

    [39] J. W. Ma, F. Xia, S. Chen, S. H. Li, J. Wang. Amplification of 18 OAM modes in a ring-core erbium-doped fiber with low differential modal gain. Opt. Express, 27, 38087-38097(2019).

    [40] I. N. Papadopoulos, S. Farahi, C. Moser, D. Psaltis. High-resolution, lensless endoscope based on digital scanning through a multimode optical fiber. Biomed. Opt. Express, 4, 260-270(2013).

    [41] S. M. Popoff, G. Lerosey, R. Carminati, M. Fink, A. C. Boccara, S. Gigan. Measuring the transmission matrix in optics: an approach to the study and control of light propagation in disordered media. Phys. Rev. Lett., 104, 100601(2010).

    [42] R. Di Leonardo, S. Bianchi. Hologram transmission through multi-mode optical fibers. Opt. Express, 19, 247-254(2011).

    [43] T. Cizmar, K. Dholakia. Shaping the light transmission through a multimode optical fibre: complex transformation analysis and applications in biophotonics. Opt. Express, 19, 18871-18884(2011).

    [44] M. Ploschner, T. Tyc, T. Cizmar. Seeing through chaos in multimode fibres. Nat. Photonics, 9, 529-535(2015).

    [45] J. Carpenter, B. J. Eggleton, J. Schroder. Observation of Eisenbud-Wigner-Smith states as principal modes in multimode fibre. Nat. Photonics, 9, 751-758(2015).

    [46] G. Lopez-Galmiche, Z. S. Eznaveh, J. E. Antonio-Lopez, A. M. V. Benitez, J. R. Asomoza, J. J. S. Mondragon, C. Gonnet, P. Sillard, G. Li, A. Schulzgen, C. M. Okonkwo, R. A. Correa. Few-mode erbium-doped fiber amplifier with photonic lantern for pump spatial mode control. Opt. Lett., 41, 2588-2591(2016).

    [47] J. Montoya, C. Aleshire, C. Hwang, N. K. Fontaine, A. Velazquez-Benitez, D. H. Martz, T. Y. Fan, D. Ripin. Photonic lantern adaptive spatial mode control in LMA fiber amplifiers. Opt. Express, 24, 3405-3413(2016).

    [48] R. Florentin, V. Kermene, J. Benoist, A. Desfarges-Berthelemot, D. Pagnoux, A. Barthelemy, J. P. Huignard. Shaping the light amplified in a multimode fiber. Light Sci. Appl., 6, e16208(2017).

    [49] D. Flamm, D. Naidoo, C. Schulze, A. Forbes, M. Duparre. Mode analysis with a spatial light modulator as a correlation filter. Opt. Lett., 37, 2478-2480(2012).

    [50] https://www.alglib.net/. https://www.alglib.net/

    [51] M. Pasienski, B. DeMarco. A high-accuracy algorithm for designing arbitrary holographic atom traps. Opt. Express, 16, 2176-2190(2008).

    [52] S. H. Tao, W. X. Yu. Beam shaping of complex amplitude with separate constraints on the output beam. Opt. Express, 23, 1052-1062(2015).

    [53] D. Flamm, C. Schulze, D. Naidoo, S. Schroter, A. Forbes, M. Duparre. All-digital holographic tool for mode excitation and analysis in optical fibers. J. Lightwave Technol., 31, 1023-1032(2013).

    [54] B. Ndagano, H. Sroor, M. McLaren, C. Rosales-Guzman, A. Forbes. Beam quality measure for vector beams. Opt. Lett., 41, 3407-3410(2016).

    Di Lin, Joel Carpenter, Yutong Feng, Yongmin Jung, Shaif-ul Alam, David J. Richardson. High-power, electronically controlled source of user-defined vortex and vector light beams based on a few-mode fiber amplifier[J]. Photonics Research, 2021, 9(5): 856
    Download Citation