Information transmission through parallel multi-task-based recognition of high-resolution multiplexed orbital angular momentum

Jingwen Zhou, Yaling Yin, Jihong Tang, Yong Xia, Jianping Yin

Front. Phys. ›› 2024, Vol. 19 ›› Issue (5) : 52202.

PDF(7779 KB)
Front. Phys. All Journals
PDF(7779 KB)
Front. Phys. ›› 2024, Vol. 19 ›› Issue (5) : 52202. DOI: 10.1007/s11467-024-1402-y
RESEARCH ARTICLE

Information transmission through parallel multi-task-based recognition of high-resolution multiplexed orbital angular momentum

Author information +
History +

Abstract

Orbital angular momentums (OAMs) greatly enhance the channel capacity in free-space optical communication. However, demodulation of superposed OAM to recognize them separately is always difficult, especially upon multiplexing more OAMs. In this work, we report a directly recognition of multiplexed fractional OAM modes, without separating them, at a resolution of 0.1 with high accuracy, using a multi-task deep learning (MTDL) model, which has not been reported before. Namely, two-mode, four-mode, and eight-mode superposed OAM beams, experimentally generated with a hologram carrying both phase and amplitude information, are well recognized by the suitable MTDL model. Two applications in information transmission are presented: the first is for 256-ary OAM shift keying via multiplexed fractional OAMs; the second is for OAM division multiplexed information transmission in an eightfold speed. The encouraging results will expand the capacity in future free-space optical communication.

Graphical abstract

Keywords

information transmission / orbital angular momentum / multi-task deep learning / holographic multiplexing / structured light

Cite this article

Download citation ▾
Jingwen Zhou, Yaling Yin, Jihong Tang, Yong Xia, Jianping Yin. Information transmission through parallel multi-task-based recognition of high-resolution multiplexed orbital angular momentum. Front. Phys., 2024, 19(5): 52202 https://doi.org/10.1007/s11467-024-1402-y

1 Introduction

Two-dimensional (2D) semiconductor and related van der Waals heterostructures (vdWHs) have attracted tremendous attention because they possess excellent characteristics [1-5]. Also, the 2D vdWHs can form different band alignments, such as type-I, II and III, which is useful to diverse device applications [6-9]. Among them, type-II band alignment plays a critical role in facilitating the electron and hole separation, resulting in long photogenerated carrier lifetimes and tunable interlayer coupling [10-13]. However, the type-II band alignments of most 2D vdWHs are sensitive to external conditions [14-16], which brings a challenge for nanodevices to achieve stable properties. Therefore, it is valuable to find 2D vdWHs with robust type-II band alignment.
More recently, MoSi2N4 monolayer, an ultrathin transition metal nitride, is in the spotlight for its successful synthesis via passivating MoN2 monolayer with silicon [17]. The MoSi2N4 monolayer is a seven-layer nanosheet consisting of one MoN2 inner layer and two Si-N outer layers [18]. Furthermore, MoSi2N4 is reported to show great potential application due to its superior performance [19-31], including the impressive ambient stability [27], large theoretical electron/hole mobility (~270/1200 cm2·V–1·s–1) [17], excellent thermoelectric property [28,29], and remarkable optical absorption in the visible light range [30,31]. Stimulated by the excellent performance, the 2D vdWHs based on MoSi2N4 have become an emerging field [32-41]. For instance, C3N4/MoSi2N4 vdWHs have achieved a tunable type-II band alignment through interlayer coupling and external electric field [38]. The MoSi2N4/NbS2 vdWHs have robust ultra-low p-type Schottky barrier heights (SBH), while the SBH of MoSi2N4/graphene can be modulated via the interlayer distance and electric fields [32]. Compared to isolated transition metal dichalogens (TMDs), the MoSi2N4/TMDs vdWHs possess higher mobility, improved optical absorption, and tunable band structures [37, 40, 41]. Additionally, Jin et al. [42] have found that the vdWHs based on MoSi2N4 and wide-band gap semiconductors, such as ZnO and GaN, have retained the type-II band alignment when the electric field is smaller, but there is a transition from type-II to type-I band alignment with the application of out-of-plane strain. Among the wide-band gap semiconductors, monolayer GeO2 features an excitonic gap in the deep ultraviolet (UV) region with an energy of 6.24 eV [43]. It is found that the GeO2 has a relatively high work function, which is higher than that of reported results for the thin films of transition metal oxides [44] and traditional monolayers of TMDs [45], suggesting that GeO2 is advisable to be integrated into optical and electronic devices as charge extraction and injection layers that provide improvements in the device ability. Also, the 2D GeO2 material is fabricated by strictly controlling the degree of oxidation at the metal-gas interface [46].
Considering better lattice mismatch and interesting performance between the GeO2 and MoSi2N4 monolayers, and there are few works on GeO2-based vdWHs, so we select GeO2 as representative one of wide-band gap semiconductors to construct the heterojunction with MoSi2N4 and study the corresponding performance. In this work, the most stable structures are obtained by the global structure search method of passing through 12 × 12 displacement grids. Additionally, the robust type-II band arrangement of GeO2/MoSi2N4 vdWHs is insensitive to the regulation of biaxial strain, interlayer distance and external electric field, which opens up more possibilities for developing the stable optoelectronic devices.

2 Computational details

All the calculations are gotten via the Vienna ab initio simulation package (VASP) based on DFT [47]. With the framework of generalized gradient approximation (GGA), the Perdew−Burke−Ernzerhof (PBE) [48] functional is employed to deal with the exchange-correlation energy, and the Heyd−Scuseria−Ernzerhof (HSE06) method is utilized to gain more precise band gap value [49]. In addition, the projector augmented wave (PAW) pseudopotentials are adopted to simulate the electron-ion potential [50]. The vdW interlayer interaction in the GeO2/MoSi2N4 heterostructure is corrected by DFT-D3 correction method. Besides, the out-of-plane vacuum region, about 20 Å, is used, and the plane-wave cutoff is employed as 500 eV. A 11×11×1 k-point mesh is adopted for the sampling in reciprocal space. The structures are relaxed until the Hellmann−Feynman forces and the total energy converge to 0.01 eV/Å and 10−5 eV, respectively. Moreover, we would like to point out that the dipole correction is utilized in our calculations.

3 Results and discussion

3.1 Crystal structure and stability of isolated layer and heterojunctions

The geometric structures, band structures and phonon dispersion curves of GeO2 and MoSi2N4 monolayers are listed in Fig.1. Both GeO2 and MoSi2N4 monolayers have hexagonal structures, as shown in Fig.1(a) and (d), and Tab.1 demonstrates that the optimized lattice parameters are a = b = 2.909 Å and a = b = 2.911 Å, respectively. All these results are consistent with previous reports [22,43]. The bond lengths of monolayers are 1.95 Å for Ge-O, 2.09 Å for Mo-N, and 1.75 Å for Si-N. Fig.1(b) and (e) indicate that the GeO2 and MoSi2N4 monolayers are semiconductors with indirect band structure, and the band gap of monolayer GeO2 and MoSi2N4 is 3.56 (5.32) and 1.79 (2.35) eV for PBE (HSE06) calculation, respectively. In addition, Fig.1(c) and (f) depict the phonon dispersion curves of GeO2 and MoSi2N4 monolayers. Obviously, there are no imaginary phonon modes, which confirms the dynamical stability of the two kinds of 2D materials.
Fig.1 (a, d) The atomic structures, (b, e) band structures, (c, f) phonon dispersion curves of GeO2 and MoSi2N4 monolayers, respectively. The Fermi level is defined as zero. In the diagram of the atomic structures, different colored balls represent different atoms.

Full size|PPT slide

Tab.1 The structural and electronic parameters. The lattice constant (a), interlayer distance (d), total energy (Etotal), and band gap of PBE (EgPBE) and HSE06 (EgHSE06).
a (Å) d (Å) Etotal (eV) EgPBE (eV) EgHSE06 (eV)
MoSi2N4 2.911 −61.603 1.79 2.35
GeO2 2.909 −18.88 3.56 5.32
GeO2/MoSi2N4 2.909 2.78 −82.816 0.275 0.946
In order to obtain the credible structures of 2D GeO2/MoSi2N4 vdWHs, a global structure search method is implemented. The total energy as a function of the in-plane shift is presented in Fig.2(a). We consider 144 stacking structures Am,n passing through a 12 × 12 displacement grids. The Am,n means the displacement of the GeO2 layer relative to the MoSi2N4 layer is (m12)a+(n12)b, where a and b are the lattice vectors of the vdWHs. Among them, the A3,6 stacking pattern is the most stable structure with the interlayer distance between GeO2 and MoSi2N4 layer of 2.78 Å. Also, Fig.2(b) shows the side view of the most stable vdWHs structure consisting of GeO2 and MoSi2N4 layer, in which the lattice mismatch is 0.07%. From the bright red region in Fig.2(a), we find that there are four metastable structures around the A3,6 stacking pattern, which feature that the total energy is close to that of the most stable structure. The total energy of these four metastable structures is shown in Table S1. In addition, Fig. S1 presents the projected band structures of the metastable and most stable structures calculated by the PBE and HSE06 methods. Similar total energy and electronic properties are given for the considered five configurations, and thus only the most stable structure is discussed in the following section. The HSE06 method is adopted for all the next calculations. Fig.2(c) indicates that GeO2 and MoSi2N4 can form a type-II heterojunction with direct band gap of 0.946 eV. To further determine the dynamical stability of the structure with the lowest energy, Fig.2(d) shows the phonon dispersion curves of the vdWHs, and all the phonon frequencies are positive, illustrating that the most stable GeO2/MoSi2N4 vdWHs is also dynamically stable.
Fig.2 (a) Color contour plot of total energy versus in-plane shift for GeO2/MoSi2N4 vdWHs, considering the 12 × 12 grids. (b) The geometric structure, (c) band structure and (d) phonon dispersion curves of the most stable GeO2/MoSi2N4 vdWHs. The Fermi level is defined as zero.

Full size|PPT slide

To further explore the electronic characteristics of the most stable structure GeO2/MoSi2N4 vdWHs, Fig.3 draws the orbital-projected electronic band structures, partial charge densities, plane-averaged differential charge density and electrostatic potential. As illustrated in Fig.3(a) and (b), the CBM is mostly contributed by the Ge_s and O_p orbitals of the GeO2 layer, while the VBM mainly comes from the N_p and Mo_d orbitals in the MoSi2N4 layer. Besides, the band shapes of GeO2 and MoSi2N4 are well preserved after forming the heterojunction due to the vdWHs interaction. The charge distribution of Fig.3(c) shows the CBM is mainly contributed by Ge and O atoms, and the VBM comes from Mo and N atoms, which coincides with the above discussion.
Fig.3 The orbital-projected band structures of GeO2/MoSi2N4 vdWHs on GeO2 (a) and MoSi2N4 (b) layer. Partial charge density (c) and plane-averaged charge density difference (d) of GeO2/MoSi2N4 vdWHs. Yellow (green) colored isosurfaces represent the charge accumulation (depletion), respectively. (e) Electrostatic potential of GeO2/MoSi2N4 vdWHs. The isosurface is set to be 0.008 e/bohr3 for (c) and 0.0001 e/ Å3 for (d).

Full size|PPT slide

The plane-averaged charge density difference is calculated to assess the charge redistribution in the GeO2/MoSi2N4 vdWHs, which can be calculated as Δρ= ρvdWHsρGeO2ρMoSi2N4, where the ρvdWHs, ρGeO2, and ρMoSi2N4 represent the charge density of the vdWHs, isolated GeO2 and MoSi2N4 monolayers, respectively. In the inset of Fig.3(d), the charge transfer mainly occurs at the interface of the vdWHs. The yellow and green region corresponds to the ρ > 0 and ρ < 0, indicating the accumulation and depletion of charge, respectively. Meanwhile, the charge is depleted near the MoSi2N4 layer and accumulated in the vicinity of the GeO2 layer, which implies that the charge can be transferred from the MoSi2N4 layer to the GeO2 layer.
In addition, the electrostatic potential along the perpendicular direction of the vdWHs is explored in Fig.3(e). The potential drop (ΔV) across the two isolated layers is 10.51 eV, which indicates the presence of a strong built-in field across the interface. The electric field strength of GeO2/MoSi2N4 vdWHs exceeds that of other MoSi2N4-based heterostructures, such as the C3N4/MoSi2N4 (6.28 eV) [38] and MoSe2/MoSi2N4 (4.03 eV) [40]. The MoSi2N4 possesses a much deeper potential than GeO2, owing to the stronger electronegativity of the MoSi2N4 monolayer. Thus, the deeper potential may facilitate the separation of electron-hole pairs, which is beneficial for photoelectric detection applications based on the GeO2/MoSi2N4 vdWHs.

3.2 Biaxial strain tunable electronic properties of the vdWHs

To understand the strain influence on the GeO2/MoSi2N4 vdWHs, we define the biaxial strain as ε = (aa0)/a0 on GeO2/MoSi2N4 vdWHs, where a0 and a represent the lattice parameter of heterostructures without and with biaxial strain, respectively. In Fig.4, the negative and positive values are used to denote the compressive and tensile strain, respectively. Fig.4(a−c) display the variation of the total energy, band gap, and interlayer charge transfer of GeO2/MoSi2N4 vdWHs under the biaxial strains. It is obvious from Fig.4(a) that the total energy reaches the minimum value under the pristine condition. The band gap decreases for GeO2/MoSi2N4 vdWHs with increasing tensile, which can be seen in Fig.4(b). On the contrary, as the compressive strain is applied, the band gap value tends to increase. When the tensile strain increases up to 6%, the band gap of the vdWHs decreases from 0.946 eV to 0.124 eV. And the variation trend of the band gap gradually slows down and approaches saturation at 4% strain. Additionally, when the compressive strain range is 0 ~ −6%, the band gap increases linearly from 0.946 eV to 2.806 eV. Moreover, it can realize a transition from direct to indirect band gap when the compressive strain is greater than -3%. As shown in Tab.2, we can see that with increasing the tensile strain for the GeO2/MoSi2N4 vdWHs, the atomic bond lengths of Ge-O, Mo-N and Si-N can be raised, while the bond angles exhibit the reduced trend. In contrast to the case of tensile strain, the increasing compressive strain can make the Ge-O, Mo-N and Si-N bond lengths decrease, and the bond angles consequently enhance. However, it is worth noting that the type-II band arrangement is well preserved in the GeO2/MoSi2N4 vdWHs under all the studied biaxial strain cases.
Tab.2 The bond lengths as a function of biaxial strain for GeO2/MoSi2N4 vdWHs.
Strain dGe-O (Å) dMo-N (Å) dSi-N (Å)
−6% 1.90 2.05 1.68
−5% 1.91 2.06 1.70
−4% 1.91 2.06 1.71
−3% 1.92 2.07 1.72
−2% 1.93 2.08 1.73
−1% 1.94 2.08 1.74
0 1.95 2.09 1.75
1% 1.96 2.10 1.76
2% 1.96 2.11 1.78
3% 1.97 2.12 1.79
4% 1.99 2.13 1.81
5% 2.00 2.14 1.82
6% 2.01 2.15 1.83
Fig.4 The variation of the total energy (a), band gap (b), and interlayer charge transfer (c) of GeO2/MoSi2N4 vdWHs as a function of biaxial strain, respectively. (d−g) Projected band structures of GeO2/MoSi2N4 vdWHs with the biaxial strains of −3%, −2%, 2% and 3%.

Full size|PPT slide

Fig.4(c) exhibits the evolution of interlayer charge transfer as a function of biaxial strain. The results show that the interlayer charge transfer of GeO2/MoSi2N4 vdWHs decreases with increasing tensile and compressive strains. Under the case of tensile and compressive strain, the interlayer charge transfer declines by 0.018 e and 0.005 e, respectively. The charge transfer tends to diminish when the compressive strain is larger than −5%. Furthermore, it reaches a maximum value of 0.031 e under the 1% tensile strain.
To better understand the above-mentioned behavior, Fig.4(d)−(g) present the projected band structures of GeO2/MoSi2N4 vdWHs under different biaxial strains. When tensile strain is applied, the conduction band edge moves to the Fermi level significantly, while the valence band edge barely moves, which is consistent with the decrease in the band gap. On the other side, the compressive strain causes the conduction band edge to move away from the Fermi level substantially, corresponding to the increase of the band gap. As the −2% compressive strain is applied, both CBM and VBM remain at Γ point. However, when the compressive strain is greater than −3%, the CBM of GeO2/MoSi2N4 vdWHs is at Γ point, while the VBM is at K point, thus an indirect band gap can be observed. In particular, for all the studied strain cases, the CBM and VBM of GeO2/MoSi2N4 vdWHs are always formed by the GeO2 and MoSi2N4 layers, respectively, confirming that it always maintains type-II band alignment characteristics. In addition, we notice that the application of strain can induce the band alignment transition from type-II to type-I or type-III in many other 2D vdWHs, such as GaSe/SnX2 (X = S, Se) [14], MoSSe/MBP [51] and BP/β-AsP vdWHs [52].

3.3 Interlayer distance tunable electronic properties of the vdWHs

The interlayer interaction has obvious effects on the electronic structures of the vdWHs. Thus, Fig.5 plots the interlayer distance change on the electronic properties of the GeO2/MoSi2N4 vdWHs. We can see from Fig.5(a) that the curve of the total energy has the minimum energy value when the interlayer distance is 2.78 Å. From Fig.5(b), the band gap increases for GeO2/MoSi2N4 vdWHs with increasing interlayer distance. Also, the band gap changes slightly when the interlayer distance is larger than 3.28 Å. In addition, Fig.5(c) shows that the interlayer charge transfer declines gradually with the increase of the interlayer distance. To reveal the related physical mechanism, in Fig.5(d−g), we give the interlayer distance effects on the projected band structure of GeO2/MoSi2N4 vdWHs. For different distances, the CBM and VBM contributions to GeO2/MoSi2N4 vdWHs are provided by the GeO2 and MoSi2N4 layers separately, showing that this type-II band alignment is insensitive to the interlayer interactions. Furthermore, the variation of the band gap is mainly due to the up and down movement of CBM.
Fig.5 The total energy (a) and band gap (b), interlayer charge transfer (c) and projected band structures (d−g) of GeO2/MoSi2N4 vdWHs at different interlayer distances, respectively.

Full size|PPT slide

3.4 Electric field tunable electronic properties of the vdWHs

The external vertical electric field provides another effective strategy for controlling the electronic structure. In the following, the variation of the band gap for GeO2/MoSi2N4 vdWHs as a function of the electric field is shown in Fig.6 (a). It should be noted that the electric field is perpendicular to the surface of the GeO2/MoSi2N4 vdWHs, and the positive direction is defined from MoSi2N4 layer to GeO2 layer, as illustrated in Fig. S2. When the electric field increases from −0.5 to 0.5 V/Å, the band gap decreases linearly, maintaining the type-II band alignment and direct band structures. Here, the band edge and projected band structures are also shown, and the reasons for the band alignment transition are discussed. In Fig.6(b), with the increase of the external electric field, the CBM of GeO2/MoSi2N4 vdWHs decreases significantly, while the VBM shifts upward slightly. Moreover, when the positive electric field reaches 0.4 V/Å, the VBM decreases more slightly, which causes the value of band gap to decrease slowly as the electric field changes from 0.4 V/Å to 0.5 V/Å. As can be seen from Fig.6(c−f), with the increase of the positive electric field, the CBM of GeO2 layer moves down gradually, while the CBM and VBM of MoSi2N4 keep almost unchanged. Compared with the 0.2 V/Å, the CBM of GeO2 layer moves towards the Fermi level under the electric field of 0.4 V/Å. On the other hand, Fig.6(f) indicates that with increasing the negative electric field, the CBM of GeO2 layer moves away from the Fermi level when the electric field reaches −0.4 V/Å. However, further increased electric field does not change the type of band arrangement. Thus, the external electric field can change the band gap, while the band alignment remains type-II for the GeO2/MoSi2N4 vdWHs. What’s more, for many other vdWHs, the band alignment is sensitive to the external electric field [16, 51, 53-56]. For example, when the positive electric field is larger than 0.2 V/Å, the band alignment of the AlN/VSe2 vdWHs changes from type-II to type-I [16]. The MoXY/WXY (X, Y = S, Se, Te; X ≠ Y) vdWHs maintain robust type-II band alignment under positive electric field, while a reverse electric field can make the band alignment change [53].
Fig.6 The band gap (a) and band alignment (b) of GeO2/MoSi2N4 vdWHs as a function of external electric field, respectively. (c−f) The projected band structures of GeO2/MoSi2N4 vdWHs with different external electric fields. The red (green) line indicates the contribution of the GeO2 (MoSi2N4) layer.

Full size|PPT slide

4 Conclusions

In conclusion, by using a global structure search approach, we design the GeO2/MoSi2N4 vdWHs to exactly explore the band structures, and the corresponding modulations of biaxial strain, electric field, and interlayer distance are considered to improve the performance of GeO2/MoSi2N4 vdWHs. The results show that the monolayer GeO2, MoSi2N4 and GeO2/MoSi2N4 vdWHs are dynamically stable at room temperature. The most stable GeO2/MoSi2N4 vdWHs possess direct band gap of 0.946 eV and type-II band alignment with the CBM and VBM from the GeO2 and MoSi2N4 layers. When the biaxial strain is applied, the direct-to-indirect band gap transition can be achieved. More interestingly, the robust type-II band arrangement of GeO2/MoSi2N4 vdWHs remains unchanged when the external electric field, interlayer distance, and biaxial strain are applied. These results reveal that the GeO2/MoSi2N4 vdWHs can be in promising optoelectronic devices with stable characteristics due to their robust type-II band arrangement.

References

[1]
A. E. Willner , H. Huang , Y. Yan , Y. Ren , N. Ahmed , G. Xie , C. Bao , L. Li , Y. Cao , Z. Zhao , J. Wang , M. P. J. Lavery , M. Tur , S. Ramachandran , A. F. Molisch , N. Ashrafi , S. Ashrafi . Optical communications using orbital angular momentum beams. Adv. Opt. Photonics, 2015, 7(1): 66
CrossRef ADS Google scholar
[2]
H. Rubinsztein-Dunlop , A. Forbes , M. V. Berry , M. R. Dennis , D. L. Andrews , M. Mansuripur , C. Denz , C. Alpmann , P. Banzer , T. Bauer , E. Karimi , L. Marrucci , M. Padgett , M. Ritsch-Marte , N. M. Litchinitser , N. P. Bigelow , C. Rosales-Guzmán , A. Belmonte , J. P. Torres , T. W. Neely , M. Baker , R. Gordon , A. B. Stilgoe , J. Romero , A. G. White , R. Fickler , A. E. Willner , G. Xie , B. McMorran , A. M. Weiner . Roadmap on structured light. J. Opt., 2017, 19(1): 013001
CrossRef ADS Google scholar
[3]
L. Allen , M. Beijersbergen , R. Spreeuw , J. Woerdman . Orbital angular momentum of light and transformation of Laguerre‒Gaussian laser modes. Phys. Rev. A, 1992, 45(11): 8185
CrossRef ADS Google scholar
[4]
J. P. Yin , W. J. Gao , Y. F. Zhu . Generation of dark hollow beams and their applications. Prog. Opt., 2003, 45(11): 119
CrossRef ADS Google scholar
[5]
M.J. Padgett, Orbital angular momentum 25 years on, Opt. Express 25(10), 11265 (2017) (Invited)
[6]
Y. J. Shen , X. J. Wang , Z. W. Xie , C. J. Min , X. Fu , Q. Liu , M. L. Gong , X. C. Yuan . Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities. Light Sci. Appl., 2019, 8(1): 90
CrossRef ADS Google scholar
[7]
A. Forbes , S. Ramachandran , Q. W. Zhan . Photonic angular momentum: Progress and perspectives. Nanophotonics, 2022, 11(4): 625
CrossRef ADS Google scholar
[8]
S. J. Li , Z. Y. Li , G. S. Huang , X. B. Liu , R. Q. Li , X. Y. Cao . Digital coding transmissive metasurface for multi-OAM-beam. Front. Phys., 2022, 17(6): 62501
CrossRef ADS Google scholar
[9]
L. Jin , Y. W. Huang , Z. W. Jin , R. C. Devlin , Z. G. Dong , S. T. Mei , M. H. Jiang , W. T. Chen , Z. Wei , H. Liu , J. H. Teng , A. Danner , X. P. Li , S. M. Xiao , S. Zhang , C. Y. Yu , J. K. W. Yang , F. Capasso , C. W. Qiu . Dielectric multi-momentum meta-transformer in the visible. Nat. Commun., 2019, 10(1): 4789
CrossRef ADS Google scholar
[10]
X. Y. Fang , H. J. Wang , H. C. Yang , Z. L. Ye , Y. M. Wang , Y. Zhang , X. P. Hu , S. N. Zhu , M. Xiao . Multichannel nonlinear holography in a two-dimensional nonlinear photonic crystal. Phys. Rev. A, 2020, 102(4): 043506
CrossRef ADS Google scholar
[11]
J. J. Guo , Y. P. Zhang , H. Ye , L. Y. Wang , P. C. Chen , D. J. Mao , C. Z. Xie , Z. H. Chen , X. W. Wu , M. Xiao , Y. Zhang . Spatially structured-mode multiplexing holography for high-capacity security encryption. ACS Photonics, 2023, 10(3): 757
CrossRef ADS Google scholar
[12]
A. E. Willner , Z. Zhao , C. Liu , R. Zhang , H. Song , K. Pang , K. Manukyan , H. Song , X. Su , G. Xie , Y. Ren , Y. Yan , M. Tur , A. F. Molisch , R. W. Boyd , H. Zhou , N. Hu , A. Minoofar , H. Huang . Perspectives on advances in high-capacity, free-space communications using multiplexing of orbital-angular-momentum beams. APL Photonics, 2021, 6(3): 030901
CrossRef ADS Google scholar
[13]
J. Lin , X. C. Yuan , S. H. Tao , R. E. Burge . Multiplexing free-space optical signals using superimposed collinear orbital angular momentum states. Appl. Opt., 2007, 46(21): 4680
CrossRef ADS Google scholar
[14]
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, 2012, 6(7): 488
CrossRef ADS Google scholar
[15]
N. Bozinovic , Y. Yue , Y. Ren , M. Tur , P. Kristensen , H. Huang , A. E. Willner , S. Ramachandran . Terabit-scale orbital angular momentum mode division multiplexing in fibers. Science, 2013, 340(6140): 1545
CrossRef ADS Google scholar
[16]
G. Vallone , V. D’Ambrosio , A. Sponselli , S. Slussarenko , L. Marrucci , F. Sciarrino , P. Villoresi . Free-space quantum key distribution by rotation-invariant twisted photons. Phys. Rev. Lett., 2014, 113(6): 060503
CrossRef ADS Google scholar
[17]
H.HuangG.XieY.YanN.AhmedY.RenY.YueD.RogawskiM.J. WillnerB.I. ErkmenK.M. BirnbaumS.J. DolinarM.P. J. LaveryM.J. PadgettM.TurA.E. Willner, 100 Tbit/s free-space data link enabled by three-dimensional multiplexing of orbital angular momentum, polarization, and wave-length, Opt. Lett. 39(2), 197 (2014)
[18]
M. Krenn , R. Fickler , M. Fink , J. Handsteiner , M. Malik , T. Scheidl , R. Ursin , A. Zeilinger . Communication with spatially modulated light through turbulent air across Vienna. New J. Phys., 2014, 16(11): 113028
CrossRef ADS Google scholar
[19]
A. J. Willner , Y. Ren , G. Xie , Z. Zhao , Y. Cao , L. Li , N. Ahmed , Z. Wang , Y. Yan , P. Liao , C. Liu , M. Mirhosseini , R. W. Boyd , M. Tur , A. E. Willner . Experimental demonstration of 20 Gbit/s data encoding and 2 ns channel hopping using orbital angular momentum modes. Opt. Lett., 2015, 40(24): 5810
CrossRef ADS Google scholar
[20]
M. Krenn , J. Handsteiner , M. Fink , R. Fickler , R. Ursin , M. Malik , A. Zeilinger . Twisted light transmission over 143 km. Proc. Natl. Acad. Sci. USA, 2016, 113(48): 13648
CrossRef ADS Google scholar
[21]
H. Zhang , J. Zeng , X. Y. Lu , Z. Y. Wang , C. L. Zhao , Y. J. Cai . Review on fractional vortex beam. Nanophotonics, 2022, 11(2): 241
CrossRef ADS Google scholar
[22]
Z. W. Liu , S. Yan , H. G. Liu , X. F. Chen . Superhigh-resolution recognition of optical vortex modes assisted by a deep-learning method. Phys. Rev. Lett., 2019, 123(18): 183902
CrossRef ADS Google scholar
[23]
Y. Na , D. K. Ko . Adaptive demodulation by deep-learning-based identification of fractional orbital angular momentum modes with structural distortion due to atmospheric turbulence. Sci. Rep., 2021, 11(1): 23505
CrossRef ADS Google scholar
[24]
J. Leach , M. Padgett , S. Barnett , S. Franke-Arnold , J. Courtial . Measuring the orbital angular momentum of a single photon. Phys. Rev. Lett., 2002, 88(25): 257901
CrossRef ADS Google scholar
[25]
G. Berkhout , M. Lavery , J. Courtial , M. Beijersbergen , M. Padgett . Efficient sorting of orbital angular momentum states of light. Phys. Rev. Lett., 2010, 105(15): 153601
CrossRef ADS Google scholar
[26]
M. Mirhosseini , M. Malik , Z. Shi , R. Boyd . Efficient separation of the orbital angular momentum eigenstates of light. Nat. Commun., 2013, 4(1): 2781
CrossRef ADS Google scholar
[27]
Y. H. Wen , I. Chremmos , Y. J. Chen , J. B. Zhu , Y. F. Zhang , S. Y. Yu . Spiral transformation for high-resolution and efficient sorting of optical vortex modes. Phys. Rev. Lett., 2018, 120(19): 193904
CrossRef ADS Google scholar
[28]
G. Gibson , J. Courtial , M. J. Padgett , M. Vasnetsov , V. Pas’ko , S. M. Barnett , S. Franke-Arnold . Free-space information transfer using light beams carrying orbital angular momentum. Opt. Express, 2004, 12(22): 5448
CrossRef ADS Google scholar
[29]
T. Doster , A. T. Watnik . Machine learning approach to OAM beam demultiplexing via convolutional neural networks. Appl. Opt., 2017, 56(12): 3386
CrossRef ADS Google scholar
[30]
S. Lohani , E. M. Knutson , M. O’Donnell , S. D. Huver , R. T. Glasser . On the use of deep neural networks in optical communications. Appl. Opt., 2018, 57(15): 4180
CrossRef ADS Google scholar
[31]
S. Park , L. Cattell , J. Nichols , A. Watnik , T. Doster , G. Rohde . De-multiplexing vortex modes in optical communications using transport-based pattern recognition. Opt. Express, 2018, 26(4): 4004
CrossRef ADS Google scholar
[32]
J. Li , M. Zhang , D. S. Wang . Adaptive demodulator using machine learning for orbital angular momentum shift keying. IEEE Photonics Technol. Lett., 2017, 29(17): 1455
CrossRef ADS Google scholar
[33]
Q. S. Zhao , S. Q. Hao , Y. Wang , L. Wang , X. F. Wan , C. L. Xu . Mode detection of misaligned orbital angular momentum beams based on convolutional neural network. Appl. Opt., 2018, 57(35): 10152
CrossRef ADS Google scholar
[34]
S. Lohani , R. T. Glasser . Turbulence correction with artificial neural networks. Opt. Lett., 2018, 43(11): 2611
CrossRef ADS Google scholar
[35]
Q. H. Tian , Z. Li , K. Hu , L. Zhu , X. L. Pan , Q. Zhang , Y. J. Wang , F. Tian , X. L. Yin , X. J. Xin . Turbo-coded 16-ary OAM shift keying FSO communication system combining the CNN-based adaptive demodulator. Opt. Express, 2018, 26(21): 27849
CrossRef ADS Google scholar
[36]
J. Li , M. Zhang , D. S. Wang , S. J. Wu , Y. Y. Zhan . Joint atmospheric turbulence detection and adaptive demodulation technique using the CNN for the OAM‒FSO communication. Opt. Express, 2018, 26(8): 10494
CrossRef ADS Google scholar
[37]
Y. Z. Shi , Z. H. Ma , H. Y. Chen , Y. G. Ke , Y. Chen , X. X. Zhou . High-resolution recognition of FOAM modes via an improved EfficientNet V2 based convolutional neural network. Front. Phys., 2024, 19(3): 32205
CrossRef ADS Google scholar
[38]
H.LuanD.LinK.LiW.MengM.GuX.Fang, 768-ary Laguerre‒Gaussian-mode shift keying free-space optical communication based on convolutional neural networks, Opt. Express 29(13), 19807 (2021)
[39]
A. Maurer , M. Pontil , B. Paredes . The benefit of multitask representation learning. J. Mach. Learn. Res., 2016, 17(1): 2853
CrossRef ADS Google scholar
[40]
Z. X. Mao , H. Y. Yu , M. Xia , S. Z. Pan , D. Wu , Y. L. Yin , Y. Xia , J. P. Yin . Broad bandwidth and highly efficient recognition of optical vortex modes achieved by the neural-network approach. Phys. Rev. Appl., 2020, 13(3): 034063
CrossRef ADS Google scholar
[41]
J. Davis , D. Cottrell , J. Campos , M. Yzuel , I. Moreno . Encoding amplitude information onto phase-only filters. Appl. Opt., 1999, 38(23): 5004
CrossRef ADS Google scholar
[42]
E. Bolduc , N. Bent , E. Santamato , E. Karimi , R. W. Boyd . Exact solution to simultaneous intensity and phase encryption with a single phase-only hologram. Opt. Lett., 2013, 38(18): 3546
CrossRef ADS Google scholar
[43]
K.ThungC.Wee, A brief review on multi-task learning, Multimedia Tools Appl. 77(22), 29705 (2018)
[44]
Y. Zhang , Q. Yang . An overview of multi-task learning. Natl. Sci. Rev., 2018, 5(1): 30
CrossRef ADS Google scholar
[45]
C.SzegedyL.WeiJ.YangqingP.SermanetS.ReedD.AnguelovD.ErhanV.VanhouckeA.Rabinovich, Going deeper with convolutions, in: IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Boston, MA, USA, 1 (2015)
[46]
S.XieR.B. GirshickP.DollárZ.TuK.He, Aggregated residual transformations for deep neural networks, in: 2017 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Honolulu, HI, USA, 5987 (2017)
[47]
R.CollobertJ.Weston, A unified architecture for natural language processing: deep neural networks with multitask learning, in: Proceedings of the 25th International Conference on Machine Learning, Association for Computing Machinery: Helsinki, Finland, 160–167 (2008)
[48]
X. Cui , W. Zhang , U. Finkler , G. Saon , M. Picheny , D. Kung . Distributed training of deep neural network acoustic models for automatic speech recognition: A comparison of current training strategies. IEEE Signal Process. Mag., 2020, 37(3): 39
CrossRef ADS Google scholar
[49]
R.B. GirshickR.C. N. N. Fast, IEEE International Conference on Computer Vision (ICCV), 1440 (2015)
[50]
B. L. Li , H. T. Luan , K. Y. Li , Q. Y. Chen , W. J. Meng , K. Cheng , M. Gu , X. Y. Fang . Orbital angular momentum optical communications enhanced by artificial intelligence. J. Opt., 2022, 24(9): 094003
CrossRef ADS Google scholar
[51]
Z. S. Wan , H. Wang , Q. Liu , X. Fu , Y. J. Shen . Ultra-degree-of-freedom structured light for ultracapacity information carriers. ACS Photonics, 2023, 10(7): 2149
CrossRef ADS Google scholar

Declarations

The authors declare that they have no competing interests and there are no conflicts.

Acknowledgements

Financial supports are from the National Natural Science Foundation of China (Grant Nos. 12174115, 91836103, and 11834003).

RIGHTS & PERMISSIONS

2024 Higher Education Press
AI Summary AI Mindmap
PDF(7779 KB)

Supplementary files

fop-21216-OF-xiacongxin_suppl_1 (520 KB)

Accesses

Citations

Detail

Sections
Recommended

/