Abstract
Computer-generated holography is a promising technique that modulates user-defined wavefronts with digital holograms. Computing appropriate holograms with faithful reconstructions is not only a problem closely related to the fundamental basis of holography but also a long-standing challenge for researchers in general fields of optics. Finding the exact solution of a desired hologram to reconstruct an accurate target object constitutes an ill-posed inverse problem. The general practice of single-diffraction computation for synthesizing holograms can only provide an approximate answer, which is subject to limitations in numerical implementation. Various non-convex optimization algorithms are thus designed to seek an optimal solution by introducing different constraints, frameworks, and initializations. Herein, we overview the optimization algorithms applied to computer-generated holography, incorporating principles of hologram synthesis based on alternative projections and gradient descent methods. This is aimed to provide an underlying basis for optimized hologram generation, as well as insights into the cutting-edge developments of this rapidly evolving field for potential applications in virtual reality, augmented reality, head-up display, data encryption, laser fabrication, and metasurface design.
Similar content being viewed by others
Introduction
Holography is a long-existing concept first raised by Dennis Gabor in the late 1940s, which aimed at improving resolution in electron microscopy1. In the 1960s, the development of laser technology enabled practical optical holography2,3. Early demonstration of optical holography can be described with two steps: interferential recording of an object wavefront and diffractive reconstruction from a hologram. Recent advancements in digital devices have enabled both the recording and the reconstruction processes to be performed computationally. One branch of holography involves optically recording an object wavefront and computationally reconstructing it from a digital hologram4,5, commonly referred to as digital holography6. This approach enables promising applications such as imaging, measurement, and detection. Another branch of holography involves computationally generating a hologram and optically reconstructing an object’s wavefront, commonly referred to as computer-generated holography (CGH), which provides an approach to digitally modulate a volumetric wavefront7. This technology, half inherited from optical holography and half advanced by computer science, has become an emerging focus of academia and industry8,9,10.
Computer-generated holograms, encoded on various types of holographic media, enable a wide range of applications. Holograms fabricated as diffractive optical elements (DOEs)11 or metasurfaces12,13,14 can reproduce specific spatial light fields, achieving structured light projection15,16,17, data storage18,19, and optical encryption20,21,22,23,24. With refreshable devices like spatial light modulators (SLMs)25,26,27, as is shown in Fig. 1, CGH is able to assist many fields of investigations, including three-dimensional display, holographic lithography28, optical trapping29, and optogenetics30,31,32. In recent years, CGH also boosts the birth and growth of potential markets of virtual reality (VR)33,34,35,36,37,38, augmented reality (AR)39,40,41,42, head-up display43,44,45, holographic printing46, optical communication47, and optical computing48. Although these applications and fields of investigation involve the encoding of holograms with various elements49,50 and devices51,52, the algorithms for hologram synthesis are similar and can be universally applied53. Therefore, computing a hologram for faithful reconstructions is an issue closely linked to fundamental physics and is widely investigated in general fields of optics. The computation of holograms is a strict adherence to the physics basis of holography while being a compromise to the existing hardware54, which requires holograms to be phase-only, amplitude-only, or rarely complex-amplitude. The object wave used for CGH can be reconstructed from either an interference pattern with a reference beam55,56 or a diffraction pattern of the object57. In early investigations58, holograms computed from diffractive wave propagations were found to be more energy-efficient compared to those computed from interferences. These wave-propagation methods yielded complex holograms (CHs) using computer-guided plotters employing the detour phase principle59. Subsequently, phase-only holograms (POHs) were proposed60, generated on the assumption that a scattered wavefront can be reconstructed with only the phase term. Proper computation of a diffraction-based hologram is essential for the optical reconstruction of CHs and POHs, and it gradually becomes vital for generating amplitude-only holograms61.
References17,18,20,27,49,208 are reprinted with permission from Springer Nature. References39,50,68,141 are reprinted with permission from © Optical Society of America. References33,40,132,133 are reprinted with permission from © ACM. Reference215 is reprinted with permission from © AAAS. References25,26,41,51,52,61,66,69,213 are reprinted under a Creative Commons Attribution 4.0 International License
The fundamental concern of the diffraction-based hologram computation can be neatly described as solving a hologram from a given intensity distribution of the object62, as is shown in Fig. 2a, which is an inverse problem with constraints imposed by physical basis and hardware implementation. This inverse problem is relatively different from the phase retrieval problems in imaging because a hologram satisfying all the constraints and reconstructing an artificial intensity distribution is not ensured to exist63. Nevertheless, the relation between the reconstructed wavefront and its intensity is ill-conditioned64. Multiple candidate holograms that approximately satisfy the constraints can be solved through the non-convex optimization65. Various optimization algorithms are thus introduced in hologram synthesis and rapidly bring about breakthroughs for CGH in noise reduction66, contrast enhancement67, crosstalk suppression68, and computing acceleration69. With the significant enhancement of CGH capabilities through optimization algorithms, achieving appropriate hologram optimization tailored to the physics process emerges as a central target, which ensures desired optical reconstructions. The motivation of computing all kinds of holograms with faithful reconstructions has pushed forward the advancement of optimization algorithms for CGH.
a The synthesis of computer-generated holograms can be described as an inverse problem. b Constraints, frameworks, and the initialization condition need to be considered in hologram optimization. c Constraints are imposed by physics and hardware in hologram optimization. Some optimization frameworks are widely used in hologram optimization: d alternating projections, where elementary projections occur to approach the intersection points of two or more enclosed feasible sets; e first-order gradient descent methods, where a local optimal solution is searched along the gradient descent direction; f second-order gradient descent methods, where the “steepest” gradient descent direction is further found with the assistance of second-order gradient
This review article focuses on optimization algorithms for hologram generation in CGH, which have become essential tools in hologram synthesis over the past decades. Concluded from existing optimization algorithms appropriately adapted to CGH inverse problem solving, as is shown in Fig. 2b, some crucial factors need to be considered, including constraints, framework, and initialization. Constraints refer to the conditions that a candidate solution should satisfy, usually modeled with the physical basis and the hardware implementation of CGH. Restricted by the constraints, possible solutions of the inverse problem constitute an enclosed feasible set γ, within which the optimization algorithms search for a solution locally or globally. The optimization framework chosen determines the path of searching within the feasible set, leading to variations in computing time, storage, and sometimes even convergence points. Initialization, especially the choice of the initialization condition, has a significant impact on the final convergence point. With a comprehensive overview of the original and recent designs of optimization algorithms, suitable handling of core issues concerning feasible hologram optimization is illustrated, offering a basis for applying optimized holograms to various fields of investigation.
Constraints
Optimization algorithms are designed to retrieve a computer-generated hologram \(H({\bf{u}})=A({\bf{u}}){e}^{i\phi ({\bf{u}})}\) from a specific object wave \(h({\bf{r}})=a({\bf{r}}){e}^{i\varphi ({\bf{r}})}\) in an inverse problem solving situation of CGH, where a(r) is the amplitude, φ(r) is the phase, \({\bf{r}}=(x,y)\) and \({\bf{u}}=(u,v)\) are the position vectors on the object plane and the hologram plane, respectively. As shown in Fig. 2c, such an optimization is restricted by several constraints64,65:
-
①
Object intensity \({a}^{2}({\bf{r}})={I}_{{\rm{obj}}}({\bf{r}})\);
-
②
Finite bandwidth \({\Delta }_{H} \,<\, \infty\) and \({\Delta }_{h} \,<\, \infty\);
-
③
Finite spatial scale \({L}_{u} \,<\, \infty\) and \({L}_{v} \,<\, \infty\);
-
④
Hologram intensity for POHs \({A}^{2}({\bf{u}})={E}_{{\rm{uni}}}\).
Constraint ① describes the pre-defined object intensity, which is fed as a constraint in optimization to ensure a faithful reconstruction. Constraint ② describes a finite bandwidth throughout the process of propagation, where ΔH and Δh represent the hologram and object wave bandwidths, respectively. Constraint ③ describes the finite spatial scale of the hologram, which is usually conducted by zero padding in practical computation. Lu and Lv represent the length and the width of the hologram to be optimized respectively. Constraint ④ is optional and specially imposed due to a constant hologram intensity of POHs, where \({E}_{{\rm{uni}}}\) is the hologram intensity.
These constraints ensure a correct optimization adapting to the physical model of holography and enable a faithful reveal of the reconstruction performance in numerical simulations. Errors and omissions of the constraints imposed on optimization can lead to a mismatch between the numerical and optical reconstructions70, usually shown as obviously excessive speckle noise and artifacts appearing in the optical reconstructions compared to the numerical ones71. Given all these constraints imposed, we can deduce that the object phase φ(r) is the only unsettled input factor in optimization, which provides a degree of freedom to design the hologram72. With different initializations of φ(r), the path and the outcome of optimization can differ greatly, leading to various performances of optical reconstructions.
The effectiveness of optimization is largely decided by whether the constraints are correctly imposed or not. The way the constraints are imposed is determined not only by the type of holograms but also by the diffractive calculation method utilized to implement computational free-space propagation. Existing diffractive calculation methods demonstrated for hologram optimization are mainly conducted by parallel computation based on the Fourier transform73. Among them, using a single fast Fourier Transform (FFT) to compute Fraunhofer diffraction at infinity is one of the mostly used propagation strategies74, which is easy and simple enough to implement in research. Fresnel diffraction is often chosen to generate far-field holograms75, with diffractive calculation methods including the single Fourier-transform Fresnel (SFT-FR) approach76, the Fresnel transfer function approach, and the Fresnel impulse response approach77. Besides, some methods derived directly from the first Rayleigh-Sommerfeld integral theory present high practicability in CGH78, covering both near-field and far-field propagations, among which the representative examples are the Rayleigh-Sommerfeld convolution method79 and its corresponding Fourier form angular spectrum method (ASM)80.
Here, we uniformly utilize the single FFT method81,82 to illustrate 2D hologram optimization algorithms and the band-limited ASM83 for the 3D case. Given different diffractive calculation methods, the most controversial part of imposing constraints lies in defining a proper bandwidth restriction to describe constraint ②, which is discussed in the following sections.
Framework
The optimization algorithm searches for an optimal solution as the desired hologram within a closed set restricted by the constraints. Several effective optimization frameworks are widely used, including alternating projection methods, which are depicted in Fig. 2d, and gradient descent methods. More specifically, the gradient descent methods also incorporate the first-order gradient descent as depicted in Fig. 2e, and the second-order gradient descent as depicted in Fig. 2f, of which the representative examples are stochastic gradient descent (SGD) and the quasi-Newton method respectively. Some other optimization frameworks are also applied to CGH, like iterative shrinkage-thresholding algorithms84 and simulated annealing85. Yet, these frameworks are less widely used in hologram synthesis.
Alternating projections
The alternating projections describe a category of optimization methods to approximate a point in the intersection of two or more enclosed feasible sets separately restricted by different constraints86. This is achieved by a pair of elementary projections repeatedly occurring in the optimization, which construct an iterative computation loop87. The Gerchberg-Saxton algorithm (G-S algorithm)88 was published in 1972 as the first account of alternating projections solving a phase retrieval problem, in which a pair of projections were used to reconstruct phase from two intensity measurements. As is shown in Fig. 3a, the phase of a specific wavefront is retrieved with the amplitudes on both the object plane \(a^{\prime} ({\bf{r}})\) and the Fourier plane \(A^{\prime} ({\bf{u}})\) repeatly replaced by known distributions, which are discretized into a \({M}_{x}\times {N}_{y}\) matrix and a \({M}_{u}\times {N}_{v}\) matrix respectively. Later presented by Fienup, this alternating scheme was improved to solve the problem of finding a Fourier transform pair satisfying the constraints in both planes, which was known as the error reduction algorithm89. As is shown in Fig. 3b, the error reduction algorithm enhances the flexibility of the alternating projections and enables various non-negative constraints to be imposed on dual planes. After that, a series of input-output algorithms90 were further developed to accelerate the convergence of the error reduction algorithm. Input-output algorithms differ from error reduction algorithms during the nonlinear operation on the object plane. In the basic input-output (IO) algorithm, the object constraints are imposed only on the points that violate the object constraints \(h({\bf{r}})\notin {\gamma }_{o}\). Such a change is based on the idea that the distribution of \(h^{\prime} ({\bf{r}})\) for the next iteration does not have to satisfy the object constraints but is given to drive a desired and constrained output. Based on this thought, different nonlinear operations can be introduced on the object plane, as is shown in Fig. 3c, through which the output-output (OO) algorithm and the hybrid input-output (HIO) algorithm were developed. The HIO algorithm possesses an advantage in avoiding the stagnation problem, which occurs sometimes in the OO algorithm and prevents the optimization from getting close to the solution. In some situations, the unique characteristics of the HIO algorithm set it apart from the input-output algorithm series for its imaging applications91,92. All these alternating projection algorithms are now widely studied as early solutions to phase retrieval problems. They have been derived into different forms for specific applications93.
a Gerchberg-Saxton algorithm (G-S algorithm)88. b Error reduction algorithm89. c Input-output algorithm series, including the basic input-output algorithm (IO), the output-output algorithm (OO), and the hybrid input-output algorithm (HIO)90. d Iterative Fourier-transform algorithm (IFTA)70. e Signal window based soft-encoding IFTA71. f Phase mask based on IFTA106
Concurrent with the advancement of algorithm flows, numerous efforts are made to adapt alternating projection algorithms into the physical model of CGH. Specially in hologram optimizations, alternating projections are applied to two enclosed sets associated with potential object solutions γo and potential hologram solutions γh, which are restricted by object constraints and hologram constraints respectively. The preliminary applications of alternating projections to solve the hologram synthesis problem in CGH appeared just before the first report of the G-S algorithm. Gabel and Liu proposed a flow chart with Fourier transform pairs to minimize binary hologram reconstruction errors in 197094. Hirsch et al.95 invented a technique for optimizing Kinoforms with a loop of forward and backward propagations in 1971. In the early studies on optimization algorithms for CGH96, Fourier transform pairs were widely used for diffractive calculation since they could easily simulate Fraunhofer diffractions at infinity. When this scheme became known as the error-reduction algorithm and was extended into an input-output algorithm series, Fienup clarified that such an iterative Fourier transform loop is the alternating projections to solve a hologram from specific constraints on the object plane and hologram plane, rather than simple exchanges of amplitude values on dual planes62,97.
With the extension of the categories of constraints, holograms can be computed well in line with the theoretical model of holography. In 1988, an iterative Fourier transform algorithm (IFTA) specially applied to CGH was proposed by Wyrowski and Bryngdahl70. As is shown in Fig. 3d, the IFTA scheme additionally includes the bandwidth constraint ② and the spatial scale constraint ③ in computation. A dense sampling on the object plane was proposed by applying the bandwidth constraint of \({\Delta }_{h}={\Delta }_{I}/2\), where ΔI was the bandwidth of the object intensity field79. The theoretical basis of this operation is that the spectra of the object wave and the spectra of the object intensity are related by autocorrelation. Minor errors and noises were able to be revealed in the numerical reconstruction, as a joint result of physical constraints and dense sampling. The performance of an optimized hologram can be sufficiently improved in optical reconstruction, with a desired object intensity fed as a constraint on the object plane. In the subsequent studies on IFTA, signal windows, energy operators, and soft encoding are introduced into the algorithm. The signal window utilizes the redundancy of hologram design and distributes the noises outside the signal window98. Because the object and the hologram are both finite in spatial scale, the object wave is defined within a window WO in computation, and the corresponding hologram is restricted within a window WH, as shown in Fig. 3e. On the object plane, a weighting factor is usually introduced to adjust the energy distribution ratio in and out of the window WO71, so that the object intensity constraint is only imposed within the signal window. On the hologram plane, with WH introduced as the bandwidth and special scale constraint on the hologram. The definition of the signal window allows for the use of amplitude freedom other than phase freedom on the object plane while optimizing holograms99,100. Introducing the double freedoms loosens the object intensity constraint ①, which exchanges for higher reconstructing accuracy by allowing noises in less-concerned areas101,102. Coordinated with quadratic phase initializations, IFTA algorithms with double freedoms can achieve speckle-free reconstructions103,104. The introduction of signal windows makes use of the dark area outside the object window to accommodate minor errors produced in the optimization of a POH, which also accelerates the process of reaching to the convergence. To ensure an appropriate distribution of the signal and the noise, energy operators E and 1/E are both used to normalize the intensities in the transformation between the object plane and the hologram plane. The soft encoding strategy was introduced to control the direction of optimization in IFTA. One way of achieving soft encoding is to apply a varying operation on the hologram plane71. This varying operation leads to minimal changes in the object phase and results in a smooth transition towards the desired optimizing direction. The signal window and the soft encoding, when combined, work for various types of initial object phases in IFTA, offering this algorithm a great deal of flexibility. Apart from the random phase, which was originally used as the object phase for Kinoform, some remarkable attempts at various phase initialization also include the quadratic phase71, the constant phase105, and the pre-optimized phase64. The utilization of constant phase and quadratic phase is driven by a demand for speckle suppression in CGH. The pre-optimized phase is utilized to develop a less iterative106,107 or noniterative108,109,110 approach to quickly generate holograms, which greatly accelerates the computation. As is shown in Fig. 3f, such algorithms can be divided into two steps: the first step optimizes a random phase mask within the signal window, which to the largest extent satisfies the constraints on the dual planes; the second step optimizes the hologram with the pre-optimized phase input as the initial object phase. The pre-optimized phase can be utilized to initialize object waves with different intensity patterns. And the computation can thus be simplified into a loop of optimizing holograms with only several iterations, or even without iterations.
Despite the introduction of alternating projection algorithms, the reconstruction of CGH can still be plagued by various unwanted effects, including speckle noise, artifacts, and stripes. These effects may trap the optimization into stagnation and remain in reconstruction even after convergence. Thus, some modifications applied to optimization algorithms are also studied. Since the complex distributions on both the object plane and the hologram plane are calculated in iterations, the framework of the alternating projections possesses the flexibility to include manipulations on the object phase111, Gaussian illumination112, weighting factors113, internal processing in the loop114, partitioned optimization115,116 and multi-loop optimization117. With highly promoted reconstructing accuracy and computation stability enabled by accumulated improvements on alternating projection algorithms, alternating projection algorithms occupy an indispensable part in optimizations of CGH, especially for the synthesis of holograms encoded on static holographic media, like DOEs118,119 and metasurfaces120,121,122.
The highly developed schemes enable alternating projection algorithms to be applied to hologram generation based on neural networks in the early stages123, which provide training datasets of ground-truth holograms. Despite the quick advancements of CGH combined with various neural networks in recent years124,125, a pre-optimization carried out by the alternating projections is still a widely adopted choice for supervised learning126,127,128,129.
Stochastic Gradient Descent (SGD)
Because of the existence of specific object intensity distributions, the inverse problem of hologram synthesis in CGH can also be cast as the optimization of a parameterized objective function requiring minimization with respect to its parameters. Since the choice of the objective function is often stochastic and differentiable with respect to its parameters, SGD is considered as an efficient and effective first-order gradient descent (\(\nabla {\mathcal{L}}\)) framework for optimization. Although such a scheme searching for a local optimal solution along the gradient descent direction has been early applied to phase retrieval130,131, SGD was hardly applied to CGH till recent years. In 2019, Chakravarthula et al. chose such a framework to demonstrate the loss functions they introduced for Wirtinger holography132. In 2020, Peng et al. utilized SGD to carry out a camera-in-the-loop optimization133.
Feasible gradient descent requires the gradient of the object function to be comparable, or in other words, the gradient with respect to the variable is a real value. Yet, a scalar wavefront is complex, which creates a major difficulty in applying gradient descent strategies in hologram optimization. Since POHs only concern the phase component of the holographic wavefront on the hologram plane, the optimization of POHs becomes relatively easy for implementation. With this regard, a POH can be synthesized by constructing an objective function \({ {\mathcal{L}}}_{P}\) associated with a forward model and solving the minimization problem:
\({ {\mathcal{L}}}_{P}\) is also referred as a loss function, which usually describes the difference between the reconstructed intensity I(ϕ) and the object intensity Iobj. I(ϕ) represents the reconstructed intensity as a function with respect to the POH ϕ. Since the reconstructed intensity I(ϕ) and reconstructed wave h(ϕ) are linked by an ill-conditioned relation \(I(\phi )={|h(\phi )|}^{2}\), the optimization for this hologram synthesis problem is non-convex. In this case, SGD is a relatively efficient optimization method to solve this problem because only first-order partial derivatives concerning all the variables are computed, which have the same computational complexity as just evaluating the objective function. Adam algorithm134, shown in Fig. 4a, is one of the most widely implemented SGD-based algorithms for solving the hologram synthesis problem.
a SGD optimization of a phase-only hologram (POH). The Adam algorithm updates the POH by calculating the 1st-moment vector b and the 2nd-moment vector p. b SGD optimization of a complex hologram (CH). The Adam algorithm updates the object phase of the CH. c Quasi-Newton optimization of a POH. The L-BFGS algorithm updates the POH by calculating the first-order gradient g and the inverse Hessian matrix \(\bar{H}\). d Quasi-Newton optimization of a CH. The L-BFGS algorithm updates the object phase of the CH
To minimize the expected value of \({ {\mathcal{L}}}_{P}\), the Adam algorithm updates the exponential moving averages of the partial derivative \({\nabla }_{\phi }{ {\mathcal{L}}}_{P}\) and the elementwise square gradient \({({\nabla }_{\phi }{ {\mathcal{L}}}_{P})}^{2}\), which are described as the 1st-moment vector b and the 2nd-moment vector p. The optimization only requires the first-order gradient of the objective function \({ {\mathcal{L}}}_{P}\) with respect to the POH ϕ, for which reason efficient computation can be achieved with small memory. In such a framework, the object constraint is imposed through the minimization problem itself, while the hologram constraints are applied using the partial derivative \({\nabla }_{\phi }{ {\mathcal{L}}}_{P}\), which is calculated in each update of optimization. Although we write \({ {\mathcal{L}}}_{P}\) in a general form of stochastic scalar function with respect to the reconstructed intensity I(ϕ) and the object intensity Iobj, the choice of \({ {\mathcal{L}}}_{P}\) is actually of great diversity for hologram synthesis135,136. In many CGH implementations, \({ {\mathcal{L}}}_{P}\) is composed of a sum of subfunctions evaluating reconstructing errors137, and a normalization term is occasionally added to balance other reconstruction parameters132,138. This flexibility allows SGD to achieve a number of functional holographic display demonstrations, as is shown in Fig. 5.
a Energy envelope expansion139. b Speckle suppression and contrast enhancement67. c Improvement in accommodation response138. d Optimization of high diffraction orders140. e Uneven liquid crystal modulation143. f Étendue expansion84. References140,143 are reprinted with permission from © Optical Society of America. Reference138 is reprinted with permission from © ACM. References67,84,139 are reprinted under a Creative Commons Attribution 4.0 International License
For the synthesis of CHs, some actions need to be taken to ensure the calculated gradient is a real value. Here we accordingly provide an extended scheme for the optimization of CHs using SGD. This is achieved simply by switching the output onto the object plane instead of the hologram plane, given that the object phase is the only floating parameter of the optimization model. The minimization problem is accordingly changed, which can be given as:
where \({ {\mathcal{L}}}_{C}\) is the loss function, and \(I(\varphi ^{\prime} )\) is the reconstructed intensity function with respect to the object phase \(\varphi ^{\prime}\). Equation (2) determines that the output of this optimization is the object phase, as shown in Fig. 4b, which is again turned into an optimization on a phase distribution. The basis of this scheme is the fact that with a given propagation model and a given object intensity, the complex hologram is only decided by the object phase. Different from that of POHs, the optimization of CHs includes both forward and backward propagations, resulting in a different expression of the gradient \({\nabla }_{\varphi ^{\prime} }{ {\mathcal{L}}}_{C}\). With the optimized object phase \({\varphi }_{{\rm{opt}}}\) and the known object intensity Iobj, the desired CH can be further calculated through an additional diffractive propagation.
Through the combination of a well-performed physical model, a well-chosen initial object phase, and a well-defined loss function, SGD itself has been verified to achieve good reconstructions of POHs, which leads to impressive accomplishments in 3D display with less or without speckles67,139. And, because the early adoption of SGD in CGH was linked with camera-in-the-loop optimization133, hardware involution was developed in tandem with the promotion of SGD-based hologram synthesis, in which the pattern captured by the camera replaced the reconstructed intensity in optimizations. With camera-in-the-loop optimization, subsequent reconstructions of SGD were further improved significantly by suppressing noises in optical setups, including high diffraction orders140, DC term141, speckles66,142, phase uniformity143, and aberrations144.
Because gradient descent contributes one of the most common optimizers to unsupervised learning of neural networks, the introduction of SGD neatly builds up a connection between CGH and unsupervised learning145,146,147. As the investigations in Fig. 6 have presented, unsupervised neural networks based on SGD accelerate the high-accuracy generation of holograms to ~0.01 s when trained by comparing I and Iobj through a defined loss function rather than large datasets148,149,150. As a step already taken further, unsupervised learning coordinated camera-in-the-loop training perfectly solves the problem of the long processing time that comes with camera-in-the-loop optimization133. This scheme pushes the boundaries of optimization-based hologram synthesis by demonstrating a trade-off between computation time, reconstruction quality, and system integration151,152,153,154.
a DeepCGH149. b Holo-encoder148. c Learned hardware-in-the-loop phase retrieval151. d Neural holography with camera-in-the-loop training133. e Neural 3D holography for AR/VR display152. f Time-multiplexed neural holography153. References148,149 are reprinted with permission from © Optical Society of America. References133,151,152 are reprinted with permission from © ACM. Reference153 is reprinted with permission from the authors
It is worth noting that the proposal of hologram optimization based on the alternating direction method of multipliers (ADMM) is closely associated with SGD. The ADMM breaks the optimization problem into subproblems with respect to multiple variables, which is achieved by modifying the minimization formula into an augmented Lagrangian form155,156. The final hologram is acquired by solving subproblems alternatingly157, which requires an inner loop of optimization in the hologram synthesis. The existing scheme of this loop is achieved by the Adam algorithm152, which incorporates the diffractive propagation model implemented by matrix manipulation.
Quasi-Newton method
The quasi-Newton method has become a widely used optimization framework for hologram synthesis in recent years. The quasi-Newton method is a second-order gradient descent method, which minimizes the loss function by constructing and storing a series of matrices that approximate the Hessian or inverse Hessian matrix of the loss function158. This operation is associated with the calculation and storage of the second-order gradient (\({\nabla }^{2} {\mathcal{L}}\)) of the loss function, which requires a larger amount of computation and storage compared with SGD. However, the calculation of the second-order gradient allows for a further search on the optimal direction, providing a possible “steepest” path on gradient descent. The combination of CGH and the quasi-Newton method was introduced by Zhang et al. when they proposed the non-convex optimization for volumetric CGH (NOVO-CGH)68. Later in 2019, the quasi-Newton method was also used in Wirtinger holography132.
The limited memory BFGS (L-BFGS) algorithm occupies the mainstream in the existing implementation of the quasi-Newton method in hologram synthesis159,160, which is chosen for its relatively high efficiency using a limited amount of storage. As depicted in Fig. 4c, \({ {\mathcal{L}}}_{P}\) is the loss function to be minimized given in Eq. (1) for POH synthesis. Similar to SGD, constraint ① in the quasi-Newton method is also imposed through the minimization problem itself, while constraints ②, ③, and ④ are imposed in the expression of the partial derivative \({\nabla }_{\phi }{ {\mathcal{L}}}_{P}\). The POH is updated by calculating both the first-order gradient \(g={\nabla }_{\phi }{ {\mathcal{L}}}_{P}\) and the inverse Hessian matrix \(\bar{H}\). The inverse Hessian is related to the second-order gradient \({g}_{i+1}-{g}_{i}\), which is estimated by adding up a series of correction matrices of each update in computation161. The accumulation of correction matrices is the major reason why the second-order gradient descent requires a greater amount of computation and storage. As a practical approach to implementing the quasi-Newton method, the L-BFGS algorithm has an upper limit for the number of correction matrices that can be stored. Once the number of updates reaches the storage limit, the stored correction matrices are discarded, and the correction matrices are reaccumulated.
Similarly, the quasi-Newton method is able to synthesize CHs by switching the output to the object phase φ and calculating the corresponding hologram through diffractive propagation, as is illustrated in Fig. 4d. The minimization problem can also be described with the expression in Eq. (2).
Although the quasi-Newton method and SGD are the second-order gradient descent and first-order gradient descent methods, respectively, they share the same derivation for \({\nabla }_{\phi }{ {\mathcal{L}}}_{P}\) and \({\nabla }_{\varphi ^{\prime} }{ {\mathcal{L}}}_{C}\). Since a second-order gradient is introduced, the quasi-Newton method usually presents higher reliability in searching for the direction of gradient descent, which enables generally higher reconstructing accuracy of holograms with diverse propagation models in practice68. However, as a consequence of the large amount of computation in optimization, the quasi-Newton method requires much longer computing time and larger memory compared with other optimization frameworks149,162.
Optimization pipelines for 3D holography
In this section, we describe the way how various optimization frameworks are applied to 3D holography. The flexibility of the optimization frameworks also brings with it a diversity of pipelines for the hologram synthesis of 3D objects. Many diffractive propagation models are applicable for volumetric optimization. Here, we unify them into the band-limited ASM83, for a better illustration of different optimization pipelines. Three major existing optimization pipelines for 3D holography are discussed in detail. These pipelines were proposed alongside the advancement of alternating projection algorithms and have been combined with other optimization frameworks.
Despite the existence of the point-cloud sampling strategy163 and the polygon-based sampling strategy164, in the optimization pipelines for 3D holograms, the object wave is mainly partitioned by layers165,166,167. One ancient approach to optimization for 3D holography is oriented from wavefront superposition58,105, which is a very conventional principle of holography168. As is shown in Fig. 7a, the alternating projections occur in each single layer of the object wave \({h}_{l}({\bf{r}})\) with the object intensity constraint ① imposed on each \({h}_{l}({\bf{r}})\). \({z}_{l}\) represents the propagation distance for the \({l}^{{\rm{th}}}\) layer. Each projection generates an optimized diffractive wavefront \({H}_{z}({\bf{u}})\) for every sampled depth z. The final hologram is synthesized by the superposition of all the wavefronts.
The superposition method has been applied to the optimization of holograms, which is especially preferred in the synthesis of CHs169 and AOMs170. This is due to the fact that the superposition itself does not break any constraints on CHs, while the hologram intensity constraint ④ for POHs can be violated after superposition171,172,173. To enable a better generation of POHs with the superposition method, a middle complex-amplitude plane is usually added to receive the superposition of waves from multiple depths136,174,175. Superposition optimization enables the preservation of full-depth cues of the object wave and possesses a competitive edge in applications like 3D display176. However, it brings about a degradation in reconstructed accuracy for POHs. The optimization part of the superposition method is essentially about solving planer optimization problems, which can be implemented by alternating projections, SGD, and the quasi-Newton method.
After the wide utilization of superposition optimization, the idea of global optimization is developed177, which enables a broader range of applications of CGH, especially those based on POHs. As is shown in Fig. 7b, the projections occur between the object volumetric matrices h(r, z) and the hologram plane, with the object intensity constraint ① imposed on the 3D matrix of the object wave h(r, z). Because the hologram intensity constraint ④ restricts the hologram \(H^{\prime} ({\bf{u}})\) throughout the optimization, the capability of such a global method in generating appropriate POHs is highlighted. A distinct characteristic of global optimization is that the dark voxels on the front layer are also involved in the optimization, blocking the desired intensity on the back layer. The crosstalk breaking performance becomes more remarkable when an object wave is scattered and reconstructed with speckles, but still requires slow phase changes between consecutive planes. This characteristic makes POHs generated by the global method more applicable in fields like optical encryption178, 3D beam shaping179,180,181, and data storage176,182. However, the global method is yet less suitable for 3D display183, in which the crosstalk breaking disturbs the visibility of distant scenes and results in an unreasonable energy distribution100. Intensive modifications have been made to improve the performance of the global methods184,185,186, and a corresponding pre-optimization scheme has also been proposed187. The 3D optimization of POHs with the global method can be carried out by alternating projections, SGD, and the quasi-Newton method68,188. However, due to the difficulty in selecting appropriate outputs for gradient-descent frameworks, the feasible global optimization of 3D CHs is currently more restricted within algorithms achieved by alternating projections189.
Sequential optimization is another viable scheme that can satisfy the hologram intensity constraint ④ for POHs. An earlier form of the sequential method was the optimization between object planes, called the ping-pong algorithm190. Then the hologram plane was also included in optimization191,192. As is shown in Fig. 7c, the sequential method is characterized by sequential propagations from distant object planes \(h^{\prime} ({\bf{r}},{z}_{l+1})\) to less distant object planes \(h^{\prime} ({\bf{r}},{z}_{l})\), and from the nearest object plane \(h^{\prime} ({\bf{r}},{z}_{1})\) to the hologram plane \(H^{\prime} ({\bf{u}})\). The hologram \(H^{\prime} ({\bf{u}})\) is subject to the hologram intensity constraint ④. In order to form the loop, the constrained hologram H(u) is then propagated sequentially through each object plane and to the most distant one, with the object intensity constraint ① imposed on every \(h({\bf{r}},{z}_{l})\). Because of the existence of the dark voxels, sequential optimization possesses a crosstalk-breaking characteristic like global optimization, which makes this method suitable for encryption and storage193. The specific problem of the sequential method is that its corresponding reconstructions have decreasing accuracy as the light propagates. And because sequential propagations are required, the implementation of the sequential method is mostly achieved by alternating projections.
Apart from the optimization pipelines, loss function formulation is another crucial factor affecting the performance of 3D holographic reconstruction. In 3D holographic applications, especially for holographic display, not only the accuracy of focus reconstruction is important, but the visual effect of defocus blur is also a great concern in evaluation. The defocused performance of 3D holographic reconstruction is related to the object phase distribution and the visible parallax within the maximum diffraction angle. On this basis, a well-defined loss function can model the spatial variation of diffractive patterns. Representative approaches to cast loss functions include the voxel distribution method, the focal stack method, and the light field method. As is shown in Fig. 8a, the voxel distribution method voxelizes and remaps the object into a 3D volume containing bright voxels and dark voxels68,69,149, which is further transferred into the object matrix associated with the loss function. The hologram pixels diffract light to focus on points in space to create bright voxels under illumination. However, it is difficult or even impossible to create a dark voxel in space immediately after a focused bright voxel, because it is physically infeasible to have the phase of light changed rapidly during the free space propagation over a very small distance. For this reason, the reconstruction of the voxel distribution method usually suffers from severe noise and greatly degraded image quality132. The focal stack method, as is shown in Fig. 8b, constructs 3D volume by rendering focal stacks at different propagation depths133,153. These computationally generated focal stacks at multiple depths are jointly involved in optimization together with the target object, which enables the diffractive reconstruction to better imitate the natural defocus blur, especially in holographic display194. The light field method originates from holographic stereograms which reconstruct an array of directional views of the object with a single hologram139,195,196,197, as is shown in Fig. 8c. Although some strategies of light field computation without hogel-based structure have been developed198, the application of the light field method is still relatively limited in 3D hologram optimization due to current band limitation existing in the pixelized holographic media like SLMs199,200.
Here we provide a computational comparison between different optimization frameworks. As presented in Fig. 9a, the optimization of CHs and POHs is demonstrated by basic frameworks of the alternating projections, the SGD method (implemented by the Adam algorithm), and the quasi-Newton method (implemented by the L-BFGS algorithm), respectively. The object phase is initialized with a random matrix. The 512 × 512 holograms are computed on a PC with an Intel Core i9-9900K 3.6 GHz CPU and 32.0 GB of RAM, and an NVIDIA GeForce RTX 2080Ti GPU. Two diffractive propagation methods including the FFT and the ASM are compared, where the propagation distance of ASM is 50 mm. In general, the quasi-Newton method outperforms other optimization frameworks in reconstructing accuracy, which is measured by root-mean-square error (RMSE, \({E}_{{\rm{RMSE}}}=\sqrt{{\sum }_{m,n}{[{a}^{2}({{\bf{r}}}_{m,n})-{I}_{{\rm{obj}}}({{\bf{r}}}_{m,n})]}^{2}}\)). However, much more computation time is required for the convergence. Compared with the alternating projections, the holograms optimized by SGD present a slightly lower RMSE and longer time. However, the flexibility in editing the loss functions and the superiority in computing efficiency make SGD unreplaceable in a number of implementations of CGH, especially for the training stage of hologram synthesis based on deep learning methods. The different performances of the first-order and second-order gradient descent methods while approaching convergence are also illustrated. Since SGD is a first-order gradient descent method, biased moment vectors are continuously introduced into the optimization when the algorithm approaches its convergence, resulting in the fluctuation around the optimal solution. On the other hand, the second-order gradient descent method searches for the optimal direction of gradient descent, the output of optimization varies at a limited range around the optimal solution. For this reason, the RMSE curves of SGD optimization in Fig. 9a bounce up when relatively weaker constraints are imposed on the optimization. The optimization convergence is also related to the diffractive calculation approaches utilized. It should be noted that the optimization of FFT-based holograms gives slightly worse quality of reconstruction here. This is because the bandwidth constraint imposed on FFT-based holograms is merged with spatial scale constraint, and thus the searching range of the FFT-based hologram optimization is less restricted around the optimal solution when approaching the convergence. Table 1 provides a review of the performance of these algorithms based on the equal testing of different frameworks.
The performance of different frameworks for 3D holography is presented in Fig. 9b. Both the superposition method and the global method are demonstrated with the reconstructions of CHs and POHs. Three optimization frameworks perform similarly with 3D holograms generated with the superposition method. As predicted, the diffractive cues for different layers are well preserved after superposition, enabling the generation of defocus dispersion for 3D display. For the global method, however, different optimization frameworks present distinct differences in reconstruction. The quasi-Newton method and SGD are better at breaking the crosstalk between layers and showing more uniform energy distributions, among which SGD performs better at suppressing severe speckle noise. Yet, the multi-depth reconstructions by the alternating projections cannot avoid the disturbance from adjacent layers. This impact is more severe with the sequential method, which is currently only applicable to alternating projection algorithms.
Initialization
The initialization condition is a crucial factor affecting the optimum solution sought out by an optimization algorithm. Specifically in a hologram synthesis problem, the object phase \(\varphi ({\bf{r}})\) is the only free parameter floating in the optimization. Given that the hologram optimization is non-convex, as is depicted in Fig. 10a, the initial value of this free parameter decides at which point the optimization path starts and to which local optimum point the optimization path is heading. Since the closet local optimal solution differs with different initialization conditions, illustrated in Fig. 10b, the choice of the object phase is closely associated with the performance of the in-focus reconstrued intensity. And the initial object phase determines the defocus pattern through the reconstructed phase. Various attempts are made to manipulate the object phase for initialization of the optimization, adapting computer-generated holograms for diverse applications. In this section, different object phase initializations, such as random phase, constant phase, and quadratic phase, and their corresponding reconstructing performances are discussed.
a A schematic diagram of the non-convex optimization in an actual hologram synthesis problem. Different optimization paths are triggered by different initialization conditions. b The variation of the loss function when a local optimal solution is searched. The closest optimum point is decided by the start point of optimization. c A rendered ideal reconstruction for a 3D scene. The reconstructions of CHs optimized with different initialization conditions. d Random object phase. e Constant object phase. f Quadratic object phase
Random phase
In the experiments of conventional optical holography, the real-world objects that generate the object waves to be recorded have scattering surfaces201. This is based on the fact that most materials encountered in the real world are rough on the scale of object wavelength. Various microscopic facets of the rough scattering surfaces of an object contribute randomly distributed phases to the recorded object wave. The random phase is widely used in CGH because the scattering surfaces of real-world objects produce scattered wavelets, which in turn leads to speckles202. Although CGH generates holograms through wave computation instead of optical interferometry, the phase term of the object wave needs to be defined according to the phase generated by the real-world object. An artificial object wave computed with a random phase distribution is thus more in line with the real object wave generated from scatterers.
It can be seen that the phase of the object wave is closely related to the defocus effects in reconstruction. Initialized with a random object phase, as shown in Fig. 10d, the reconstructed wave of optimization can generate a natural blur of defocus like that of the scatterers in the real world, which slowly varies on spatial distribution. This capability is essential because reconstructing a 3D scene is the core strength of holography and reproducing the object image faithfully at various positions in space is necessary. From this perspective, random or to some extent random distributions conform better to the phases of waves modulated by real-world objects in most cases. Since scattering materials occupy a considerable proportion in the natural world203, preserving the propagating characteristic of a scattered wave and at the same time removing speckles is unavoidable for CGH.
Apart from this consideration, there is an additional reason to use a random phase. In a great number of applications, holograms need to be encoded onto phase-only elements, such as DOEs, metasurfaces, and liquid crystal on silicon (LCoS). The POH operates only on the phase of an incident wave and is generated based on the assumption that a scattered wavefront can be reconstructed with only the phase information. Therefore, using random phase is also an essential procedure to enable the encoding of POHs in specific applications of holography.
However, a significant drawback of introducing a random phase into optimization is the degraded optical reconstruction severely affected by speckle noise. Careful studies have identified that speckles in CGH are closely associated with phase singularities and are regarded as the results of optical vortices71. Speckles in a scattered wavefront are found to be intertwined with optical vortices204, indicating the existence of phase singularities on which amplitudes are precisely zero and phases are indeterminate. Although scattered wavefronts are artificially generated in CGH, speckles can be suppressed optically when phase singularities are eliminated205. Computational approaches to suppress speckles are essentially making modifications to the random phase introduced in the generation of holograms. Optimization algorithms are able to restrict the scattering caused by the random phase to suppress speckles, which is achieved by feeding the object intensity and bandwidth limitation as constraints67,68. Some optimizing or training methods, such as camera-in-the-loop training66,133,142, can produce more remarkable results by involving optical systems in computation. However, it has been discovered that optimization algorithms are not inherently capable of eliminating all the speckles205. Given the connection between the random phase and speckles, intensive research is carried out to replace the random phase with other phase formats, including constant phase, quadratic phase, and some other artificially defined phase distributions, which will be introduced in the following sections.
Constant phase
Constant phase is increasingly utilized to synthesize holograms due to the invention of double-phase holograms206, delivering highly photorealistic reconstructions free from speckles207,208. As a choice of initialization, constant phase is more frequently used in 3D hologram optimization with the superposition method, especially for the synthesis of POHs optimized in virtue of middle complex amplitude planes136. The constant phase physically describes the phase distribution of extremely smooth surfaces in the natural world, like mirrors or objects with surfaces as smooth as mirrors. Suchlike objects cause less scattering and fewer speckles naturally even when they are illuminated by coherent beams. As is shown in Fig. 10e, when initialized with a constant object phase, the optimized object wave is reconstructed with undiffused diffracting patterns at defocus positions. This kind of defocus effect with diffracting patterns greatly differs from defocus blurs which are generated by scatterers. Constructing object waves with constant phases means replacing the materials of the object with the ones that naturally cannot cause light scattering and produce speckles. Consequently, the object wave would lose the propagating characteristic as a scattered wave and fail to provide a faithful appearance of the object in 3D space. Observed along the direction of propagation, however, constant phase generates diffracting outlines instead of defocus blurs, weakening the depth sensation of 3D scenes.
Quadratic phase
The quadratic phase enables the production of faithful, speckle-free, planer reconstructions, especially when it is coordinated with alternating projection algorithms. With the improvements enabled by the double freedoms103,209 and proper bandwidth limitation102,210, alternating projection algorithms are able to overcome ring artifacts brought about by the quadratic phase, which highly promotes the reconstructing performance in holographic projection.
Nevertheless, the intensity pattern attached to the quadratic phase cannot keep its feature size with the propagation of light172, which strictly restricts its application within two-dimensional (2D) holographic projections and beam shaping211. As is shown in Fig. 10f, the object wave can be constructed by such a quadratic phase for 3D hologram optimization. In the optimization pipeline, the diffractive wavefronts are converted into CHs and reconstructed at various distances. With the convergent quadratic phase, the object wave is reconstructed with undiffused diffracting patterns at multiple defocus positions. And the reconstructed object intensity cannot maintain its feature size with the enlargement of the reconstructing distance. Because of this enlarging characteristic, the application of the reconstruction for the quadratic phase is limited within 2D holographic projection.
Defining the object phase directly as a quadratic phase has a huge advantage in that it can generate holograms efficiently with a single propagation and avoids speckles. And this quadratic phase approach is indeed applicable to 2D projection. However, it has a consequence that the object wave would lose the propagating characteristic as a scattered wave and fail to provide a faithful appearance of the object and maintain the feature size in 3D space. This is why random phase is still widely used in CGH even after such methods were proposed. Another drawback of introducing well-defined phases, like the constant phase and the quadratic phase, is that any perturbation in the optical configuration of reconstruction, including element damage, dust, or varying illumination, can cause severe degradation in the image quality71. This property is concluded as the “robustness” brought about by the initial phase and compared in Table 2.
Other phase patterns
Some other attempts at adjusting random phase also include imposing constraints on the phase gradient, controlling the standard deviation of phase radius212, and limiting the range of phase radius213,214. A regularization term may be added to the objective function to restrict the variation of some parameters, like the gradient of the object phase, in optimization. With these modifications on the object phase, scattering characteristics of the object wave, which is ensured by sufficient randomness and fluctuating ranges, are weakened synchronously with the suppression of speckles, resulting in various distortions in depth dimension138. Table 2 provides a general comparison of 3D reconstructing performances of different initial phases.
Conclusion
In this review article, we have focused on the inverse problem in CGH and provided an overview of various frameworks based on non-convex optimization for appropriate hologram generation. The corresponding algorithms synthesize holograms by searching for optimal solutions within the feasible set restricted by physical models and implementation schemes. Although the computation of holograms is a numerical process, purely computational strategies cannot guarantee the generation of appropriate holograms with faithful reconstruction unless the underlying physics of holography is carefully considered in the calculation. Therefore, this review incorporates an understanding of the fundamental physics in CGH and provides an explanation of the various optimization frameworks, constraints imposed on optimization, 2D/3D optimization pipelines, and object phase initializations utilized in CGH algorithms. Recent advancements in calculation principles and computational strategies have further extended the capabilities of hologram optimization algorithms. By involving cameras or other hardware sets in optimization, the accuracy of planar reconstructions from CGH has been improved significantly. Furthermore, optimization-based hologram synthesis has been enhanced by introducing neural networks, resulting in a significant reduction in computation time to ~1 s208,215. Further improvement in the computing pipeline has made the optimization of 3D holograms feasible in practice, resulting in natural-like depth perceptions. These progresses have laid a foundation for the fast advance of CGH and been implemented on a wide range of holographic media, including digital micromirror devices, LCoS, DOEs119,216, and metasurfaces217,218,219,220. Overall, the optimization algorithms have significantly improved the accuracy and efficiency when generating appropriate holograms for faithful reconstruction, while considering the fundamental physics of holography, leading to their deployment in various fields.
Data availability
The open-source code for the above-mentioned numerical simulations on 2D and 3D hologram optimization is provided as supplemental materials221.
References
Gabor, D. A new microscopic principle. Nature 161, 777–778 (1948).
Leith, E. N. & Upatnieks, J. Reconstructed wavefronts and communication theory. J. Opt. Soc. Am. 52, 1123–1130 (1962).
Denisyuk, Y. N. Photographic reconstruction of the optical properties of an object in its own scattered radiation field. Soviet Phys. Doklady 7, 543 (1962).
Goodman, J. W. & Lawrence, R. W. Digital image formation from electronically detected holograms. Appl. Phys. Lett. 11, 77–79 (1967).
Situ, G. H. Deep holography. Light Adv. Manuf. 3, 13 (2022).
Fratz, M. et al. Digital holography in production: an overview. Light Adv. Manuf. 2, 15 (2021).
Brown, B. R. & Lohmann, A. W. Complex spatial filtering with binary masks. Appl. Opt. 5, 967–969 (1966).
Blanche, P. A. Holography, and the future of 3D display. Light Adv. Manuf. 2, 28 (2021).
Blinder, D. et al. The state-of-the-art in computer generated holography for 3D display. Light Adv. Manuf. 3, 35 (2022).
Wang, D. et al. Large viewing angle holographic 3D display system based on maximum diffraction modulation. Light Adv. Manuf. 4, 18 (2023).
Sandford O’Neill, J. et al. 3D switchable diffractive optical elements fabricated with two‐photon polymerization. Adv. Opt. Mater. 10, 2102446 (2022).
Neshev, D. N. & Miroshnichenko, A. E. Enabling smart vision with metasurfaces. Nat. Photonics 17, 26–35 (2023).
Eliezer, Y. et al. Suppressing meta-holographic artifacts by laser coherence tuning. Light Sci. Appl. 10, 104 (2021).
Li, L. L. et al. Intelligent metasurfaces: control, communication and computing. eLight 2, 7 (2022).
Forbes, A., De Oliveira, M. & Dennis, M. R. Structured light. Nat. Photonics 15, 253–262 (2021).
Shen, Y. J. et al. Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities. Light Sci. Appl. 8, 90 (2019).
Dorrah, A. H. et al. Light sheets for continuous-depth holography and three-dimensional volumetric displays. Nat. Photonics 17, 427–434 (2023).
Makey, G. et al. Breaking crosstalk limits to dynamic holography using orthogonality of high-dimensional random vectors. Nat. Photonics 13, 251–256 (2019).
Hu, Y. Q. et al. 3D-integrated metasurfaces for full-colour holography. Light Sci. Appl. 8, 86 (2019).
Fang, X. Y., Ren, H. R. & Gu, M. Orbital angular momentum holography for high-security encryption. Nat. Photonics 14, 102–108 (2020).
Li, X. et al. Independent light field manipulation in diffraction orders of metasurface holography. Laser Photonics Rev. 16, 2100592 (2022).
Kim, I. et al. Pixelated bifunctional metasurface-driven dynamic vectorial holographic color prints for photonic security platform. Nat. Commun. 12, 3614 (2021).
Javidi, B. et al. Roadmap on optical security. J. Opt. 18, 083001 (2016).
Hou, J. F. & Situ, G. H. Image encryption using spatial nonlinear optics. eLight 2, 3 (2022).
Li, Y. L. et al. Tunable liquid crystal grating based holographic 3D display system with wide viewing angle and large size. Light Sci. Appl. 11, 188 (2022).
An, J. et al. Slim-panel holographic video display. Nat. Commun. 11, 5568 (2020).
Smalley, D. E. et al. Anisotropic leaky-mode modulator for holographic video displays. Nature 498, 313–317 (2013).
Xue, G. P. et al. Polarized holographic lithography system for high-uniformity microscale patterning with periodic tunability. Microsyst. Nanoeng. 7, 31 (2021).
Otte, E. & Denz, C. Optical trapping gets structure: structured light for advanced optical manipulation. Appl. Phys. Rev. 7, 041308 (2020).
Adesnik, H. & Abdeladim, L. Probing neural codes with two-photon holographic optogenetics. Nat. Neurosci. 24, 1356–1366 (2021).
Carmi, I. et al. Holographic two-photon activation for synthetic optogenetics. Nat. Protoc. 14, 864–900 (2019).
Pégard, N. C. et al. Three-dimensional scanless holographic optogenetics with temporal focusing (3D-SHOT). Nat. Commun. 8, 1228 (2017).
Maimone, A., Georgiou, A. & Kollin, J. S. Holographic near-eye displays for virtual and augmented reality. ACM Transac. Gr. 36, 85 (2017).
He, Z. H. et al. Progress in virtual reality and augmented reality based on holographic display. Appl. Opt. 58, A74–A81 (2019).
Chang, C. L. et al. Toward the next-generation VR/AR optics: a review of holographic near-eye displays from a human-centric perspective. Optica 7, 1563–1578 (2020).
Xiong, J. H. et al. Augmented reality and virtual reality displays: emerging technologies and future perspectives. Light Sci. Appl. 10, 216 (2021).
Park, J. H. & Lee, B. Holographic techniques for augmented reality and virtual reality near-eye displays. Light Adv. Manuf. 3, 9 (2022).
Kress, B. C. & Pace, M. Holographic optics in planar optical systems for next generation small form factor mixed reality headsets. Light Adv. Manuf. 3, 42 (2022).
Martinez, C. et al. See-through holographic retinal projection display concept. Optica 5, 1200–1209 (2018).
Jang, C. et al. Holographic near-eye display with expanded eye-box. ACM Transac. Gr. 37, 195 (2018).
Wakunami, K. et al. Projection-type see-through holographic three-dimensional display. Nat. Commun. 7, 12954 (2016).
Sando, Y. et al. Holographic augmented reality display with conical holographic optical element for wide viewing zone. Light Adv. Manuf. 3, 12 (2022).
Skirnewskaja, J. & Wilkinson, T. D. Automotive holographic head‐up displays. Adv. Mater. 34, 2110463 (2022).
Coni, P., Damamme, N. & Bardon, J. L. The future of holographic head-up display. IEEE Consumer Electron. Mag. 8, 68–73 (2019).
Cheng, D. W. et al. Design and manufacture AR head-mounted displays: a review and outlook. Light Adv. Manuf. 2, 24 (2021).
Xiong, J. H. et al. Holo-imprinting polarization optics with a reflective liquid crystal hologram template. Light Sci. Appl. 11, 54 (2022).
Blanche, P. A. et al. Holographic three-dimensional telepresence using large-area photorefractive polymer. Nature 468, 80–83 (2010).
Wetzstein, G. et al. Inference in artificial intelligence with deep optics and photonics. Nature 588, 39–47 (2020).
Yu, H. et al. Ultrahigh-definition dynamic 3D holographic display by active control of volume speckle fields. Nat. Photonics 11, 186–192 (2017).
Huang, Z. Q., Marks, D. L. & Smith, D. R. Out-of-plane computer-generated multicolor waveguide holography. Optica 6, 119–124 (2019).
Park, J., Lee, K. & Park, Y. Ultrathin wide-angle large-area digital 3D holographic display using a non-periodic photon sieve. Nat. Commun.10, 1304 (2019).
Makowski, M. et al. Dynamic complex opto-magnetic holography. Nat. Commun. 13, 7286 (2022).
Pi, D. P., Liu, J. & Wang, Y. T. Review of computer-generated hologram algorithms for color dynamic holographic three-dimensional display. Light Sci. Appl. 11, 231 (2022).
He, Z. H. et al. Optimal quantization for amplitude and phase in computer-generated holography. Optics Express 29, 119–133 (2021).
Burch, J. J. A computer algorithm for the synthesis of spatial frequency filters. Proc. IEEE 55, 599–601 (1967).
Bryngdahl, O. & Lohmann, A. Single-sideband holography. J. Opt. Soc. Am. 58, 620–624 (1968).
Lohmann, A. W. & Paris, D. P. Binary fraunhofer holograms, generated by computer. Appl. Opt. 6, 1739–1748 (1967).
Lesem, L. B., Hirsch, P. M. & Jordan, J. A. The kinoform: a new wavefront reconstruction device. IBM J. Res. Dev. 13, 150–155 (1969).
Brown, B. R. & Lohmann, A. W. Computer-generated binary holograms. IBM J. Res. Dev. 13, 160–168 (1969).
Burckhardt, C. B. Use of a random phase mask for the recording of Fourier transform holograms of data masks. Appl.Opt. 9, 695–700 (1970).
Li, J., Smithwick, Q. & Chu, D. P. Holobricks: modular coarse integral holographic displays. Light Sci. Appl. 11, 57 (2022).
Fienup, J. R., Crimmins, T. R. & Holsztynski, W. Reconstruction of the support of an object from the support of its autocorrelation. J. Opt. Soc. Am. 72, 610–624 (1982).
Barakat, R. & Newsam, G. Necessary conditions for a unique solution to two‐dimensional phase recovery. J. Mathe. Phys. 25, 3190–3193 (1984).
Bräuer, R., Wyrowski, F. & Bryngdahl, O. Diffusers in digital holography. J. Opt. Soc. Am. A 8, 572–578 (1991).
Barakat, R. & Newsam, G. Algorithms for reconstruction of partially known, band-limited Fourier-transform pairs from noisy data. J. Opt. Soc. Am. A 2, 2027–2039 (1985).
Peng, Y. F. et al. Speckle-free holography with partially coherent light sources and camera-in-the-loop calibration. Sci. Adv. 7, eabg5040 (2021).
Lee, B. et al. High-contrast, speckle-free, true 3D holography via binary CGH optimization. Sci. Rep. 12, 2811 (2022).
Zhang, J. Z. et al. 3D computer-generated holography by non-convex optimization. Optica 4, 1306–1313 (2017).
Shi, L., Li, B. C. & Matusik, W. End-to-end learning of 3D phase-only holograms for holographic display. Light Sci. Appl. 11, 247 (2022).
Wyrowski, F. & Bryngdahl, O. Iterative Fourier-transform algorithm applied to computer holography. J. Opt. Soc. Am. A 5, 1058–1065 (1988).
Aagedal, H. et al. Theory of speckles in diffractive optics and its application to beam shaping. J. Modern Opt. 43, 1409–1421 (1996).
Dallas, W. J. Deterministic diffusers for holography. Appl. Opt. 12, 1179–1187 (1973).
Voelz, D. G. & Roggemann, M. C. Digital simulation of scalar optical diffraction: revisiting chirp function sampling criteria and consequences. Appl. Opt. 48, 6132–6142 (2009).
Tricoles, G. Computer generated holograms: an historical review. Appl. Opt. 26, 4351–4360 (1987).
Kelly, D. P. Numerical calculation of the Fresnel transform. J. Opt. Soc. Am. A 31, 755–764 (2014).
Muffoletto, R. P., Tyler, J. M. & Tohline, J. E. Shifted Fresnel diffraction for computational holography. Opt. Express 15, 5631–5640 (2007).
Voelz, D. G. Computational Fourier Optics: A MATLAB Tutorial. (SPIE Press: Bellingham, 2011).
Born, M. & Wolf, E. Principles of Optics: Electromagnetic Theory of Propagation, Interference and Diffraction of Light 4th edn, (New York: Elsevier, 2013).
Goodman, J. W. Introduction to Fourier Optics 3rd edn, (Englewood: Roberts and Company Publishers, 2005).
Matsushima, K., Schimmel, H. & Wyrowski, F. Fast calculation method for optical diffraction on tilted planes by use of the angular spectrum of plane waves. J. Opt. Soc. Am. A 20, 1755–1762 (2003).
Wyrowski, F. & Bryngdahl, O. Speckle-free reconstruction in digital holography. J. Opt. Soc. Am. A 6, 1171–1174 (1989).
Hecht, E. Optics. (Pearson Education India: Noida, 2012).
Matsushima, K. & Shimobaba, T. Band-limited angular spectrum method for numerical simulation of free-space propagation in far and near fields. Opt. Express 17, 19662–19673 (2009).
Kuo, G. et al. High resolution étendue expansion for holographic displays. ACM Transac. Gr. 39, 66 (2020).
Yoshikawa, N. & Yatagai, T. Phase optimization of a kinoform by simulated annealing. Appl. Opt. 33, 863–868 (1994).
Boyd, S. & Dattorro, J. Alternating Projections. (Stanford University, 2003).
Elser, V. Phase retrieval by iterated projections. J. Op. Soc. Am. A 20, 40–55 (2003).
Gerchberg, B. R. W. & Saxton, W. O. A practical algorithm for the determination of phase from image and diffraction plane pictures. Optik 35, 237–246 (1972).
Fienup, J. R. Reconstruction of an object from the modulus of its Fourier transform. Opt. Lett. 3, 27–29 (1978).
Fienup, J. R. Phase retrieval algorithms: a comparison. Appl. Opt. 21, 2758–2769 (1982).
Takajo, H. et al. Study on the convergence property of the hybrid input–output algorithm used for phase retrieval. J. Opt. Soc. Am. A 15, 2849–2861 (1998).
Takajo, H., Takahashi, T. & Shizuma, T. Further study on the convergence property of the hybrid input–output algorithm used for phase retrieval. J. Opt. Soc. Am. A 16, 2163–2168 (1999).
Fienup, J. R. Reconstruction of a complex-valued object from the modulus of its Fourier transform using a support constraint. J. Opt. Soc. Am. A 4, 118–123 (1987).
Gabel, R. A. & Liu, B. Minimization of reconstruction errors with computer generated binary holograms. Appl. Opt. 9, 1180–1191 (1970).
Hirsch, P. M., Jordan, J. A. & Lesem, L. B. Jr. Method of making an object dependent diffuser. U.S. Patent No. 3619022A (1971).
Gallagher, N. C. & Liu, B. Method for computing kinoforms that reduces image reconstruction error. Appl. Opt. 12, 2328–2335 (1973).
Fienup, J. R. Iterative method applied to image reconstruction and to computer-generated holograms. Opt. Eng. 19, 193297 (1980).
Akahori, H. Spectrum leveling by an iterative algorithm with a dummy area for synthesizing the kinoform. Appl. Opt. 25, 802–811 (1986).
Liu, J. S., Caley, A. J. & Taghizadeh, M. R. Symmetrical iterative Fourier-transform algorithm using both phase and amplitude freedoms. Opt. Commun. 267, 347–355 (2006).
Chang, C. L. et al. Speckle-suppressed phase-only holographic three-dimensional display based on double-constraint Gerchberg-Saxton algorithm. Appl. Opt. 54, 6994–7001 (2015).
Pang, H. et al. High-accuracy method for holographic image projection with suppressed speckle noise. Opt. Express 24, 22766–22776 (2016).
Chen, L. Z. et al. Weighted constraint iterative algorithm for phase hologram generation. Appl. Sci. 10, 3652 (2020).
Liu, K. X., He, Z. H. & Cao, L. C. Double amplitude freedom Gerchberg–Saxton algorithm for generation of phase-only hologram with speckle suppression. App. Phys. Lett. 120, 061103 (2022).
Chen, L. Z. et al. Phase hologram optimization with bandwidth constraint strategy for speckle-free optical reconstruction. Opt. Express 29, 11645–11663 (2021).
Leseberg, D. Computer-generated three-dimensional image holograms. Appl. Opt. 31, 223–229 (1992).
Velez Zea, A., Ramirez, J. F. B. & Torroba, R. Optimized random phase only holograms. Opt. Lett. 43, 731–734 (2018).
Velez-Zea, A. & Torroba, R. Optimized random phase tiles for non-iterative hologram generation. Appl. Opt. 58, 9013–9019 (2019).
Velez-Zea, A., Barrera-Ramírez, J. F. & Torroba, R. Improved phase multiplexing using iterative and non-iterative hologram generation. Opt. Lasers Eng. 151, 106921 (2022).
Chen, L. Z. et al. Non-iterative phase hologram generation with optimized phase modulation. Opt. Express 28, 11380–11392 (2020).
Zhang, C. et al. Non-iterative phase hologram generation for color holographic display. Opt. Express 30, 195–209 (2022).
Wyrowski, F., Hauck, R. & Bryngdahl, O. Computer-generated holography: hologram repetition and phase manipulations. J. Opt. Soc. Am. A 4, 694–698 (1987).
Amako, J., Miura, H. & Sonehara, T. Speckle-noise reduction on kinoform reconstruction using a phase-only spatial light modulator. Appl. Opt. 34, 3165–3171 (1995).
Kuzmenko, A. V. Weighting iterative Fourier transform algorithm of the kinoform synthesis. Opt. Lett. 33, 1147–1149 (2008).
Bigué, L. & Ambs, P. Optimal multicriteria approach to the iterative Fourier transform algorithm. Appl. Opt. 40, 5886–5893 (2001).
Wang, D. et al. Holographic capture and projection system of real object based on tunable zoom lens. PhotoniX 1, 6 (2020).
Wu, L. & Zhang, Z. Y. Domain multiplexed computer-generated holography by embedded wavevector filtering algorithm. PhotoniX 2, 1 (2021).
Zhai, T. T. et al. An approach for holographic projection with higher image quality and fast convergence rate. Optik 159, 211–221 (2018).
Peng, Y. F. et al. Mix-and-match holography. ACM Transac. Gr. 36, 191 (2017).
Soifer, V. A., Kotlar, V. & Doskolovich, L. Iteractive Methods for Diffractive Optical Elements Computation. (CRC Press: London, 2014).
Overvig, A. C. et al. Dielectric metasurfaces for complete and independent control of the optical amplitude and phase. Light Sci. Appl. 8, 92 (2019).
Rubin, N. A. et al. Jones matrix holography with metasurfaces. Sci. Adv. 7, eabg7488 (2021).
Deng, L. G. et al. Bilayer‐metasurface design, fabrication, and functionalization for full‐space light manipulation. Adv. Opt. Mater. 10, 2102179 (2022).
Yamauchi, S., Chen, Y. W. & Nakao, Z. Optimization of computer-generated holograms by an artificial neural network. In Proc. Second International Conference. Knowledge-Based Intelligent Electronic Systems. Proceedings KES’98 (Cat. No. 98EX111). Adelaide, SA, Australia: IEEE, 220-223, 1998.
Shimobaba, T. et al. Deep-learning computational holography: a review. Front. Photonics 3, 854391 (2022).
Zhang, Y. X. et al. Progress of the computer-generated holography based on deep learning. Appl. Sci. 12, 8568 (2022).
Horisaki, R., Takagi, R. & Tanida, J. Deep-learning-generated holography. Appl. Opt. 57, 3859–3863 (2018).
Goi, H., Komuro, K. & Nomura, T. Deep-learning-based binary hologram. Appl. Opt. 59, 7103–7108 (2020).
Liu, S. C. & Chu, D. P. Deep learning for hologram generation. Opt. Express 29, 27373–27395 (2021).
Ishii, Y. et al. Optimization of phase-only holograms calculated with scaled diffraction calculation through deep neural networks. Appl. Phys. B 128, 22 (2022).
Chen, Y. X. et al. Gradient descent with random initialization: fast global convergence for nonconvex phase retrieval. Mathe. Programm. 176, 5–37 (2019).
Ma, C. et al. Implicit regularization in nonconvex statistical estimation: gradient descent converges linearly for phase retrieval and matrix completion. In Proc. 35th International Conference on Machine Learning. Stockholm, Sweden: PMLR, 3345-3354, 2018.
Chakravarthula, P. et al. Wirtinger holography for near-eye displays. ACM Transac. Gr. 38, 213 (2019).
Peng, Y. F. et al. Neural holography with camera-in-the-loop training. ACM Transac. Gr. 39, 1185 (2020).
Kingma, D. P. & Ba, J. Adam: a method for stochastic optimization. In Proc. 3rd International Conference on Learning Representations. San Diego, CA, USA, 2015.
Yang, F. et al. Perceptually motivated loss functions for computer generated holographic displays. Sci. Rep. 12, 7709 (2022).
Wang, D. et al. Curved hologram generation method for speckle noise suppression based on the stochastic gradient descent algorithm. Opt. Express 29, 42650–42662 (2021).
Chen, C. et al. Multi-depth hologram generation using stochastic gradient descent algorithm with complex loss function. Opt. Express 29, 15089–15103 (2021).
Kim, D. et al. Accommodative holography: improving accommodation response for perceptually realistic holographic displays. ACM Transac. Gr. 41, 111 (2022).
Lee, D. et al. Expanding energy envelope in holographic display via mutually coherent multi-directional illumination. Sci. Rep. 12, 6649 (2022).
Gopakumar, M. et al. Unfiltered holography: optimizing high diffraction orders without optical filtering for compact holographic displays. Opt. Lett. 46, 5822–5825 (2021).
Choi, S. et al. Optimizing image quality for holographic near-eye displays with Michelson Holography. Optica 8, 143–146 (2021).
Chen, L. Z., Zhu, R. Z. & Zhang, H. Speckle-free compact holographic near-eye display using camera-in-the-loop optimization with phase constraint. Opt. Express 30, 46649–46665 (2022).
Sui, X. et al. Polarimetric calibrated robust dual-SLM complex-amplitude computer-generated holography. Opt. Lett. 48, 3625–3628 (2023).
Chen, C. et al. Off-axis camera-in-the-loop optimization with noise reduction strategy for high-quality hologram generation. Opt. Lett. 47, 790–793 (2022).
Horisaki, R. et al. Three-dimensional deeply generated holography [Invited]. Appl. Opt. 60, A323–A328 (2021).
Liu, C. et al. Intelligent coding metasurface holograms by physics-assisted unsupervised generative adversarial network. Photonics Res. 9, B159–B167 (2021).
Khan, A. et al. GAN-Holo: generative adversarial networks-based generated holography using deep learning. Complexity 2021, 6662161 (2021).
Wu, J. C. et al. High-speed computer-generated holography using an autoencoder-based deep neural network. Opt. Lett. 46, 2908–2911 (2021).
Hossein Eybposh, M. et al. DeepCGH: 3D computer-generated holography using deep learning. Opt. Express 28, 26636–26650 (2020).
Shui, X. H. et al. Diffraction model-informed neural network for unsupervised layer-based computer-generated holography. Opt. Express 30, 44814–44826 (2022).
Chakravarthula, P. et al. Learned hardware-in-the-loop phase retrieval for holographic near-eye displays. ACM Transac. Gr. 39, 186 (2020).
Choi, S. et al. Neural 3D holography: learning accurate wave propagation models for 3D holographic virtual and augmented reality displays. ACM Transac. Gr. 40, 240 (2021).
Choi, S. et al. Time-multiplexed neural holography: a flexible framework for holographic near-eye displays with fast heavily-quantized spatial light modulators. In Proc. ACM SIGGRAPH 2022 Conference Proceedings. Vancouver, BC, Canada: ACM, 32, 2022.
Xia, X. X. et al. Investigating learning-empowered hologram generation for holographic displays with ill-tuned hardware. Opt. Lett. 48, 1478–1481 (2023).
Boyd, S. et al. Distributed optimization and statistical learning via the alternating direction method of multipliers. Foundations Trends® Machine Learn. 3, 1–122 (2011).
Deng, W., Yin, W. T. & Zhang, Y. Group sparse optimization by alternating direction method. In Proc. SPIE 8858, Wavelets and Sparsity XV. San Diego, California, United States: SPIE, 242-256, 2013.
Huang, Z. Z. et al. Aberration-free synthetic aperture phase microscopy based on alternating direction method. Opt. Lasers Eng. 160, 107301 (2023).
Dennis, J. E. Jr. & Schnabel, R. B. Numerical Methods for Unconstrained Optimization and Nonlinear Equations. (SIAM: Philadelphia, 1996).
Nocedal, J. Theory of algorithms for unconstrained optimization. Acta Numerica 1, 199–242 (1992).
Liu, D. C. & Nocedal, J. On the limited memory BFGS method for large scale optimization. Mathe. Programm. 45, 503–528 (1989).
Nocedal, J. Updating quasi-Newton matrices with limited storage. Mathe. Comput. 35, 773–782 (1980).
Sui, X. et al. Spectral-envelope modulated double-phase method for computer-generated holography. Opt. Express 30, 30552–30563 (2022).
Tsang, P. W. M., Poon, T. C. & Wu, Y. M. Review of fast methods for point-based computer-generated holography. Photonics Res. 6, 837–846 (2018).
Matsushima, K. & Nakahara, S. Extremely high-definition full-parallax computer-generated hologram created by the polygon-based method. Appl. Opt. 48, H54–H63 (2009).
Bayraktar, M. & Özcan, M. Method to calculate the far field of three-dimensional objects for computer-generated holography. Appl. Opt. 49, 4647–4654 (2010).
Zhao, Y. et al. Accurate calculation of computer-generated holograms using angular-spectrum layer-oriented method. Opt. Express 23, 25440–25449 (2015).
Jia, J., Si, J. & Chu, D. P. Fast two-step layer-based method for computer generated hologram using sub-sparse 2D fast Fourier transform. Opt. Express 26, 17487–17497 (2018).
McCutchen, C. W. Generalized aperture and the three-dimensional diffraction image. J. Opt. Soc. Am. 54, 240–244 (1964).
Piestun, R., Spektor, B. & Shamir, J. Unconventional light distributions in three-dimensional domains. J. Modern Opt. 43, 1495–1507 (1996).
Liu, J. P., Wu, M. H. & Tsang, P. W. M. 3D display by binary computer-generated holograms with localized random down-sampling and adaptive intensity accumulation. Opt. Express 28, 24526–24537 (2020).
Sun, P. et al. Computer-generated holographic near-eye display system based on LCoS phase only modulator. In Proc. SPIE 10396, Applications of Digital Image Processing XL. San Diego, California, United States: SPIE, 294-300, 2017.
Sun, P. et al. Holographic near-eye display system based on double-convergence light Gerchberg-Saxton algorithm. Opt. Express 26, 10140–10151 (2018).
Pang, Y. F. et al. Simple encoding method of phase-only hologram for low crosstalk full-color multi-plane holographic projection. Opt. Lasers Eng. 147, 106748 (2021).
Chang, C. L., Xia, J. & Jiang, Y. Q. Holographic image projection on tilted planes by phase-only computer generated hologram using fractional Fourier transformation. J. Display Technol. 10, 107–113 (2014).
Hu, Y. L. et al. An improved method for computer generation of three-dimensional digital holography. J. Opt. 15, 125704 (2013).
Piestun, R. & Shamir, J. Synthesis of three-dimensional light fields and applications. Proc. IEEE 90, 222–244 (2002).
Piestun, R., Spektor, B. & Shamir, J. Wave fields in three dimensions: analysis and synthesis. J. Opt. Soc. Am. A 13, 1837–1848 (1996).
Stepien, P. J., Gajda, R. & Szoplik, T. Distributed kinoforms in optical security applications. Opt. Eng. 35, 2453–2458 (1996).
Kotlyar, V. V., Khonina, S. N. & Soifer, V. A. Iterative calculation of diffractive optical elements focusing into a three-dimensional domain and onto the surface of the body of rotation. J. Modern Opt. 43, 1509–1524 (1996).
Levy, U. et al. Iterative algorithm for determining optimal beam profiles in a three-dimensional space. Appl. Opt. 38, 6732–6736 (1999).
Shamir, J., Piestun, R. & Spektor, B. 3D light structuring and some applications. In Proc. SPIE 3729, Selected Papers from International Conference on Optics and Optoelectronics’ 98 (SPIE). Dehradun, India: SPIE, 222-228, 1999.
Zhou, P. C. et al. Dynamic compensatory Gerchberg-Saxton algorithm for multiple-plane reconstruction in holographic displays. Opt. Express 27, 8958–8967 (2019).
Haist, T., Schönleber, M. & Tiziani, H. J. Computer-generated holograms from 3D-objects written on twisted-nematic liquid crystal displays. Opt. Commun. 140, 299–308 (1997).
Velez-Zea, A. & Torroba, R. Mixed constraint in global and sequential hologram generation. Appl. Opt. 60, 1888–1895 (2021).
Velez-Zea, A. et al. Alternative constraints for improved multiplane hologram generation. Appl. Opt. 61, B8–B16 (2022).
Velez-Zea, A., Barrera-Ramírez, J. F. & Torroba, R. Improved phase hologram generation of multiple 3D objects. App. Opt. 61, 3230–3239 (2022).
Velez-Zea, A. & Torroba, R. Noniterative multiplane holographic projection. Appl. Opt. 59, 4377–4384 (2020).
Curtis, V. R. et al. DCGH: dynamic computer generated holography for speckle-free, high fidelity 3D displays. In Proceedings of the 2021 IEEE Virtual Reality and 3D User Interfaces (VR). Lisboa, Portugal: IEEE, 1-9, 2021.
Shabtay, G. et al. Optimal synthesis of three-dimensional complex amplitude distributions. Opt. Lett. 25, 363–365 (2000).
Dorsch, R. G., Lohmann, A. W. & Sinzinger, S. Fresnel ping-pong algorithm for two-plane computer-generated hologram display. Appl. Opt. 33, 869–875 (1994).
Makowski, M. et al. Three-plane phase-only computer hologram generated with iterative Fresnel algorithm. Opt. Eng. 44, 125805 (2005).
Makowski, M. et al. Iterative design of multiplane holograms: experiments and applications. Opt. Eng. 46, 045802 (2007).
Zhou, P. C. et al. Image quality enhancement and computation acceleration of 3D holographic display using a symmetrical 3D GS algorithm. App. Opt. 53, G209–G213 (2014).
Shi, L., Ryu, D. H. & Matusik, W. Ergonomic-centric holography: optimizing realism, immersion, and comfort for holographic display. Laser Photonics Rev 18, 2300651 (2024).
Shi, L. et al. Near-eye light field holographic rendering with spherical waves for wide field of view interactive 3D computer graphics. ACM Transac. Gr. 36, 236 (2017).
Li, J., Smithwick, Q. & Chu, D. P. Full bandwidth dynamic coarse integral holographic displays with large field of view using a large resonant scanner and a galvanometer scanner. Opt. Express 26, 17459–17476 (2018).
Chen, J. S., Smithwick, Q. Y. J. & Chu, D. P. Coarse integral holography approach for real 3D color video displays. Opt. Express 24, 6705–6718 (2016).
Park, J. H. & Askari, M. Non-hogel-based computer generated hologram from light field using complex field recovery technique from Wigner distribution function. Opt. Express 27, 2562–2574 (2019).
Lu, T. X. et al. Pixel-level fringing-effect model to describe the phase profile and diffraction efficiency of a liquid crystal on silicon device. Appl. Opt. 54, 5903–5910 (2015).
Yang, H. N. & Chu, D. P. Phase flicker in liquid crystal on silicon devices. J. Phys. Photonics 2, 032001 (2020).
Oliver, B. M. Sparkling spots and random diffraction. Proc. IEEE 51, 220–221 (1963).
Jones, R. & Wykes, C. Holographic and Speckle Interferometry. 2nd edn. (Cambridge: Cambridge University Press, 1989).
Goodman, J. W. Speckle Phenomena in Optics: Theory and Applications. (Englewood: Roberts and Company Publishers, 2007).
Nye, J. F. Natural Focusing and Fine Structure of Light: Caustics and Wave Dislocations. (Bristol: Institute of Physics Publishing, 1999).
Senthilkumaran, P., Wyrowski, F. & Schimmel, H. Vortex stagnation problem in iterative Fourier transform algorithms. Opt. Lasers Eng. 43, 43–56 (2005).
Lee, W. H. Sampled Fourier transform hologram generated by computer. Appl. Opt. 9, 639–643 (1970).
Maimone, A. & Wang, J. R. Holographic optics for thin and lightweight virtual reality. ACM Transac. Gr. 39, 67 (2020).
Shi, L. et al. Towards real-time photorealistic 3D holography with deep neural networks. Nature 591, 234–239 (2021).
Wu, Y. et al. Adaptive weighted Gerchberg-Saxton algorithm for generation of phase-only hologram with artifacts suppression. Opt. Express 29, 1412–1427 (2021).
Tian, S. Z., Chen, L. Z. & Zhang, H. Optimized Fresnel phase hologram for ringing artifacts removal in lensless holographic projection. Appl. Opt. 61, B17–B24 (2022).
Pang, H. et al. Speckle-reduced holographic beam shaping with modified Gerchberg–Saxton algorithm. Opt. Commun. 433, 44–51 (2019).
Yoo, D. et al. Optimization of computer-generated holograms featuring phase randomness control. Opt. Lett. 46, 4769–4772 (2021).
Yang, D. et al. Diffraction-engineered holography: beyond the depth representation limit of holographic displays. Nat. Commun. 13, 6012 (2022).
He, Z. H. et al. Frequency-based optimized random phase for computer-generated holographic display. Appl. Opt. 60, A145–A154 (2021).
Ren, H. R. et al. Three-dimensional vectorial holography based on machine learning inverse design. Sci. Adv. 6, eaaz4261 (2020).
Wyrowski, F. & Bryngdahl, O. Digital holography as part of diffractive optics. Rep. Progress Phys. 54, 1481–1571 (1991).
Song, Q. H. et al. Vectorial metasurface holography. Appl. Phys. Rev. 9, 011311 (2022).
Zhao, R. Z., Huang, L. L. & Wang, Y. T. Recent advances in multi-dimensional metasurfaces holographic technologies. PhotoniX 1, 20 (2020).
Jiang, Q., Jin, G. F. & Cao, L. C. When metasurface meets hologram: principle and advances. Adv. Opt. Photonics 11, 518–576 (2019).
Gao, H. et al. Recent advances in optical dynamic meta-holography. Opto-Electronic Adv. 4, 210030 (2021).
https://github.com/THUHoloLab/Optimization_algorithms_for_CGH.
Acknowledgements
This work is supported by the National Science Foundation of China (62035003) and the Tsinghua University Initiative Scientific Research Program (20193080075) as well as the Cambridge Tsinghua Joint Research Initiative.
Author information
Authors and Affiliations
Corresponding authors
Ethics declarations
Conflict of interest
The authors declare no competing interests.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Sui, X., He, Z., Chu, D. et al. Non-convex optimization for inverse problem solving in computer-generated holography. Light Sci Appl 13, 158 (2024). https://doi.org/10.1038/s41377-024-01446-w
Received:
Revised:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41377-024-01446-w
This article is cited by
-
Rapid in-air ultrasound holography measurement and camera-in-the-loop generation using thermography
Communications Engineering (2025)
-
High contrast holography through dual modulation
Scientific Reports (2025)
-
Real-time holographic camera for obtaining real 3D scene hologram
Light: Science & Applications (2025)
-
Development of a universal method for quantization of computer-generated holograms in the optical image reconstruction
Measurement Techniques (2025)