Open Access
9 December 2023 Wavefront shaping improves the transparency of the scattering media: a review
Chunxu Ding, Rongjun Shao, Qiaozhi He, Lei S. Li, Jiamiao Yang
Author Affiliations +
Abstract

Significance

Wavefront shaping (WFS) can compensate for distortions by optimizing the wavefront of the input light or reversing the transmission matrix of the media. It is a promising field of research. A thorough understanding of principles and developments of WFS is important for optical research.

Aim

To provide insight into WFS for researchers who deal with scattering in biomedicine, imaging, and optical communication, our study summarizes the basic principles and methods of WFS and reviews recent progress.

Approach

The basic principles, methods of WFS, and the latest applications of WFS in focusing, imaging, and multimode fiber (MMF) endoscopy are described. The practical challenges and prospects of future development are also discussed.

Results

Data-driven learning-based methods are opening up new possibilities for WFS. High-resolution imaging through MMFs can support small-diameter endoscopy in the future.

Conclusion

The rapid development of WFS over the past decade has shown that the best solution is not to avoid scattering but to find ways to correct it or even use it. WFS with faster speed, more optical modes, and more modulation degrees of freedom will continue to drive exciting developments in various fields.

1.

Introduction

Light can carry and process information in parallel in multiple degrees of freedom, such as space, time, wavelength, amplitude, phase, and polarization. Therefore, it has been an essential tool for information observation, transmission, and compilation in the last several decades. Optical techniques are indispensable, from tracking the trajectory of stars in astronomy, to observing the microscopic structure of cells in biomedicine. But if light propagates in a scattering environment, such as disordered materials, biological tissues, and multimode fibers (MMFs), the inhomogeneous distributions of refractive indices will add random distortions to the carried information.13 This phenomenon significantly deteriorates the performance of traditional optical techniques. Biomedicine is particularly affected. High-resolution imaging and high-precision laser therapy often rely on ballistic photons or quasiballistic photons46 because they can achieve coherent and tight focusing. However, scattering in biological tissues causes the number of ballistic photons to decay exponentially with propagation depth, limiting optical focusing and direct imaging to a depth of 1 to 2 mm.7 At the depths of more than one optical transport mean-free path, ballistic photons are rare, and fully developed speckles become dominant. Thus for a long time, overcoming scattering to make the turbid medium “transparent” has been urgently needed.

The phenomenon of scattering events exhibiting determinism with the speckle decorrelation window is called the memory effect.810 Although the scattering transmission of ballistic photons is highly complex, there is no irreversible loss of the carried information. If the scrambled information in the scattered light can be successfully retrieved and rearranged, the “transparency” of the medium may be achieved. Based on this, Vellekoop and Mosk developed wavefront shaping (WFS).11 The phase control pattern of the spatial light modulator is optimized by an iterative algorithm, and the precompensation information is introduced into the input light.1215 A focused spot can be successfully obtained behind the scattering medium. A transmission matrix (TM) model describing the linear relationship between the input and output light has also been established.1619 By measuring the TM, the scattering distortion can be precompensated to obtain the desired output. In addition, time-reversing the scattered field can also restore the speckle to the original incident wavefront.2023 In the past decade, thanks to the development of highly adjustable spatial light modulators, such as liquid crystal,11 microelectromechanical,24 and acousto-optic modulators,25 especially the increase in the number of controllable modes and modulation rates, a large amount of research on WFS has emerged.

WFS has provided a powerful tool for biomedicine. By manipulating large-scale optical modes with high resolution, WFS allows the construction of diffraction-limited spatiotemporal focusing behind opaque biological tissues.26 Furthermore, guidance with ultrasound, which produces a focus much deeper than light in the medium, can help to focus light inside turbid tissues without invasive manipulation.27 The high-resolution and high-contrast antiscattering focusing based on WFS increases the penetration depth of two-photon microscopy,28 optical coherence tomography,29 and fluorescence imaging,30 promising noninvasive diagnosis and treatment,31 and benefiting the development of phototherapy and optogenetics. MMF only has a diameter of a few hundred microns but can support tens of thousands of optical modes. But the inherent mode crosstalk makes it impossible to directly image but produces speckles similar to multiple scattering. WFS can compensate for the crosstalk, thereby subtly converting MMFs into endoscopic fibers for minimally invasive imaging of living cells and tissues. Thus WFS accelerates the deployment of MMFs to endoscopic imaging.32

In this review, we introduce the principles of WFS, including the concept of light scattering and three basic methods. Then we review the latest developments in the field of focusing, imaging, and MMF. Finally, we discuss the current challenges and prospects of WFS.

2.

Basic Principles of WFS

2.1.

Multiple Scattering

Affected by the inhomogeneity of the microscopic refractive indices of the medium, photons on different paths experience different phase delays, and the transmission directions are deflected to different degrees [Fig. 1(a)]. Therefore, ballistic photons decrease sharply during propagation, and the scattered light of different paths interferes randomly, forming a speckle pattern with chaotic intensity distribution.1 Generally, the absorption coefficient μa is used to describe a fractional decrease in the light intensity per unit distance traveled due to absorption, and the scattering mean free path of scattering ls is used to represent the average distance between two adjacent scattering events.33 ls quantifies the scattering strength of the scattering medium. After the distance known as the transport mean free path l=ls/(1g), the propagation path of the scattered light becomes sufficiently random, completely losing correlation with the original trajectory.2 g is the anisotropy factor of the scattering medium and represents the average value of the cosine of the light deflection angle θ during the scattering events. The anisotropy factor of biological tissue is typically 0.9,31 which means that the forward scattering is dominant. The transport mean free path is typically 0.1 to 1 mm, and the observed speckle patterns are basically the result of multiple scattering.

Fig. 1

Principles of WFS: (a) propagation of light in a scattering medium. (b) Visualization of TM. The input and output optical fields are expanded as vectors. The field Ekout at the k’th output mode is linked to the input Enin by a column tmn of TM. (c) The optimization method. The output field of the scattering medium is used as the feedback to optimize the modulation pattern of the spatial light modulator by the algorithm, and the desired output is finally obtained. (d) OPC. Time-reverse scattered field and recover the input field.

JBO_29_S1_S11507_f001.png

Theoretically, the transport of light can be described by the wave equation, as can multiple scattering. However, the complexity of multiple scattering makes it nearly impossible to obtain a complete numerical solution. However, there are correlations between multiple scatterings. The same input can result in an unchanged speckle response. When the input changes slightly in tilt, the speckle pattern maintains a high degree of similarity, i.e., memory effect.34 It should be noted that the memory effect is valid within the speckle correlation time.1 Once there is a disturbance, the scattering transformation will quickly decorrelate. The transformation of the light field before and after scattering is deterministic. If the optimal inputs can be constructed according to the transformation, it will be possible to obtain the desired output through scattering.

Mesoscopic transport35 and random matrix theory36 provide a solution for the quantitative analysis of multiple scattering. In the spatial domain, the input and output fields of the scattering medium can be represented by Enin and Emout, respectively (with n=1,,N and m=1,,M). n and m are the optical channels or modes. For a system with a surface area A, theoretically, the maximum number of spatial modes allowed for the input and output is Nm2πA/λ2, where λ is the wavelength of light.37 In practice, N and M depend on the discrete resolution and the finite field. The transformation between input and output can be described by a linear complex TM T, i.e., Emout=tmnEnin or Eout=TEin. The TM indicates that a channel of the scattered output is the weighted sum of all input channels.16 When the input is a plane wave, the output depends on the entire matrix. Since the results of multiple scattering are fully developed speckles, where channels or modes are highly uncorrelated, each element of the TM is close to a random variable, generally obeying a circularly symmetric Gaussian distribution. The TM expression can also be modified to represent a reflective matrix not limited to the transmissive forward scattering.38

2.2.

WFS via TM Engineering

The TM describes the linear relationship between the light before and after scattering [Fig. 1(b)]. If each element of the matrix can be accurately obtained, all the scattering events can be analyzed, and any scattering output can be achieved using a specially designed input. Therefore, measuring the TM to compensate for the scattering distortion is the focus of WFS. At present, there are two major methods for measuring the TM of scattering media, namely the interferometric method3942 and the noninterferometric method.4346 The former is first demonstrated by Popoff et al.16 They loaded the orthonormal basis as the input to the spatial light modulator sequentially. The camera detected the interference between the speckle output of each basis and the additional reference beam. The complex amplitude information of the detection was analyzed by the phase-shifting method, and finally the TM was obtained by specific calculation.16 In Popoff’s method, since the reference beam and the signal beam came from the same path, it was also called self-reference measurement. Additionally, introducing an external reference for phase shifting or holographic measurement can also be used for the measurement. The noninterferometric method is first demonstrated by Drémeau et al., mainly relying on phase retrieval algorithms to infer the elements of the TM from a large number of input patterns and corresponding speckle outputs.47 The phase retrieval algorithms are mostly driven by Bayes’ theorem,43 and the measurement of TM is completed with many iterations. Due to the antinoise performance, the noninterferometric method is currently a popular research.

Once measured, the TM T can be used to analyze and control the light propagating through the scattering medium. By analyzing the eigenvectors of TT, the transmission eigenchannel, complete control of the light transmission in the scattering medium can be achieved.48 By inverting the TM via Ein=T1Eout, the input wavefront that should be modulated can be calculated according to the target output, or the hidden input can be recovered from the output speckle.49,50 In practice, the inverse of the matrix is often replaced by the conjugate transpose T or some other operators.51 Although not exactly equivalent, it is more stable than direct inversion. Because the goal of antiscattering focusing is to form a single focus at one fixed target location, it is not necessary to measure the entire TM. The optimal input wavefront is only related to a certain column of the TM. Therefore, the measurement error tolerance is high. However, recovering the input image from the speckle, or constructing a specific output requires more accurate measurement, because almost all elements of the TM are used in the calculation. But due to various issues, such as the limited number of optical modes and unstable interferometry, the implementation of these complex tasks is quite difficult and often requires additional help from iterative or learning-based methods.52,53

In order to control more degrees of freedom of the scattered light, a lot of expansions of the complex amplitude TM method have been developed. For example, the control of polarization allows the construction of vector beams for optical tweezers,54 increasing the information capacity of MMFs.55 Based on the polarization coupling effect of scattering, and the fact that the scattered vector field is the summation of the intensities of two fully developed speckles in two orthogonal polarization states,51 researchers extend the linear transformation relationship to a two by two dimension,56 which means a pair of inputs in two orthogonal polarization states becomes a pair of scattered outputs. In addition to the spatial degrees of freedom, spectral and time resolved transmission matrices can also be obtained using monochromatic measurement at multiple frequencies,57,58 or low-coherence measurement in a broad spectrum of light.59,60 Then frequency-dependent focusing and spatiotemporal focusing can be achieved.

2.3.

Iteration-Based WFS

Iteration-based WFS uses the field distribution at the target position as the feedback to search and optimize the modulation pattern of the spatial light modulator with iterative algorithms [Fig. 1(c)]. It can gradually optimize the incident light and guide the scattered light to generate the target field through the scattering medium.11,61 This method was the first demonstration of WFS. In the implementation of iteration-based WFS, taking focusing as an example, detection devices, such as cameras or photomultiplier tubes, are used to monitor and detect the light intensity of the target focusing position. The iterative algorithm takes the detected light intensity as feedback and completes the optimization.6165 Finally, the one with the highest feedback is determined as the best modulation pattern, and a highly enhanced antiscattering focusing can be obtained.

The iteration-based method can be used to overcome the diffraction limit of a conventional imaging system and generate enhanced focusing,66,67 without complex interference techniques. For iteration-based WFS, imaging a large field-of-view often requires focus scanning.1,68 It generally relies on the memory effect. However, the effective range of memory effect is small, which leads to a limited imaging field of view and depth. Constructing a movable guide star, such as ultrasound,69 can achieve deep imaging without the aid of memory effects, but the focal spot is several orders of magnitude larger than the optical diffraction limited focus.66,70 The imaging resolution and contrast are typically insufficient. Recently, a model-based method has been proposed to achieve the focusing at any spatial location by calculating the refractive index model of the scattering medium.71 But this method requires strong prior knowledge, and its practicability needs further exploration.

The exploration of advanced noninvasive feedback is another important direction for iteration-based methods.27 In practice, cameras cannot be used to provide feedback inside the scattering medium. However, there is an urgent need to achieve optical focusing and imaging inside biological tissues. Therefore, internal guide stars have been proposed, such as fluorescence markers,72 photoswitchable proteins,73 magnetic particles,74 and microbubbles.75 The fluorescent marker was employed in the medium as the earliest guide star.72 The fluorescence signal outside the medium is detected as feedback to optimize the focused spot. In addition, people have proposed several noninvasive guide stars, such as coherence gating76 and particle displacement.77 Recently, acousto-optic feedback methods have been extensively studied,66,69,7880 as ultrasound can propagate deep in biological tissue. In addition, the acoustic signal generated via the photoacoustic effect inside the scattering medium as the feedback was also explored.29,81 Moreover, people can also focus ultrasound in the medium to induce acousto-optic modulation to shift the light frequency and detect the shift as the feedback to achieve noninvasive focusing.69

2.4.

Optical Phase Conjugation

The analytical solution of the wave equation has time-reversal symmetry.37 It means that if the phase conjugate field of the scattered field propagates back to the scattering medium, the initial incident field can be restored [Fig. 1(d)]. When the incident light comes from a specific point source, the time-reversed light is refocused at that location. This method based on time-reversal is called optical phase conjugation (OPC). OPC was first demonstrated in analog devices, such as acoustic time-reversal mirrors82,83 and optical nonlinear crystal.84,85 Digital methods of interference recording and spatial light modulation can be alternatives to analog methods to achieve optical focusing through turbid media.20

Compared with the TM and iteration-based WFS, the wavefront information to be modulated can be determined by OPC with only one measurement, significantly reducing the time consumption. It can achieve a fast response of millisecond or submillisecond level.86 But OPC also introduces a system alignment challenge. This challenge stems from two major implementation steps of OPC: recording of scattered optical fields, and playback of phase-conjugate optical fields. In a typical implementation, the camera records the interference patterns of the scattered light and the reference light, and the complex amplitude of the scattered light is calculated. Then the spatial light modulator displays the corresponding pattern to generate the phase-conjugate field and trace it back to the scattering medium. In this process, the pixels of the recording camera and the playback modulator need to be matched one by one, so that the time-reversal can be accurately achieved,23 adding complexity to the implementation of OPC. One possible solution is to design a unique device with both detection and modulation functions.1 In addition, to achieve the focus inside the scattering medium, the guide stars, such as ultrasound, are also required.73,76,87

3.

Application of WFS

3.1.

Focusing and Imaging

Clinical imaging technologies, such as ultrasound imaging29 and magnetic resonance imaging,1 can penetrate biological tissues to obtain macroscopic images. But their spatial resolution is orders of magnitude lower than optical microscopy. In general, the scattering properties in biological tissues are wavelength-dependent.88 As the wavelength increases, the transport mean free path becomes longer. Near-infrared light can reach a penetration depth of several millimeters to centimeters.89 But conventional optical techniques using visible or near-infrared light, such as multiphoton microscopy,90 still rely on ballistic photons, limiting the imaging depth to 1 to 2 mm. Using scattered light rather than the almost nonexistent ballistic light is a better choice to solve the problem in biological tissues. The rapidly developed WFS makes it possible to achieve focusing and imaging through thick scattering media and even in living tissues.

Fast and highly enhanced focusing inside scattering media has been the focus of research. In general, the energy gain of light focusing is affected by the number of effective modulation modes. However, increasing the number of modes reduces the speed of focusing. This trade-off is particularly prominent in TM and iteration-based methods.1 Digital micromirror devices (DMDs) with high refresh rates are used to replace the traditional liquid crystal modulator, greatly improving the focusing speed.18 However, DMDs can only control the binary amplitude of light, then the focusing quality is reduced. A variety of binary coding methods have been proposed, such as Lee holographic method91 and superpixel method,92,93 to improve the modulation dimension of DMD to the complex amplitude. But these methods sacrifice the number of modes. In addition, polarization has also been taken into account, and WFS is carried out in two orthogonal polarization directions.23,94,95 On the one hand, the number of modes was cleverly increased by two times. On the other hand, it enables WFS to control polarization, another important degree of freedom of light. Recently, a WFS technique that simultaneously achieves high-speed, high-energy gain, and high-modulation mode number has been proposed96 [Fig. 2(a)]. By combining photorefractive crystal-based analog OPC and stimulated emission amplification, the focusing energy gain is increased by three orders of magnitude, and the speed is improved by more than 50 times. Light focusing in a living mouse ear has been achieved. It promises optical focusing for biomedical applications.

Fig. 2

Fast and high-enhanced optical focusing. (a) Fast OPC: (a1) forward amplification and recording process based on stimulated emission light amplification; (a2) playback process based on time-reversal; and (a3) optical focusing results in a living-mouse ear. (b) Ultrasound-guided in vivo optical focusing: (b1) ultrasonic guide star switch off; (b2) ultrasonic guide star switch on; (b3) playback the differential field to achieve in vivo focusing; (b4) energy profile of focusing; and (b5) experimental focusing results. Panels (a1)–(a3) are from Ref. 96. Panels (b1)–(b5) are from Ref. 80.

JBO_29_S1_S11507_f002.png

Focusing light in the scattering medium, such as biological tissue, requires guide stars in vivo.27 Since the scattering effect of ultrasound is about 1000 times weaker than that of light, a new concept of optical focusing based on ultrasonic encoding was first proposed and applied to WFS.69 It is noninvasive and moveable. Because this method uses the focus of ultrasound in the tissue to guide the focus of coherent light, the resolution of optical focus is limited by the ultrasonic focus.42 In order to improve the resolution, some methods, including variance coding time inversion97 and nonlinear ultrasonic guided iterative optimization66 have been proposed. These methods break the ultrasonic resolution limit and realize the optical diffraction-limited focusing. Since only a small amount of frequency-shifted first-order diffracted photons are useful for focusing, it reduces the focusing efficiency and peak-to-background ratio. Recently, it has been found that the brighter zeroth-order diffracted light, which was previously considered useless or even harmful, has a field perturbation effect80 [Fig. 2(b)]. When switching the ultrasonic guide star on and off alternately, the differential field can be detected, thereby guiding the light to focus to the location of the field disturbance, improving the efficiency and flexibility of light focusing.

The most straightforward method of translating focusing to imaging is raster scanning, which can be combined with fluorescent labeling to realize the imaging of microstructure by scanning the focus.98 It can also replace the time-gated method by modulating the light to compensate for the scattering of the wall and form a focus to scan hidden objects, to achieve non-line-of-sight imaging99 [Fig. 3(a)]. Because the focus of WFS can be close to the diffraction limit, the imaging resolution can be significantly improved. In addition, the memory effect makes it possible to achieve imaging without scanning focus100 [Fig. 3(b)]. This method can reconstruct the hidden objects in the scattering environment using the speckle correlation calculation method without complex optimization of the wavefront. However, it should be noted that such methods only can image small objects in the narrow-angle range where the memory effect takes effect.

Fig. 3

Optical imaging of WFS. (a) Non-line-of-sight imaging based on WFS: (a1) principle and (a2) imaging results and the scanning focus. (b) Fluorescence imaging based on memory effect: (b1) fluorescence images of different pollen grains without scattering and (b2) imaging results with scattering. (c) Image-guided WFS: (c1) principle and (c2) imaging results. (d) Learning-based WFS: (d1) principle and (d2) imaging results of dynamic objects under dynamic scattering aberrations. Panels (a1) and (a2) are from Ref. 99. Panels (b1) and (b2) are from Ref. 100. Panels (c1) and (c2) are from Ref. 101. Panels (d1) and (d2) are from Ref. 102.

JBO_29_S1_S11507_f003.png

In recent years, the development of deep learning has made a breakthrough in WFS.103,104 An image-guided WFS method was developed101 [Fig. 3(c)], which iteratively optimizes the phase of the spatial light modulator with the goal of maximizing the image quality metric (improved combination of entropy and variance). Diffractive-limited, high-contrast images of hidden objects from the very low-contrast speckle pattern can be recovered, even with the object outside the field of view of the memory effect. In addition, introducing neural networks to calculate complex optical aberrations from scattering media is also a promising direction102 [Fig. 3(d)]. Through backpropagation without guide stars, the antiscattering imaging of a wide field of view of dynamic scenes at diffraction-limited resolution can also be achieved.

3.2.

WFS in MMF

Correcting aberrations inside MMF has also been studied via WFS. In general, an MMF has a diameter of only a few hundred microns but can support tens of thousands of optical modes. This advantage makes it an ideal device for endoscopic imaging. It also holds great potential to replace the coarse glass fiber bundles105 currently used for medical purposes. However, due to the mode coupling, meaning different modes occur in different phase delays, MMF shows the characteristics of the scattering medium. The light through MMF cannot be directly used to form an image but produces a disordered speckle. The conversion of MMF to endoscope imaging fiber using WFS has been widely studied.

For the linear mode coupling process in MMF, the optical input and output can still be connected by the TM. Therefore, WFS is effective in MMF. In fact, methods, such as OPC,28 TM,106 and iterative optimization,107 have been applied in MMF-based endoscopy imaging, and in vivo imaging of animal neurons108 and blood cells109 has also been realized. However, these methods require MMF not to be deformed or bent, which limits applications in medical diagnosis. A spatial-frequency tracking adaptive beacon light-field-encoded endoscope has been proposed recently110 [Fig. 4(a)]. The special beacon intensity that reflects the correlation between the current bending states and the precalibrated TM is detected by the single-pixel detector. Adaptive tracking is completed by searching the TM that maximizes the intensity from the premeasured dataset. This method can improve the resolution of endoscopic imaging to subdiffraction-limited value with high stability and robustness against the movement and deformation of the fiber. In addition to endoscopic microstructure imaging, MMF has also been used in recent years for far-field imaging of three-dimensional (3D) real scenes111,112 [Fig. 4(b)]. Using WFS to correct the aberration of MMF, combined with the time-of-flight method, high-speed 3D imaging can be achieved, benefiting motion detection and tracking for industrial applications.

Fig. 4

Imaging through an MMF. (a) Adaptive MMF endoscope: (a1) principle and (a2) endoscopic imaging results. (b) Three-dimensional far-field imaging through an MMF: (b1) principle; (b2) imaging target; and (b3) imaging results. Panels (a1) and (a2) are from Ref. 110. Panels (b1)–(b3) are from Ref. 111.

JBO_29_S1_S11507_f004.png

However, there are nonlinear effects32 between the speckle output and the input of MMF, which greatly reduce the effectiveness of the TM based on the linear premise. Deep neural networks have been shown to be useful for learning nonlinear processes in MMF.104 By training a conventional convolutional neural network,113 the input image can be recovered from the speckle without the unstable TM measurement [Fig. 5(a)]. However, the generalization of the neural network is limited, and it is usually only effective for the images highly similar to the training data set. Thus the performance is limited for the general purpose of image transmission. A method to reconstruct the complex inversion matrix of MMF by statistical means successfully realizes the transmission and imaging of full-color natural images with high frame rate and high resolution, without data-based learning114 [Fig. 5(b)]. Because this approach is based on a physically informed model of the imaging system that retrieves an approximation to TM, its generalization is better than that of the deep neural network approaches.

Fig. 5

Smart imaging of MMF. (a) Learning-based imaging: (a1) flowchart and (a2) imaging results. (b) Statistics-based imaging: (b1) principle and (b2) input image and reconstruction results. Panels (a1) and (a2) are from Ref. 113. Panels (b1) and (b2) are from Ref. 114.

JBO_29_S1_S11507_f005.png

The mode coupling of MMF occurs in various degrees of freedom.51 In principle, to achieve full control of MMF, the modulated light needs to be completely reversed in time. This means that the modulated light is a volume field that should have arbitrarily specified amplitude, phase, and polarization at every point in space and time. However, this kind of full degree of freedom controlled field has been difficult to produce in research.115,116 Recently, a time-reversal device that can independently and simultaneously measure and control all classical degrees of freedom has been invented117 (Fig. 6). This method can generate a series of arbitrary two-dimensional vector fields at a high rate and is used for spatiotemporal control of light in MMF, including imaging, nonlinear optical phenomena, and micromanipulation. The modulation of all degrees of freedom could be an important direction of WFS development.

Fig. 6

Multidegrees of freedom light control in MMF: (a) principle; (b) vector spectral light generation; and (c) optical field with space and polarization state changing with time. Panels (a)–(c) are from Ref. 117.

JBO_29_S1_S11507_f006.png

4.

Discussion and Outlook

So far, WFS has shown unique advantages and potential in solving scattering problems. In this paper, the basic concepts and general methods of WFS, as well as the contributions and developments in focusing, imaging, and optical fiber endoscopy, are reviewed. It is worth mentioning that although dealing with scattering in biomedicine is the current focus, the applications of WFS are not limited to this.118121 As mentioned in the introduction, in recent years, holographic display122124 and optical communication125127 have also been supported by the rapid development of WFS. However, at present, they are still in the experimental verification stage. For example, the holographic display using scattering can only generate a simple pattern composed of focused points, which is far from the commercial 3D digital holographic display.128

A key problem preventing WFS from practical applications is the trade-off between the speed and the number of modulation modes. The TM of the scattering medium changes with time, environmental disturbance, or bending deformation.129,130 In general, the decorrelation time of biological tissue is approximately milliseconds and decreases with the increase of tissue depth.131 In addition, for MMF, bending is basically unavoidable. Despite a lot of exploration, such as the development of modulators from liquid crystal modulators to DMDs with refresh rates of up to kilohertz,24,132 there is still a long way to go to achieve real-time correction of dynamic scattering. WFS that takes effect within the speckle correlation time will be the key to determining whether it can be applied.

In addition, the number of input and output optical modes of a scattering medium is huge, but the modes WFS can control are often limited or even insufficient. On the one hand, the number of input modes often depends on the number of effective modulation units, which is much smaller than the theoretical maximum number,37 which limits the performance of WFS. For example, the enhancement of antiscattering focusing is approximately linear with the number of modulation modes.66 On the other hand, the number of pixels of the camera is also limited, and the speckle imaging in the camera is a lowpass filtering process.32 Thus the measurement of all output modes is almost impossible. Currently, people often choose to trade the number of optical modes for the speed,1 because the increase of modes in most methods puts a significant burden on speed. WFS outside the speckle correction time window is no longer useful to correct the scattering.

Apart from the speed and mode numbers, the degree of freedom of control is also a key to WFS. Multiple scattering not only confuses light into speckles of randomly distributed intensity but also disturbs the information of almost all degrees of freedom of light.51 The ultimate ideal of WFS needs to control all the information of the light, including but not limited to spatiotemporal information in the scattering process, which will truly enable the “transparency” of the scattering medium. Currently, it is still a challenge to control the complex information of all the degrees of freedom independently for existing spatial light modulators. With the development of vector vortex beams,54 spectroscopy,133 and nonlinear optics,134 this challenge can potentially be overcome.

Data-driven learning-based methods bring new vitality to the development of WFS. Instead of relying on physical models, such as TM, a deep neural network model is used to learn compensating scattering aberrations from a large training data set and can achieve better performance. For example, it can learn how to correct dynamic scattering aberrations102 and the deformation of MMF.114 It can also enlarge the field of view limited by the memory effect, lower the requirement of scattering calibration, and improve the stability of the system.135 The relevant research is in the initial stage, and future engineering efforts are still needed.1 For example, training a neural network requires a large and diverse set of targeted data, resulting in expensive training and insufficient generalization. Learning-based imaging mostly stays in the two-dimensional plane, and how to translate it to depth-resolved volume imaging is still challenging. Meanwhile, since the training process requires direct access to the input and output planes, imaging objects inside the scattering medium presents another difficulty. Nevertheless, learning-based WFS is becoming the focus of research and is promising to bring about a new technological revolution in this field.

5.

Conclusion

Scattering is a fundamental problem in the field of optics. Overcoming the limitation of scattering and achieving optical “transparency” in turbid media is always an urgent goal. The rapid development of WFS over the past decade has shown that the best solution is not to avoid scattering but to find ways to correct it or even use it. So far, there are many well-established methods to compensate scattering to construct a bright focus136,137 or to transmit images.138,139

In this paper, the basic principles, methods of WFS, and the latest applications of WFS in focusing, imaging, and MMF endoscopy are described. In addition, the present challenges and outlooks of WFS are discussed. Currently, there are three major aspects for the development of WFS: faster to adapt to dynamic scattering,140 more modes to meet the needs of actual scenarios,96 and more degrees of freedom to be controlled simultaneously.95 Fortunately, the introduction of data-driven learning-based methods may open up a new possibility for WFS.141 Although WFS is still not mature enough for practical applications, the development of this field is very encouraging. Various exciting results show the great potential that WFS will drive major developments in biomedicine, imaging, optical communication, and other fields.

Disclosures

Lei S. Li has a financial interest in BLOCH Quantum Imaging Solutions, although they did not support this work. The other authors declare no competing financial interests.

Code and Data Availability

The code and data that support the findings of this study are available from the corresponding authors upon reasonable request.

Acknowledgments

The authors would like to acknowledge the funding from the National Natural Science Foundation of China (Grant Nos. 62375171 and 62205189), the Fundamental Research Funds for the Central Universities, Shanghai Pujiang Program (Grant No. 22PJ1407500), Shanghai Jiao Tong University 2030 Initiative (Grant No. WH510363001-10), the Oceanic Interdisciplinary Program of Shanghai Jiao Tong University (Grant No. SL2022ZD205), the Science Foundation of Donghai Laboratory (Grant No. DH-2022KF01001), and the internal fund from Rice University (Grant No. F10000205).

References

1. 

S. Gigan et al., “Roadmap on wavefront shaping and deep imaging in complex media,” J. Phys.: Photonics, 4 (4), 042501 https://doi.org/10.1088/2515-7647/ac76f9 (2022). Google Scholar

2. 

J. W. Goodman, Speckle Phenomena in Optics: Theory and Applications, Roberts and Company Publishers( (2007). Google Scholar

3. 

A. Ishimaru, Wave Propagation and Scattering in Random Media, 2 148 –166 Academic Press, New York (1978). Google Scholar

4. 

L. V. Wang and H. Wu, Biomedical Optics: Principles and Imaging, John Wiley & Sons( (2012). Google Scholar

5. 

A. Escobet-Montalbán et al., “Three-photon light-sheet fluorescence microscopy,” Opt. Lett., 43 5484 –5487 https://doi.org/10.1364/OL.43.005484 OPLEDP 0146-9592 (2018). Google Scholar

6. 

K. Maslov et al., “Optical-resolution photoacoustic microscopy for in vivo imaging of single capillaries,” Opt. Lett., 33 929 –93 https://doi.org/10.1364/OL.33.000929 OPLEDP 0146-9592 (2008). Google Scholar

7. 

V. Ntziachristos, “Going deeper than microscopy: the optical imaging frontier in biology,” Nat. Methods, 7 603 –614 https://doi.org/10.1038/nmeth.1483 1548-7091 (2010). Google Scholar

8. 

I. Freund, M. Rosenbluh and S. Feng, “Memory effects in propagation of optical waves through disordered media,” Phys. Rev. Lett., 61 (20), 2328 https://doi.org/10.1103/PhysRevLett.61.2328 PRLTAO 0031-9007 (1988). Google Scholar

9. 

S. Feng et al., “Correlations and fluctuations of coherent wave transmission through disordered media,” Phys. Rev. Lett., 61 (7), 834 https://doi.org/10.1103/PhysRevLett.61.834 PRLTAO 0031-9007 (1988). Google Scholar

10. 

G. Osnabrugge et al., “Generalized optical memory effect,” Optica, 4 (8), 886 –892 https://doi.org/10.1364/OPTICA.4.000886 (2017). Google Scholar

11. 

I. M. Vellekoop and A. P. Mosk, “Focusing coherent light through opaque strongly scattering media,” Opt. Lett., 32 (16), 2309 –2311 https://doi.org/10.1364/OL.32.002309 OPLEDP 0146-9592 (2007). Google Scholar

12. 

D. B. Conkey et al., “Genetic algorithm optimization for focusing through turbid media in noisy environments,” Opt. Express, 20 (5), 4840 –4849 https://doi.org/10.1364/OE.20.004840 OPEXFF 1094-4087 (2012). Google Scholar

13. 

H. Li et al., “Adaptive optical focusing through perturbed scattering media with a dynamic mutation algorithm,” Photonics Res., 9 (2), 202 –212 https://doi.org/10.1364/PRJ.412884 (2021). Google Scholar

14. 

L. Liu et al., “High-contrast light focusing through scattering media with multipixel encoding,” Appl. Phys. Express, 14 (9), 092009 https://doi.org/10.35848/1882-0786/ac200e APEPC4 1882-0778 (2021). Google Scholar

15. 

H. Jin et al., “Limiting the incident NA for efficient wavefront shaping through thin anisotropic scattering media,” Optica, 8 (4), 428 –437 https://doi.org/10.1364/OPTICA.413174 (2021). Google Scholar

16. 

S. M. Popoff et al., “Measuring the transmission matrix in optics: an approach to the study and control of light propagation in disordered media,” Phys. Rev. Lett., 104 (10), 100601 https://doi.org/10.1103/PhysRevLett.104.100601 PRLTAO 0031-9007 (2010). Google Scholar

17. 

L. Devaud et al., “Temporal light control in complex media through the singular-value decomposition of the time-gated transmission matrix,” Phys. Rev. A, 105 (5), L051501 https://doi.org/10.1103/PhysRevA.105.L051501 (2022). Google Scholar

18. 

T. Zhao et al., “High-speed photoacoustic-guided wavefront shaping for focusing light in scattering media,” Opt. Lett., 46 (5), 1165 –1168 https://doi.org/10.1364/OL.412572 OPLEDP 0146-9592 (2021). Google Scholar

19. 

E. Seo et al., “Near-field transmission matrix microscopy for mapping high-order eigenmodes of subwavelength nanostructures,” Nat. Commun., 11 (1), 2575 https://doi.org/10.1038/s41467-020-16263-z NCAOBW 2041-1723 (2020). Google Scholar

20. 

Z. Yaqoob et al., “Optical phase conjugation for turbidity suppression in biological samples,” Nat. Photonics, 2 (2), 110 –115 https://doi.org/10.1038/nphoton.2007.297 NPAHBY 1749-4885 (2008). Google Scholar

21. 

T. R. Hillman et al., “Digital optical phase conjugation for delivering two-dimensional images through turbid media,” Sci. Rep., 3 (1), 1909 https://doi.org/10.1038/srep01909 SRCEC3 2045-2322 (2013). Google Scholar

22. 

Z. Li et al., “Edge enhancement through scattering media enabled by optical wavefront shaping,” Photonics Res., 8 (6), 954 –962 https://doi.org/10.1364/PRJ.388062 (2020). Google Scholar

23. 

L. Liu et al., “Anti-scattering light focusing with full-polarization digital optical phase conjugation based on digital micromirror devices,” Opt. Express, 30 (18), 31614 –31622 https://doi.org/10.1364/OE.467444 OPEXFF 1094-4087 (2022). Google Scholar

24. 

J. Yang et al., “Anti-scattering light focusing by fast wavefront shaping based on multi-pixel encoded digital-micromirror device,” Light Sci. Appl., 10 (1), 149 https://doi.org/10.1038/s41377-021-00591-w (2021). Google Scholar

25. 

M. Xu and L. V. Wang, “Photoacoustic imaging in biomedicine,” Rev. Sci. Instrum., 77 (4), https://doi.org/10.1063/1.2195024 RSINAK 0034-6748 (2006). Google Scholar

26. 

I. M. Vellekoop, A. Lagendijk and A. P. Mosk, “Exploiting disorder for perfect focusing,” Nat. Photonics, 4 (5), 320 –322 https://doi.org/10.1038/nphoton.2010.3 NPAHBY 1749-4885 (2010). Google Scholar

27. 

R. Horstmeyer, H. Ruan and C. Yang, “Guidestar-assisted wavefront-shaping methods for focusing light into biological tissue,” Nat. Photonics, 9 (9), 563 –571 https://doi.org/10.1038/nphoton.2015.140 NPAHBY 1749-4885 (2015). Google Scholar

28. 

E. E. Morales-Delgado, D. Psaltis and C. Moser, “Two-photon imaging through a multimode fiber,” Opt. Express, 23 (25), 32158 –32170 https://doi.org/10.1364/OE.23.032158 OPEXFF 1094-4087 (2015). Google Scholar

29. 

L. V. Wang and S. Hu, “Photoacoustic tomography: in vivo imaging from organelles to organs,” Science, 335 (6075), 1458 –1462 https://doi.org/10.1126/science.1216210 SCIEAS 0036-8075 (2012). Google Scholar

30. 

J. Bertolotti et al., “Noninvasive imaging through opaque scattering layers,” Nature, 491 (7423), 232 –234 https://doi.org/10.1038/nature11578 (2012). Google Scholar

31. 

H. Yu et al., “Recent advances in wavefront shaping techniques for biomedical applications,” Curr. Appl. Phys., 15 (5), 632 –641 https://doi.org/10.1016/j.cap.2015.02.015 1567-1739 (2015). Google Scholar

32. 

Z. Yu et al., “Wavefront shaping: a versatile tool to conquer multiple scattering in multidisciplinary fields,” The Innovation, 3 (5), 100292 https://doi.org/10.1016/j.xinn.2022.100292 (2022). Google Scholar

33. 

H. Cao, A. P. Mosk and S. Rotter, “Shaping the propagation of light in complex media,” Nat. Phys., 18 (9), 994 –1007 https://doi.org/10.1038/s41567-022-01677-x NPAHAX 1745-2473 (2022). Google Scholar

34. 

S. Li et al., “Memory effect assisted imaging through multimode optical fibres,” Nat. Commun., 12 (1), 3751 https://doi.org/10.1038/s41467-021-23729-1 NCAOBW 2041-1723 (2021). Google Scholar

35. 

P. Sheng and B. van Tiggelen, Introduction to Wave Scattering, Localization and Mesoscopic Phenomena, Taylor & Fransis( (2007). Google Scholar

36. 

C. W. J. Beenakker, “Random-matrix theory of quantum transport,” Rev. Mod. Phys., 69 731 –808 https://doi.org/10.1103/RevModPhys.69.731 RMPHAT 0034-6861 (1997). Google Scholar

37. 

A. P. Mosk et al., “Controlling waves in space and time for imaging and focusing in complex media,” Nat. Photonics, 6 (5), 283 –292 https://doi.org/10.1038/nphoton.2012.88 NPAHBY 1749-4885 (2012). Google Scholar

38. 

H. Yu, J. H. Park and Y. Park, “Measuring large optical reflection matrices of turbid media,” Opt. Commun., 352 33 –38 https://doi.org/10.1016/j.optcom.2015.04.073 OPCOB8 0030-4018 (2015). Google Scholar

39. 

X. Tao et al., “High-speed scanning interferometric focusing by fast measurement of binary transmission matrix for channel demixing,” Opt. Express, 23 (11), 14168 –14187 https://doi.org/10.1364/OE.23.014168 OPEXFF 1094-4087 (2015). Google Scholar

40. 

S. R. Huisman et al., “Programmable multiport optical circuits in opaque scattering materials,” Opt. Express, 23 (3), 3102 –3116 https://doi.org/10.1364/OE.23.003102 OPEXFF 1094-4087 (2015). Google Scholar

41. 

S. Leedumrongwatthanakun et al., “Programmable linear quantum networks with a multimode fibre,” Nat. Photonics, 14 (3), 139 –142 https://doi.org/10.1038/s41566-019-0553-9 NPAHBY 1749-4885 (2020). Google Scholar

42. 

T. Chaigne et al., “Controlling light in scattering media non-invasively using the photoacoustic transmission matrix,” Nat. Photonics, 8 (1), 58 –64 https://doi.org/10.1038/nphoton.2013.307 NPAHBY 1749-4885 (2014). Google Scholar

43. 

T. Zhao et al., “Bayes’ theorem-based binary algorithm for fast reference-less calibration of a multimode fiber,” Opt. Express, 26 20368 –20378 https://doi.org/10.1364/OE.26.020368 OPEXFF 1094-4087 (2018). Google Scholar

44. 

G. Huang et al., “Generalizing the Gerchberg-Saxton algorithm for retrieving complex optical transmission matrices,” Photonics Res., 9 (1), 34 –42 https://doi.org/10.1364/PRJ.406010 (2021). Google Scholar

45. 

M. N’Gom et al., “Mode control in a multimode fiber through acquiring its transmission matrix from a reference-less optical system,” Opt. Lett., 43 (3), 419 –422 https://doi.org/10.1364/OL.43.000419 OPLEDP 0146-9592 (2018). Google Scholar

46. 

G. Huang et al., “Retrieving the optical transmission matrix of a multimode fiber using the extended Kalman filter,” Opt. Express, 28 (7), 9487 –9500 https://doi.org/10.1364/OE.389133 OPEXFF 1094-4087 (2020). Google Scholar

47. 

A. Drémeau et al., “Reference-less measurement of the transmission matrix of a highly scattering material using a DMD and phase retrieval techniques,” Opt. Express, 23 (9), 11898 –11911 https://doi.org/10.1364/OE.23.011898 OPEXFF 1094-4087 (2015). Google Scholar

48. 

M. Kim et al., “Maximal energy transport through disordered media with the implementation of transmission eigenchannels,” Nat. Photonics, 6 (9), 581 –585 https://doi.org/10.1038/nphoton.2012.159 NPAHBY 1749-4885 (2012). Google Scholar

49. 

H. Lee et al., “High-throughput volumetric adaptive optical imaging using compressed time-reversal matrix,” Light Sci. Appl., 11 (1), 16 https://doi.org/10.1038/s41377-021-00705-4 (2022). Google Scholar

50. 

K. Lee and Y. Park, “Exploiting the speckle-correlation scattering matrix for a compact reference-free holographic image sensor,” Nat. Commun., 7 (1), 13359 https://doi.org/10.1038/ncomms13359 NCAOBW 2041-1723 (2016). Google Scholar

51. 

S. Rotter and S. Gigan, “Light fields in complex media: mesoscopic scattering meets wave control,” Rev. Mod. Phys., 89 (1), 015005 https://doi.org/10.1103/RevModPhys.89.015005 RMPHAT 0034-6861 (2017). Google Scholar

52. 

H. Yu, K. Lee and Y. Park, “Ultrahigh enhancement of light focusing through disordered media controlled by mega-pixel modes,” Opt. Express, 25 (7), 8036 –8047 https://doi.org/10.1364/OE.25.008036 OPEXFF 1094-4087 (2017). Google Scholar

53. 

Y. He et al., “Genetic-algorithm-assisted coherent enhancement absorption in scattering media by exploiting transmission and reflection matrices,” Opt. Express, 29 (13), 20353 –20369 https://doi.org/10.1364/OE.426496 OPEXFF 1094-4087 (2021). Google Scholar

54. 

Q. Zhao et al., “Creation of cylindrical vector beams through highly anisotropic scattering media with a single scalar transmission matrix calibration,” Photonics Res., 10 (7), 1617 –1623 https://doi.org/10.1364/PRJ.457928 (2022). Google Scholar

55. 

W. Xiong et al., “Complete polarization control in multimode fibers with polarization and mode coupling,” Light Sci. Appl., 7 (1), 54 https://doi.org/10.1038/s41377-018-0047-4 (2018). Google Scholar

56. 

S. Tripathi et al., “Vector transmission matrix for the polarization behavior of light propagation in highly scattering media,” Opt. Express, 20 (14), 16067 –16076 https://doi.org/10.1364/OE.20.016067 OPEXFF 1094-4087 (2012). Google Scholar

57. 

D. Andreoli et al., “Deterministic control of broadband light through a multiply scattering medium via the multispectral transmission matrix,” Sci. Rep., 5 (1), 10347 https://doi.org/10.1038/srep10347 SRCEC3 2045-2322 (2015). Google Scholar

58. 

M. Mounaix et al., “Spatiotemporal coherent control of light through a multiple scattering medium with the multispectral transmission matrix,” Phys. Rev. Let., 116 (25), 253901 https://doi.org/10.1103/PhysRevLett.116.253901 (2016). Google Scholar

59. 

M. Mounaix et al., “Deterministic light focusing in space and time through multiple scattering media with a time-resolved transmission matrix approach,” Phys. Rev. A, 94 (4), 041802 https://doi.org/10.1103/PhysRevA.94.041802 (2016). Google Scholar

60. 

S. Kang et al., “Imaging deep within a scattering medium using collective accumulation of single-scattered waves,” Nat. Photonics, 9 (4), 253 –258 https://doi.org/10.1038/nphoton.2015.24 NPAHBY 1749-4885 (2015). Google Scholar

61. 

X. Wei et al., “Harnessing a multi-dimensional fibre laser using genetic wavefront shaping,” Light Sci. Appl., 9 149 https://doi.org/10.1038/s41377-020-00383-8 (2020). Google Scholar

62. 

Y. Zhao et al., “Gradient-assisted focusing light through scattering media,” Opt. Lett., 46 (7), 1518 –1521 https://doi.org/10.1364/OL.417606 OPLEDP 0146-9592 (2021). Google Scholar

63. 

C. M. Woo et al., “Optimal efficiency of focusing diffused light through scattering media with iterative wavefront shaping,” APL Photonics, 7 (4), 046109 https://doi.org/10.1063/5.0085943 (2022). Google Scholar

64. 

Q. Zhao et al., “Parameter-free optimization algorithm for iterative wavefront shaping,” Opt. Lett., 46 (12), 2880 –2883 https://doi.org/10.1364/OL.427215 OPLEDP 0146-9592 (2021). Google Scholar

65. 

L. Liu et al., “Anti-scattering light focusing by full-polarization wavefront shaping based on digital micromirror devices,” Appl. Phys. Express, 15 (9), 092001 https://doi.org/10.35848/1882-0786/ac851e APEPC4 1882-0778 (2022). Google Scholar

66. 

P. Lai et al., “Photoacoustically guided wavefront shaping for enhanced optical focusing in scattering media,” Nat. Photonics, 9 (2), 126 –132 https://doi.org/10.1038/nphoton.2014.322 NPAHBY 1749-4885 (2015). Google Scholar

67. 

O. Katz et al., “Noninvasive nonlinear focusing and imaging through strongly scattering turbid layers,” Optica, 1 (3), 170 –174 https://doi.org/10.1364/OPTICA.1.000170 (2014). Google Scholar

68. 

A. M. Caravaca-Aguirre et al., “High contrast three-dimensional photoacoustic imaging through scattering media by localized optical fluence enhancement,” Opt. Express, 21 (22), 26671 –26676 https://doi.org/10.1364/OE.21.026671 OPEXFF 1094-4087 (2013). Google Scholar

69. 

X. Xu, H. Liu and L. V. Wang, “Time-reversed ultrasonically encoded optical focusing into scattering media,” Nat. Photonics, 5 (3), 154 –157 https://doi.org/10.1038/nphoton.2010.306 NPAHBY 1749-4885 (2011). Google Scholar

70. 

D. B. Conkey, “Super-resolution photoacoustic imaging through a scattering wall,” Nat. Commun., 6 (1), 7902 https://doi.org/10.1038/ncomms8902 NCAOBW 2041-1723 (2015). Google Scholar

71. 

A. Thendiyammal et al., “Model-based wavefront shaping microscopy,” Opt. Lett., 45 (18), 5101 –5104 https://doi.org/10.1364/OL.400985 OPLEDP 0146-9592 (2020). Google Scholar

72. 

I. M. Vellekoop et al., “Demixing light paths inside disordered metamaterials,” Opt. Express, 16 67 –80 https://doi.org/10.1364/OE.16.000067 OPEXFF 1094-4087 (2008). Google Scholar

73. 

J. Yang et al., “Focusing light inside live tissue using reversibly switchable bacterial phytochrome as a genetically encoded photochromic guide star,” Sci. Adv., 5 eaay1211 https://doi.org/10.1126/sciadv.aay1211 STAMCV 1468-6996 (2019). Google Scholar

74. 

Z. Yu et al., “Time-reversed magnetically controlled perturbation (TRMCP) optical focusing inside scattering media,” Sci. Rep., 8 2927 https://doi.org/10.1038/s41598-018-21258-4 SRCEC3 2045-2322 (2018). Google Scholar

75. 

H. Ruan, M. Jang and C. Yang, “Optical focusing inside scattering media with time-reversed ultrasound microbubble encoded light,” Nat. Commun., 6 8968 https://doi.org/10.1038/ncomms9968 NCAOBW 2041-1723 (2015). Google Scholar

76. 

M. Rueckel, J. A. Mack-Bucher and W. Denk, “Adaptive wavefront correction in two-photon microscopy using coherence-gated wavefront sensing,” Proc. Natl. Acad. Sci. U. S. A., 103 17137 –17142 https://doi.org/10.1073/pnas.0604791103 (2006). Google Scholar

77. 

E. H. Zhou et al., “Focusing on moving targets through scattering samples,” Optica, 1 227 –232 https://doi.org/10.1364/OPTICA.1.000227 (2014). Google Scholar

78. 

R. Gao et al., “Breaking acoustic limit of optical focusing using photoacoustic-guided wavefront shaping,” Laser Photonics Rev., 15 (8), 2000594 https://doi.org/10.1002/lpor.202000594 (2021). Google Scholar

79. 

J. Yang et al., “Angular-spectrum modeling of focusing light inside scattering media by optical phase conjugation,” Optica, 6 (3), 250 –256 https://doi.org/10.1364/OPTICA.6.000250 (2019). Google Scholar

80. 

Z. Cheng and L. V. Wang, “Focusing light into scattering media with ultrasound-induced field perturbation,” Light Sci. Appl., 10 159 https://doi.org/10.1038/s41377-021-00605-7 (2021). Google Scholar

81. 

F. Kong et al., “Photoacoustic-guided convergence of light through optically diffusive media,” Opt. Lett., 36 (11), 2053 –2055 https://doi.org/10.1364/OL.36.002053 OPLEDP 0146-9592 (2011). Google Scholar

82. 

M. Fink, “Time-reversed acoustics,” Sci. Am., 281 (5), 91 –97 https://doi.org/10.1038/scientificamerican1199-91 SCAMAC 0036-8733 (1999). Google Scholar

83. 

A. Derode, P. Roux and M. Fink, “Robust acoustic time reversal with highorder multiple scattering,” Phys. Rev. Lett., 75 4206 –4209 https://doi.org/10.1103/PhysRevLett.75.4206 PRLTAO 0031-9007 (1995). Google Scholar

84. 

D. A. B. Miller, “Time reversal of optical pulses by four-wave mixing,” Opt. Lett., 5 300 –302 https://doi.org/10.1364/OL.5.000300 OPLEDP 0146-9592 (1980). Google Scholar

85. 

A. Yariv, “Four wave nonlinear optical mixing as real time holography,” Opt. Commun., 25 23 –25 https://doi.org/10.1016/0030-4018(78)90079-2 OPCOB8 0030-4018 (1978). Google Scholar

86. 

Y. Liu et al., “Focusing light inside dynamic scattering media with millisecond digital optical phase conjugation,” Optica, 4 (2), 280 –288 https://doi.org/10.1364/OPTICA.4.000280 (2017). Google Scholar

87. 

J. Yang et al., “Fighting against fast speckle decorrelation for light focusing inside live tissue by photon frequency shifting,” ACS Photonics, 7 (3), 837 –844 https://doi.org/10.1021/acsphotonics.0c00027 (2020). Google Scholar

88. 

A. Bashkatov et al., “Optical properties of human skin, subcutaneous and mucous tissues in the wavelength range from 400 to 2000 nm,” J. Phys. D Appl. Phys., 38 2543 https://doi.org/10.1088/0022-3727/38/15/004 (2005). Google Scholar

89. 

P. Theer, M. T. Hasan and W. Denk, “Two-photon imaging to a depth of 1000 mm in living brains by use of a Ti:Al2O3 regenerative amplifier,” Opt. Lett., 28 1022 –1024 (2003). https://doi.org/10.1364/ol.28.001022 Google Scholar

90. 

P. T. C. So et al., “Two-photon excitation fluorescence microscopy,” Annu. Rev. Biomed. Eng., 2 (1), 399 –429 https://doi.org/10.1146/annurev.bioeng.2.1.399 ARBEF7 1523-9829 (2000). Google Scholar

91. 

W. H. Lee, “Binary computer-generated holograms,” Appl. Opt., 18 3661 –3669 https://doi.org/10.1364/AO.18.003661 APOPAI 0003-6935 (1979). Google Scholar

92. 

S. A. Goorden, J. Bertolotti and A. P. Mosk, “Superpixel-based spatial amplitude and phase modulation using a digital micromirror device,” Opt. Express, 22 (15), 17999 –18009 https://doi.org/10.1364/OE.22.017999 OPEXFF 1094-4087 (2014). Google Scholar

93. 

R. Shao et al., “Volumetric random-access multi-focus scanning based on fast light modulation,” Opt. Lasers Eng., 158 107128 https://doi.org/10.1016/j.optlaseng.2022.107128 (2022). Google Scholar

94. 

Y. Shen et al., “Focusing light through scattering media by full-polarization digital optical phase conjugation,” Opt. Lett., 41 (6), 1130 –1133 https://doi.org/10.1364/OL.41.001130 OPLEDP 0146-9592 (2016). Google Scholar

95. 

C. Ding et al., “Spatial full degree-of-freedom scattered optical field modulation,” Laser Photonics Rev., 17 2300104 https://doi.org/10.1002/lpor.202300104 (2023). Google Scholar

96. 

Z. Cheng et al., “High-gain and high-speed wavefront shaping through scattering media,” Nat. Photonics, 17 299 –305 https://doi.org/10.1038/s41566-022-01142-4 NPAHBY 1749-4885 (2023). Google Scholar

97. 

B. Judkewitz et al., “Speckle-scale focusing in the diffusive regime with time reversal of variance-encoded light (TROVE),” Nat. Photonics, 7 (4), 300 –305 https://doi.org/10.1038/nphoton.2013.31 NPAHBY 1749-4885 (2013). Google Scholar

98. 

I. M. Vellekoop and C. M. Aegerter, “Scattered light fluorescence microscopy: imaging through turbid layers,” Opt. Lett., 35 (8), 1245 –1247 https://doi.org/10.1364/OL.35.001245 OPLEDP 0146-9592 (2010). Google Scholar

99. 

R. Cao et al., “High-resolution non-line-of-sight imaging employing active focusing,” Nat. Photonics, 16 462 –468 https://doi.org/10.1038/s41566-022-01009-8 NPAHBY 1749-4885 (2022). Google Scholar

100. 

A. Boniface, J. Dong and S. Gigan, “Non-invasive focusing and imaging in scattering media with a fluorescence-based transmission matrix,” Nat. Commun., 11 6154 https://doi.org/10.1038/s41467-020-19696-8 NCAOBW 2041-1723 (2020). Google Scholar

101. 

T. Yeminy and O. Katz, “Guidestar-free image-guided wavefront shaping,” Sci. Adv., 7 (21), eabf5364 https://doi.org/10.1126/sciadv.abf5364 STAMCV 1468-6996 (2021). Google Scholar

102. 

B. Y. Feng et al., “NeuWS: neural wavefront shaping for guidestar-free imaging through static and dynamic scattering media,” Sci. Adv., 9 (26), eadg4671 https://doi.org/10.1126/sciadv.adg4671 STAMCV 1468-6996 (2023). Google Scholar

103. 

Y. Luo et al., “Towards smart optical focusing: deep learning-empowered dynamic wavefront shaping through nonstationary scattering media,” Photonics Res., 9 (8), B262 –B278 https://doi.org/10.1364/PRJ.415590 (2021). Google Scholar

104. 

U. Teğin et al., “Controlling spatiotemporal nonlinearities in multimode fibers with deep neural networks,” APL Photonics, 5 (3), 030804 https://doi.org/10.1063/1.5138131 (2020). Google Scholar

105. 

A. Porat et al., “Widefield lensless imaging through a fiber bundle via speckle correlations,” Opt. Express, 24 16835 –16855 https://doi.org/10.1364/OE.24.016835 OPEXFF 1094-4087 (2016). Google Scholar

106. 

M. Plöschner, T. Tyc and T. Čižmár, “Seeing through chaos in multimode fibres,” Nat. Photonics, 9 (8), 529 –535 https://doi.org/10.1038/nphoton.2015.112 NPAHBY 1749-4885 (2015). Google Scholar

107. 

T. Čižmár and K. Dholakia, “Shaping the light transmission through a multimode optical fibre: complex transformation analysis and applications in biophotonics,” Opt. Express, 19 (20), 18871 –18884 https://doi.org/10.1364/OE.19.018871 OPEXFF 1094-4087 (2011). Google Scholar

108. 

T. Zhong et al., “Optically selective neuron stimulation with a wavefront shaping-empowered multimode fiber,” Adv. Photonics Res., 3 (3), 2100231 https://doi.org/10.1002/adpr.202100231 (2022). Google Scholar

109. 

S. Turtaev et al., “High-fidelity multimode fibre-based endoscopy for deep brain in vivo imaging,” Light Sci. Appl., 7 92 https://doi.org/10.1038/s41377-018-0094-x (2018). Google Scholar

110. 

Z. Wen et al., “Single multimode fibre for in vivo light-field-encoded endoscopic imaging,” Nat. Photonics, 17 679 –687 https://doi.org/10.1038/s41566-023-01240-x NPAHBY 1749-4885 (2023). Google Scholar

111. 

D. Stellinga et al., “Time-of-flight 3D imaging through multimode optical fibers,” Science, 374 (6573), 1395 –1399 https://doi.org/10.1126/science.abl3771 SCIEAS 0036-8075 (2021). Google Scholar

112. 

I. T. Leite et al., “Observing distant objects with a multimode fiber-based holographic endoscope,” APL Photonics, 6 (3), 036112 https://doi.org/10.1063/5.0038367 (2021). Google Scholar

113. 

B. Rahmani et al., “Multimode optical fiber transmission with a deep learning network,” Light Sci. Appl., 7 69 https://doi.org/10.1038/s41377-018-0074-1 (2018). Google Scholar

114. 

P. Caramazza et al., “Transmission of natural scene images through a multimode fibre,” Nat. Commun., 10 2029 https://doi.org/10.1038/s41467-019-10057-8 NCAOBW 2041-1723 (2019). Google Scholar

115. 

B. Sun et al., “Four-dimensional light shaping: manipulating ultrafast spatiotemporal foci in space and time,” Light Sci. Appl., 7 17117 https://doi.org/10.1038/lsa.2017.117 (2018). Google Scholar

116. 

O. Hernandez et al., “Three-dimensional spatiotemporal focusing of holographic patterns,” Nat. Commun., 7 11928 https://doi.org/10.1038/ncomms11928 NCAOBW 2041-1723 (2016). Google Scholar

117. 

M. Mounaix et al., “Time reversed optical waves by arbitrary vector spatiotemporal field generation,” Nat. Commun., 11 5813 https://doi.org/10.1038/s41467-020-19601-3 NCAOBW 2041-1723 (2020). Google Scholar

118. 

M. W. Matthès et al., “Optical complex media as universal reconfigurable linear operators,” Optica, 6 (4), 465 –472 https://doi.org/10.1364/OPTICA.6.000465 (2019). Google Scholar

119. 

Q. Zhao et al., “Speckle-based optical cryptosystem and its application for human face recognition via deep learning,” Adv. Sci., 9 (25), 2202407 https://doi.org/10.1002/advs.202202407 (2022). Google Scholar

120. 

I. T. Leite et al., “Three-dimensional holographic optical manipulation through a high-numerical-aperture soft-glass multimode fibre,” Nat. Photonics, 12 33 –39 https://doi.org/10.1038/s41566-017-0053-8 NPAHBY 1749-4885 (2017). Google Scholar

121. 

A. M. Caravaca-Aguirre and R. Piestun, “Single multimode fiber endoscope,” Opt. Express, 25 (3), 1656 –1665 https://doi.org/10.1364/OE.25.001656 OPEXFF 1094-4087 (2017). Google Scholar

122. 

P. Yu et al., “Ultrahigh-density 3D holographic projection by scattering-assisted dynamic holography,” Optica, 10 481 –490 https://doi.org/10.1364/OPTICA.483057 (2023). Google Scholar

123. 

P. Yu et al., “Dynamic polarization holographic projection enabled by a scattering material-based reconfigurable hologram,” ACS Photonics, 9 (11), 3712 –3719 https://doi.org/10.1021/acsphotonics.2c01303 (2022). Google Scholar

124. 

H. Yu et al., “Ultrahigh-definition dynamic 3D holographic display by active control of volume speckle fields,” Nat. Photonics, 11 (3), 186 –192 https://doi.org/10.1038/nphoton.2016.272 NPAHBY 1749-4885 (2017). Google Scholar

125. 

Q. Zhang et al., “Learning the matrix of few-mode fibers for high-fidelity spatial mode transmission,” APL Photonics, 7 (6), 066104 https://doi.org/10.1063/5.0088605 (2022). Google Scholar

126. 

Y. Zhou et al., “High-fidelity spatial mode transmission through a 1-km-long multimode fiber via vectorial time reversal,” Nat. Commun., 12 1866 https://doi.org/10.1038/s41467-021-22071-w NCAOBW 2041-1723 (2021). Google Scholar

127. 

L. Gong et al., “Optical orbital-angular-momentum-multiplexed data transmission under high scattering.,” Light Sci. Appl., 8 (1), 27 https://doi.org/10.1038/s41377-019-0140-3 (2019). Google Scholar

128. 

J. H. Park and B. Lee, “Holographic techniques for augmented reality and virtual reality near-eye displays,” Light: Adv. Manuf., 3 (1), 137 –150 (2022). Google Scholar

129. 

B. Rudolf et al., “Thermal stability of wavefront shaping using a DMD as a spatial light modulator,” Opt. Express, 29 (25), 41808 –41818 https://doi.org/10.1364/OE.442284 OPEXFF 1094-4087 (2021). Google Scholar

130. 

D. E. B. Flaes et al., “Robustness of light-transport processes to bending deformations in graded-index multimode waveguides,” Phys. Rev. Lett, 120 (23), 233901 https://doi.org/10.1103/PhysRevLett.120.233901 (2018). Google Scholar

131. 

M. M. Qureshi et al., “In vivo study of optical speckle decorrelation time across depths in the mouse brain,” Biomed. Opt. Express, 8 (11), 4855 –4864 https://doi.org/10.1364/BOE.8.004855 BOEICL 2156-7085 (2017). Google Scholar

132. 

G. Zou et al., “Random-access multi-focus manipulation through superpixel-encoding wavefront engineering,” Appl. Phys. Express, 15 (11), 112004 https://doi.org/10.35848/1882-0786/ac99b7 APEPC4 1882-0778 (2022). Google Scholar

133. 

B. W. Pogue and M. S. Patterson, “Review of tissue simulating phantoms for optical spectroscopy, imaging and dosimetry,” J. Biomed. Opt., 11 041102 https://doi.org/10.1117/1.2335429 JBOPFO 1083-3668 (2006). Google Scholar

134. 

L. G. Wright, D. N. Christodoulides and F. W. Wise, “Controllable spatiotemporal nonlinear effects in multimode fibres,” Nat. Photonics, 9 306 –310 https://doi.org/10.1038/nphoton.2015.61 NPAHBY 1749-4885 (2015). Google Scholar

135. 

S. Cheng et al., “Artificial intelligence-assisted light control and computational imaging through scattering media,” J. Innov. Opt. Health, 12 (04), 1930006 https://doi.org/10.1142/S1793545819300064 (2019). Google Scholar

136. 

C. M. Hsieh, M. O. Malik and Q. Liu, “High-speed feedback based wavefront shaping for spatiotemporal enhancement of incoherent light through dynamic scattering media,” Opt. Lett., 48 2313 –2316 https://doi.org/10.1364/OL.491457 OPLEDP 0146-9592 (2023). Google Scholar

137. 

L. Valzania and S. Gigan, “Online learning of the transmission matrix of dynamic scattering media,” Optica, 10 708 –716 https://doi.org/10.1364/OPTICA.479962 (2023). Google Scholar

138. 

X. Hu et al., “Adaptive inverse mapping: a model-free semi-supervised learning approach towards robust imaging through dynamic scattering media,” Opt. Express, 31 14343 –14357 https://doi.org/10.1364/OE.484252 OPEXFF 1094-4087 (2023). Google Scholar

139. 

X. Zhang et al., “Different channels to transmit information in scattering media,” PhotoniX, 4 10 https://doi.org/10.1186/s43074-023-00087-3 (2023). Google Scholar

140. 

E. Guo et al., “Dynamic imaging through random perturbed fibers via physics-informed learning,” Opt. Laser Technol., 158 108923 https://doi.org/10.1016/j.optlastec.2022.108923 OLTCAS 0030-3992 (2023). Google Scholar

141. 

X. Zhang et al., “Physical origin and boundary of scalable imaging through scattering media: a deep learning-based exploration,” Photonics Res., 11 1038 –1046 https://doi.org/10.1364/PRJ.490125 (2023). Google Scholar

Biographies of the authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Chunxu Ding, Rongjun Shao, Qiaozhi He, Lei S. Li, and Jiamiao Yang "Wavefront shaping improves the transparency of the scattering media: a review," Journal of Biomedical Optics 29(S1), S11507 (9 December 2023). https://doi.org/10.1117/1.JBO.29.S1.S11507
Received: 14 October 2023; Accepted: 22 November 2023; Published: 9 December 2023
Advertisement
Advertisement
KEYWORDS
Light scattering

Biomedical optics

Wavefronts

Scattering

Scattering media

Modulation

Transparency

Back to Top