Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Design and characterization of a hyperbolic-elliptical lens pair in a rapid beam steering system for single-pixel terahertz spectral imaging of the cornea

Open Access Open Access

Abstract

Terahertz (THz) time-domain spectroscopy has been investigated for assessment of the hydration levels in the cornea, intraocular pressure, and changes in corneal topography. Previous efforts at THz imaging of the cornea have employed off-axis parabolic mirrors to achieve normal incidence along the spherical surface. However, this comes at the cost of an asymmetric field-of-view (FOV) and a long scan time because it requires raster-scanning of the collimated beam across the large mirror diameter. This paper proposes a solution by designing a pair of aspheric lenses that can provide a larger symmetric spherical FOV (9.6 mm) and reduce the scan time by two orders of magnitude using a novel beam-steering approach. A hyperbolic-elliptical lens was designed and optimized to achieve normal incidence and phase-front matching between the focused THz beam and the target curvature. The lenses were machined from a slab of high-density polyethylene and characterized in comparison to ray-tracing simulations by imaging several targets of similar sizes to the cornea. Our experimental results showed excellent agreement in the increased symmetric FOV and confirmed the reduction in scan time to about 3-4 seconds. In the future, this lens design process can be extended for imaging the sclera of the eye and other curved biological surfaces, such as the nose and fingers.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Terahertz time-domain spectroscopy (THz-TDS) has shown significant promise in the field of biomedical diagnostic imaging and sensing not only due to its non-ionizing photon-energy levels, but thanks to its unique sensitivity to hydration changes [13]. One example is in determination of skin burn injuries [46]. Classification of burn severity is possible by detecting changes in the THz reflectivity, and using machine learning algorithms [7] such as neural networks [8] and support vector machines [9], the wound-healing outcome can be predicted with high accuracy levels. Additionally, THz imaging has been used for diagnosis of brain cancer tumors [1013], breast cancer biopsies [14,15], skin cancer delineation [1618] and in several ophthalmological applications [1928]. In most of these applications, the only viable imaging modality for clinical applications is spectroscopic imaging in the reflection mode.

THz reflection imaging of a biological surface using a single-pixel detector requires ensuring normal incidence on the target over all scanned pixels. In prior THz imaging studies of burned skin, an f-theta lens was used for this purpose to focus and raster-scan the THz beam on a flat image plane [29,30]. To ensure normal incidence, a fused-silica imaging window was placed in contact with the skin. Curved biological surfaces such as the cornea, cannot rely on traditional raster-scanning techniques to achieve normal incidence due to the target curvature. In addition to normal incidence, a noncontact approach is required given the cornea’s sensitivity to physical touch. For instance, THz hydration sensing of the sclera has been investigated but has required the use of a Golay cell for detection of scattered THz radiation due to a lack of normal incidence [31]. However, the use of diffuse spectroscopy techniques to in vivo corneal imaging applications is not practical due to low signal to noise ration and long data acquisition times [32]. Moreover, the field-of-view (FOV) of the imaging optic should be large enough to cover the entire corneal surface for mapping hydration gradients.

To satisfy these imaging requirements, previous efforts for THz imaging of spherical targets have employed 90° off-axis parabolic mirrors (OAPM) as the focusing optic. As shown in Fig. 1(a), aligning the center of curvature of the cornea to the focal point of an OAPM and steering the collimated beam perpendicular to the mirror aperture result in normal incidence on all pixels scanned along the target surface [33,34]. Using the same optical beam steering layout, we have demonstrated a broadband THz-TDS topographic scanner with spectroscopic bandwidth between 0.1 and 1 THz for imaging spherical targets [35]. Converting variations in the measured time-of-arrival (TOA) of the reflected THz beam to a spatial height profile allows for topographic mapping of the cornea. In addition, analysis of the detected signal in the frequency domain reveals further information about the cornea’s physiological condition through mapping hydration gradients. Recently it has been shown that elevations in intraocular pressure [36] and endothelial layer damage [37] in the cornea can be detected using this system by evaluating changes in the reflected THz spectral area and slope, respectively.

 figure: Fig. 1.

Fig. 1. (a) Illustration of beam-steering across the cornea using an OAPM. Placing the center of the spherical target at the focus of the OAPM ensures normal incidence. (b) Non-uniform scanning trajectory of the collimated THz beam, shown by black circles, is calculated across the OAPM aperture with constant azimuthal (blue lines) and elevation (red circles) angles on the sphere. Beam steering approach in (b) results in an equiangular sampling across the corneal surface (c). The resultant FOV is approximately 8 mm on the corneal surface. The asymmetry of the FOV in the X-axis can be seen. (d) The spot size variation of the THz beam focused by the OAPM on a spherical target is shown by ABCD ray tracing simulations and is in agreement with previous published results in [33], leading to a dramatic change in spatial resolution on the target surface.

Download Full Size | PDF

One advantage of using OAPMs is that they do not suffer from broadband attenuation losses or introduce chromatic aberrations. However, as shown in Fig. 1(b, c), beam steering using a 3-inch diameter OAPM yields a limited asymmetric field-of-view (FOV) of approximately 8 mm. Figure 1(b) demonstrates the scanning trajectory of the center of the collimated THz beam on the aperture of the OAPM to achieve a uniform mapping of the corneal surface, shown in Fig. 1(c). In addition, prior quasi-optical simulations have indicated that the focused beam spot size on the corneal surface changes significantly as a function of beam deflection along the asymmetric axis of the OAPM [33], which is also shown in Fig. 1(d). Finally, the current method of 2D mapping across the cornea requires raster-scanning the collimated THz beam across the large mirror diameter (see Fig. 1(a)). This process takes several minutes and is not ideal for clinical imaging, where complete coverage of the cornea should be completed in under a few seconds. OAPM are also difficult to align in the THz regime and custom parabolic mirror designs with desired aperture and focal length are expensive to manufacture. To address this problem, we investigate the use of scanning lenses for coupling light to the corneal surface with matching spherical phase fronts. Replacing the OAPM with a custom scanning lens for focusing the THz beam onto the cornea would allow for a symmetric FOV while still achieving normal incidence for a curved target centered at the focus. In addition, using a pair of lenses of the same design to collimate and focus the THz beam allows for a faster beam-steering approach using a gimballed mirror, which reduces the total scan time from minutes to seconds [38].

Most commercial THz lenses (i.e., plano-convex lenses, such as TPX50, Menlo Systems, Newton, NJ) suffer from spherical aberrations at the focal plane and yield a small FOV because they are not designed for imaging applications via beam-steering. Axicon lenses have been proposed for extracting corneal parameters in the THz regime by extending the depth of focus using a Bessel beam, however the beam’s first-order sidelobes are sensitive to variations in the target’s radius of curvature, making hydration mapping challenging [3941]. A 220-330 GHz quasioptical system for corneal sensing using a pair of optimized biconvex aspheric lenses has been proposed previously [42,43]. These lenses were designed to achieve a high coupling coefficient between the focused beam and curved target when the corneal surface is displaced to superconfocal and subconfocal positions. However, the application of this quasioptical system was limited to hydration sensing at the center of the cornea, and it cannot be used for 2D imaging over a large surface area, where one may wish to investigate local changes in corneal hydration and topography.

In this paper, we design a hyperbolic-elliptical aspheric lens to image the cornea with normal incidence angle over the entire span of the FOV. The lens’ faces are designed independently from each other to provide a FOV, which is at least as large as that of OAPM, but now rotationally symmetric. The lens design is also optimized to achieve phase-front matching between the THz beam and the target curvature. Upon machining the lenses, they were aligned in a collocated THz-TDS reflection imaging setup. Performance of the lenses were compared to ray-tracing simulations and the previous OAPM measurements by imaging several spherical targets of similar size to the cornea, where the reduction in scan speed and ability to perform topographic imaging was also demonstrated. We show that using a spherical beam steering approach using a pair of such lenses the scan time of corneas can be significantly reduced from minutes to about 1-4 seconds, depending on the THz-TDS trace acquisition rate, which is now compatible with clinical applications. While a pair of OAPM could also be used to collimate and focus the THz beam [28], in addition to the challenges of fabricating custom OAPMs, this would increase the amount of space required to build a portable corneal scanner for clinical imaging, where the total optical footprint should be minimized. In the future, this beam-steering approach with aspheric lenses can potentially be used for imaging other curved biological tissue, such as the sclera of the eye, or skin imaging application on surfaces such as the nose or fingers.

2. Methodology

2.1 THz Reflection geometry

A schematic of the optical layout and a picture of the system is shown in Fig. 2. A broadband THz beam is generated using 1560-nm femtosecond pulses incident upon a GaAs Photoconductive Antenna (PCA) with a 100-MHz repetition rate (Menlo Systems, Newton, NJ). Using a TPX50 lens (Menlo Systems, Newton, NJ), the diverging beam is collimated and incident upon on a silicon beam-splitter (Si BS). The transmitted beam is focused by another TPX50 lens onto a 45° mirror housed within a motorized gimbal. The gimbal, consisting of a stacked combination of a rotation and a goniometer stage (Zaber technologies, Vancouver, BC), allows for a heliostat design that enables beam-steering along the azimuthal and elevation axes, respectively [38]. The gimbal directs the THz beam across the aperture of the aspheric lens pair, which collimates and focuses the beam onto an aligned spherical target at the focus of the second lens. For all scanned pixels, if the incident beam is normal to the surface of the sphere, the reflected beam follows the same optical path back to the Si beam-splitter, where it is focused onto the PCA detector by a third TPX50 lens. Using an Asynchronous Optical sampling technique [44], a time-domain trace of the THz pulse is collected with 0.01-picoseconds temporal resolution.

 figure: Fig. 2.

Fig. 2. (a) Layout of the single-pixel scanner for imaging curved targets, separated into a THz-TDS collocated reflection setup (red box) and beam-steering arrangement for directing the beam across the aspheric lenses, which focuses the beam onto the aligned spherical target (blue box). (b) A photograph of the setup aligned on an optical table is shown with labelled components.

Download Full Size | PDF

2.2 Lens design

An aspheric lens design is implemented for focusing the beam without introducing spherical aberrations, because the front and back lens’ faces can be designed independently from one another for different imaging applications. Both lenses for the scanner have the same profile, where each lens surface is modeled separately as an aspheric curve with lens height, $z(r )$, defined for each radial position, r, using the following equation [45],

$$z(r )= \; \frac{{{r^2}/R}}{{1 + \sqrt {1 - ({1 + k} ){{(r/R)}^2}} }} + \; {a_4}{r^4} + {a_6}{r^6} + \; {a_8}{r^8} + {a_{10}}{r^{10}} + \ldots , $$
where R is the initial radius of curvature, ${a_i}$’s are higher-order aspheric coefficients which adjust the height of the lens, and k is a conic constant describing the nature of the aspheric profile. Several aspheric lenses have been previously proposed for THz subwavelength-resolution imaging using the same relationship as in Eq. (1) [46]. The planar-hyperbolic model ($k\; < \; - 1$) succeeds in correcting for spherical aberrations at the focus but suffers from total internal reflection losses [46]. To account for this, an elliptical-aspheric design was proposed that adjusts the first lens face to follow an ellipsoid curve ($ - 1\; < \; k\; < \; 0$). An alternative symmetric-pass lens design allows for a large numerical aperture [47] by having a short focal length but suffers from high attenuation losses due to an increased central lens thickness.

In this work, we chose a hyperbolic-elliptical design with a lens diameter of 3 inches (76.2 mm) which is the same diameter as the OAPM used in our previous THz corneal scanner. The parameters in Eq. (1) need to be chosen to image the target with a large enough FOV and achieve normal incidence across the target surface, which requires phase-front matching between the focused THz beam and target curvature. For optimizing an aspheric lens surface, these parameters include $R$, k and ${a_i}$, which ranges from the 4th to 12th power in increments of 2, giving rise to a total of 14 parameters (i.e., degrees of freedom) for a single aspheric lens design. For determining these aspheric parameters, Nelder-Mead optimization was employed [48]. A block diagram detailing the optimization process is shown in Fig. 3, where a simplex of 15 lenses was created using the rand function in MATLAB for determining initial aspheric parameters. For the hyperbolic face of each lens, the search space for the conic constant was limited to $ - 3\; < \; k\; < \; - 1$ and for the elliptical face, $ - 1\; < \; k\; < \; 0$. For each design, the center thickness was limited to a maximum of 1 inch (25.4 mm) to constrain the attenuation losses. The same aspheric parameters were used for both lenses in the hyperbolic-elliptical pair to limit the number of degrees of freedom for the optimization, allowing for faster convergence. In addition, having identical lenses allows for a linear relationship between the beam deflection from the gimbal and the displacement of the focused THz beam on the target surface. This permits easier mapping of pixels when forming images without requiring additional geometric transformations.

 figure: Fig. 3.

Fig. 3. Block diagram detailing the process for the lens design. The initial inputs to the optimization, including the extracted index of refraction, are used to create a 15-lens simplex. The simplex is updated iterativly using Nelder-Mead optimization, where the cost function is derived from several performance metrics obtained via ray-tracing simulations.

Download Full Size | PDF

Each design in the simplex is evaluated using ray-tracing simulations, which requires knowledge of the complex index of refraction, $n$, of the lens material. High density Polyethylene (HDPE) was chosen for the material due to its low absorption coefficient and dispersion in our system’s usable bandwidth of 0.1-1 THz [49]. The index of refraction of HDPE in the THz regime has been previously reported to be between 1.53 [50] and 1.54 [51]. However, a sensitivity analysis (which will be explained in section 3) concluded that small variations in the value of n can impact the optical performance for a specific lens design. Before optimizing the aspheric parameters, the value of n for the material sample used for machining the lens needs to be known. An 1-inch thick, 12’’ x 12’’ slab of HDPE (W. W. Grainger, Lake Forest, IL) was placed at the focus of a 4-f THz transmission setup, where the complex value of n was determined using a material parameter extraction method [52]. For twenty spatially-disjointed THz-TDS measurements obtained across the slab, as shown in Fig. 4(a), the measured and theoretical impulse response functions were compared using a total variation error function and the value of n was updated iteratively using gradient descent [52]. With this method, the average value of n for the HDPE was determined to be 1.532, as shown below in Fig. 4(b).

 figure: Fig. 4.

Fig. 4. (a) The real index of refraction n of the HDPE slab at 20 different locations extracted using THz-TDS transmission measurements. Plotted in each box is the average value for a bandwidth of .1 to 1 THz. (b) Average extracted index of refraction of HDPE over the twenty spatially-disjointed measurements as a function of frequency with error bars showing the standard error.

Download Full Size | PDF

Given the refractive index, the focal length f of each design in the simplex was calculated using the lens maker’s equation for a thick lens,

$$\frac{1}{f} = ({n - 1} )\left[ {\frac{1}{{{R_1}}} - \; \frac{1}{{{R_2}}} + \; \frac{{({n - 1} )d}}{{n{R_1}{R_2}}}} \right],$$
where d is the center lens thickness and ${R_1}$ and ${R_2}$ are the radius of curvature parameters for the front and back surfaces of the lens, respectively. While the previously-discussed quasioptical system [42] relied on a separate penalty function to optimize the lenses for reducing aberrations at the focus, the focal length for each design in our simplex was used as an input in a spherical aberration correction model based on Fermat’s principle [53] to adjust the hyperbolic face of the lens, with an example shown in the block diagram (see Fig. 3).

The ray-tracing simulation for evaluating each design is separated into several stages, starting at the back focus of the first aspheric lens with rays propagating towards its hyperbolic surface. Located at this point would be the pivot center of the gimballed mirror responsible for beam-steering. In the simulation, the THz beam is treated as the superposition of 250 single rays directed to equally spaced positions along the lens aperture. After collimation, each ray is focused by the second lens onto a spherical target of similar size to the cornea (8-mm radius) centered at the focus. All rays that strike the target with normal incidence are reflected back through both lenses via the same optical path. Since the distance of our target surface from the focus (8 mm) is greater than the calculated Rayleigh length, it is adequately sufficient to rely on geometrical optics for a simpler initial design of these lenses. However, future optimization and refinement of this design should utilize physical optics methods. Figure 5 shows the different stages of the ray-tracing simulation, where the blue rays represent the incident beam for different initial beam deflection positions, including for angles that miss the lens. The red rays show the path of the reflected rays from the spherical target. The phase of the incident chief ray (${\textrm{s}_0}$) at the corneal surface is determined using the ray’s propagation path through free space and the aspheric lenses, assuming a center frequency of 0.5 THz. For all other rays, the phase, $\textrm{s}$, is similarly calculated at different locations from the second aspheric lens surface to the focal point. In order to calculate the beam’s phase front, represented by the green curves in Fig. 5, we minimized the parameter $\Delta \textrm{s}$ defined as the sum of $|{\textrm{s} - \textrm{}{\textrm{s}_0}} |$ for all rays. This calculation confirms that for our optimized lenses, the spherical phase-front matches the corneal surface.

 figure: Fig. 5.

Fig. 5. Ray-tracing simulation for two aspheric lenses of the same design. (a) A set of incident rays (blue) are steered across and focused by the pair of lenses onto an aligned spherical target (purple). Also plotted is the spherical beam phase front at 0.5 THz for various locations (green), including at the target surface. (b) Reflected rays (red) from the target are collimated and focused by the lens back to the starting point of the simulation.

Download Full Size | PDF

For each beam-deflection angle defined from $\theta = \; $0° to ${\theta _{max}}$, four optical metrics were used for evaluating a lens’ performance in the simplex; (1) deviation of $\Delta s$ between the focused THz beam phase-front and target curvature, (2) deviation of the angles of the incident rays ${\mathrm{\varphi }_i}$ from 90°, (3) average arc-length spot size $\bar{z}\; $ of the beam on the target and (4) FOV(${\theta _{max}}$) on the spherical target such that the reflected beam returns to the starting point. The initial diverging THz beam has a Gaussian distribution; therefore, a bundle of rays is evaluated for each deflection angle when calculating these four performance metrics. Each ray in the bundle is given a weight as part of a normalized Gaussian distribution defined by the system’s measured beam divergence of 18.4°. Incident/reflected rays that miss either lenses or the target are given a weight of zero. Once the Gaussian-weighted values are calculated, each performance metric is given a subsequent weighting term and added together in a cost-function, CF, given by,

$$CF = ({w1\ast \Delta s} )+ \left( {w2\ast \mathop \sum \limits_{i = 0}^{{\theta_{max}}} |{{{90}^\circ } - {\mathrm{\varphi }_i}} |} \right) + ({w3\ast \bar{z}} )+ \; \left( {w4\ast \frac{1}{{FOV({{\theta_{max}}} )}}} \right),$$
where ${\textrm{w}_{i\; }}$ refers weight of each term in the cost function. This cost function is used for evaluating and updating each lens in the simplex as part of the optimization process. The simplex is updated until all 14 aspheric parameters converge to a design with a minimized CF value. The weight terms for the phase-front and incident angle parameters, i.e., $w1$ and $w2$, respectively, were given the same value of 10 as these quantities are related. Moreover, the contribution of the error terms associated with these two weights, as well as the spot size term i.e., $w3$, are numerically comparable values. Therefore, the weight term for the spot size parameter was also set to 10. Importantly, it should be noted that these three error terms constitute quantities that are also dependent on the numbers of rays used in the Gaussian beam bundle ray tracing. In other words, the arbitrary choice of the number of rays used can effectively scale the values of $w1$, $w2$ and $w3$. In contrast, the FOV parameter is calculated for the entire THz beam and is not scaled by the number of rays used in the analysis. Therefore, the weight of the FOV parameter was adjusted during multiple iterations of the optimization to obtain the largest coverage on the cornea without impacting the other cost function parameters. In this case, a final FOV weight value of 1500 was used to generate the lenses. Table 1 provides the optimized values for the aspheric parameters. A 2-mm thick mounting lip was added to the optimized lens design before machining it on a CNC mill with 1-µm resolution.

Tables Icon

Table 1. Optimized aspheric parameters for front (1st row) and back (2nd row) lens surfaces

3. Results

3.1 Simulation results

Using the described optimization and ray-tracing simulations, a pair of 3-inch diameter HDPE aspheric lenses were designed. The center thickness and front focal length for each lens is approximately 22 mm and 23 mm, respectively. Figure 6 shows the performance metrics of the optimized lens pair on an 8-mm radius target aligned at the focus of the second lens. Figure 6(a) shows a larger symmetric FOV of 9.6 mm can be achieved with a beam deflection angle ${\theta _{max}}\; $ of ${\pm} $40° to scan the lens aperture. With a 3-inch OAPM, a similar FOV of 8-8.8 mm was possible, but was asymmetric in one axes due to the mirror’s parabolic profile [35].

 figure: Fig. 6.

Fig. 6. The performance metrics of the optimized aspheric lens design was calculated using the Gaussian-weighted ray-tracing simulations as a function of the beam deflection angle. (a) The FOV, defined as the arc-length distance on the spherical target from the apex, and the incident power of the Gaussian beam as a function of deflection angle. The power drop after ∼30° is due to clipping of the beam towards the edges of the lenses. (b) Angle of incidence of the incoming beam on the target. (c) Expected beam spot size on the target, defined as an arc length. The dashed lines indicate the starting point where the beam is clipped by the edge of the lenses.

Download Full Size | PDF

Similar beam analysis in Fig. 6(b) indicates that for all deflection angles, the maximum deviation from normal incidence for the incoming beam is less than 0.03°. Finally, the expected average arc-length spot size of the beam on the cornea was determined to be 1.3 mm as shown in Fig. 6(c), which is comparable to the 1-mm spatial resolution measured previously using an OAPM to image a spherical Bohler star target [35]. This analysis shows that beyond a certain deflection angle (dashed lines in Fig. 6(c)), part of the THz beam is clipped by the edge of the lenses; as part of the process for calculating the performance metrics, any rays that extend past this deflection angle are given a weight of zero. For determining the threshold for acceptable beam cutoff, an SNR of 5 (20% of the incident power remaining after clipping) was used.

Compared to a plano-convex lens, our aspheric lenses do not suffer from spherical aberrations or a significant deviation from normal incidence across the entire FOV. Ray-tracing simulations for the optimized lenses indicate excellent matching between the focused THz beam’s spherical phase-front and the target’s curvature across the FOV as shown in Fig. 7. As mentioned before, a sensitivity analysis was performed to evaluate the impact on the optical performance of the lens due to changes in the refractive index, $n$, from design values. As seen in Fig. 8, adjusting the real part of n by ${\pm} $0.003 from the extracted value causes the incoming rays to diverge from normal incidence on the sphere. In addition, towards the edges of the scanning range, the spherical phase-front begins to deviate from the target curvature. This indicates that for any future lens design, the index of refraction of the exact batch of material used for machining must be measured beforehand within a certain degree of accuracy.

 figure: Fig. 7.

Fig. 7. (a) Simulation of the focused THz beam (blue) on the cornea (purple) after passing through a pair of TPX50 plano-convex lenses. The spherical phase front (green) of the beam is not coupled with the target curvature and the reflected beam (red) does not follow the same optical path back to the detector. (b) The focused THz beam on the cornea after passing through our custom designed lenses, which shows phase-front matching along the FOV and no spherical aberrations.

Download Full Size | PDF

 figure: Fig. 8.

Fig. 8. (a) Ray-tracing simulation of the optimized lens design using the average extracted value of the index of refraction, $n$ = 1.532. (b-c) Ray-tracing simulations showing the impact of adjusting n by ${\pm} $0.003, i.e., $n$ = 1.529 and $n$ = 1.535 in (b) and (c), respectively, for the same lens design. (d-f) Deviation of the incident/reflected beam from normal incidence towards the edges of the scanning range as well as the mismatching between the spherical THz phase front and target curvature at the edge of the FOV, marked by the black boxes in (a-c) respectively.

Download Full Size | PDF

3.2 Optical performance characterization

Both lenses were aligned to a co-located THz reflection system as previously shown in Fig. 2. To determine the axial resolution of the scanner and compare the lenses to ray-tracing simulations, an 8-mm radius metallic sphere, shown in Fig. 9(a), was aligned to the focus of the second aspheric lens. THz-TDS imaging of the sphere was performed using the gimballed mirror stage, which consisted of a 360° rotation stage and a goniometer with a maximum deflection angle limit of ${\pm} $20° for steering the beam in the azimuthal and elevation axes, respectively. Given that the maximum deflection angle needed to scan the entire aperture of the lens is ${\pm} $40°, the current iteration of the scanner is limited in the elevation axes only. Future designs will replace the goniometer to utilize the full symmetric FOV of the lenses. To form topographic images, the time-of-arrival (TOA) of the THz-TDS pulse was obtained for all pixels and compared to the TOA of an apex measurement that corresponds to 0° beam deflection. A 200-µm-thick tape cross was placed on the metallic sphere, as shown in Fig. 9(a). The width of the cross strips was 2 mm and the length of the vertical strip was 8 mm. Figure 9(b) shows a topographic image of this target with a 3°-pixel size. As expected, at the pixels located near the edges of the cross, a portion of the focused beam is reflected from the metallic sphere while the rest of the beam is reflected from the air/tape interface. Due to the difference in the surface height, a superposition of the two reflected pulses is detected in these pixels. This superposition results in a decreased amplitude and increased pulse width (FWHM). A 2D Gaussian filter was applied to the THz topographic image to smooth the edges, which is shown in Fig. 9(b), where the tape-cross is clearly distinguishable from the metallic backing within the scanner’s FOV. For extracting surface height variation traces from the image, the FWHM edge-detection filter used in [35] was implemented. Figure 9(c) shows an example trace along the azimuthal axis (red line in Fig. 9(a)) and a reference trace measured on the metallic sphere before mounting the tape (dashed blue line in Fig. 9(a)). This reference measurement, shown as the blue line in Fig. 9(c), can be used to determine the axial resolution of the corneal scanner. Assuming the metallic sphere is smooth, variations in the reference surface profile indicate that this system can measure surface height changes that are larger than 15-20 µm. This minimum surface height detection threshold matches the result previously obtained using an OAPM.

 figure: Fig. 9.

Fig. 9. (a) Photograph of a smooth metallic sphere with a 200-µm-thick tape cross. Also, two azimuthal trajectories are marked with solid red line and dashed blue line for subsequent surface profile measurements, (b) THz topographic image of the cross, which shows surface height variations between the tape and metal backing with respect to a baseline TOA measurement taken at the apex. (c) Surface height variation profiles along the marked traces from the photograph, (a), which show the symmetric FOV and the thickness of the tape in the red trace and the minimum surface height measurement threshold in blue trace.

Download Full Size | PDF

To demonstrate the improvement in the system’s imaging speed, a resin sphere of similar size to the cornea was aligned to the aspheric lenses. In two quadrants contained within the FOV, aluminum foil was taped to the sphere as indicated in Fig. 10(a). Using the new beam-steering approach [38], the normalized peak-peak amplitude of the reflected THz beam from the target was mapped out in about 3-4 seconds, where contrast was shown between the resin and highly-reflective aluminum foil as seen in Fig. 10(b). The total imaging time is determined by several predetermined settings, including the pixel size (set to 3° in both azimuthal and elevation axes), the number of averages taken of the THz-TDS waveform (5 averages), the time per measurement at each pixel (here, 0.05 seconds, limited by ASOPS trace acquisition rate of 100 waveforms per second) and the speed settings of the rotation (253°/second) and goniometer (87°/second) stages used for the gimbal. Future versions of the scanner will replace the ASOPS system with a 2 kHz electronically controlled optical sampling (ECOPS) technique [38], which will reduce the time per measurement at each pixel and reduce the total imaging time to under 2 seconds by reducing the time per pixel to about 2.5 milliseconds.

 figure: Fig. 10.

Fig. 10. (a) A resin ball target with 2 patches of reflective aluminum taped underneath to the surface. (b) THz image of the peak-peak amplitude for all scanned pixels on the resin ball, obtained in about 3-4 seconds (see Visualization 1).

Download Full Size | PDF

4. Conclusion

THz imaging of curved targets, such as the cornea, has previously used parabolic mirrors for focusing the THz beam with normal incidence along the target surface. However, using OAPM provides an asymmetric FOV and results in an extended scan time due to the required beam-steering approach. Here, we present a new hyperbolic-elliptical lens design for focusing and scanning the THz beam with normal incidence onto a spherical target. In addition to ensuring phase-front matching to the surface curvature of the target and achieving an improved symmetric FOV, the total scan time is reduced from minutes to a few seconds. To accomplish this, we introduced a novel spherical beam steering strategy composed of a pair of these hyperbolic-elliptical lenses and a daisy-chained stack of a rotational and a goniometer stage. Our approach to custom-design a spherical scanning lens pair system can be generalized to other imaging applications, for example in detection of skin carcinomas on surfaces such as nose or fingers, and diagnosis of ocular diseases on the sclera of the eye.

Funding

Stony Brook University; National Institute of General Medical Sciences (R01GM112693).

Acknowledgments

The authors thank Dr. Brandon Kovarovic for printing the resin sample and the Stony Brook University Physics Machine Shop for fabricating the lenses.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. X. Chen, H. Lindley-Hatcher, R. I. Stantchev, J. Wang, K. Li, A. Hernandez Serrano, Z. D. Taylor, E. Castro-Camus, and E. Pickwell-MacPherson, “Terahertz (THz) biophotonics technology: Instrumentation, techniques, and biomedical applications,” Chem. Phys. Rev. 3(1), 011311 (2022). [CrossRef]  

2. A. Leitenstorfer, A. S. Moskalenko, T. Kampfrath, J. Kono, E. Castro-Camus, K. Peng, N. Qureshi, D. Turchinovich, K. Tanaka, and A. G. Markelz, “The 2023 terahertz science and technology roadmap,” J. Phys. D: Appl. Phys. 56(22), 223001 (2023). [CrossRef]  

3. A. G. Markelz and D. M. Mittleman, “Perspective on Terahertz Applications in Bioscience and Biotechnology,” ACS Photonics 9(4), 1117–1126 (2022). [CrossRef]  

4. M. H. Arbab, T. C. Dickey, D. P. Winebrenner, A. Chen, M. B. Klein, and P. D. Mourad, “Terahertz reflectometry of burn wounds in a rat model,” Biomed. Opt. Express 2(8), 2339–2347 (2011). [CrossRef]  

5. M. H. Arbab, D. Winebrenner, T. Dickey, A. Chen, M. Klein, and P. Mourad, “Terahertz spectroscopy for the assessment of burn injuries in vivo,” J. Biomed. Opt. 18(7), 077004 (2013). [CrossRef]  

6. O. B. Osman, T. Jack Tan, S. Henry, A. Warsen, N. Farr, A. M. McClintic, Y.-N. Wang, S. Arbabi, and M. H. Arbab, “Differentiation of burn wounds in an in vivo porcine model using terahertz spectroscopy,” Biomed. Opt. Express 11(11), 6528–6535 (2020). [CrossRef]  

7. M. E. Khani, O. Osman, Z. Harris, A. Chen, J.-W. Zhou, A. J. Singer, and M. H. Arbab, “Accurate and early prediction of the wound healing outcome of burn injuries using the wavelet Shannon entropy of terahertz time-domain waveforms,” J. Biomed. Opt. 27(11), 116001 (2022). [CrossRef]  

8. O. B. Osman, Z. B. Harris, M. E. Khani, J. W. Zhou, A. Chen, A. J. Singer, and M. Hassan Arbab, “Deep neural network classification of in vivo burn injuries with different etiologies using terahertz time-domain spectral imaging,” Biomed. Opt. Express 13(4), 1855–1868 (2022). [CrossRef]  

9. M. E. Khani, Z. B. Harris, O. B. Osman, J. W. Zhou, A. Chen, A. J. Singer, and M. H. Arbab, “Supervised machine learning for automatic classification of in vivo scald and contact burn injuries using the terahertz Portable Handheld Spectral Reflection (PHASR) Scanner,” Sci. Rep. 12(1), 5096 (2022). [CrossRef]  

10. A. Gavdush, N. Chernomyrdin, K. Malakhov, S.-I. Beshplav, I. Dolganova, A. Kosyrkova, P. Nikitin, G. Musina, G. Katyba, I. Reshetov, O. Cherkasova, G. Komandin, V. Karasik, A. Potapov, V. Tuchin, and K. Zaytsev, “Terahertz spectroscopy of gelatin-embedded human brain gliomas of different grades: a road toward intraoperative THz diagnosis,” J. Biomed. Opt. 24(02), 027001 (2019). [CrossRef]  

11. A. A. Gavdush, N. V. Chernomyrdin, G. A. Komandin, I. N. Dolganova, P. V. Nikitin, G. R. Musina, G. M. Katyba, A. S. Kucheryavenko, I. V. Reshetov, A. A. Potapov, V. V. Tuchin, and K. I. Zaytsev, “Terahertz dielectric spectroscopy of human brain gliomas and intact tissues ex vivo: double-Debye and double-overdamped-oscillator models of dielectric response,” Biomed. Opt. Express 12(1), 69–83 (2021). [CrossRef]  

12. L. Wu, D. Xu, Y. Wang, B. Liao, Z. Jiang, L. Zhao, Z. Sun, N. Wu, T. Chen, H. Feng, and J. Yao, “Study of in vivo brain glioma in a mouse model using continuous-wave terahertz reflection imaging,” Biomed. Opt. Express 10(8), 3953–3962 (2019). [CrossRef]  

13. S. Yamaguchi, Y. Fukushi, O. Kubota, T. Itsuji, T. Ouchi, and S. Yamamoto, “Brain tumor imaging of rat fresh tissue using terahertz spectroscopy,” Sci. Rep. 6(1), 30124 (2016). [CrossRef]  

14. P. C. Ashworth, E. Pickwell-MacPherson, E. Provenzano, S. E. Pinder, A. D. Purushotham, M. Pepper, and V. P. Wallace, “Terahertz pulsed spectroscopy of freshly excised human breast cancer,” Opt. Express 17(15), 12444–12454 (2009). [CrossRef]  

15. A. Fitzgerald, V. Wallace, S. Pinder, A. Purushotham, P. O’Kelly, and P. Ashworth, “Classification of terahertz-pulsed imaging data from excised breast tissue,” J. Biomed. Opt. 17(1), 016005 (2012). [CrossRef]  

16. V. P. Wallace, A. J. Fitzgerald, E. Pickwell, R. J. Pye, P. F. Taday, N. Flanagan, and T. Ha, “Terahertz Pulsed Spectroscopy of Human Basal Cell Carcinoma,” Appl. Spectrosc. 60(10), 1127–1133 (2006). [CrossRef]  

17. V. P. Wallace, A. J. Fitzgerald, S. Shankar, N. Flanagan, R. Pye, J. Cluff, and D. D. Arnone, “Terahertz pulsed imaging of basal cell carcinoma ex vivo and in vivo,” Br. J. Dermatol. 151(2), 424–432 (2004). [CrossRef]  

18. C. S. Joseph, R. Patel, V. A. Neel, R. H. Giles, and A. N. Yaroslavsky, “Imaging of ex vivo nonmelanoma skin cancers in the optical and terahertz spectral regions Optical and Terahertz skin cancers imaging,” J. Biophotonics 7(5), 295–303 (2014). [CrossRef]  

19. Z. D. Taylor, J. Garritano, S. Sung, N. Bajwa, D. B. Bennett, B. Nowroozi, P. Tewari, J. Sayre, J. P. Hubschman, S. Deng, E. R. Brown, and W. S. Grundfest, “THz and mm-Wave Sensing of Corneal Tissue Water Content: Electromagnetic Modeling and Analysis,” IEEE Trans. THz Sci. Technol. 5(2), 170–183 (2015). [CrossRef]  

20. Z. D. Taylor, J. Garritano, S. Sung, N. Bajwa, D. B. Bennett, B. Nowroozi, P. Tewari, J. W. Sayre, J. P. Hubschman, S. X. Deng, E. R. Brown, and W. S. Grundfest, “THz and mm-Wave Sensing of Corneal Tissue Water Content: In Vivo Sensing and Imaging Results,” IEEE Trans. THz Sci. Technol. 5(2), 184–196 (2015). [CrossRef]  

21. A. Chen, O. B. Osman, Z. B. Harris, A. Abazri, R. Honkanen, and M. H. Arbab, “Investigation of water diffusion dynamics in corneal phantoms using terahertz time-domain spectroscopy,” Biomed. Opt. Express 11(3), 1284–1297 (2020). [CrossRef]  

22. E. A. Kekkonen, A. A. Konovko, E. S. Lee, I. M. Lee, I. A. Ozheredov, K. H. Park, T. N. Safonova, E. I. Sikach, and A. P. Shkurinov, “Assessment of the degree of hydration of ocular surface tissues using THz reflectometry,” Quantum Electron. 50(1), 61–68 (2020). [CrossRef]  

23. I. Ozheredov, M. Prokopchuk, M. Mischenko, T. Safonova, P. Solyankin, A. Larichev, A. Angeluts, A. Balakin, and A. Shkurinov, “In vivo THz sensing of the cornea of the eye,” Laser Phys. Lett. 15(5), 055601 (2018). [CrossRef]  

24. A. Tamminen, M. Baggio, I. I. Nefedova, Q. Sun, S. A. Presnyakov, J. Ala-Laurinaho, E. R. Brown, V. P. Wallace, E. Pickwell-MacPherson, T. Maloney, N. P. Kravchenko, M. Salkola, S. X. Deng, and Z. D. Taylor, “Extraction of Thickness and Water-Content Gradients in Hydrogel-Based Water-Backed Corneal Phantoms Via Submillimeter-Wave Reflectometry,” IEEE Trans. THz Sci. Technol. 11(6), 647–659 (2021). [CrossRef]  

25. M. Mizuno, H. Kitahara, K. Sasaki, M. Tani, M. Kojima, Y. Suzuki, T. Tasaki, Y. Tatematsu, M. Fukunari, and K. Wake, “Dielectric property measurements of corneal tissues for computational dosimetry of the eye in terahertz band in vivo and in vitro,” Biomed. Opt. Express 12(3), 1295–1307 (2021). [CrossRef]  

26. I. Ozheredov, T. Safonova, E. Sikach, M. Mischenko, M. Prokopchuk, A. Larichev, Y. Listopadskaya, and A. Shkurinov, “Potential clinical applications of terahertz reflectometry for the assessment of the tear film stability,” Opt. Eng. 59(06), 1 (2020). [CrossRef]  

27. L. Ke, L. Zhang, N. Zhang, Q. Y. S. Wu, H. S. Leong, A. Abdelaziem, J. S. Mehta, and Y.-C. Liu, “Corneal elastic property investigated by terahertz technology,” Sci. Rep. 12(1), 19229 (2022). [CrossRef]  

28. Y. Hu, M. Baggio, S. O. Dabironezare, A. Tamminen, B. Toy, J. Ala-Laurinaho, E. R. Brown, N. Llombart, S. X. Deng, V. P. Wallace, and Z. D. Taylor, “650 GHz Imaging as Alignment Verification for Millimeter Wave Corneal Reflectometry,” IEEE Trans. THz Sci. Technol. 12(2), 151–164 (2022). [CrossRef]  

29. Z. B. Harris, S. Katletz, M. E. Khani, A. Virk, and M. H. Arbab, “Design and characterization of telecentric f-θ scanning lenses for broadband terahertz frequency systems,” AIP Adv. 10(12), 125313 (2020). [CrossRef]  

30. Z. B. Harris, A. Virk, M. E. Khani, and M. H. Arbab, “Terahertz time-domain spectral imaging using telecentric beam steering and an f-θ scanning lens: distortion compensation and determination of resolution limits,” Opt. Express 28(18), 26612–26622 (2020). [CrossRef]  

31. E. Iomdina, G. Goltsman, S. Seliverstov, A. Sianosyan, K. Teplyakova, and A. Rusova, “Study of transmittance and reflectance spectra of the cornea and the sclera in the THz frequency range,” J. Biomed. Opt. 21(9), 097002 (2016). [CrossRef]  

32. M. E. Khani, O. B. Osman, and M. H. Arbab, “Diffuse terahertz spectroscopy in turbid media using a wavelet-based bimodality spectral analysis,” Sci. Rep. 11(1), 22804 (2021). [CrossRef]  

33. S. Sung, S. Dabironezare, N. Llombart, S. Selvin, N. Bajwa, S. Chantra, B. Nowroozi, J. Garritano, J. Goell, A. Li, S. X. Deng, E. Brown, W. S. Grundfest, and Z. D. Taylor, “Optical System Design for Noncontact, Normal Incidence, THz Imaging of in vivo Human Cornea,” IEEE Trans. THz Sci. Technol. 8(1), 1–12 (2018). [CrossRef]  

34. S. Sung, S. Selvin, N. Bajwa, S. Chantra, B. Nowroozi, J. Garritano, J. Goell, A. D. Li, S. X. Deng, E. R. Brown, W. S. Grundfest, and Z. D. Taylor, “THz Imaging System for in vivo Human Cornea,” IEEE Trans. THz Sci. Technol. 8(1), 27–37 (2018). [CrossRef]  

35. A. S. Virk, Z. B. Harris, and M. H. Arbab, “Development of a terahertz time-domain scanner for topographic imaging of spherical targets,” Opt. Lett. 46(5), 1065–1068 (2021). [CrossRef]  

36. A. Chen, A. Virk, Z. Harris, A. Abazari, R. Honkanen, and M. H. Arbab, “Non-contact terahertz spectroscopic measurement of the intraocular pressure through corneal hydration mapping,” Biomed. Opt. Express 12(6), 3438–3449 (2021). [CrossRef]  

37. A. Chen, Z. B. Harris, A. Virk, A. Abazari, K. Varadaraj, R. Honkanen, and M. H. Arbab, “Assessing Corneal Endothelial Damage Using Terahertz Time-Domain Spectroscopy and Support Vector Machines,” Sensors 22(23), 9071 (2022). [CrossRef]  

38. Z. B. Harris and M. H. Arbab, “Terahertz PHASR Scanner With 2 kHz, 100 ps Time-Domain Trace Acquisition Rate and an Extended Field-of-View Based on a Heliostat Design,” IEEE Trans. THz Sci. Technol. 12(6), 619–632 (2022). [CrossRef]  

39. M. Baggio, A. Tamminen, J. Lamberg, R. Grigorev, S. V. Pälli, J. Ala-Laurinaho, I. Nefedova, J. L. Bourges, S. X. Deng, E. R. Brown, V. P. Wallace, and Z. D. Taylor, “Submillimeter-Wave Cornea Phantom Sensing Over an Extended Depth of Field With an Axicon-Generated Bessel Beam,” IEEE Trans. THz Sci. Technol. 13(2), 132–144 (2023). [CrossRef]  

40. M. Baggio, A. Tamminen, F. Zarrinkhat, I. Nefedova, A. Aspelin, J. Ala-Laurinaho, V. Wallace, E. Brown, and Z. Taylor, “Optical path length measurement of spherical dielectric shells with axicon-generated Bessel beams; applications to corneal sensing,” in 46th International Conference on Infrared and Millimeter Waves (IRMMW-THz, 2021), pp. 1–2.

41. A. Tamminen, S. V. Pälli, J. Ala-Laurinaho, M. Salkola, A. V. Räisänen, and Z. Taylor, “Axicon-hyperbolic lens for reflectivity measurements of curved surfaces,” in 14th European Conference on Antennas and Propagation (EuCAP, 2020) pp. 1–5.

42. A. Tamminen, S. V. Pälli, J. Ala-Laurinaho, M. Salkola, A. V. Räisänen, and Z. D. Taylor, “Quasioptical System for Corneal Sensing at 220–330 GHz: Design, Evaluation, and Ex Vivo Cornea Parameter Extraction,” IEEE Trans. THz Sci. Technol. 11(2), 135–149 (2021). [CrossRef]  

43. F. Zarrinkhat, J. Lamberg, A. Tamminen, M. Baggio, I. Nefedova, J. Ala-Laurinaho, E. E. M. Khaled, J. Rius, J. Romeu, and Z. Taylor, “Vector spherical harmonic analysis and experimental validation of spherical shells illuminated with broadband, millimeter wave Gaussian beams: applications to corneal sensing,” Biomed. Opt. Express 13(7), 3699–3722 (2022). [CrossRef]  

44. T. Yasui, E. Saneyoshi, and T. Araki, “Asynchronous optical sampling terahertz time-domain spectroscopy for ultrahigh spectral resolution and rapid data acquisition,” Appl. Phys. Lett. 87(6), 061101 (2005). [CrossRef]  

45. H. Qin, “Aberration correction of a single aspheric lens with particle swarm algorithm,” Opt. Commun. 285(13-14), 2996–3000 (2012). [CrossRef]  

46. Y. H. Lo and R. Leonhardt, “Aspheric lenses for terahertz imaging,” Opt. Express 16(20), 15991–15998 (2008). [CrossRef]  

47. N. V. Chernomyrdin, M. E. Frolov, S. P. Lebedev, I. V. Reshetov, I. E. Spektor, V. L. Tolstoguzov, V. E. Karasik, A. M. Khorokhorov, K. I. Koshelev, A. O. Schadko, S. O. Yurchenko, and K. I. Zaytsev, “Wide-aperture aspherical lens for high-resolution terahertz imaging,” Rev. Sci. Instrum. 88(1), 014703 (2017). [CrossRef]  

48. F. Gao and L. Han, “Implementing the Nelder-Mead simplex algorithm with adaptive parameters,” Comput. Optim. Appl. 51(1), 259–277 (2012). [CrossRef]  

49. K. Zhou, W. Miao, B. Fan, Y. Delorme, S. Caroopen, M. Batrung, and S. Shi, “Transmittance of high-density polyethylene from 0.1 THz to 15 THz,” in Proc. SPIE (SPIE, 2019).

50. R. Piesiewicz, C. Jansen, S. Wietzke, D. Mittleman, M. Koch, and T. Kürner, “Properties of Building and Plastic Materials in the THz Range,” J. Infrared Millim. Terahertz Waves 28(5), 363–371 (2007). [CrossRef]  

51. M. Naftaly, R. E. Miles, and P. J. Greenslade, “THz transmission in polymer materials — a data library,” in Joint 32nd International Conference on Infrared and Millimeter Waves and the 15th International Conference on Terahertz Electronics (IRMMW-THz, 2007) pp. 819–820.

52. T. D. Dorney, R. G. Baraniuk, and D. M. Mittleman, “Material parameter estimation with terahertz time-domain spectroscopy,” J. Opt. Soc. Am. A 18(7), 1562–1571 (2001). [CrossRef]  

53. R. G. González-Acuña and H. A. Chaparro-Romo, “General formula for bi-aspheric singlet lens design free of spherical aberration,” Appl. Opt. 57(31), 9341–9345 (2018). [CrossRef]  

Supplementary Material (1)

NameDescription
Visualization 1       Video demonstrates our rapid beam steering approach, which is used to image a target similar to the cornea with the custom hyperbolic-elliptical scanning lens pair in about 3-4 seconds.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1.
Fig. 1. (a) Illustration of beam-steering across the cornea using an OAPM. Placing the center of the spherical target at the focus of the OAPM ensures normal incidence. (b) Non-uniform scanning trajectory of the collimated THz beam, shown by black circles, is calculated across the OAPM aperture with constant azimuthal (blue lines) and elevation (red circles) angles on the sphere. Beam steering approach in (b) results in an equiangular sampling across the corneal surface (c). The resultant FOV is approximately 8 mm on the corneal surface. The asymmetry of the FOV in the X-axis can be seen. (d) The spot size variation of the THz beam focused by the OAPM on a spherical target is shown by ABCD ray tracing simulations and is in agreement with previous published results in [33], leading to a dramatic change in spatial resolution on the target surface.
Fig. 2.
Fig. 2. (a) Layout of the single-pixel scanner for imaging curved targets, separated into a THz-TDS collocated reflection setup (red box) and beam-steering arrangement for directing the beam across the aspheric lenses, which focuses the beam onto the aligned spherical target (blue box). (b) A photograph of the setup aligned on an optical table is shown with labelled components.
Fig. 3.
Fig. 3. Block diagram detailing the process for the lens design. The initial inputs to the optimization, including the extracted index of refraction, are used to create a 15-lens simplex. The simplex is updated iterativly using Nelder-Mead optimization, where the cost function is derived from several performance metrics obtained via ray-tracing simulations.
Fig. 4.
Fig. 4. (a) The real index of refraction n of the HDPE slab at 20 different locations extracted using THz-TDS transmission measurements. Plotted in each box is the average value for a bandwidth of .1 to 1 THz. (b) Average extracted index of refraction of HDPE over the twenty spatially-disjointed measurements as a function of frequency with error bars showing the standard error.
Fig. 5.
Fig. 5. Ray-tracing simulation for two aspheric lenses of the same design. (a) A set of incident rays (blue) are steered across and focused by the pair of lenses onto an aligned spherical target (purple). Also plotted is the spherical beam phase front at 0.5 THz for various locations (green), including at the target surface. (b) Reflected rays (red) from the target are collimated and focused by the lens back to the starting point of the simulation.
Fig. 6.
Fig. 6. The performance metrics of the optimized aspheric lens design was calculated using the Gaussian-weighted ray-tracing simulations as a function of the beam deflection angle. (a) The FOV, defined as the arc-length distance on the spherical target from the apex, and the incident power of the Gaussian beam as a function of deflection angle. The power drop after ∼30° is due to clipping of the beam towards the edges of the lenses. (b) Angle of incidence of the incoming beam on the target. (c) Expected beam spot size on the target, defined as an arc length. The dashed lines indicate the starting point where the beam is clipped by the edge of the lenses.
Fig. 7.
Fig. 7. (a) Simulation of the focused THz beam (blue) on the cornea (purple) after passing through a pair of TPX50 plano-convex lenses. The spherical phase front (green) of the beam is not coupled with the target curvature and the reflected beam (red) does not follow the same optical path back to the detector. (b) The focused THz beam on the cornea after passing through our custom designed lenses, which shows phase-front matching along the FOV and no spherical aberrations.
Fig. 8.
Fig. 8. (a) Ray-tracing simulation of the optimized lens design using the average extracted value of the index of refraction, $n$ = 1.532. (b-c) Ray-tracing simulations showing the impact of adjusting n by ${\pm} $0.003, i.e., $n$ = 1.529 and $n$ = 1.535 in (b) and (c), respectively, for the same lens design. (d-f) Deviation of the incident/reflected beam from normal incidence towards the edges of the scanning range as well as the mismatching between the spherical THz phase front and target curvature at the edge of the FOV, marked by the black boxes in (a-c) respectively.
Fig. 9.
Fig. 9. (a) Photograph of a smooth metallic sphere with a 200-µm-thick tape cross. Also, two azimuthal trajectories are marked with solid red line and dashed blue line for subsequent surface profile measurements, (b) THz topographic image of the cross, which shows surface height variations between the tape and metal backing with respect to a baseline TOA measurement taken at the apex. (c) Surface height variation profiles along the marked traces from the photograph, (a), which show the symmetric FOV and the thickness of the tape in the red trace and the minimum surface height measurement threshold in blue trace.
Fig. 10.
Fig. 10. (a) A resin ball target with 2 patches of reflective aluminum taped underneath to the surface. (b) THz image of the peak-peak amplitude for all scanned pixels on the resin ball, obtained in about 3-4 seconds (see Visualization 1).

Tables (1)

Tables Icon

Table 1. Optimized aspheric parameters for front (1st row) and back (2nd row) lens surfaces

Equations (3)

Equations on this page are rendered with MathJax. Learn more.

z ( r ) = r 2 / R 1 + 1 ( 1 + k ) ( r / R ) 2 + a 4 r 4 + a 6 r 6 + a 8 r 8 + a 10 r 10 + ,
1 f = ( n 1 ) [ 1 R 1 1 R 2 + ( n 1 ) d n R 1 R 2 ] ,
C F = ( w 1 Δ s ) + ( w 2 i = 0 θ m a x | 90 φ i | ) + ( w 3 z ¯ ) + ( w 4 1 F O V ( θ m a x ) ) ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.