throbber
Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Fundamentals of macro axial gradient index
`optical design and engineering
`
`Paul K. Manhart, MEMBER SPIE
`Richard Blankenbecler, MEMBER SPIE
`Cheshire Optics
`974 Cottrell Way
`Stanford, California 94305
`E-mail: pmanhart@worldnet.att.net,
`rblankenbecler@worldnet.att.net
`
`Abstract. Homogeneous lens material is characterized by an index of
`refraction and a point on the glass map nd⫽f(␷d). Gradient refractive
`index (GRIN) lenses have a spatially varying index and dispersion and
`are represented by a line on the glass map. GRIN lenses open the door
`to a wide variety of optical design applications incorporating entire lenses
`of axial gradient refractive material (macro-AGRIN). Axial gradient mate-
`rial essentially gives biaspheric behavior to lenses with spherical sur-
`faces and exhibits a controlled gradient in both index and dispersion.
`Thus, the applications for this material range from simple singlet lens
`used for imaging laser light, in which spherical aberration is eliminated,
`to complex multielement lens systems, where improved overall perfor-
`mance is desired. The fusion/diffusion process that produces this mate-
`rial is surprisingly simple, repeatable, and applicable to mass production.
`The advantages of AGRIN technology coupled with the recent advances
`in material development and its accessibility in commercially available
`lens design programs provides optical designers with the opportunity to
`push the performance of optical systems farther than with conventional
`optics.
`© 1997 Society
`of Photo-Optical
`Instrumentation Engineers.
`[S0091-3286(97)00406-6]
`
`Subject terms: optical materials; optical systems; gradient index; axial gradient
`refractive index; gradient refractive index; radial gradient refractive index; macro
`axial gradient, radial gradient; optical design; glass lines; unique optical systems;
`high performance; asphere.
`
`Paper MAT-04 received Aug. 1, 1996; revised manuscript received Nov. 21,
`1996; accepted for publication Dec. 16, 1996.
`
`1 Introduction
`Traditional optical systems are based on the use of homo-
`geneous refractive materials with spherical surfaces. A
`spherical surfaced lens will produce images with intrinsic
`optical aberrations. Basic optical principles lead to the use
`of multiple elements in a lens system to provide imagery
`across a required field with acceptably low aberration con-
`tent. Fabrication of spherical surfaces can be carried out
`using several traditional methods to extremely high accu-
`racy and smoothness.
`Improved images with reduced aberrations can be ob-
`tained by using aspheric surfaces. The production and test-
`ing of nonspherical surfaces is more complex and consid-
`erably more
`expensive
`than for
`spherical
`surfaces.
`Generally, aspheric surfaces will cost about 10 times that of
`a spherical surface. Nevertheless, cost efficient mass pro-
`duction of optical elements with aspheric surfaces can be
`accomplished with plastics or glasses, usually to acceptable
`accuracy for some applications. While mass production of
`molded aspherics can be cost competitive, it is subject to
`extremely high startup costs, limited material selection and
`small diameters.
`Axial gradient refractive index 共AGRIN兲 offers an ap-
`proach that would build specific aspheric effects into the
`refractive properties of the material while permitting the
`use of traditional spherical surface finishing methods. The
`refractive power at each point on a surface is determined by
`
`Snell’s law, given in Eq. 共1兲, where ␪ is the angle of inci-
`dence the ray makes
`
`n sin ␪⫽n⬘ sin ␪⬘
`
`共1兲
`
`with the surface normal, ␪⬘ is the angle after refraction, n is
`the index of refraction of the incident medium and n⬘ is the
`index of the refractive medium. The shape of the surface
`determines the progressive change in angle of incidence
`across the aperture that leads to image formation. This
`same relation determines the intrinsic aberration that will
`be produced by the surface. Additional control of the ray
`passage through a surface can be gained by using a material
`in which the index of refraction varies with the location on
`the surface and/or within the medium. If the refractive in-
`dex varies along the optical axis, it is known as an axial
`gradient. If the index varies with radius, it is known as a
`radial gradient.
`
`2 Axial Gradients
`The term macro-AGRIN refers to a solid piece of axial
`gradient optical material with relatively large diameter,
`thickness and change in index (⌬n兲, and is used inter-
`changeably with the term AGRIN throughout this paper.
`
`Opt. Eng. 36(6) 1607–1621 (June 1997)
`
`0091-3286/97/$10.00
`
`© 1997 Society of Photo-Optical Instrumentation Engineers
`
`1607
`
`1
`
`APPLE 1020
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`Macro-AGRIN is made via a high temperature fusion/
`diffusion process described later in this text that turns a
`series of discrete glasses on the glass map into glass lines
`共see Fig. 1兲. An axial gradient is typically defined by a
`polynomial function in z,
`the distance along the optical
`axis. A common expression for this type of gradient is
`
`n 0⫽n 00⫹n 01z⫹n 02z 2⫹n 03z 3⫹ . . .
`
`.
`
`共2兲
`
`The most significant way an axial gradient affects aberra-
`tions is by the variation of index over the curved surface of
`the lens. Snell’s law states that there are only two factors
`that determine how light is bent when it hits a glass surface.
`The first factor is the angle of incidence ␪of the ray on the
`surface, and the second factor is the index of refraction
`n⬘ at the point of contact. An aspheric surface controls
`aberrations by changing the localized radius of curvature
`共surface normal兲 on the lens surface, thus changing the
`angle of incidence for a particular ray. An axial gradient
`lens affects aberrations by leaving the surface spherical and
`changing the index of refraction through the material.
`When a curved surface is put on a plane parallel piece of
`this material, the gradient is exposed because the curve of
`the surface exposes different depths z in the material. This
`results in a changing index value tracking radially outward
`along the curved surface. A lens made from a solid block of
`axial gradient material has two surfaces exposing the gra-
`dient 共front and back兲, thereby giving each lens the equiva-
`lent of two aspheric surfaces. The contribution to aberration
`at the surface of a lens, due to a gradient material is re-
`ferred to as a surface term. In addition to the two surface
`terms, there is a transfer term as the wavefront propagates
`through the lens, between the two surfaces. In this region,
`the light travels in a curve as the wavefront encounters a
`continuous change in index and dispersion, identified by
`the Abbe number,
`
`␯d⫽
`
`n d⫺1
`n F⫺n C
`
`.
`
`共3兲
`
`It is important to think in terms of wavefronts, as opposed
`to rays, when understanding how light travels through a
`gradient material. A wavefront always bends toward the
`high index, where light travels slower, similar to a march-
`ing band turning a corner.
`Adding up the contributions from the two surface terms
`and the transfer term can give AGRIN an advantage over
`aspheric lenses. In addition, the spherical surfaces of lenses
`made of AGRIN are self generating, easy to test, plate fit to
`high accuracy and no more expensive to fabricate than con-
`ventional surfaces. The effect on aberrations from a macro-
`AGRIN lens can be adjusted by changing the lens shape
`factor or gradient profile1 关Eq. 共2兲兴. A multielement lens
`system can benefit from a gradient index lens by relaxing
`the requirements of the other lenses. The gradient material
`essentially acts like a ‘‘work sponge,’’ soaking up a portion
`of the work that would have previously been distributed
`among all the elements. Designing functionality into an op-
`tical material helps relax optical systems and paves the way
`to improved performance. Improved performance can be in
`the guise of smaller, cheaper, sharper, lighter or stronger
`products.
`
`1608 Optical Engineering, Vol. 36 No. 6, June 1997
`
`The importance of gradient index materials have been
`realized for over 40 years.2–4 Extensive theoretical explo-
`ration and conceptual designs can be found in many scien-
`tific articles.5–8 The poor reputation of GRIN material,
`however, comes from stagnation in the material develop-
`ment. Small size 共up to 3 mm兲, poor repeatability, and high
`cost are associated with GRIN materials. Four major differ-
`ent processes, i.e., ion exchange, sol-gel, chemical vapor
`deposition 共CVD兲, and high temperature fusion/diffusion
`are currently used to produce GRIN materials. Recogniz-
`able commercial products are Selfoc™ produced by Nip-
`pon Sheet Glass via an ion-exchange process, graded-index
`optical fibers via the CVD process and axial gradient laser
`singlets via fusion/diffusion. Theoretically, radial GRIN is
`more versatile than AGRIN because the radial refractive
`index gradient directly contributes to the optical power,
`Petzval and spherical aberration of the optical system.
`However, by the same token, this makes the performance
`of the radial GRIN lens more sensitive to manufacturing
`variations and repeatability in the gradient index profile
`than AGRIN lenses. A radial gradient lens will directly
`affect the first and higher order properties of an optical
`system because the gradient is perpendicular to the wave-
`front propagation. Repeatability is an important issue with
`radial GRIN.
`The contribution to aberration correction from the trans-
`fer term of a thin axial gradient lens is secondary in that the
`gradient does not directly contribute to the optical power or
`aberrations of the lens, but rather, it is the exposure of the
`gradient along a curved surface that is responsible for most
`of the aberration correction. The gradient in this case is
`essentially parallel to the wavefront propagation, so the
`transfer term is not as dominant as it is with radial gradient.
`Therefore, the AGRIN lens performance is more tolerant to
`small manufacturing variations in the profile compared to
`the radial GRIN lens.
`
`3 Fusion/Diffusion Process (Macro-AGRIN)
`A process to produce solid blocks of AGRIN is based on
`controlled fusion/diffusion of different glasses at a high
`temperature. The process, developed by Hagerty and
`Pulsifer,9 starts by identifying a specific family of glass
`compositions that span the desired range in optical index
`and dispersion. AGRIN glass is created by fusing together a
`stack of discrete molten glass layers, where each constitu-
`ent layer has a distinctive composition and desired optical
`property. After a controlled diffusion process at high tem-
`perature 共see Fig. 2兲, the glass layers fuse and diffuse into
`one piece of gradient index glass with a smooth variation of
`optical properties throughout. Prescribable index profiles
`共linear, quadratic and cubic兲, in lead and crown glass com-
`positions have been achieved in large diameters, thickness
`and index variation.10–13 The initial step function index
`profile of the layers is eliminated via the diffusion of con-
`stituent atoms within and across the layer boundaries. This
`process results in a smooth variation of index. By varying
`the index and thickness of each layer, and by controlling
`diffusion temperature and time, AGRIN can be produced
`for arbitrary monotonic index profiles. The profiles are
`typically monotonic because of the separation of the con-
`stituent molten glasses based on density. High temperature
`diffusion 共1100 °C兲 yields high diffusion rates for all com-
`
`2
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`Fig. 1 Glass map showing refractive index as a function of Abbe number.
`
`ponents in the glass, therefore, a wide variety of glass fami-
`lies can be used to produce AGRIN with different index/
`dispersion relationships 关n d⫽ f (␷d)兴. Profiles with steep
`slopes (dn/dz), experience more stress than profiles with
`small slopes. This stress is due to mismatches in the coef-
`ficient of thermal expansion of the constituent glasses after
`diffusion. The resultant stress manifests itself as residual
`asymmetric radial terms in the AGRIN material, limiting
`the useful aperture over which an optical quality wavefront
`can be obtained. Steeper sloped gradients have smaller use-
`ful diameter than shallower sloped gradients because of the
`residual stress associated with a more rapid change in ma-
`terial composition.
`Two key factors of the fusion/diffusion process are 共1兲
`the selection of base glass compositions and 共2兲 the diffu-
`sion process. The glass compositions must be compatible
`with each other, so that the compositional changes due to
`the diffusion between glass layers will not cause undesir-
`able property changes, such as thermal expansion, phase
`separation and devitrification. The step function-like index
`profile between layers is eliminated via the diffusion of
`constituent atoms within and between the layers during the
`process; this eventually results in smooth variation in index
`
`as well as other physical and material properties. The pro-
`cess is illustrated in Fig. 2.
`
`4 Diffusion Software
`To prescribe the index profile, it is necessary to 共1兲 estab-
`lish the relation between glass properties, such as refractive
`index, dispersion,
`thermal expansion coefficients, glass
`transition temperature, and glass composition, and 共2兲 un-
`derstand and simulate the diffusion process. The design
`problem for prescribing the refractive index profile is to
`find the desirable glass family and an achievable initial in-
`dex distribution n(z,0) that will yield the desired index
`profile n(z,t) to the accuracy required after diffusion at
`temperature T for a time t. In the fabrication methods al-
`ready described, the index of each layer can be fixed for a
`given sample batch; the design variables are then the thick-
`ness of each layer, the diffusion temperature T and time
`t. Diffusion simulation software14,15 determine these pro-
`cess variables quickly. The diffusion model is based on
`multicomponent interdiffusion theory. All constituents of
`
`Optical Engineering, Vol. 36 No. 6, June 1997
`
`1609
`
`3
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`Fig. 2 Fusion/diffusion process for axial gradient material fabrication.
`
`the glass participate in diffusion at high temperature. The
`diffusion constants of these constituents are also functions
`of their elemental concentrations.
`The change in the index distribution during the diffusion
`process was simulated using the diffusion software and is
`illustrated in Fig. 3 for a seven-layer sample. It shows that
`diffusion at the high index region is much faster than at the
`low index side. This is expected because of the higher lead
`densities in the higher index material. After being diffused
`at high temperature for the required time, the index profile
`becomes smooth and has a linear index gradient region be-
`tween 1.580 and 1.750. Using the parameters established
`by the diffusion software, the calculated index profiles are
`
`in excellent agreement with the produced index profiles,
`demonstrating the effectiveness of the software in prescrib-
`ing the index profile for specific optical designs.10–12 Figure
`4 plots the calculated index profile along with the measured
`data of the produced profile. The results from different dif-
`fusion experiments also show excellent repeatability in
`both the index profile and optical quality.
`The unique aspect of this process for producing macro
`gradient index glass is that layers of glass possessing dif-
`ferent physical properties can be fused together to produce
`a repeatable predetermined index gradient. The controlled
`diffusion of glass components such as Pb, Ba and La, pro-
`vides a gradual transition in properties from layer to layer.
`
`Fig. 3 Diffusion simulation for seven-layer profile as a function of diffusion time t at 1000 °C.
`
`1610 Optical Engineering, Vol. 36 No. 6, June 1997
`
`4
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`Fig. 4 Comparison of the produced index profile with the profile calculated from diffusion software.
`
`Carefully chosen glasses enable this high temperature dif-
`fusion process to produce gradient
`index glass having
`macro size, high ⌬n and good wavefront quality.
`
`5 Optical Design
`Macro-AGRIN offers optical designers with a unique op-
`portunity to incorporate lenses made from solid blocks of
`gradient material into their optical systems. Gradient mate-
`rials are described by profiles. For axial gradients, the pro-
`file is a plot of the index of refraction at a reference wave-
`length 共typically 0.58756 ␮m兲, as a function of axial
`position within the blank 关Eq. 共2兲兴. Profiles can be for either
`increasing or decreasing index as a function of position.
`The slope is negative for a decreasing index and positive
`for an increasing index. Some fundamental insights into the
`behavior of lenses made from this material follow, along
`with a process to define and model the dispersive properties
`of the gradient.
`
`5.1 LinearVersusNonlinearProfile
`New issues have emerged while designing optical systems
`with macro-AGRIN materials. For instance, it has been
`found that nonlinear profiles are more advantageous than
`linear profiles because the relative slope of the gradient
`over each lens surface can be different.1 Figure 5共a兲 shows
`calculated third order spherical aberration, 共SA3兲 of an
`F/3 singlet lens, plotted as a function of lens shape factor
`X, for a linear gradient range of ⫺0.4⭐⌬n⭐0.4. The dif-
`ferent curves represent different slopes, or ⌬n’s. For a lens
`with the object at infinity, minimum SA3 occurs near a
`shape factor, or bending, of ⫹1.0, which corresponds to a
`convex first surface and a flat rear surface. To correct an
`
`F/3 singlet for SA3 using a linear gradient, ⌬n must be
`between ⫺0.03 and ⫹0.25. Once that condition is met,
`there exists a solution for zero SA3. The most prominent
`aspect of this plot however, is that the curves tend to pivot
`about a point that has a shape factor of zero 共equiconvex/
`convex兲. The phenomena that we call the pivot point has
`been observed for all cases analyzed, including root mean
`square 共rms兲 spot size, coma, astigmatism and distortion.
`The pivot point also exists for spherical type gradient ge-
`ometry’s 共but shifts as a function of the exact radius of the
`isoindicial surfaces兲. In essence, the existence of the pivot
`point implies that a monotonic linear profile throughout the
`material has no effect in controlling SA3 when the shape
`factor of the lens is near zero. Wang hints at this in his
`dissertation.16 For an equiconvex lens with a linear axial
`gradient, the net effect of the gradient on spherical aberra-
`tion is zero because the two spherical surface curvatures, of
`equal and opposite sign, have the same gradient slope over
`the saggita z 共sag兲, given by
`
`z⫽
`
`cr 2
`1⫹ 共1⫺c 2r 2兲1/2 .
`
`共4兲
`
`The contribution to spherical aberration from one gradient
`surface is canceled by the other gradient surface. For a lens
`bending ⫺1⭐X⭐⫹1, a linear axial gradient lens’ optical
`performance is compromised by one of the surfaces. Me-
`niscus lenses, however, have the property that the gradient
`contribution to spherical aberration is additive on both sur-
`faces because their curvatures are of the same sign.
`The existence of the transfer term, and the fact that the
`marginal ray heights are different on each surface for a
`
`Optical Engineering, Vol. 36 No. 6, June 1997
`
`1611
`
`5
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`Fig. 5 Third order spherical aberration versus lens shape factor for (a) a linear gradient profile and (b)
`for a quadratic gradient profile.
`
`1612 Optical Engineering, Vol. 36 No. 6, June 1997
`
`6
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`dient slope over the two surfaces of a lens. The first, Fig.
`6共a兲, shows two separate slopes, of opposite sign, over each
`surface. In this case, the gradient surface terms are additive
`for shape factors between ⫾1 and subtractive for meniscus
`shapes. Processes such as ion exchange will produce such
`profiles, of equal but opposite slope, but the small penetra-
`tion depth of gradient into the material limits the useful
`diameter of the optics. Figure 6共c兲 shows a grin cap,10
`which essentially puts a gradient over one side of the lens
`and makes the other side homogeneous. In the case of a
`grin cap lens there is only one surface term. To make a grin
`cap lens, two lenses would be made 共one gradient and one
`homogeneous兲 and cemented together. This adds complex-
`ity to the fabrication process. Figure 6共b兲 shows a natural
`quadratic profile that starts out shallow on the first surface
`and ends up steep on the second surface. The fusion/
`diffusion process that makes macro gradient naturally
`comes out in a sigmoid profile, flat on the top, steepest in
`the center, and flat at the end. Careful positioning of the
`lens within this type of profile allows for many more solu-
`tions to SA3 than with a linear gradient profile.
`
`5.2 AsphereVersusAGRIN
`To correct for a given amount of spherical aberration, the
`steepness of a gradient slope scales inversely as a function
`of lens aperture. The smaller the lens, the steeper the slope
`required to do the same amount of work as a larger lens.
`For example, the amount of ⌬n required to correct the
`spherical aberration of a 10 mm diameter f /2 singlet is the
`same as that required for a 50 mm diameter f /2 singlet.
`Since ⌬n is a constant, but ⌬z, the difference in sag be-
`tween the two lenses is not, the slope ⌬n/⌬z scales in-
`versely with aperture. Some of the issues relating the lens
`design process with the material fabrication process include
`practical limits to lens diameter and the quality of the trans-
`mitted wavefront. From a manufacturing perspective,
`steeper sloped materials are currently better suited to
`smaller aperture optics and shallow sloped materials may
`be used for larger diameter optics. The reason for this, is
`that residual thermal stress in the material is a function of
`⌬n and slope. Steeper sloped materials have higher stress
`which results in stronger residual radial terms distorting the
`transmitted wavefront in an asymmetric way. The higher
`stress results from the material changing rapidly over a
`short distance . To maintain comparable wavefront quality
`to that of shallower sloped material, the aperture must be
`reduced. Since this is a material and manufacturing issue,
`and not a fundamental
`limitation,
`these issues may be
`worked out in time.
`To compare a lens with an aspheric surface to a macro-
`AGRIN lens requires some definitions. A direct comparison
`of AGRIN material to an asphere is difficult because of the
`transfer term associated with gradient material. To isolate
`the surface term of an AGRIN lens, it is necessary to ex-
`amine a plano-convex lens, with the plano side facing a
`collimated wavefront. In this special case, the transfer term
`is zero and the plane wave propagates through the axial
`gradient undeviated. The wavefront is only affected by the
`gradient on refraction at the second surface, where the
`variation of index over the curved surface affects the bend-
`ing of light differently than a constant index would. A
`
`Optical Engineering, Vol. 36 No. 6, June 1997
`
`1613
`
`Fig. 6 Three methods of altering gradient slope over each curved
`surface of a lens: (a) two cemented linear gradients, (b) nonlinear
`profile, and (c) grin cap.
`
`given ray, shifts the pivot point slightly away from the zero
`shape factor. The proximity of the pivot point to the zero
`shape factor is direct evidence that the inhomogeneous
`transfer term is small compared to the surface term, and
`thus not a significant variable for monochromatic designs
`of thin AGRIN lenses. One consequence of the pivot point
`is that SA3 can be made to change sign with bending if
`⌬n is large enough. Another, is that less ⌬n is required for
`positive bendings than for negative bendings to achieve this
`sign reversal. This is a direct consequence of the minimum
`SA3 occurring for a shape factor near ⫹1.0 for a homoge-
`neous lens. The pivot point provides useful insight when
`designing optical systems with this material. To maximize
`the potential of macro-AGRIN, lens shape factors near zero
`should be avoided, and lenses with shape factors between
`⫾1 should be viewed with suspicion. There are situations
`however, in which it is desirable to have a lens with a shape
`factor near zero. Such situations might exist when trying to
`incorporate gradients into existing designs with minimal
`amount of redesign, when trying to balance other aberra-
`tions, or when working with finite conjugates. To effec-
`tively apply axial gradients to near equiconvex or concave
`lenses the gradient contribution over each surface must be
`changed by adjusting the relative slope of the gradient over
`each of the lenses curved surface.
`Nonlinear sigmoid type gradient profiles can help ad-
`dress the problem of the pivot point cancellation. The slope
`of a nonlinear gradient may start out high and end low or
`visa versa. The nature of a nonlinear gradient therefore al-
`lows the GRIN aberration contribution on each surface to
`be changed by altering the relative slope of the gradient
`over each surface. Thus, nonlinear axial gradients are more
`effective than linear axial gradients because there is less
`compromise for lens bendings between ⫾1. Figure 5共b兲
`shows the same plot as Fig. 5共a兲 for a nonlinear, monotonic
`共quadratic兲, gradient profile. There are many more solutions
`for zero SA3 using nonlinear gradients than for linear gra-
`dients. Note in Fig. 5共b兲 that SA3 does not change sign as it
`does for the linear gradients in Fig. 5共a兲. Once a particular
`⌬n hits zero SA3, it stays there for the remaining shape
`factors. Although the pivot point does not change with non-
`linear gradients, there are many more useful solutions at
`other shape factors.
`Figure 6 shows several ways to change the relative gra-
`
`7
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`Fig. 7 Delta index required to correct spherical aberration as a function of lens f-number.
`
`plano-convex lens such as this, with a shape factor
`X⫽⫺1 has a fair amount of negative spherical aberration.
`The shape factor for minimum SA3 of a homogeneous lens
`is near a convex-plano lens (X⫽⫹1), where the convex
`side faces the infinite conjugate. The amount of gradient
`and/or asphericity required to correct a plano-convex lens
`for SA3 will be more than it would for a convex-plano, but
`the relative difference between ⌬n and maximum aspheric-
`ity reflect the work that the isolated surface terms are con-
`
`tributing to spherical aberration. Figure 7 shows a plot of
`⌬n required to correct spherical aberration as a function of
`lens f -number for surface contribution only. The ⌬n values
`in these plots show what is required to correct all orders of
`spherical aberration for a one inch diameter, f /2 singlet of
`shape factor X⫽⫾1.
`Figure 8 shows the same plot as Fig. 7, for an aspheric
`lens of shape factor X⫽⫾1. The y axis represents the
`
`Fig. 8 Normalized aspheric sag required to correct spherical aberration as a function of
`f-number.
`
`lens
`
`1614 Optical Engineering, Vol. 36 No. 6, June 1997
`
`8
`
`

`

`Downloaded From: https://www.spiedigitallibrary.org/journals/Optical-Engineering on 28 Dec 2022
`Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
`
`Manhart and Blankenbecler: Fundamentals of macro axial gradient index . . .
`
`maximum amount of aspheric sag deviation required to cor-
`rect spherical aberration. Since the maximum aspheric sag
`of a surface scales with radius of curvature, the aspheric
`sag values plotted in Fig. 8 are normalized to the radius of
`curvature of the aspheric surface.
`
`6 Dispersion Modeling: A Walk Along the Glass
`Line
`All points within homogeneous elements have the same
`value of index and are the same function of wavelength.
`GRIN elements have a spatially varying index function.
`The GRIN transfer term depends on the spatial derivative
`of the index. The index and its spatial derivative have dif-
`ferent dispersion properties. This difference allows chro-
`matic corrections to be dealt with in new ways in systems
`with GRIN elements.
`Commercially available optical design software has
`been able to trace rays through an inhomogeneous material
`for many years. Some of the more popular profiles and
`geometries are programmed for user convenience. When a
`gradient geometry is selected 共such as axial, radial, or
`spherical兲, the user may define the coefficients to describe a
`specific index variation 共profile兲, throughout the material.
`These coefficients may also be variables during optimiza-
`tion.
`It is generally known that index of refraction for optical
`materials is a function of wavelength.17 This is also true for
`macro-AGRIN material where the index n is a function of
`Abbe number ␷ 关Eq. 共3兲, Fig. 1兴. The index profile is typi-
`cally described by a polynomial 关Eq. 共2兲兴. Because of dis-
`persion, the coefficients describing this profile vary accord-
`ing to wavelength. Designing white light optical systems
`with gradients requires the designer to control the relation-
`ship between ⌬n and ⌬␷if the resultant gradient material is
`to be realistic. Since AGRIN is represented by glass lines,
`or monotonic curves, on the glass map, this relationship can
`be predefined for each specific family of glasses and thus
`transparent to the optical designer. Each constituent glass
`comprising the family has its own unique set of Buchdahl
`coefficients describing its dispersion. Fitting these Buch-
`dahl coefficients to a function in index through the full
`range of the material links the dispersion coefficients to
`material position within the material.
`Given the Buchdahl dispersion expression as follows,
`
`␦n⫽V 1␻⫹V 2␻2⫹V 3␻3⫹V 4␻4,
`
`where
`
`␦n⫽n i⫺n d , ␻⫽
`
`␦␭
`1⫹␤共␦␭ 兲 , ␤⫽
`
`1.0
`␭ d⫺0.187
`
`,
`
`␦␭⫽␭ i⫺␭ d ,
`
`␭ d⫽0.58756 ␮m.
`
`Each glass in the family has its own set of coefficients,
`V 1 , V 2 , V 3 , and V 4 , to describe its dispersion. To model
`the gradual variation in chromatic properties within the gra-
`dient material after each glass has been fit to Buchdahl, the
`coefficients V i are fit to a cubic function in n d , defined as
`the index at the reference wavelength 共typically 587.56
`nm兲. The result looks like
`
`
`
`2兲⫹␥i共n d兲⫹␦i .V i⫽␣i共n d3兲⫹␤i共n d
`
`
`
`共5兲
`
`Since the Buchdahl fit in this case is carried out to the
`fourth term, there are four values of alpha, beta and gamma
`and delta each. Once n d(z) is known from the reference
`profile through the material 关Eq. 共2兲兴, the Buchdahl coeffi-
`cients can be calculated at that location. Once the Buchdahl
`coefficients are known, so is the index at any wavelength.
`This methodology has been implemented in several of the
`commercially available software codes where each AGRIN
`glass line has its own unique name that the designer can
`call out. Once the type of AGRIN glass is chosen 共flint,
`crown etc.兲, the designer only needs to be concerned with
`the index profile at 587.56 nm wavelength. The software
`will fill in the indices at all the system wavelengths. This
`model contributes only 1⫻10⫺5 uncertainty17 in index and
`may be implemented from catalog data of constituent
`glasses prior to diffusion, or measured index data at several
`wavelengths after diffusion.17
`For lens designers who want to use off-the-shelf AGRIN
`profiles, a gradient surface type has been developed allow-
`ing the use of predefined profiles that can be called by
`name. This capability can be found in CODE V 共CODE V
`Optical Design Software is a product of Optical Research
`Associates, Pasadena, California兲, OSLO 共OS

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket