You are on page 1of 16

Research Article Vol. 60, No.

3 / 20 January 2021 / Applied Optics 681

Advancing lightweight mirror design: a paradigm


shift in mirror preforms by utilizing design for
additive manufacturing
Nicholas Horvath1, * AND Matthew Davies2
1
Energy and Transportation Science Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37830, USA
2
Department of Mechanical Engineering and Engineering Science, The University of North Carolina at Charlotte, Charlotte, North Carolina 28223,
USA
*Corresponding author: horvathnw@ornl.gov

Received 16 September 2020; revised 16 December 2020; accepted 17 December 2020; posted 17 December 2020 (Doc. ID 410350);
published 18 January 2021

Additive manufacturing is a disruptive technology that can be leveraged by the redesign of components in most
engineering fields. Fundamental engineering resources for lightweight mirrors were developed more than 30 years
ago with a main design limitation, state of the art manufacturing. Here, we present two design methodologies for
the design of lightweight mirrors. The first method utilizes analytical expressions to design a traditional isogrid
mirror, which provided the foundation for most lightweight mirrors to date. The second method employs a combi-
nation of topology optimization, lattice infill, and analytical estimation to develop an advanced lightweight mirror
designed for additive manufacturing. The advanced mirror design outperforms the traditional design for each
functional requirement, including a 94% reduction in predicted surface quilting and a higher specific stiffness. The
manufacturing of the advanced mirror is only possible with an additive manufacturing process. © 2021 Optical
Society of America

https://doi.org/10.1364/AO.410350

1. INTRODUCTION over an isogrid unit cell [4]. Mehta expanded on this work to
Open back mirrors are one of the most often used designs for include the equivalent bending thickness to calculate flexural
lightweight mirrors in telescope design because of their high rigidity for both open and closed back mirrors, along with opti-
stiffness to mass ratio. Most designs utilize an isogrid support mizations based on state of the art manufacturing limitations at
structure on the backside of the mirror to achieve the stiffness that time [5]. Vukobratovich and Valente began implementing
to mass ratio requirements. These traditional isogrid mirror finite element (FE) methods to further analyze designs that used
designs are often two-dimensional (2D) geometries that were analytical expressions for lightweight mirrors. Additionally, the
developed such that they are readily milled or cast to net shape. previous work included a review of arch back mirror designs
However, introducing additive manufacturing (AM) as a new and the trade-offs, including manufacturing difficulty [6,7].
process chain for mirror preforms could offer profound advan- An intrinsic trade-off with isogrid designs is residual surface
tages in both the mechanical design of the mirror by using lattice deformations created during the manufacturing process, known
structures and topology optimization, along with a reduction as “quilting.” These deformations are caused by the reduced
in lead time. Sweeney et al. provided case studies using multiple stiffness of the unsupported facesheet, which is localized over
AM methods with varying material choices for mirror designs the unit cells, between the rib structure on the optical surface
and discussed findings such as stress relief during polishing, isot- [8,9]. The quilting of an isogrid structure has been addressed
ropy of thermo-mechanical properties, and reduced lead time in various ways thus far, one method includes extra support
prototyping [1]. While other work has shown other successes between the isogrid rib structure, known as cathedral ribbing
in additively manufactured mirror preforms that were then [10,11]. Cathedral ribs reduce the unsupported area of the
post-polished [2,3], discussions on the design methodology open cell but still maintain a discrete anisotropic open cell for
leveraging AM have not been discussed in detail. which residual quilting has still been shown. Other methods
Isogrid lightweight mirror designs were developed decades include post-machining techniques that do not involve a large
ago with fundamental design work developed by Barnes, where polishing pressure, such as ion-beam figuring [12]. However,
he introduced analytical solutions for self-weight deflections such machines are inherently slow, expensive, and require a
of symmetric sandwich mirrors and local surface deformations specialized skillset. The deformations over the aperture from

1559-128X/21/030681-16 Journal © 2021 Optical Society of America


682 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

the array of unit cells cause the surface to have a quilted texture to valley (P-V). A minimum thickness limitation of 1 mm was
with a lower-order radial and azimuthal spatial wavelength also set as a general manufacturing constraint.
variation than the typical surface finish; however, they are higher To constrain the mirror design, initial volumetric boundary
order than the surface figure (or form). These are known as conditions must be set. First, the diameter is prescribed by the
midspatial frequencies (MSF). Excessive amplitudes over these optical design, with the second being the mirror overall thick-
spatial wavelengths can have deleterious effects on optical per- ness. The mirror design methodologies assume the thin plate
formance, even while meeting both figure and surface finish relationship for self-weight deflection, where the deflection is
requirements [13]. proportional to r 4 /h 2 , where r is the semi-diameter, and h is
The common limitations for the mirror design in the preced- the overall mirror thickness. Normalization of the thickness to
ing work were the state of the art manufacturing and advanced the diameter (φ) show that the increase in compliance below
computer modeling. Analogously, the implementation of com- 0.1φ resulted in significant self-weight deflection with a three
plex freeform optics was limited by the manufacturing methods point mount with an increased sensitivity of the mount location
available. However, with modern machines and multi-axis [18]. The proportionality ratio becomes less sensitive starting
manufacturing, these freeform shapes are now realizable and at 0.2φ−0.3φ. Thus, a design decision must be made to either
transformative for an optical designer [14,15]. Freeform optics require stiffness in the mirror substrate or a separate support
have proven to provide equal or better performance with a structure, such as a whiffletree. Due to other influences, includ-
smaller volume, thus reducing the overall weight of optical sys- ing iterative manufacturing and athermalization, which are
tems [16,17]. Combining complex surface shapes with complex outside of the scope of this document, the mirrors would be sup-
mechanical geometries for mirror preforms can result in a new ported at three points and kinematically mounted to the optical
standard for optical system performance. bench, without an additional complex support structure, thus
Therefore, the purpose of this document is both (i) to revisit requiring adequate stiffness in the mirror [19]. For these reasons,
traditional analytical expressions and provide a systematic the mirror starting thickness was set to 0.2φ or 36.8 mm.
design method and (ii) to develop a new design methodol-
ogy, utilizing a combination of analytical and FE methods, 3. MATERIAL SELECTION
considering current state of the art manufacturing, to create
higher performance mirrors. Thus, the document comprises One of the more attractive materials used for lightweight mir-
rors in telescope design is SiC because of the material’s high
two main sections on the design and analysis methodology for
thermal stability and high specific stiffness [20–23]. Thermal
lightweight mirrors. Section 5 describes the design method of
stability is characterized both by the time for the stabilization
an isogrid open back mirror that can be made by traditional
of thermal gradients and the distortion caused by a steady state
manufacturing methods, followed by a FE model for design
thermal gradient. To visualize this, materials are often compared
confirmation. Section 6 then provides the design method for a
by putting them on the same graphic as a function of important
mirror design considering AM for the mirror preform fabrica-
criteria or figures of merit. For example, Fig. 1 displays the tran-
tion, also followed by a FE model for design confirmation. The
sient and steady state thermal-dimensional stability of various
methodologies are presented in general form, and then both are
materials (horizontal axis) against specific stiffness E /ρ (vertical
validated with a case study. axis), where E is the elastic modulus, and ρ is the density. In
Fig. 1(a), the horizontal axis is D/α, where D is the thermal
2. GENERAL DESIGN CONDITIONS AND diffusivity in mm2 /s, and α is the thermal expansion coefficient
FUNCTIONAL REQUIREMENTS in 10−6 /K . Greater diffusivity and a lower thermal expansion
coefficient correspond to more rapid smoothing of transient
To provide a real world case study, boundary conditions and temperature gradients and reduced strain gradient (stresses,
functional requirements were defined. In previous work by the distortion) for an instantaneous thermal gradient. In Fig. 1(b),
authors, mirrors were designed for a freeform, three mirror, the horizontal axis is λ/α, where λ is the thermal conductivity in
anastigmatic telescope in both aluminum and silicon carbide W/m-K, and α is the thermal expansion coefficient in 10−6 /K .
(SiC) [18]. The case study for this document is on the design of Greater thermal conductivity and a lower thermal expansion
the 184 mm tertiary mirror to meet design functional require- coefficient result in reduced steady state temperature gradient
ments. The low curvature freeform mirror was simplified to and reduced strain gradient for a given temperature.
a flat mirror for analysis, as the focus of this document is to SiC far outperforms other materials according to both cri-
introduce the design methodology for the mirror support struc- teria. However, manufacturing complex geometries in SiC has
ture, which is applied to all flat and low curvature mirrors on been challenging and often result in high costs and lead times.
this scale. The design requirements for the telescope comprise Moreover, grinding and polishing of SiC is time consuming
maximum overall weight and mirror deflection limitations from and challenging, thus near net shapes are preferable to reduce
self-weight loading and final surface polishing. The weight of unduly long post-processing times. Terrani et al. introduced
the telescope was distributed between the housing, mounting a novel process to additively manufacture complex SiC com-
cells, and mirrors, leading to target mass per unit area of less than ponents to a near net shape, with very high purity and density
20 kg/m2 . A system level error budget was developed, with the [24]. Expanding on this work, Horvath et al. addressed the
surface figure error (SFE) allowance for the self-weight deflec- post-processing of this material compared to chemical vapor
tion set to less than 5 nm RMS. The MSF subaperture surface deposition (CVD) SiC and found that equivalent surface fin-
deflections (quilting) from polishing were limited to 43 nm peak ishes could be achieved with additional work to reduce residual
Research Article Vol. 60, No. 3 / 20 January 2021 / Applied Optics 683

Fig. 1. Figures of merit for common materials used in telescopes. (a) Specific stiffness (E /ρ) versus diffusivity (D) per thermal expansion (α).
(b) Specific stiffness (E /ρ) versus conductivity (λ) per thermal expansion (α).

surface porosity [25]. Because of these prior studies, the mirror rib thickness to meet the solidity ratio requirements, while also
design was revisited, specifically in SiC; however, the design designing it to meet the quilting requirements. In addition, the
methodology is agnostic to the material selection. The mate- location of the neutral axis varies as a function of the solidity
rial properties used for SiC were 3000 kg/m3 for the density, ratio and facesheet thickness. Thus, an analytical expression
300 GPa for the elastic modulus, and 0.17 for Poisson’s ratio. was developed to approximate the location of the neutral axis
using the defined design parameters. Because variations of the
design parameters change the outcome for multiple functional
4. NOMENCLATURE requirements and are therefore not independent, a series of
The list of symbols used in the paper is given. Other variables are graphics plotting the design parameters were developed to better
defined throughout the document. visualize the trade-offs and limiting values to achieve the mirror
functional requirements. Figure 2 provides a flowchart for the
traditional mirror design methodology.
5. TRADITIONAL ANALYTICAL MIRROR DESIGN
METHODOLOGY A. Areal Mass Density and Self-Weight Deflection
In this section, we implement the design methodology using Mirror apertures can vary by orders of magnitude in size based
analytical methods to design a lightweight, open back, isogrid the specific application; thus, a better metric for mirror mass is
mirror. Imposing the boundary conditions set in Section 2, areal mass density or mass per unit area. The general form for
the mirror areal mass density functional requirement is first the estimation of the mass per unit area of the mirror is shown as
introduced, which prescribes the facesheet thickness limits. Eq. (1):
From this, the mass reduction is accomplished by introducing
m
the solidity ratio, where the solidity ratio range is defined by the = ρ(t f + ηh c ), (1)
facesheet thickness and self-weight deflection requirement. The A
isogrid support structure is designed by varying the cell size and where m/A is the areal mass density, ρ is the material density, η
is the solidity ratio, and h c is the isogrid cell depth. The overall
δq = Peak to valley surface quilting deflection mirror thickness, defined as 36.8 mm for this case, is simply
δRMS = Root mean square mirror self-weight deflection t f + h c . From Eq. (1), we show in Fig. 3 the facesheet thickness
η= Areal solidity ratio required as a function of the solidity ratio to achieve two areal
κ= Rib effectiveness factor mass density requirements. Here, the limits of the facesheet
ν= Poisson’s ratio thickness and required solidity ratio are shown to meet the areal
φ= Mirror diameter mass density functional requirement. The maximum facesheet
ρ= Material density thickness to meet the 20 kg/m2 requirement is approximately
A= Mirror area 6.5 mm, leaving, however, no material for an isogrid support
B= Inscribed cell diameter structure. At the lower limit for the facesheet thickness of 1 mm,
D= Flexural rigidity of a circular disc the manufacturing constraint mentioned, the solidity ratio
E= Modulus of elasticity
required is approximately 0.16.
h= Overall mirror thickness
Using the facesheet and solidity ratio limits defined by Eq. (1)
hc = Cell depth
and shown in Fig. 3, the first functional requirement that can
r= Mirror semi-diameter
tf = Facesheet thickness
be met is the self-weight RMS surface deflection. From the thin
tw = Rib/web thickness plate assumption, an analytical estimate of the mirror RMS
surface deflection is shown in Eq. (2):
684 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

Fig. 2. Design methodology flowchart for the analytical design of a lightweight mirror considering traditional manufacturing techniques.

ρg r 4 Mehta for both open and closed back mirrors [5]. The equiv-
δRMS = C (1 − ν 2 ), (2)
E hb2 alent bending thickness for an open back mirror is shown by
Eq. (3):
where C is a constant from the mounting condition, ρ is density,
g is gravitational acceleration, r is the semi-diameter of the ! 31
mirror, ν is Poisson’s ratio, and h b is the equivalent bending (1 − κη)(t f 4 − κηh c 4 ) + κη(t f + h c )4
hb = . (3)
thickness. The equivalent bending thickness was derived by t f + κηh c
Research Article Vol. 60, No. 3 / 20 January 2021 / Applied Optics 685

Fig. 3. Achievable areal mass density considering the facesheet Fig. 4. Self-weight mirror deflection with a three point mount
thickness and solidity ratio with a mirror overall thickness set as (C = 0.323 [26]).
36.8 mm.

the cell diameter and discrete cell web thicknesses. Thus, the cell
It is worth emphasizing at this point that Mehta introduced size and web thickness can be selected to meet the solidity ratio
a rib effectiveness factor, κ, to account for buckling and shear requirements, however additional constraints on the cell size
of the ribs in the isogrid. Mehta used an effectiveness factor of must be introduced.
0.5 for triangular and square cells and described the hexagonal The first constraint is an upper geometric limit due to mount-
pattern as “significantly smaller” [5]. The effectiveness factor ing point location. As mentioned, this study is constrained to
provides a conservative estimate of the flexural rigidity of the a subaperture three point mount, which is known to be at an
mirror, thus having the potential for a design heavier than optimal location at ≈ 66% of the diameter. Moreover, for an
required. However, with most designs having partial cells and open back mirror, the triangular cell geometry provides the
geometric variations around mounting point locations, further highest flexural rigidity compared to square or hexagonal cells.
analysis is required by numerical FE methods. After the creation An appropriate design will have the mounting points at a nodal
of a solid model for FE verification, small changes can be made location, where the interconnected webs meet to provide the
to the design parameters to meet the functional requirements. least compliant design at the mounting point, shown in Fig. 5.
The self-weight deflection for multiple facesheet thicknesses To achieve this using triangular cells, starting from a central
as a function of the solidity ratio using a three point mount is node and maintaining symmetry, the maximum length of the
shown in Fig. 4. From the data in Fig. 3, it is shown that, for any triangular cell wall is 0.66r , with r being the semi-diameter of
facesheet thickness that meets the requirement of 20 kg/m2 , the mirror. Thus the maximum inscribed circle to calculate the
the solidity ratio to achieve this is too large to also meet the self- solidity ratio is defined by Eq. (5), which results in a 35 mm cell
weight deflection requirement. As mentioned, the conservative size for a 184 mm mirror:
effectiveness factor could result in a stiffer mirror than designed.
For this reason, values for the facesheet thickness of 2 mm and 0.66r
B= √ . (5)
the solidity ratio of 0.15 were chosen. These values exceed the 3
areal mass density requirement by a small margin and would be
The second constraint is from the P-V quilting functional
analyzed further after the creation of a solid model.
requirement. From this surface deformation caused by the
pressure applied during mirror polishing, which is from Barnes
B. Isogrid Cell Geometry [4], we can estimate using the cell size and unit cell geometry in
To design the isogrid structure, the solidity ratio must be decom- Eq. (6):
posed into physical design parameters. The solidity ratio is
P B4 P B4
defined as the ratio between the area of an open geometry over δq = ψ = 12ψ (1 − ν 2 ), (6)
the area of a solid geometry and is shown in Eq. (4): E D E tf 3

(2B + tw )tw where δq is the deflection over the unit cell, P is the polishing
η= , (4) pressure, E is the modulus of elasticity, D is the flexural rigid-
(B + tw )2
ity, and ψ is a cell geometry constant with the triangular cell
where B and tw are the geometry dependent inscribed circle cell ψ = 0.00151. Note that the equation is independent from the
size and interconnecting web thickness, respectfully. Figure 5 depth of the cell. To quantify the effect from these localized
shows the three main geometries used for isogrid support struc- surface deformations, Vukobratovic derives the Strehl ratio as a
tures with the parameters for calculation of the solidity ratio. function of the wavelength and quilting deflection from Eq. (6),
Here, Fig. 6 shows the effect on the solidity ratio as a function of as shown in Eq. (7):
686 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

Fig. 5. Examples of typical isogrid geometry. Left to right: triangular cell, square cell, hexagonal cell. The central circle represents the central node
for the connection points for the web with the triangular cell having six connections at the central node, the square and hexagonal with 4 and 3, respec-
tively. Design parameters web or rib thickness, tw , and inscribed cell size, B, are used to calculate the solidity ratio.

Fig. 6. Solidity ratio as a function of the cell diameter at multiple Fig. 7. Quilting deflection as a function of facesheet thickness for
web thickness values. multiple isogrid cell sizes under loading from polishing pressure.

 2  n
δ 0.66r 1
Ii 4π 2 2λq B= √ , (8)
S= =  2    2  . (7) 3 2
Io δq δ
2
1 − 2π 2λ 1 − 4π 2λq
2
where n is a positive integer. From this, the next lowest size for
this mirror is 17.5 mm, which meets the quilting requirement
For a test wavelength of 632.8 nm and a Strehl ratio constraint for a facesheet thickness greater than 1.5 mm. Using Fig. 6, the
of 0.95 to allow for MSF wavefront error over multiple mirrors, cell wall size is selected to meet the solidity ratio at the defined
this equates to a P-V quilting deflection allowance of 43 nm. cell diameter, which was selected as 1.5 mm.
Because of the hardness of SiC, higher polishing pressures are
required to efficiently remove material [9]; thus, the polishing
pressure of 42 kPa was used as the load condition. The quilting C. Neutral Axis Location for Isogrid Mirror
deflection as a function of facesheet thickness and cell size, with The location of the neutral axis of an open back lightweight
a polishing pressure of 42 kPa, is shown in Fig. 7. mirror is not at the geometric centerline of the mirror. However,
From these two additional cell size constraints and using it is known that the neutral axis is between the geometric center-
Fig. 7, the appropriate cell size can be selected to satisfy both the line of the mirror and geometric centerline of the facesheet. The
quilting deflection allowance and, subsequently, the solidity effect of the solidity ratio and facesheet thickness will change this
ratio to achieve the areal mass density functional requirement. location between these bounds. Placing the mounting features
From the previous figures and constraints, it is shown for this at the neutral axis reduces deformations at the optical surface by
application that the maximum cell size of 35 mm will result in reducing the length of the moment arm of the mount geometry.
non-acceptable quilting deflections for any facesheet thickness Forces at the mounting points could be caused by either thermal
less than 4.5 mm; thus, a smaller size must be used. However, to or mechanical external loads; thus, it is important to know
ensure that each mounting point is located at the web nodes, the the approximate location to reduce the bending moment. An
cell sizes are limited to discrete dimensions that can be calculated analytical equation using the previous design parameters for the
by Eq. (8): mirror was developed and confirmed by FE analysis (FEA).
Research Article Vol. 60, No. 3 / 20 January 2021 / Applied Optics 687

Fig. 8. Diagram for the derivation of the neutral axis using parame-
ters for the lightweight mirror.

Fig. 9. Location of the neutral axis as a function of the solidity ratio


The approximation of the neutral axis of the circular plate for multiple facesheet thickness values, normalized to the overall mirror
was viewed in the cross section and simplified to a beam model. thickness.
The model was treated as a two-component composite struc-
ture with a facesheet and lightweight back. Figure 8 provides a
diagram of the model, including the mirror design parameters. then satisfies the neutral axis boundary conditions mentioned
This model assumes flat or low curvature mirrors, where, for previously.
higher curvature mirrors, there is a small shift in the neutral axis A FE simulation was performed on a series of fixed free
location that is inversely proportional to the natural logarithm beam models with an end loaded moment. The series of mod-
of the ratio of front and back curvatures. els comprised a solid facesheet and open back structure with
Assuming pure bending and linearly elastic material, the loca- defined solidity ratios and facesheet thicknesses. The location
tion of the neutral axis can be found by analytic methods by set- of minimum strain in the model after loading was recorded at
ting the first moment of the area equal to zero. For a composite multiple parameters in order to perform a least squares fit for h 2
structure, this results in Eqs. (9)–(11): by adjusting the exponential term n. The exponential term was
Z Z found to be 1.40 ± 0.08 over a 95% confidence interval. The
analytical equation results for the location of the neutral axis,
E 1 y dA + E 2 y dA = 0, (9)
1 2 normalized to the mirror thickness, as a function of the solidity
  ratio with the overlaid FE sample points is found in Fig. 9. From
hc
Z
this, the neutral axis location can be approximated after selecting
y dA = φh c − h1 , (10) the facesheet thickness, cell depth, and solidity ratio. For the
1 2
mirror design in this document, the location of the neutral axis
is approximately 12 mm from the top of the facesheet or the
 
tf
Z
y dA = φt f + hc − h1 . (11) optical surface.
2 2
By treating the open back portion of the mirror as a cellular solid,
D. CAD Model and Finite Element Analysis of
we know from Ashby that the relative stiffness is proportional to
Analytical Designed Mirror
the relative density by a power law, as shown in Eq. (12):
 n Using the values calculated from Sec. 5, the mirror was designed
E1 ρ1 in computer aided design (CAD) software and is shown in
=C , (12)
E2 ρ2 Fig. 10. Once designed, the areal mass density of the mirror was
higher than estimated. This is a result of the mass estimation
where C is the proportionality scaling constant, and n ranges in Eq. (1), which does not account for partial cells, the actual
between 1 and 2 for most cellular solids [27]. Because the areal mount geometry, fillets to reduce stress concentrations and
solidity ratio is constant as a function of cell depth, it was treated access for manufacturing tooling, and the outer shell. To meet
equivalent to relative density, resulting in Eq. (13): the functional requirement, the web thickness was reduced from
1.5 mm to 1 mm, and the facesheet was reduced from 2 mm to
E 1 = C E 2 ηn . (13)
1.5 mm, which reduced the areal mass density to 20.6 kg/m2 .
Substituting Eqs. (10), (11), and (13) into Eq. (9) and reorgan- The final solidity ratio of this mirror as designed in CAD, which
izing in terms of distance from the top of the facesheet, h 2 , the is calculated by the area of the open isogrid structure divided by
location of the neutral axis is shown in Eq. (14): the area of the face sheet, is 0.15. The neutral axis location was
then recalculated based on the actual solidity ratio and updated
1 ηn h c 2 + t f (2h c + t f ) facesheet thickness, adjusting the mounting plane to 13 mm
h 2 = N A = (t f + h c ) − , (14)
2 ηn h c + t f from the optical surface.

which is in terms of the design parameters used in the mirror


1. Finite Element Analysis of the Analytical Mirror Design
design. The exponential value is assumed to be between 1 and
2 and is solved by a least squares curve fit to a series of FE mod- FE methods are used extensively for design analysis of mirrors to
els. The scaling constant was set to one such that the equation characterize surface shape changes from the expected load cases.
688 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

polishing load case, the mirror was constrained by the back


surface, which represents the constraint if using traditional
polishing techniques. For each load case, the nodal points from
the deformed FEA model were extracted for further surface
processing.

2. FEA results
Figures 11(a) and 12(a) show the results in the native CAE
software from the gravitational load and polishing pressure,
Fig. 10. CAD model of analytical designed mirror with updated
parameters to meet the areal mass density requirement.
respectively. The surface nodes were extracted and least squares
fit with Zernike polynomials. All terms up to the 7th-order
radial and azimuthal, Z77 , were removed from the surfaces to
However, FE software lacks the ability to quantify the shape of reveal the MSF surface deformations from the isogrid structure.
the deformation, only reporting nodal displacements. Thus, the Figure 11(b) shows the imported surface, 11(c) the 7th-order
deformed nodes are extracted and post-processed to quantify Zernike polynomial fit, and 11(d) the residual map after sub-
the shape change. The changes to the surface shapes can be tracting the Zernike map. The RMS deflection of the surface is
described by any number of polynomials; however, the more found to be 1.2 nm, with higher-order residual terms an order
common basis sets used are the Zernike polynomials, which are of magnitude less. While an order of magnitude less than the
a standard basis set for both optical design and optical metrology requirement, the isogrid structure is visible in the residual map
due to the orthonormality over circular apertures. just from the gravity load. Thus, the mirror functional require-
The designed model was imported into computer aided ment of 5 nm for self-weight deflection has been achieved. The
engineering (CAE) software for FEA. A volume mesh with a analytical solution estimates that the self-weight deflection
maximum 4 mm edge length and quadratic tetrahedral ele- with the selected mirror parameters would be 6 nm RMS. As
ments were used for the entire mirror. An additional symmetric, mentioned, the rib effectiveness factor results in a conservative
quadrangulate mesh was used with shell elements on the optical estimate of the flexural rigidity; thus, it had a slight over estimate
surface, with the elements connected to the main body as depen- of the expected self-weight deflection.
dent elements by the element faces. This method allows for a The surface processing from simulating the polishing pres-
robust FEA while providing a uniform, symmetric mesh at the sure was treated in the same fashion and is shown in Fig. 12.
optical surface for further processing. However, the main interest is the residual map in Fig. 12(d),
Two load cases were run in the FEA simulation: the first was with the quilting functional requirement set as 43 nm P-V. The
a gravitational acceleration load with the force vector gener- residual map shows that the P-V deflection is 42 nm, which
ated by the 1 g acceleration value acting on the mass of each just meets the functional requirement. This value also agrees
tetrahedral element. The second load case was a full aperture with the analytical solution, Eq. (6), with the updated facesheet
pressure simulating polishing. In the gravitational load case, the thickness of 1.5 mm.
mirror was kinematically constrained by creating local oriented In summary, the lightweight mirror has been analyzed to
coordinate systems at each of the three mounting points and have a self-weight deflection estimate of 1.2 nm RMS, a quilting
only constraining 2 deg of freedom per mount point. For the deflection from polishing of 42 nm P-V, and a total mass of

Fig. 11. (a) FEA results from a 1g acceleration normal to the optical surface. The displacement magnitude is shown. (b) Raw imported surface data
from quadrangulate mesh, (c) Zernike fit, and (d) Zernike fit removal from the data set, or residual map.
Research Article Vol. 60, No. 3 / 20 January 2021 / Applied Optics 689

Fig. 12. (a) Traditional mirror surface deflection from polishing pressure in native FEA output, (b) raw imported surface data from quadrangulate
mesh, (c) Zernike fit, and (d) Zernike fit removal from the data set, or residual map.

0.55 kg, resulting in an areal mass density of 20.6 kg/m2 . The The model is imported into CAE software for a topology opti-
areal mass density result is above the initial functional require- mization using the appropriate load cases, mounting restraints,
ment by a small fraction; however, it is considered acceptable material properties, and functional requirements. The manda-
with all other requirements and manufacturing constraints tory geometries are selected such that the optimization does
being met. not suggest material removal in these volumes. For a mirror, the
mandatory geometry should be the facesheet and mounting
points. The topology optimization results in a model comprised
6. MIRROR DESIGN METHODOLOGY FOR of threshold elements that are ranked between 0 and 1. In this
ADVANCED MANUFACTURING case, threshold elements define which elements are contribut-
AM is the latest addition in the advanced manufacturing ing to the overall stiffness required to meet the displacement
portfolio and is often described as a leading disruptive and requirements, with one as a full contribution and zero as little
transformative technology for the next industrial revolution. to no contribution. Multiple topologies can be investigated
Unlike traditional subtractive manufacturing methods, such as by varying the acceptable threshold range. Depending on the
milling or grinding, AM does not require line of sight to create limiting factors of the AM method chosen, which is both mate-
a geometry. Because of this, AM enables new design freedoms rial and printing machine specific, the appropriate topology
can be selected. Once the topology optimized geometry is
not previously investigated. Combining commercially available
selected, further weight reduction is performed by introducing
CAE software with AM, the design of the mirror should be
three-dimensional (3D) lattice structures, as opposed to 2D
revisited to leverage these benefits.
isogrids. Along with areal mass density reduction, the 3D lattice
This section uses the same initial boundary conditions and
structures will have an effect on the flexural rigidity of the mirror
functional requirements found in Section 2. The mirror initial
and the surface quilting. Here, we utilize the previous analyti-
conditions began with the same volumetric design space as
cal equations combined with CAE to discuss the trade-offs in
the traditional mirror, a 184 mm diameter with a thickness of
selecting a lattice structure such that an increase in the specific
36.8 mm and three mounting pads at 0.66φ. The functional stiffness of the mirror is achieved while also reducing the quilt-
requirements are also identical, with the areal mass density ing deformations. After the design of the mirror is completed, a
requirement set as 20 kg/m2 , quilting deflection less than secondary FEA is performed to independently confirm that the
43 nm P-V, and a self-weight deflection less than 5 nm RMS. chosen design parameters meet the functional requirements.
A minimum thickness limitation was also set to 1 mm, iden-
tical to the traditional mirror. However, the similarities to the
traditional design end there. In this methodology, the func- A. Topology Optimization from External Load Cases
tional requirements are met by a combination of CAE and The initial solid model was generated and imported into CAE
fundamental understanding of the analytical methodology. software for the topology optimization process. A unique differ-
Figures 13 and 14 provide a flowchart for the advanced entiation between the analytical design method versus the CAE
mirror design methodology. This design method begins with design method at this point is the calculation of bending stiff-
the creation of a solid model defined by the initial conditions. ness or flexural rigidity. The analytical model for flexural rigidity
690 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

Fig. 13. Part 1: design methodology flowchart for the advanced mirror design considering additive manufacturing.

must assume a homogeneous, linearly elastic component, thus An interesting observation to highlight is the connected-
leading to an equal distribution of stiffening elements over the ness between the mounting pads in the topologies shown in
entire mirror. Moreover, the assumptions used in the analytical Fig. 15. This is a result of the kinematic constraint boundary
methods resulted in a mirror with more mass than required to condition. With each mounting location constraining 2 deg of
achieve the stiffness requirements, only confirmed after a FEA. freedom, the mirror requires high bending stiffness to connect
A CAE design method instead develops a compliance matrix the force loop from one mount point to the next in order to
containing all discrete elements in the mesh, which is minimized achieve the displacement requirements. A full encastre bound-
during the optimization. This enables a localized distribution of ary condition at each mount point would result in localized stiff
stiffness only where required on the mirror and mass reduction elements near the mount with only the facesheet as a connected
where the material does not have a significant contribution to structure between the mounts, similar to work previously pub-
the overall stiffness. This is represented by the threshold ele- lished [28]. Thus, care must be taken to accurately represent
ments in the optimization output. Therefore, the main purpose the actual boundary conditions, as a boundary condition not
of the topology optimization is to provide the designer with the representative of the actual mount could result in a significant
internal force loop through the mirror from the external loads, underestimation of the mirror stiffness when used in practice.
thus showing where stiff elements are required in the design.
Identical to the traditional mirror FEA, the model was B. Weight Reduction by Lattice Infill
meshed with solid tetrahedral elements with planar symmetry. The mirror with the selected topology optimization surpassed
The mount pads were kinematically constrained using local the displacement requirements; however, it exceeded the areal
oriented coordinate systems that constrain two translational mass density requirement. In similar fashion to the traditional
degrees of freedom per mount, representative of a 3 vee–3 ball design, a reduction of the remaining available volume must
kinematic mount. Combined loading was applied for both be performed in order to meet the mass requirement. The
gravitational acceleration to the solid body and polishing pres- traditional design and previous state of the art manufacturing
sure in the optimization. The topology optimization converged techniques were limited to a 2D method with discrete constant
to the displacement requirements within 25 iterations. Multiple cells, of which a solidity ratio was used as the design parameter.
topologies are shown in Fig. 15 with the minimum thresh- With state of the art AM, we can reassess this method with a vol-
old value shown, along with the input geometry. A topology umetric approach using 3D lattice structures. The solidity ratio
with a minimum threshold value of 0.6 was selected, which expanded to three dimensions is described as relative density for
contained excessive elements to ensure connectedness of the lattice structures or foams [27]. Using certain lattice geometries
structural loop, assurance that minimum thickness values were can also reduce the geometric anisotropy seen in isogrid designs.
not exceeded, and allowance for further design work to reduce We know from Eq. (6), confirmed to be accurate on the
the mass. The smoothed result combined with the solid mount traditional design, that the quilting deflection is dependent on
pads and facesheet is shown in Fig. 16. the facesheet thickness and the cell area behind the facesheet.
Research Article Vol. 60, No. 3 / 20 January 2021 / Applied Optics 691

Fig. 14. Part 2: design methodology flowchart for the advanced mirror design considering additive manufacturing.

Shown in Fig. 7, for a 1 mm facesheet in SiC, any cell size less stiffness is both a material design parameter and a geometric
than 10 mm would result in meeting the threshold requirement design parameter for a lattice structure. The combination of
of 43 nm. This creates a maximum unit cell size for the lattice. these is known as relative stiffness, which is proportional to the
Because the cell geometry far from the facesheet does not have relative density of the lattice and lattice type. SiC provides a high
a significant contribution to the quilting, the lattice geometry material specific stiffness E /ρ; however, the unit cell design and
can vary as a function of distance from the facesheet. This allows size of the lattice infill will provide the geometric contribution
for a graded relative density as a function of distance from the to the relative stiffness. Al-Ketan et al. provided empirical data
facesheet. A similar idea has been shown to be effective for on multiple lattice topologies, reporting the relative stiffness
reducing quilting deflections by using stochastic foam cored values for each [29]. Of those tested, the sheet diamond triply
mirrors. However, an intrinsic property of stochastic foam is periodic minimal surface (TMPS) lattice topology was shown to
a higher compliance compared to an ordered lattice structure, be the least sensitive in terms of stiffness reduction as a function
thus leading to a lower overall flexural rigidity [8,27]. of relative density. Similar to selecting a topology threshold,
Self-weight deflection is shown to be inversely proportional when choosing a lattice, one must understand both the geo-
to the specific stiffness, as in Eq. (2). However, the specific metric limits of the AM method and open paths for powder
692 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

Fig. 15. Topologies after convergence on the minimum compliance functional requirement at different remaining threshold values. A threshold
value of zero represents an element in the mesh with little to no contribution to minimizing the compliance, while a threshold value of one represents
an element that has a maximum contribution in minimizing compliance.

removal to avoid trapped powders. While out of the scope of this


document, Sweeney et al. provide a comprehensive foundation
on metallic mirror printing limitations using multiple print-
ing methods [1]. However, extensive work in this area is still
required and is changing as new machines are developed. The
SiC printing method developed by Terrani et al. uses a binder
jet printing method followed by a chemical vapor infiltration
[24]. Because this printing method has minimal limitations for
geometry creation and the sheet diamond TPMS lattice does
not contain enclosed pockets, it was a suitable candidate for
Fig. 17. Mirror design combining the 0.6 threshold topology
further analysis and subsequent prototyping for experimental optimization output with a cylindrical TPMS sheet diamond lattice
design confirmation in future work. structure.
Using implicit modeling software, the TPMS sheet diamond
unit cell array was defined in cylindrical coordinates to provide
a rotational symmetry of the lattice on the topology optimized quilting deflection. The slope of this graded density transition
mirror. An azimuthal frequency of 30 units were used to create was not quantified as optimal; however, it resulted in a relative
a maximum spacing of 10 mm at the aperture edge. A graded density value of 0.25 for the remaining structure after topology
relative density was also used, which varied the thickness of the optimization.
lattice elements from 4 mm to 1 mm, starting from the back of From Al-Ketan et al. [29], we can approximate the relative
the facesheet to 8 mm behind the facesheet. After 8 mm, the stiffness for the sheet diamond TMPS lattice by the square root
lattice thickness was held constant at 1 mm. This combination of the relative density. Substituting this value in Eq. (2) for the
resulted in no cells in any direction greater than 10 mm and elastic modulus and using the actual mirror thickness as the
the largest spacing behind the facesheet of 6 mm, thus meeting equivalent bending thickness, we approximate the self-weight
the threshold value derived from the analytical equation for deflection to be 4 nm RMS. This approximation does not
account for the complex topology optimization; however, the
order of magnitude approximation is a useful step to progress
towards a FEA confirmation.
The high azimuthal cell frequency of the cylindrical TPMS
model used resulted in some unrealizable artifacts for manufac-
turing at the center point of the mirror. To resolve this, the center
was removed and replaced with another sheet diamond lattice
with less azimuthal cells to ensure that the center point was well
behaved. The final model includes a 1 mm thick facesheet; how-
ever, it is hidden in the figure to better show the lattice structure
behind the solid facesheet. Figure 17 shows discrete slices of the
Fig. 16. For further design work, the 0.6 minimum threshold ele- lattice starting from the back side of the solid 1 mm facesheet in
ment geometry was selected. This geometry was selected to ensure con- Fig. 18. The final model has a mass of 0.45 kg, an estimation of
nectedness through the structural loop and allowance for further mass self-weight deflection of 4 nm RMS, and a quilting deflection of
reduction by lattice infills. less than 10 nm PV.
Research Article Vol. 60, No. 3 / 20 January 2021 / Applied Optics 693

Fig. 18. Discrete slices showing the TPMS lattice structure behind the facesheet.

C. Advanced Mirror Design FEA Confirmation Here, we can see the effect from the topology optimized
The final mirror design that combines both topology opti- structure in the residual map, similar to the isogrid structure
mization and lattice infill was post-processed with FEA to in the traditional design; however, the deflection is an order
confirm that all requirements have been met. However, it is of magnitude less than the functional requirement. Figure 19
computationally expensive to create a compliance matrix with shows that the RMS and P-V deflection is less than the func-
the amount of elements required to make a refined mesh on tional requirement and in agreement with the approximation
the lattice. Therefore, two methods were used to perform the from the substitution of relative stiffness in Eq. (2).
analysis. The first analysis confirms the self-weight deflection
requirements, and the second analysis confirms the deflection 2. Surface Quilting from the Lattice
from polishing pressure.
The higher-order quilting surface deflections do not require
a full aperture analysis to quantify the edge effects from the
1. Self-Weight Deflection FEA lattice if the lattice at the surface has symmetry. While this was
To quantify the self-weight deflection by FEA, a compliance performed on the traditional mirror design, a sub-sample of the
matrix was created by a separate FEA on a unit cell of the lattice aperture over some cells would have been sufficient to provide
used. This compliance matrix is used in a solid version of the enough information on the quilting deflection. Because the
mirror (Fig. 16), as the elastic modulus material property for ability to perform meshing on large lattice structures is limited
the volume the lattice occupies. This solid attribute in the FE to a homogenized method, a sub-sample approach was used
model is known as a homogenized material unit cell, which to analyze the quilting deflection from polishing on the lattice
provides an accurate representation of the stiffness of a lattice supported mirror, where direct meshing on the lattice could be
without having to mesh a large lattice scaffold. The unit cell performed.
homogenization method is an accepted method to perform A solid tetrahedral volume mesh was created over a 25 mm
analysis on advanced lattice structure designs [30–32]. To per- circular section of the lightweight mirror at four different
form an appropriate homogenization, the relative density of the locations. One location was over a mount pad, with two other
lattice, the lattice type, and the orientation of the lattice must locations 90 deg apart at the same radial distance, and one more
be known. The relative density can be described as the ratio of at the center of the mirror. The same conditions as the tradi-
the mass of the lattice over the solid, both of which are known. tional mirror were applied, with a 42 kPa polishing pressure over
The homogenized unit cell uses the same relative density, thus the subaperture sample. The four samples were post-processed
providing the relative stiffness, which can then be used on the with the same criteria as the traditional mirror, and the residual
larger model for analysis. An adjustment to the lattice results in a deflection maps are shown in Fig. 20.
new relative density, where the homogenized unit cell can then As a result of the rotational symmetry, each of the three maps
also be updated for analysis. However, homogenization will look near identical with an average P-V displacement of 2.3 nm
not reveal edge effects from the lattice spacing, such as quilting and 0.3 nm RMS compared to the traditional mirror with
deflections. To quantify the quilting deflection, another method 42 nm and 9 nm, respectively. Using Eq. (7), the Strehl ratio
was used and is described in the following section. from these quilting displacements would be greater than 0.98.
The homogenized model was analyzed with the same criteria The striations in the three maps match the spiral rib portion
as the traditional design in Section 5.D. The FE mesh used of the TPMS lattice structure just behind the facesheet. This
solid tetrahedral elements with 4 mm edge lengths combined is another confirmation that the geometry further from the
with a quadrangulate surface mesh used for post-processing. A facesheet has a negligible effect on surface quilting. With both
gravitational acceleration load with the force vector generated the self-weight deflection and quilting deflection within the
by the 1g acceleration value acts on the mass of each tetrahedral specifications, the design of the mirror is complete. The overall
element. The native FEA solution is shown in Fig. 19(a), with mass of the final mirror is 0.45 kg, equating to a mass per unit
the post-processed surface data shown in Figs. 19(b)–19(d). area of 15.5 kg/m2 .
694 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

Fig. 19. (a) Topology optimized mirror with TPMS lattice homogenization gravitational acceleration load FEA in native FEA output,
(b) raw imported surface data from quadrangulate mesh, (c) Zernike fit, and (d) Zernike fit removal from the data set, or residual map.

Fig. 20. Residual maps from polishing the lattice backed mirror. The locations of each map: (a) 12 o’clock, (b) 3 o’clock, (c) 6 o’clock, and
(d) center. The peripheral samples (a)–(c) are at the same radial distance, which is equal to the mounting radial distance.

7. CONCLUSION Table 1. Summary of the Functional Requirements for


Both Mirror Design Case Studies and the Resulting
Here, we provided two systematic design methodologies to Performance Improvement with the Advanced Design
create lightweight mirrors. The first design method used a series
Functional Traditional Advanced Percent
of analytical expressions with discussions on key trade-offs for
Requirement Mirror Mirror Improvement (%)
the design parameters to achieve the functional requirements, as
m/A: 20 kg/m2 20.6 kg/m2 15.5 kg/m2 24
outlined in Sec. 2. The second method established a workflow
δRMS : 5 nm RMS 1.2 nm RMS 0.9 nm RMS 25
for combining topology optimization and lattice structure infills δq : 43 nm PV 42 nm PV 2.3 nm PV 94
with analytical estimations to achieve the functional require-
ments. During this method, key waypoints were identified
The traditional mirror design, while meeting all functional
when considering design for manufacturability. Each method- requirements, was outperformed by the mirror designed for
ology was applied through a case study that concluded with an AM for each requirement. The advanced mirror design had
independent FEA for design confirmation. A summary of each an overall reduction of both areal mass density and self-weight
mirror and the evaluation in performance for each functional deflection, 24% and 25%, respectively, which is representa-
requirement are shown in Table 1. tive of a higher specific stiffness than the traditional design.
Research Article Vol. 60, No. 3 / 20 January 2021 / Applied Optics 695

The increase in specific stiffness is a direct result of the topol- publication, acknowledges that the US government retains a nonexclu-
ogy optimization with the appropriate boundary conditions, sive, paid-up, irrevocable, worldwide license to publish or reproduce the
where the optimization provided a distribution of stiffness only published form of this manuscript, or allow others to do so, for US gov-
ernment purposes. The DOE will provide public access to these results of
where required. This is contrary to the traditional design that
federally sponsored research in accordance with the DOE Public Access Plan
assumes an even distribution of stiffness over the entire aperture (http://energy.gov/downloads/doe-public-access-plan).
to develop the analytic expressions for self-weight deflection. This research is synergistic with the National Science Foundation I/UCRC
While the first natural frequency of the mirrors was not dis- Center for Freeform Optics (IIP-1338877,IIP-1338898, IIP-1822049and
cussed, because the frequency is proportional to the specific IIP-1822026). The authors acknowledge Tom Nelson and Peter Marasco from
stiffness of the mirror, it can be inferred that the mirror with a AFRL for helpful discussions.
higher specific stiffness also has a higher first natural frequency,
Disclosures. The authors declare no conflicts of interest.
considering that the aperture size, thickness, and mounting
conditions are the same.
The overall quilting deflection was reduced by 94% on the REFERENCES
advanced lattice structure design compared to the traditional
1. M. Sweeney, M. Acreman, T. Vettese, R. Myatt, and M. Thompson,
isogrid design. The quilting improvement is accomplished by
“Application and testing of additive manufacturing for mirrors and
reducing the cell size behind the facesheet or increasing the precision structures,” Proc. SPIE 9574, 957406 (2015).
facesheet thickness. However, the increase in facesheet thick- 2. E. Hilpert, J. Hartung, S. Risse, R. Eberhardt, and A. Tünnermann,
ness results in a significant weight increase with little benefit to “Precision manufacturing of a lightweight mirror body made by
overall mirror stiffness. Reducing the cell size on the traditional selective laser melting,” Precis. Eng. 53, 310–317 (2018).
design to the next discrete size results in a solidity ratio too high 3. N. Heidler, E. Hilpert, J. Hartung, H. von Lukowicz, C. Damm, T.
Peschel, and S. Risse, “Additive manufacturing of metal mirrors for
to achieve the areal mass density requirement. The lattice infill
TMA telescope,” Proc. SPIE 10692, 106920C (2018).
allowed for a decrease in cell size to a level that also allowed for 4. W. P. Barnes, “Optimal design of cored mirror structures,” Appl. Opt.
a decrease in facesheet thickness compared to the traditional 8, 1191–1196 (1969).
design, thus reducing both the quilting deflection and areal mass 5. P. K. Mehta, “Flexural rigidity characteristics of light-weighted
density. The generation of a lattice structure with 3D param- mirrors,” Proc. SPIE 0748, 158–171 (1987).
eters allowed for a smaller cell size where required and larger 6. D. Vukobratovich, B. Iraninejad, R. M. Richard, Q. M. Hansen, and
R. Melugin, “Optimum shapes for lightweighted mirrors,” Proc. SPIE
elsewhere, shown by the graded relative density.
0332, 419–423 (1982).
The lattice structure also provides unique flexibility to the 7. T. M. Valente and D. Vukobratovich, “A comparison of the merits of
designer, including but not limited to the lattice unit cell geom- open-back, symmetric sandwich, and contoured back mirrors as
etry, unit cell orientation, and graded density. A unique design light-weighted optics,” Proc. SPIE 1167, 20–36 (1989).
feature that can have a direct effect on optical performance is 8. D. Vukobratovich, “Lightweight laser communications mirrors made
the lattice unit cell frequency behind the facesheet. The case with metal foam cores,” Proc. SPIE 1044, 216–226 (1989).
9. T. Hull, M. J. Riso, J. M. Barentine, and A. Magruder, “Mid-spatial
studied showed a rotationally symmetric cell structure behind
frequency matters: examples of the control of the power spec-
the facesheet with an azimuthal spatial frequency of 30 cycles tral density and what that means to the performance of imaging
per aperture; however, depending on specific spatial frequency systems,” Proc. SPIE 8353, 835329 (2012).
requirements from the optical design, this can be modified with 10. T. L. Gray, M. W. Smith, L. E. Cohan, and D. W. Miller, “Minimizing
either more or less frequency. In a traditional design, this is an high spatial frequency residual error in active space telescope
unimaginable design parameter when considering all other mirrors,” Proc. SPIE 7436, 74360M (2009).
11. J. B. Hadaway, R. Eng, H. P. Stahl, J. R. Carpenter, J. R. Kegley, and
functional requirements. Utilization of this unique capability to
W. D. Hogue, “Cryogenic performance of lightweight SiC and C/SiC
alter image performance in a deterministic way could provide mirrors,” Proc. SPIE 5487, 1018–1029 (2004).
interesting results in future studies. 12. U. Papenburg, W. Pfrang, G. S. Kutter, C. E. Mueller, B. P. Kunkel,
The advancement of AM has proven to be a disruptive tech- M. Deyerler, and S. Bauereisen, “Optical and optomechanical
nology for component design in almost any field of engineering. ultralightweight C/SiC components,” Proc. SPIE 3782, 141–156
This advancement was leveraged to develop a new design meth- (1999).
13. R. E. Parks, “Specifications: figure and finish are not enough,” Proc.
odology and provide a paradigm shift in the mirror preform
SPIE 7071, 70710B (2008).
design, resulting in a higher performance mirror that is readily 14. P. J. Smilie, “Design and characterization of an infrared Alvarez lens,”
prototyped on the order of days. Opt. Eng. 51, 013006 (2012).
Funding. Office of the Secretary of Defense ; Office of Energy Efficiency 15. M. Beier, J. Hartung, T. Peschel, C. Damm, A. Gebhardt, S.
and Renewable Energy (NFE-20-08030). Scheiding, D. Stumpf, U. D. Zeitner, S. Risse, R. Eberhardt, and A.
Tünnermann, “Development, fabrication, and testing of an anamor-
Acknowledgment. This research was supported by the Department of phic imaging snap-together freeform telescope,” Appl. Opt. 54,
Energy, Office of Science, Office of Energy Efficiency and Renewable Energy, 3530–3542 (2015).
Office of Advanced Manufacturing, Energy and Transportation Science 16. J. Reimers, A. Bauer, K. P. Thompson, and J. P. Rolland, “Freeform
Division, under Award Number NFE-20-08030 and by the Department spectrometer enabling increased compactness,” Light Sci. Appl. 6,
of Defense, Office of Secretary of Defense, and used resources at the e17026 (2017).
Manufacturing Demonstration Facility, a DOE-EERE User Facility at 17. A. Bauer, E. M. Schiesser, and J. P. Rolland, “Starting geometry
Oak Ridge National Laboratory. The authors also acknowledge Andrew creation and design method for freeform optics,” Nat. Commun. 9,
Sartorelli and Duann Scott from nTopology for the technical software advice. 1–11 (2018).
This manuscript has been authored by UT-Battelle, LLC, under con- 18. N. W. Horvath and M. A. Davies, “Concurrent engineering of a
tract DE-AC05-00OR22725 with the US Department of Energy (DOE). next-generation freeform telescope: mechanical design and
The US government retains and the publisher, by accepting the article for manufacture,” Proc. SPIE 10998, 109980X (2019).
696 Vol. 60, No. 3 / 20 January 2021 / Applied Optics Research Article

19. N. W. Horvath, M. A. Davies, and S. R. Patterson, “Kinematic mirror 27. M. F. Ashby, “The mechanical properties of cellular solids,” Metall.
mount design for ultra-precision manufacturing, metrology, and sys- Trans. A 14A, 1755–1769 (1983).
tem level integration for high performance visible spectrum imaging 28. S. Liu, R. Hu, Q. Li, P. Zhou, Z. Dong, and R. Kang, “Topology
systems,” Precis. Eng. 60, 535–543 (2019). optimization-based lightweight primary mirror design of a
20. J. Robichaud, “SiC optics for EUV, UV, and visible space missions,” large-aperture space telescope,” Appl. Opt. 53, 8318–8325 (2014).
Proc. SPIE 4854, 39–49 (2003). 29. O. Al-Ketan, R. Rowshan, and R. K. A. Al-Rub, “Topology-mechanical
21. D. Castel, E. Sein, S. Lopez, T. Nakagawa, and M. Bougoin, “The property relationship of 3D printed strut, skeletal, and sheet based
3.2m all SiC telescope for SPICA,” Proc. SPIE 8450, 84502P (2012). periodic metallic cellular materials,” Addit. Manuf. 19, 167–183
22. E. Sein, Y. Toulemont, F. Safa, M. Duran, P. Deny, D. D. Chambure, T. (2018).
Passvogel, and G. Pilbratt, “A 3.5 M SiC telescope for HERSCHEL 30. D. Cellucci and K. C. Cheung, “Evaluation of cellular solids derived
Mission,” Proc. SPIE 4850, 606–618 (2003). from triply periodic minimal surfaces,” in ASME Manufacturing
23. J. Robichaud, D. Sampath, C. Wainer, J. Schwartz, C. Penton, and S. Science and Engineering (2016).
Mix, “Silicon carbide optics for space and ground based astronomi- 31. Y. Lu, W. Zhao, Z. Cui, H. Zhu, and C. Wu, “The anisotropic elastic
cal telescopes,” Proc. SPIE 8450, 845002 (2012). behavior of the widely-used triply-periodic minimal surface based
24. K. Terrani, B. Jolly, and M. Trammell, “3D printing of high-purity scaffolds,” J. Mech. Behav. Biomed. Mater. 99, 56–65 (2019).
silicon carbide,” J. Am. Ceram. Soc. 103, 1575–1581 (2019). 32. J. M. Guedes and N. Kikuchi, “Preprocessing and postprocessing for
25. N. Horvath, A. Honeycutt, and M. A. Davies, “Grinding of additively materials based on the homogenization method with adaptive finite
manufactured silicon carbide surfaces for optical applications,” CIRP element methods,” Comput. Methods Appl. Mech. Eng. 83, 143–198
Ann. 69, 509–512 (2020). (1990).
26. K. Schwertz and J. H. Burge, Optomechanical Design and Analysis
(SPIE, 2012).

You might also like