You are on page 1of 20

Computational Fluid Dynamics Modeling and Analysis

For the Giant Magellan Telescope (GMT)


John Ladd*a, Jeffrey Slotnickb, William Norbya, Bruce Bigelowc, William Burgettc
a
Boeing Research & Technology, The Boeing Company, St. Louis, MO, USA
b
Boeing Research & Technology, The Boeing Company, Huntington Beach, CA, USA
c
GMTO Corporation, 465 N. Halstead St., Pasadena, CA, USA 91107

ABSTRACT

The Giant Magellan Telescope (GMT) is planned for construction at a summit of Cerro Las Campanas at the Los
Campanas Observatory (LCO) in Chile. GMT will be the most powerful ground-based telescope in operation in the
world. Aero-thermal interactions between the site topography, enclosure, internal systems, and optics are complex. A key
parameter for optical quality is the thermal gradient between the terrain and the air entering the enclosure, and how
quickly that gradient can be dissipated to equilibrium. To ensure the highest quality optical performance, careful design
of the telescope enclosure building, location of the enclosure on the summit, and proper venting of the airflow within the
enclosure is essential to minimize the impact of velocity and temperature gradients in the air entering the enclosure.

High-fidelity Reynolds-Averaged Navier Stokes (RANS) Computational Fluid Dynamics (CFD) analysis of the
GMT, enclosure, and LCO terrain is performed to study (a) the impact of either an open or closed enclosure base soffit
external shape design, (b) the effect of telescope/enclosure location on the mountain summit, and (c) the effect of
enclosure venting patterns. Details on the geometry modeling, grid discretization, and flow solution are first described.
Then selected computational results are shown to quantify the quality of the airflow entering the GMT enclosure based
on soffit, site location, and venting considerations. Based on the results, conclusions are provided on GMT soffit design,
site location, and enclosure venting. The current work is not used to estimate image quality but will be addressed in
future analyses as described in the conclusions.

Keywords: GMT, Telescope, Enclosure, Los Campanas Observatory (LCO), Soffit, Magellan, CFD, RANS

1. CFD MODELING
1.1 Introduction
Goals of this work are to utilize CFD to understand the sensitivity of the flow characteristics affecting the ingestion
of low-level air from the terrain with design considerations such as site location, enclosure design type (i.e. open vs.
closed soffit), and enclosure venting configurations. The air flow is considered to be a viscous ideal gas with no
temperature variation as detailed in Section 1.4. The Reynolds-Averaged Navier Stokes (RANS) equations are solved
via a pressure-based algorithm and all cases are computed using the same procedure so that meaningful comparisons are
made.
In the CFD analysis procedure, geometry for the desired configuration is assembled and processed in order to obtain
a suitable computational mesh (or grid) for the numerical analysis. Once the grid is generated with the desired resolution
and quality, the flow solution is obtained by executing the flow solver software using an appropriate number of computer
processors, or cores. Details of each of these elements of the process are further discussed below.
1.2 Geometry
For the GMT CFD analysis, the computational model is composed of a simplified geometrical representation of the
telescope primary and secondary mirrors, enclosure, and surrounding terrain. Including a high level of geometric detail
is desirable but must be weighted with the corresponding increase in computational time (cost) to obtain the flow
solution. Boeing engineers worked closely with the Giant Magellan Telescope Organization (GMTO) to include
geometric features of interest yet exclude components considered too small and insignificant to the resulting flow field
for this initial study. The telescope components as designed and modeled in the CFD simulations are described in the
following sections.

*john.a.ladd@boeing.com phone: 314-232-1413 boeing.com

Modeling, Systems Engineering, and Project Management for Astronomy VII,


edited by George Z. Angeli, Philippe Dierickx, Proc. of SPIE Vol. 9911, 991114
2016 SPIE CCC code: 0277-786X/16/$18 doi: 10.1117/12.2231933

Proc. of SPIE Vol. 9911 991114-1

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


1.2.1 Teelescope
A side annd back view of o the GMT insstallation show
wing key geomeetric dimensionns is depicted in i Figure 1a annd 1b.
The telescopee center is defiined to be 10.7 meters from thhe observing floor,
fl and a totaal of 22.5 meteers from groundd level.

25. 5m

on Axis -,
107m

ng Floor -

11 8m

1 Level-
(a) Side view (b) Back
B view

Figuure 1. Views of th
he GMT installaation and CFD teelescope model Figure 2. CFD model of telescope andd supports

A simpliified telescope Computer Aidded Design (CA AD) geometry definition in thhe STandard for f the Exchannge of
Product moddel data (STEP P) format was obtained
o from GMTO as show wn in Figure 2.
2 Essential feattures include a
representation of the seven primary mirroor segments andd actuator houssings as conneccted slab hexaggons, the seconndary
mirror as a siingle thick hex
xagon, and the main
m truss support structure underneath
u thee telescope. The bottom of thee GMT
f observing floor inside thee enclosure. The
model is posiitioned on the flat T telescope ellevation angle in the current work w
was kept at a nominal valuee of 60 degreess.
1.2.2 Ennclosure
The simpplifications of the enclosure models
m for the GMT for bothh the open and closed enclosuure base soffit
configurationns are shown in n Figure 3. Thhe original moddels, Figure 3a and 3b, were simplified
s first by omitting thhe truss
framework within
w the enclo
osures and beloow the open sooffit enclosure as
a shown in Figgures 3c and 3d. Next, the fully
open ventinng pattern omittted the thinner vertical suppoort members beetween the top and bottom off each horizontaal vent
opening, show wn in Figure 4a4 and 4b, as thhey are not connsidered to be significant
s for the
t current floww simulations.

(a) Open soffit CAD moddel (b) Clossed soffit CAD model
m

Proc. of SPIE Vol. 9911 991114-2

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


C
EEC=
CC
111
(c) Open soffit simplified CAD model (d) Closed soffit simplified CAD model
Figure 3. GMT enclosure soffit configuration comparison

(a) Original vent openings (open soffit)

(b) Simplified vent openings (open soffit)

Figure 4. Nominal GMT venting patterns

SUBJECT: AVAILABLE PROJECT CONTOURS DATE', 09N312015

MODEL RESOLUTION
- PROVIDE 2.0m CONTOURS WITHIN 5 DIA. OF ENCLOSURE
- PROVIDE 10.0m CONTOURS FOR REMAINING AREA

PREDOMINANT
WIND DIRECTION
AVAILABLE SURVEYED CONTOURS

CENTER OF GMT TELESCOPE PIER


;.

n,)r-!

PREVIOUS SITE MODEL AREA (965m X 845m)


PROVIDED TO BOEING

NEW SITE MODEL AREA - 2,000m X 2,000m


(CENTERED ON TELESCOPE PIER)
TO BE PROVIDED TO BOEING
1250 625 0 1250 2500

SCALE IN METERS

Figure 5. Topographic map of LCO and identification of GMT CFD surface domain

Proc. of SPIE Vol. 9911 991114-3

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


1.2.3 Terrain
The terrain model used in the computational analysis is provided by M3 Engineering as shown in Figure 5. This
topographic image illustrates a 2000 square meter section of the LCO, roughly centered on the GMT enclosure, which
was extracted for use in the CFD model. To enable accurate terrain modeling in the CFD simulations, appropriate
surface sampling resolutions are chosen to adequately represent the features of the terrain around the enclosure. On
the summit, within approximately 5 enclosure diameters around the telescope, surface features of 2 meters in length or
greater are captured. Everywhere else, surface features of 10 meters in length or greater are captured.

The terrain geometry file was originally obtained as a Tagged Image File Format (tiff) file with color mapped to
elevation. The tiff file was converted in AutoCad to a Civil 3D drawing (dwg) file, and was then made available as a
STEP file as depicted in Figure 6. The blue arrow in this figure denotes the approximate location of the telescope on
the summit. Due to the large number of small triangular surface entities (~152,000) present in the file, direct
processing of this model for grid generation could not be completed with standard geometry manipulation tools. As an
alternative, a separate software tool was written to generate a regular ordered surface definition at a similar resolution.
This was accomplished by computing the intersection curve of a vertical planar surface with the terrain surface and
systematically marching in 4 meter increments from one boundary to the opposite side of the domain. For each
intersection line, points were then redistributed such that each line contained points that were equally spaced 4 meters
apart. The end product of this processing was a regular lattice of points that contained cells of 4 square meters, as
shown in Figure 7. The yellow cylinder indicates the approximate location of the baseline telescope location on the
summit. To ensure that the choice of 4 meters was sufficient to accurately capture the surface features of the terrain,
the surrogate lattice surface was overlaid onto the original CAD surface, as seen in Figure 8. As depicted, the
surrogate model retains the key geometry features of the original terrain model, including the edges of the summit
where the telescope/enclosure will be located.

Figure 6. Original surface definition of the LCO terrain

Figure 7. Surrogate geometry model of LCO terrain Figure 8. Overlay of LCO surface (red) with surrogate model (blue)

Proc. of SPIE Vol. 9911 991114-4

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


1.3 Grid Generation
The Boeing-developed Modular Aerodynamic Computational Analysis Process (MADCAP) code is used to first
combine the telescope, enclosure, and terrain geometry models into one single geometry model, as shown in Figure 9.
Once the geometry of all the components is assembled, the resolution of the computational grid discretization required
for an accurate CFD analysis is determined. The cell spacing on the surface mesh are tailored depending on their
location as described below.

Figure 9. Combined GMT, enclosure, and terrain models

1. Surface grid cell edge lengths on the enclosure are held to 1 meter everywhere except for 0.25 meters near the
vent openings to ensure generation of a minimum of 2 cells across the thickness of the opening.
2. Cell edge lengths are specified at 0.25 meters over the entire telescope, including support truss, so that there are
at least 4 to 5 cells across the thickness and width of support members.
3. Cell spacing is held to 1 meter along the perimeter of enclosure intersection with site terrain summit surface.
4. From the edges of the summit, cell spacing is allowed to grow along the terrain surface to 20 meters at the outer
boundaries of the computational domain.
5. In the direction normal to the terrain surface, the first computational cell spacing is 3 millimeters, which
provides a minimum of 25 cells in the velocity boundary layers.

With these grid rules, MADCAP is used to generate a high-quality unstructured tetrahedron surface mesh. Selected
images of the surface grid and a cutting plane near the vent slots for the closed soffit enclosure configuration are shown
in Figure 10. Once the surface mesh is generated, a volume mesh to fill in the computational domain, as depicted in
Figure 11, is computed using the Advancing Front Local Reconnection (AFLR) code developed by Mississippi State
University1. This technique utilizes the orthogonal properties of a prism layer near the surface which transitions to all
tetrahedron cells away from the boundaries. The grid cells grow significantly in size as the distance away from the
enclosure increases since flow gradients vanish and free-stream flow properties are recovered near the outer boundaries.
The nominal grid size used in the current studies totals approximately 16M cells.

Like many ground based telescopes, the GMT design will incorporate a variable height, semipermeable windscreen
over the main enclosure opening to help shield the telescope from excessive wind velocities and flow gradients that
would otherwise be present. A windscreen was modeled in the CFD simulations using both a porous-jump internal
boundary condition (BC) as well as a discrete slab windscreen with circular holes. Both of these techniques are briefly
described next.

Proc. of SPIE Vol. 9911 991114-5

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


J
2 f

:
/If!:

. \

Figure 10. Selected images of the GMT computational grid for the closed soffit configuration

(a) Entire CFD domain (b) Close-up of telescope/enclosure on summit

Figure 11. Volume mesh for CFD simulations

Proc. of SPIE Vol. 9911 991114-6

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


To model the windscreen using the porous wall BC in the flow solver, a curved grid plane is defined and embedded
in the volume grid system just inside the front opening of the enclosure, as depicted in Figure 12. Grid points are
clustered ahead of, and behind, the embedded surface to properly compute the effect of the windscreen on the flow
entering the enclosure, as shown in yellow. The geometric relationship between the embedded grid surface and the
telescope is depicted in Figure 13. The height of the embedded plane is 33 meters, which is the maximum height of a
screen which will not intrude into the optical path of the telescope at an elevation of 60. For the current study, a 25%
porous screen was assumed which is correlated to the momentum loss coefficient through empirical data for thin screens
with round holes. As will be shown in Section 2, the primary effect of the screen is to block/deflect the incoming flow
leading to reduced velocities and increased pressures on the upstream side.

`1`: . .:-!: !
..-

Figure 12. Embedded grid for porous wall BC Figure 13. Relation of windscreen surface Figure 14. Relation of windscreen
with telescope seeing path surface with telescope seeing path
For the modeling of the windscreen using the discrete CAD definition, grid clustering near the holes is illustrated
in Figure 14. This level of resolution is chosen to minimize the number of grid points and computation time for the CFD
simulation while capturing the key effect of flow blockage on the air entering the enclosure. The addition of the discrete
windscreen model increases the size of the computational grid by a factor of 3. However, this direct method should be
used for irregular porosity concepts since the porous wall boundary condition uses empirical relationships based on small
sharp-edged circular holes.
1.4 Flow Solution Boundary Conditions and Domain
The Ansys Fluent flow solver software package2 is used to perform the CFD analysis. Fluent has a variety of
modeling capabilities including fluid flow, heat transfer, and reacting gas chemistry. The computational grid system
generated with AFLR is first converted to a Fluent Case file (.cas). Once the grid is converted and read into the Fluent
code, initial and boundary conditions are set. Because of the very low speed (< 15 m/s) of the air flow for the GMT
analyses, the flow is treated as incompressible (no change in density). A pressure- based solver is used for the
momentum equations. In the current analysis, the energy equation not solved and temperature is constant. The pressure-
velocity coupling scheme used is the Semi-IMplicit Pressure-Linked Equations (SIMPLE) approach. The momentum
equations are solved using a 2nd-order upwind spatial integration. The flow is assumed fully turbulent, and the two-
equation k-omega Shear Stress Transport (SST) model is used to provide predictions of the turbulent viscosity. Further
details on the solution procedure are available3.

The computational domain including specification of boundary conditions is shown in Figure 15. The domain edges
are oriented along the primary wind direction so that the boundary conditions on the lateral side planes can be set as
inviscid and impermeable. The velocity at the inflow plane is set perpendicular to the boundary. The downstream
boundary is set to free stream pressure (zero gauge pressure) for all cases. The choice of the velocity distribution used as
the inflow boundary condition is guided by similar relationships used in wind tunnel testing4. A power-law distribution
is used to relate the local and a reference velocity (U and Uref) to the local and reference height (Z and Zref):

= (1)

Proc. of SPIE Vol. 9911 991114-7

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


For the GMT T analysis, a terrrestrial velocitty profile with exponent n=0.16, appropriatte for rough open flow, is assumed.
a
Based on thee Reference 4 wind
w tunnel datta, a reference height (Zref) off 500m was choosen. Choice of
o the referencee
velocity (Ureff) is guided by wind speed meeasurements att the Cerro Loss Campanas suummit5. The wiind speed is defined as
percentiles off velocity magnitude data colllected from a measurement
m l
location approxximately 10 meeters above graade on
the north endd of the LCO su ummit as depiccted in Figure 16. For this stuudy, the 75th percentile wind speed from nigght
measured datta is 9.8 m/s. ToT achieve this velocity, a vallue of Uref at the upstream bouundary is initiaally assumed. Using
U
this Uref, the Fluent
F CFD co ode is then run to obtain the velocity
v profilee along the tow
wer axis includiing the 10m eleevation
point locationn. Two iteratio ons were required to determinne that an infloow velocity proofile with Uref=7.3
= m/s at a Zref r =500
meters (Figurre 17) results in n a velocity off approximatelly 10 m/s at thee tower location as desired (F Figure 18). For the 25th
percentile wiind speed of 4.0 0 m/s, a value of Uref=3.0 is required.
r

Vmag (n
ni a
io
Figure 15. Flueent boundary connditions for Figure 16. Location
L of LCO
O summit velocitty
CFD simulationns measurement

-0-Tower Locatic

U, = 7.3
7f = 500
= 9.
U = 9.8
1ol
Zw.=10

./.
-IMIKEILMCCE071E
10 2.0 3.0 0 50 60 7.0 8 0 9.0 10.0 11.0
o o oax
4.1
O 10 20 30 40 50 60 7.(0 8 0 9 0 10.0
Ve locity (m/s) Velocity (m /:sl

Fiigure 17. Inflow and tower veloccity profiles Fiigure 18. Enlargged view of toweer velocity profille

A key phhysical phenom menon that is im


mportant in thee CFD analysiss is the acceleraation of the floow from the infflow
boundary aloong the terrain up
u to the summ mit. A near-surrface velocity deficit
d in the innflow profile off ~ 4 m/s (at 100m
above terrainn) increases to a near-surface velocity excesss of ~10 m/s (aat 10m above terrain)
t at the tower location on the

Proc. of SPIE Vol. 9911 991114-8

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


summit as shown in Figures 17 and 18. This occurs because the rising terrain slope contracts and accelerates the flow,
reaching a maximum above the summit.
2. CFD RESULTS
Results of the CFD simulations to assess the effect of the soffit configurations, location of the telescope on the
summit, and the presence or absence of the windscreen on the airflow entering the enclosure are presented here.
Simulation results are reviewed primarily for the 75th percentile head wind velocity and direction (azimuth=20) but also
the telescope looking 90 degrees counterclockwise (right cross wind, azimuth = 110) for both open and closed soffit
enclosure configurations as depicted in Figure 19. The 75th percentile wind speed from night measured data is 9.8 m/s.
The full open venting pattern (as shown in Fig. 4b) is used. The effect of the soffit and trends in flow with the
telescope at alternate locations is presented followed by the effects of the windscreen and alternate venting
configurations.

Primary Wind Direction Primary Wind Direction


N N

(a) Head-wind (b) Right-cross-wind

Figure 19. GMT orientations for nominal site location used in Task 1

2.1 Effect of Soffit at Baseline Site


2.1.1 Flow Characteristics
Analysis of the CFD simulations is performed to better understand the relevant flow physics that drive the effects of
soffit configuration on the quality of the airflow entering the enclosure. The CFD post-processing code Fieldview from
Intelligent Light is used to process all CFD simulations.
13i 0 --NWALnQ1JWl0 0

VIOk
c0000000000

\A
P!-NH:4P19":-'5P9
00000000000
;vv.;
'
A

(a) Closed soffit (b) Open soffit


th
Figure 20. Comparison of velocity and flow streamlines for soffit configurations, 75 percentile headwind, porous windscreen

Proc. of SPIE Vol. 9911 991114-9

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


The direct effect of the soffit geometric configuration on airflow approaching the enclosure (left to right) is shown in
Figure 20 for the 75% headwind case. In this comparison, velocity vectors in the lateral centerplane of the enclosure are
shown and colored according to the local velocity magnitude. The vertical black line on the windward side is the
location of the porous jump boundary condition plane. While the velocity magnitudes within the enclosure are similar
between the solutions of the two soffit types, there is a large recirculating region above the summit just in front of the
aperture for the closed soffit configuration not present in the open soffit. This is the center of the so called necklace
vortex resulting from the stagnation of the flow on the blunt cylindrical enclosure since the vortex hangs around the
obstacle as a necklace as shown as the red streamlines in Figure 21a for the closed soffit configuration. As will be
shown, it is this region that elevates more near-surface air up and into the enclosure compared to the open soffit. For the
open soffit, as shown in Figure 21b, the flow still stagnates on the smaller support cylinder under the enclosure, however
the flow is also contained on three sides under the enclosure which tends to break up this rotational behavior and the
flow more uniformly passes around the smaller cylindrical support column.

(a) Closed Soffit (b) Open Soffit


Figure 21. Streamlines patterns from different soffit configurations

The computed pressure contours in the same center cutting plane for the two soffit types are shown in Figure 22. The
pressure is lower in the center of the necklace vortex compared to the surrounding flow as expected. While this low
pressure region is not seen in the open soffit configuration, there is a flow expansion over the overhanging windward
edge which drops the pressure considerably. The average pressure within and above the open soffit enclosure is seen to
be approximately 5 Pa lower than for the closed soffit. The open soffit provides lower flow blockage than the closed
soffit and as a result the surrounding velocities are slightly higher and the pressure slightly lower.

(Pa) P9 (Pa)
30.0 30.0
25.0 25.0
20.0 20.0
15.0 15.0
10.0 10.0
5.0 5.0
0.0 0.0
1. 10.0
-15.0 -15.0
-20.0 -20.0
-25.0 -25.0
-30.0 -30.0

(a) Closed soffit (b) Open soffit

Figure 22. Comparison of gauge pressure on enclosure symmetry plane as a function of soffit type baseline site, 75th
percentile wind speed, head-wind, porous windscreen, vents full open

Proc. of SPIE Vol. 9911 991114-10

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


2.1.2 m Analyses
Hmin
For the current
c GMT flow
fl analysis, solution
s of the energy equatioon to compute temperature
t grradient effects was
w not
included in thhe CFD simulaations as mentiioned previouslly. However, thhe heating of the air mass onn the terrain thaat
surrounds thee telescope and d enclosure on the LCO summ mit, and the floow of that air innto the enclosuure, is of primaary
concern. In liieu of a compuuted temperaturre gradient, thee key metric ussed in this analyysis is the minnimum height above a
the terrain (H
HMIN) that a flow
w particle enteering the enclossure would expperience as it trraveled from thhe inflow bounndary to
the enclosuree. Figure 23 graaphically illusttrates the definnition of this paarameter as welll as the heightt above grade (H( AG) on
the summit.

Figure 23.
2 Definitions of
o streamline heiights relative to enclosure
e

A more quantitative
q annalysis of the CFD
C simulationns is performedd to characterizze the flow enteering the encloosure. As
mentioned prreviously, miniimum height abbove terrain, HMIN, is definedd as a proxy forr temperature gradient.
g Trackking
HMIN is used to estimate thee amount of unndesirable air enntering the encclosure. Also, it i is assumed thhat air closer too the
ground will exhibit
e larger teemperature graadients than airr further abovee the terrain. Too this end, the flow
f at points at
a the
enclosure enttrance, defined d along lines at various heightts above grade HAG (as defiined in Fig 23)), is characterizzed by
tracing stream
mlines upstream m (backwards in time) to the domain inflow w boundary andd computing thhe HMIN along each e
streamline paath. An illustraation of the lattiice of release points
p (shown for
f the enclosuure with the cloosed soffit) is given
g in
Figure 24.

Figure 244. Streamline rellease points at opening of enclossure

Proc. of SPIE Vol. 9911 991114-11

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


The streamlines are then traced backward in time so that their trajectory can be analyzed and their minimum
distance above the terrain computed. With a value HMIN determined for each release point, a contour plot of values can
be constructed from the aperture array as shown in Figure 25. A point is removed from the plot (represented by white
regions) if that streamline cannot be integrated upstream because of reversed flow in these locations. The flow entering
the enclosure with the closed soffit originates from locations closer to the terrain (darker blue/black contours) than does
flow entering the enclosure with the open soffit.

Hmin (m) Hmin (m)


44.0 1111111111111mi1 I 44.0
42.0 1111! 42.0
40.0 r
40.0
38.0 38.0
36.0 36.0
34.0 34.0
32.0 32.0
30.0 30.0
28.0 28.0
26.0 26.0
24.0 24.0
22.0 22.0
20.0 20.0
18.0 18.0
16.0 16.0
14.0 14.0
12.0
I NI
12.0
10.0 10.0
8.0 8.0
6.0 6.0
4.0 4.0

(a) Closed soffit (b) Open soffit

Figure 25. HMIN contours at enclosure entrance baseline site, 75th percentile wind speed, head-wind, porous windscreen,
vents full open
This trend illustrated in the bar chart in Figure 26. In this plot, the percentage of flow from a given HMIN value is
grouped in the following buckets: 0-8m, 8-16m, 16-32m, and 32-50m. This data quantitatively shows that the effect of
the closed soffit configuration results in more of the lower level air entering the enclosure as compared to the open soffit
configuration, for both the enclosure with and without the windscreen modeled. Specifically, of the total amount of air
entering the enclosure with the windscreen, 17% of the flow for the open soffit originates between the ground and 16m
above terrain. For the closed soffit, 32% of the flow originates between the ground and 16m above terrain. Air that
originates from between 16m and 32m represents 67% of the total air mass entering the enclosure with the open soffit
and 53% of the total air mass entering the enclosure with the closed soffit. In contrast, of the total amount of air entering
the enclosure with no windscreen, 19% of the flow for the open soffit originates between the ground and 16m above
terrain. For the closed soffit, 27% of the flow originates between the ground and 16m above terrain. However, air that
originates from between 16m and 32m represents only 42% of the total air mass entering the enclosure with the open
soffit and 34% of the total air mass entering the enclosure with the closed soffit. This data further corroborates the lifting
effect of the flow, particularly in the range of HMIN from 16m-32m, due to the presence of the windscreen.

80
Open Soffit, Windscreen Off
% Flow Entering Enclosure

70
Closed Soffit, Windscreen Off
60 Open Soffit, Windscreen On
50 Closed Soffit, Windscreen On

40
30
20
10
0
0 to 8m 8 to 16m 16 to 32m 32 to 50m
HMIN Buckets

Figure 26. Amount of flow entering enclosure sorted by HMIN buckets as a function of soffit type and windscreen baseline
site, 75th percentile wind speed, head wind, vents full open

Proc. of SPIE Vol. 9911 991114-12

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


A quantitative comparison of the magnitude of HMIN at the enclosure opening between simulations with both open
and closed soffit configurations with and without the windscreen modeled is shown in Figure 27. These simulations are
performed at the baseline site with the 75th percentile winds and the enclosure positioned in the head wind orientation
with vents full open. In this plot, the minimum distance above the terrain is computed for each line of release points,
given as height above grade (HAG), at the enclosure opening. The data shows that the flow for both closed soffit
configurations originates from closer to the terrain as compared to the open soffit cases. Additionally, as mentioned
above, the effect of the windscreen tends to lift more of the lower level air higher vertically into the enclosure opening
starting around 26m height above ground.

62
60
58
56
54
52
50
Height Above Grade (m)

48
46
44
42
40
38
36
34
32 Open Soffit, Windscreen Off
30
28 Closed Soffit, Windscreen Off
26
24
22 Open Soffit, Windscreen On
20
18 Closed Soffit, Windscreen On
16
14
4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 44 46 48
Minimum Distance Above Terrain Hmin (m)
Figure 27. Comparison of the distribution of HMIN at the enclosure opening as a function of soffit type and windscreen
baseline site, 75th percentile wind speed, head-wind, vents full open

2.2 Effect of Telescope/Enclosure Location on Summit


2.2.1 Terrain-Only CFD Simulation Results
To guide selection of alternate locations for the telescope on the summit, CFD analysis was conducted for the 75%
wind condition without any telescope or enclosure present. Contours of computed surface pressure variations and near-
surface oilflow are shown in Figure 28. Regions of flow expansions, such as around the sharp summit edge, are seen
as blue/black contours. Regions of higher relative pressure are shown as yellow/red contours. The near-surface
streamlines indicate straighter approach flow on the windward and leeward side of the eastern portion of the summit.
Results from this analysis were used to propose two alternate locations (with the telescope and enclosure) as shown in
Figure 29.

Computed velocity magnitude contours and streamlines near the top of the summit are shown in Figure 30 at each of
the three examined site location with the flow direction from left to right. The outline of the enclosure is shown for
reference but was not included in the CFD simulations. The computed data show how the upward sloping terrain near
the windward summit edge for the baseline location causes the flow to decelerate considerably compared to similar
locations at the alternate sites. The approaching flow for the Alternate #2 site is seen to be the smoothest in terms of
terrain profile, streamline pattern, and velocity gradients although also having the largest peak velocity near 12 m/s as
shown in Figure 30c.

Proc. of SPIE Vol. 9911 991114-13

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


V N
Baseline Site

Figure 28. Computed terrain surface pressures and oilflow lines

Figure 29. Locations of baseline and alternate site locations on summit

":-

(a) Baseline (b) Alternate #1 (c) Alternate #2


Figure 30. Terrain-only velocity magnitude contours at site locations
2.2.2 Flow Characteristics at Alternate Site Locations
In addition to analyzing the telescope and enclosure at the northwest side of the summit (baseline), two alternate
sites were also analyzed with CFD to understand this sensitivity. The Alternate #1 was approximately one enclosure
diameter from the Baseline while the Alternate #2 location was centered on the southeast end of the summit as shown in
Figure 29. The 75% headwind condition was assumed for all simulations and the windscreen was not modeled.

Proc. of SPIE Vol. 9911 991114-14

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


Closed Soffit

(a) Baseline Site (b) Alternate #1 Site (c) Alternate #2 Site


th
Figure 31. Velocity vectors in the lateral mid-plane of the enclosure, 75 percentile headwind, no windscreen

Despite the different site locations, the flow stagnation into the enclosure base from the simulations with the open
soffit at all three site locations is similar in shape and magnitude as shown in Figure 31. Inside the enclosure, however,
the velocity distribution across the primary mirror shows slightly different trends. In all three cases, the flow entering the
bottom of the enclosure impinges on the base of the primary mirror causing a local region of recirculating flow. At the
baseline location, some of the lower air flows over the primary mirror generating a thicker boundary layer as compared
to the predicted flow at the alternate locations. From the Alt1 and Alt2 simulations, the flow entering at the bottom of the
enclosure turns more towards the floor of the enclosure resulting in higher velocity air flowing over the primary mirror,
as well as some flow being diverted down and around the backside of the mirror.

In contrast, the simulations with the closed soffit show significantly different flow characteristics ahead of and
inside the enclosure. The size of the region of flow stagnation at the base of the closed soffit decreases significantly
based on the simulations at the Alt1 and Alt2 locations as compared to the baseline location. As a result, there is less of a
lifting effect of the flow into the enclosure at the Alt1 and Alt2 locations, which diminishes the area of flow recirculation
above the primary mirror in the optical path. In fact, this data suggests that the closed soffit velocity profile along the
beam axis for the Alt2 location is only slightly less favorable as compared to the velocity profile computed for the open
soffit configuration at the baseline location.
2.2.3 Hmin Analyses at Alternate Sites
A quantitative comparison of the magnitude of HMIN at the enclosure opening for simulations with the 75th
percentile winds for both open and closed soffit configurations, with the enclosure in the head-wind orientation
positioned at the three site locations, is shown in Figure 32. In this plot, the minimum distance above the terrain is
computed for each line of release points, given as height above grade (HAG), at the enclosure opening. The data shows
that the flow for all closed soffit configurations originates from closer to the terrain as compared to the open soffit cases.
Near the bottom of the enclosure opening, at the elevation where data for all six simulations is available (HAG=21m), the
lowest value of HMIN is computed as approximately 7m for the closed soffit at the baseline location. The highest value of
HMIN is computed as approximately 15m for the open soffit configuration at the Alt1 location. The difference between
the highest and lowest value is approximately 8m, which is more than double the closed soffit value. This data suggests
that the location of the enclosure on the summit has a larger effect on the flow entering the enclosure for the closed soffit
configurations than it does for the open soffit configurations. It is also noted that there is an 11% difference in HMIN
between the simulation with the open soffit configuration at the baseline site location and the closed soffit configuration
at the Alt2 site location. Although the uncertainty in the computational results is not known, this difference suggests that
the closed soffit configuration at the Alt2 site location may provide similar flow quality inside the enclosure compared to

Proc. of SPIE Vol. 9911 991114-15

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


the open soffit at the baseline site location. This conclusion is also supported by the data in Fig. 31 which shows
differences in velocity magnitude over the mirror along the beam axis between these two cases appear to be small.

62
Baseline - Open Soffit
58 Baseline - Closed Soffit
Height Above Grade (m) 54 Alternate #1 - Open Soffit
50 Alternate #1 - Closed Soffit
46 Alternate #2 - Open Soffit
Alternate #2 - Closed Soffit
42
38
34
30
26
22
18
14
4.0 8.0 12.0 16.0 20.0 24.0 28.0 32.0 36.0 40.0 44.0 48.0 52.0
Minimum Distance Above Terrain Hmin (m)
Figure 32. Comparison of the distribution of HMIN at the enclosure opening at GMT site locations 75th percentile wind
speed, head-wind, no windscreen, vents full open

In Figure 33, the percentage of flow from a given HMIN value is grouped in the following buckets: 0-8m, 8-16m,
16-32m, and 32-50m, consistent with the analysis presented in Fig. 24. This data quantitatively shows that the effect of
the closed soffit configuration results in more of the lower level air entering the enclosure as compared to the open soffit
configuration. Further, the smallest amount of low-level air entering the enclosure for the open soffit occurs at the Alt1
site location, while the smallest amount of low-level air entering the enclosure for the closed soffit occurs at the Alt2 site
location. Specifically, 19%, 7%, and 10% of the total flow entering the enclosure with the open soffit configuration at the
baseline, Alt1, and Alt2 site locations, respectively, originates from between 0 and 16m above the terrain. In contrast,
30%, 24%, and 18% of the total flow enters the enclosure with the closed soffit configuration at the baseline, Alt1, and
Alt2 site locations, respectively. In general, there is no significant difference between simulations by soffit type or site
location on the amount of flow entering the enclosure from between 16m and 32m above terrain. Also, small amounts of
near-ground layer air from heights between 0 and 8m are predicted to enter the enclosure for the closed soffit
configuration only at the baseline site location, but not at the Alt 1 or Alt2 site locations. Finally, more of the air
originating from heights of 32m above terrain enters the enclosure for the open soffit as compared to the closed soffit.
80
% Flow Entering Enclosure

Baseline Open
70
Baseline Closed
60 Alternate 1 Open
Alternate 1 Closed
50
Alternate 2 Open
40 Alternate 2 Closed
30

20

10

0
0 to 8m
i Fil
EI 11
8 to 16m
i
16 to 32m
HMIN Buckets
i
32 to 50m

Figure 33. Amount of flow entering enclosure sorted by HMIN buckets as a function of soffit type and site location 75th
percentile wind speed, head-wind, no windscreen, vents full open

Proc. of SPIE Vol. 9911 991114-16

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


2.3 Effect of Telescope/Enclosure Orientation

(a) 75% Headwind (b) 75% Tailwind

.
.
ta.

:::
.
.a

t
(c) 75% 90-Degree Crosswind (d) 75% 45-Degree Crosswind
Figure 34. Comparison of flow vectors colored by velocity magnitude in plane parallel to flow direction through enclosure center as
function of telescope pointing baseline site, closed soffit, 75th percentile winds, porous windscreen, vents full open
Numerous CFD simulations were conducted for the 75th percentile wind at several pointing directions of the closed-
soffit telescope with the initial ventilation scheme fully open and a porous windscreen model included. Velocity vectors
colored and scaled by total velocity magnitude from the CFD simulations for each of the four enclosure orientations, in
the plane through the mirror center point aligned with the primary flow direction, are presented in Figure 34. For the
head-wind case, the flow deceleration approaching the windscreen (represented by thin black vertical line) is evident.
Similarly, for the tail wind case, flow stagnation into the back enclosure wall is clearly seen since this vent configuration
terminates toward the rear centerplane. For the right cross wind case, flow directly enters all four open vent levels, and
high velocity air is shown streaming over and around the mirror with a considerable recirculation region directly over the
primary mirror. For the right 45 cross wind case, the flow stagnates on the outside of the enclosure door facing the
primary wind direction. For this simulation, flow enters the enclosure on the left and exits the enclosure on the right
similar to the right cross wind case, albeit at a lower overall velocity magnitude.
2.4 Effect of Telescope/Enclosure Venting Configuration
a

E
.R\ R

000000000000
000000000000
--t .K'9tiPF'P
Ai
td1R

(a) Vents full open (b) Video1, VRCW1 venting configuration


http://dx.doi.org/10.1117/12.2231933.1
Figure 35. Comparison of velocity vectors in a plane parallel to flow direction through enclosure center as function of
venting configuration baseline site, closed soffit, 75th percentile winds, right cross wind, porous windscreen

Proc. of SPIE Vol. 9911 991114-17

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


For the right cross-wind telescope orientation, improved venting configurations are explored in order to improve the
quality of the flow in and around the primary and secondary mirrors inside the enclosure. Velocity vectors in the plane
through the mirror center point aligned with the primary flow direction, for the right cross wind orientation with two
venting configurations (full open and VRCW1) are presented in Figure 35. The simulations are obtained at the baseline
site location with 75th percentile winds for the closed soffit enclosure configuration with windscreen model. The
VRCW1 venting pattern closes the level 1 and 2 vents on the right windward side of the enclosure as seen in Figure 35b.
The effect of closing these vents results in significantly improved flow quality near the primary mirror surface and lower
velocities in the region of the secondary mirror. Closing of the level 3 vent on the right side of Figure 35b would likely
further reduce the flow gradients between the primary and secondary mirrors.
2.5 Time-Accurate Flow Modeling
All computed results presented thus far have been from RANS, or steady-state flow solutions. The solutions
obtained from this approach, when sufficiently converged, provide adequate predictions of basic flow characteristics of
the various configurations. This type of analyses has been useful for top-level design considerations such as site
location, soffit type, and ventilation configurations. However for more accurate and detailed flow predictions of more
evolved telescope designs, the time-dependent approach should be utilized. In particular this allows estimations of
unsteady pressure loads possible which help size the structural elements to provide the desired stiffness needed for the
optical performance of the telescope. If the energy equation is active and temperature effects are included, the time
varying thermal behavior of the enclosure with the chosen ventilation scheme can also be investigated.
To exercise this analysis mode in the FLUENT program, a test case was computed of the closed-soffit configuration
with a 75th percentile 90-degree crosswind case and the windscreen model included. The turbulence model employed is
the Detached Eddy Simulation (DES) model that directly resolves turbulent eddies away from viscous walls and returns
to the RANS model near the wall where direct turbulent behavior cannot be resolved without adding extremely high grid
resolution and resorting to the costly Large Eddy Simulation in these regions. The same computational grid that was
used for the previous RANS analyses was used for this initial time-dependent simulation but typically the mesh is
refined to properly resolve the variety of turbulent length scales in the problem. The time step chosen for the simulation
was 0.01 seconds based on the 10 m/s nominal flow velocity and smallest grid sizes above the telescope near 15 cm.
Full solution saves were made every 20 time steps or every 0.2 seconds to create animations such as isocontours of q-
criterion of which a single frame is shown below in Figure 36. The q-criterion is defined as the second-invariant of the
deformation tensor and is written as

Q = 0.5(W*W - S*S) (2)

where W is the vorticity magnitude and S is the mean rate-of-strain. This allows an effective visualization of the
turbulent flow structures by selecting a small positive value for the contour and coloring according to a scalar function,
in this case static gauge pressure. The figure reveals good resolution of small turbulent structures near the telescope
where the grid cells are as small as 15 cm but poor resolution of the structures near areas of large (1m) grid spacing (as
expected) such as near the windward edge of the sharp enclosure doors at the top of the enclosure.

Figure 36. Isocontours of q-criterion colored by gauge pressure baseline site, closed soffit, 75th percentile winds, right
cross wind, porous windscreen

Proc. of SPIE Vol. 9911 991114-18

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


As the flow solution evolves in time, statistical parameters such as time-mean and Root-Mean-Squared (RMS)
quantities can be computed such as surface pressure shown in Figures 37a and 37b respectively for the telescope and
floor surfaces. These quantities are based on approximately 60 seconds of simulated time. While the variance in
average surface pressure on the primary mirror is relatively small at 5 Pascals, this variance has a bearing on the design
of the adaptive optics controlling the primary mirror shaping. The RMS pressure variations in Figure 37b show elevated
levels near the edge of the primary mirror and even higher levels on the secondary mirror (white contours) indicating the
largest pressure fluctuations are in these areas. The unsteady pressure environment on the secondary mirror is perhaps
one of the most important design considerations since its stability is paramount to optical performance.

(a) Average Surface Pressures (b) RMS Surface Pressures

Figure 37. Computed contours of time-averaged telescope surface pressures and RMS surfaces pressures baseline site, closed
soffit, 75th percentile winds, right cross wind, porous windscreen

3. FUTURE WORK

After initial successful RANS analyses of the most major design considerations for the GMT, future work will focus
on higher fidelity time-dependent analyses using a more detailed telescope model as shown in Figure 38. Unsteady CFD
analyses including these detailed components will provide much needed data for understanding structural design
requirements. Inclusion of the energy equation will provide spatial and temporal predictions of temperature and density
variations to aid design of thermal management techniques and ultimately the predicted optical performance of the GMT
using Optical Path Integration (OPD) procedures used routinely for aircraft turret analyses to determine optical
performance.

Figure 38. Computer Aided Design (CAD) model of detailed primary and secondary mirror components for CFD

Proc. of SPIE Vol. 9911 991114-19

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx


CONCLUSIONS
CFD analysis of the GMT and enclosure is performed to study the impact of the enclosure soffit design on the
quality of the air flow entering the enclosure. Simulations for 75th percentile winds with the telescope pointing into the
wind, away from the wind, pointing 90 degrees to the wind, and pointing 45 degrees in to the wind for both closed and
open soffit configuration concepts were completed. Since temperature gradients in the flow are not directly computed in
this CFD simulation, the minimum height above terrain is defined as a proxy to characterize the quality of air as
undesirable entering the enclosure. Both qualitative and quantitative analyses of the CFD results indicate that more of
the undesirable quality air enters the enclosure with the closed soffit than enters with the open soffit, a trend that was
seen regardless of GMT site location. This is due to an increased area of flow stagnation at the base of the enclosure with
the closed soffit, which pushes more of the lower-elevation air into the enclosure opening. The smallest amount of low-
level air entering the enclosure for the open soffit occurs at the Alt1 site location, while the smallest amount of low-level
air entering the enclosure for the closed soffit occurs at the Alt2 site location.
The effect of venting configurations is predicted to be most significant for the 90-degree and 45-degree crosswind
cases where the approach flow compresses against the enclosure and air is forced through the ventilation slots and flows
above and below the primary mirror. Closing the lowest level L1 and L2 vents (VRCW1 venting) provides a more
favorable flow field for the cross-wind cases by causing air at the higher L3 location to exit out the top of the enclosure
rather than directly over the telescope. Velocity gradients and flow recirculation over the telescope are also reduced
significantly with the VRC1W1 venting scheme.
Future work will focus on time-accurate CFD simulations using higher-fidelity geometry modeling and more refined
grid resolution to be able to predict figures of merit directly applicable to aiding the design of many of the GMT
components including primary mirror actuation, secondary mirror stability, enclosure and door unsteady loading, and
windscreen considerations such as thickness, porosity, and deployment schemes for different pointing directions.
Inclusion of the energy equation in the CFD will allow spatial and temporal variations of temperature and density so that
Optical Path Distortion (OPD) estimates can ultimately be obtained to predict potential image quality.

ACKNOWLEDGEMENTS
This work has been supported by the GMTO Corporation, a non-profit organization operated on behalf of an
international consortium of universities and institutions: Astronomy Australia Ltd, the Australian National University,
the Carnegie Institution for Science, Harvard University, the Korea Astronomy and Space Science Institute, the So
Paulo Research Foundation, the Smithsonian Institution, the University of Texas at Austin, Texas A&M University, the
University of Arizona, and the University of Chicago.
REFERENCES
[1] Marcum, D.L., Anisotropic Solution Adaptive Unstructured Grid Generation Using AFLR, Final
Report, NASA Grant No. NNL04AA91G, (March 2007).
[2] http://www.ansys.com/Products/Fluids/ANSYS+Fluent
[3] Kelecy, F.J. Coupling Momentum and Continuity Increases CFD Robustness, Ansys Advantage, Vol 2,
Issue 2, (2008).
[4] Chen, D., and Cochran, L., CPP Wind Tunnel Test Report Topography, CPP, Inc., CPP-EF-DOC-
00007, Rev A, (1 May 2011).
[5] Hardie, K., and Trancho, G., GMT Environmental Conditions, GMT-REF-00144, Rev C., (24 July
2015).
[6] Vogiatzis, K., and Thompson, H., On the Precision of Aero-Thermal Simulations for TMT, Thirty
Meter Telescope (USA), SPIE 9911-40, (2016).
[7] Teran, J., Burgett, W., and Grigel, E., GMT Site, Enclosure, and Facilities Design and Development
Overview and Update, SPIE 9906-35, (2016).
[8] McCarthy, P., Fanson, J., and Bernstein, R., Overview and Status of the Giant Magellan Telescope
Project, GMTO Corp., SPIE 9906-37, (2016).
[9] Danks, R., Smeaton, W., Initial Computational Fluid Dynamics Modeling of the Giant Magellan
Telescope Site and Enclosure, SPIE 9911-41, (2016).
[10] Farahani, A., Kolesnikov, A., and Cochran, L., GMT Enclosure Wind and Thermal Study, GMTO
Corp., Proc. 8444, 1-13 (2012).

Proc. of SPIE Vol. 9911 991114-20

Downloaded From: http://proceedings.spiedigitallibrary.org/ on 02/01/2017 Terms of Use: http://spiedigitallibrary.org/ss/termsofuse.aspx

You might also like