You are on page 1of 15

Proceedings of ASME Turbo Expo 2017: Turbomachinery Technical Conference and Exposition

GT2017
June 26-30, 2017, Charlotte, NC, USA

GT2017-63126

A SENSITIVITY STUDY OF GAS TURBINE EXHAUST DIFFUSER-COLLECTOR


PERFORMANCE AT VARIOUS INLET SWIRL ANGLES AND STRUT STAGGER ANGLES

Michal P. Siorek Stephen Guillot, Song Xue* , Wing F. NG**


Solar Turbines Incorporated Techsburg Inc.
San Diego, CA, USA Christiansburg, VA, USA
Siorek_Michal_P@solarturbines.com sguillot@techsburg.com, wng@techsburg.com,

ABSTRACT PIV Particle image velocimetry


Ps Static pressure
This paper describes studies completed using a quarter-scaled rig Ps
¯ Mass-average static pressure
to assess the impact of turbine exit swirl angle and strut stagger on a Pt Total Pressure
turbine exhaust system consisting of an integral diffuser-collector. Pt
¯ Mass-average total pressure
Advanced testing methods were applied to ascertain exhaust QPS Quadratic Pressure Strain
performance for a range of inlet conditions aerodynamically matched to SST Shear Stress Transport
flow exiting an industrial gas turbine. Flow visualization techniques TDC Top-dead-center
along with complementary Computational Fluid Dynamics (CFD) TRIT Turbine Rotor Inlet Temperature
predictions were used to study flow behavior along the diffuser x Axial Distance
endwalls. Complimentary CFD analysis was also completed with the
aim to ascertain the performance prediction capability of modern day
analytical tools for design phase and off-design analysis. The K-Epsilon 1. INTRODUCTION
model adequately captured the relevant flow features within both the
diffuser and collector, and the model accurately predicted the recovery Part of the life cycle of many industrial gas turbines often includes
at design conditions. At off-design conditions, the recovery predictions one or more performance uprates, typically defined by an increase in
were found to be pessimistic. The integral diffuser-collector exhaust pressure ratio, flow, and/or firing temperature. For example, the Solar
accommodated a significant amount of inlet swirl without a degradation Turbines TitanTM 130 two-shaft engine was initially introduced in 1998
in performance, so long as the inlet flow direction did not significantly at an ISO power rating of 13.3 MW (17,800-hp), a thermal efficiency of
deviate from the strut stagger angle. Strut incidence at the hub was 34.5% and a compressor pressure ratio of 16:1 [1]. The engine has
directly correlated with reduction in overall performance, whereas the evolved to a power rating of 16.7 MW (22,490 hp), an increase in
diffuser-collector performance was not significantly impacted by strut turbine rotor inlet temperature (TRIT) of 100ºF, and a compressor
incidence at the shroud. pressure ratio of 19:1. Similarly, the Solar Turbines Mars 90 was
initially introduced in 1980 at an ISO power rating of 8.3 MW (11,150-
hp), a thermal efficiency of 32.8%, and a compressor pressure ratio of
NOMENCLATURE 16:1 [2,3]. The product has evolved to power rating of 11.8 MW
(15,900hp), a thermal efficiency of 34.4%, an increase of TRIT by over
BDC Bottom-dead-center 200ºF, and a compressor pressure ratio of 17:1. Depending on scope,
DEHS Di-Ethyl-Hexyl_Sebacat modifications to the external airfoil geometry may not be needed to meet
DOE Design of experiments uprate performance targets. As a result, the turbine section may operate
Cp Exhaust recovery at off-design with the last stage being most impacted. The exhaust
Cpi Local recovery coefficient system is thereby subjected to variations in inlet flow conditions
Cpo Total pressure coefficient specifically the velocity and the flow angle, which adversely affect
∆Cp Difference in recovery, Cp- Cp, ref pressure recovery and affect overall engine performance. The change
DES Detached Eddy Simulation in swirl could impact exhaust diffuser performance and must be
EB Elliptical blending accounted for when conducting overall turbine performance
ISO International Organization for Standardization assessments in the engine growth design phase. Changes to diffuser
L Diffuser length inlet flow conditions, specifically variations in swirl angle, will occur
LPS Linear Pressure Strain for industrial two-shaft engines with power turbines operating at a wide

* Song Xue is currently affiliated with Concepts NREC


** Wing Ng is the Chris C. Kraft Endowed Professor at Virginia Tech

1 Copyright © 2017 Solar Turbines Incorporated


range of speeds and exhaust flow rates. The ability for a design engineer recovery when integrating a collector with an industrial gas turbine
to accurately account for changes in exhaust flow swirl angle is exhaust diffuser with profiled struts set at a 7o stagger. In this study, the
evaluated in this study. addition of inlet swirl up to 21o improved recovery, and a portion of the
loss introduced by the collector was reclaimed.
The classic work of Sovran and Klomp [4] did not consider effects
of inlet swirl angle in their detailed characterization work, however, the Several numerical investigations were completed to study the
influence of inlet swirl on diffuser performance was recognized as influence of inlet flow conditions on both the flow behavior and the
important during that era. Japikse and Baines [5] compiled performance performance of exhaust hoods used in large steam turbine [13, 14, 15,
information from several swirl angle studies completed on annular and 16] and gas turbine [17] applications. These studies mainly reported
diffuser geometries without struts. In most cases, peak diffuser on the necessity of including realistic inlet flow angles in numerical
performance was observed at an inlet swirl angle of 10º to 20º. analysis and discussed predictions of complex flow structures within the
McDonald et al. [6] conducted inlet swirl angle sensitivity studies for a collector. These numerical studies along with particle image
number of conical diffusers at incompressible flow conditions and velocimetry (PIV) data [18, 19], probe data [12, 16], and flow
Kumar and Kumar [7] completed similar work for annular diffusers. visualization data [10, 12] all document the formation of two counter-
Both authors measured an improvement in diffuser recovery with rotating vortices. Several authors [14, 16, 17] observed variations in
introduction of swirl and Kumar observed an optimum swirl angle of vortex size, shape and strength with changes in inlet swirl angles.
17º for diffuser area ratios of three and below. Both authors observed a Despite differences in geometries studied, the authors reported regions
performance benefit with the introduction of inlet swirl to be greater for of backflow at the collector exit plane. Diffuser inlet swirl was found
diffusers with aggressive area ratios with established regions of to influence both the size and location of backflow regions. In general,
separation when operating without swirl. Kumar attributed the increase the reports use Computational Fluid Dynamics (CFD) analysis to
in pressure recovery to radial pressure gradients that enhance complement test data and highlight the ability of CFD analysis in
momentum exchange between the boundary layer and mainstream flow. capturing the main flow structures and general trends of pressure loss.
McDonald indicated that the optimum length ratio (L/R1) may be Good agreement between CFD and experimental data has been found
reduced for diffusers operating with inlet swirl and this should be when optimizing geometry with flow at the design point [19]. However,
considered in the design process. limited studies were found that discuss recovery prediction capabilities
of current state-of-the art CFD codes for diffuser-collector exhausts with
Senoo et al. [8] completed similar swirl angle studies for diffusers integral struts that are operating at a wide range of incidence angles. A
with struts and concluded that the debit in performance due to the number of these studies such as Sultanian et al. [17] were completed
increase in drag can be offset by strut design considerations including when mesh sizes were limited to within a couple of million elements by
airfoil shape and stagger angle. For axial flows, setting a strut at a small computational resources available at the time. Computational
incidence angle to induce swirl was shown to yield a performance advancements have enabled greater model resolution including
benefit. For large inlet swirl angles, utilizing the strut to reduce swirl increased mesh densities and the capacity to solve unsteady flow fields
was found to reduce the performance loss associated with diffuser with models such as Detached Eddy Simulation (DES). The accuracy
operation at high swirl angles. Vassiliev et al. [9] studied the impact of of CFD predictions with these higher resolution models is of particular
inlet swirl angle and inlet Mach number on pressure recovery for an interest.
annular diffuser geometry representative of an industrial gas turbine
engine operating at full and part-load conditions. An adjustable inlet This study focusses on operation of a diffuser-collector gas turbine
guide vane was used in the experimental work with the baseline inlet exhaust at off-design conditions. Experimental measurements were
swirl angle linearly varying from -30º at the inner casing to 0º at the gathered on a quarter-scale test rig with inlet flow conditions matched
shroud. The authors provided swirl angle sensitivities for both exhaust using aerodynamic similarity parameters. Rig data was gathered for
pressure loss and pressure recovery, and concluded that in the range of both axial swirl and three angles that were similar in distribution to
interest, the pressure recovery is practically independent from Mach engine outlet conditions. A design of experiments (DOE) was
number. A reduction in pressure recovery was measured when the inlet completed for a range of strut stagger angles. Detailed measurements,
swirl angle at the hub exceeded -30º, leading to the formation of flow visualization within the diffuser section, as well as complementary
separation zones along the strut. Complimentary CFD predictions of CFD analysis were completed. The effect of swirl on diffuser-collector
diffuser recovery were accurate to 4% as compared to test data. In a performance was reported on. The accuracy of CFD analysis as an
separate report, the authors found that the realizable K-Epsilon analytical tool used for assessing exhaust diffuser-collector performance
turbulence model with two-zone near wall treatment was best suited for at design and off-design inlet flow conditions was evaluated.
predicting relative differences in diffuser recovery for a wide range of
diffuser geometries [10].
2. TEST FACILITIES
Japikse and Pampreen [11] completed scale-model rig testing to
study the effect of inlet swirl angle on the recovery of a double discharge 2.1 Wind Tunnel
exhaust collector designed for an automotive gas turbine. The addition
of a collector to a straight-walled annular diffuser operating with axial The impact of inlet swirl distribution on the performance of a
inlet flow adversely impacted the diffuser recovery by five to eleven diffuser-collector and the associated role of strut design were
points and the deficit was Mach number dependent. The swirl angle investigated in a cold-flow blowdown wind tunnel facility shown in
dependency was greatly reduced with the presence of the collector and Figure 1. The facility consisted of a Kaeser 100 hp axial screw
a recovery of 0.45 was reported at a high swirl angle of 45°. In contrast, compressor attached to a downstream Airtec dryer that was connected
annular diffuser performance diminished by more than 0.25 points when to receivers. An 8-inch globe valve regulated the flow through the rig,
operating at inlet flow angles above 45°, and the measured recovery was and the valve was adjusted with a pneumatic actuator connected to a
found to be below 0.35. Xue et al. [12] reported a similar drop in Siemens controller. Typical steady-state test durations were 10 to 12

2 Copyright © 2017 Solar Turbines Incorporated


seconds and a 15-minute recharge period was required between tests.
Twelve consecutive tests were required to gather data for each test case.

2.2 Diffuser Collector Geometry

The test rig, as shown in Figure 2, consisted of a quarter-scale


diffuser-collector geometry representative of an industrial gas turbine
exhaust system. A cross section of the rig was provided in Figure 3.
The upstream flow conditioning section consisted of a bellmouth, a
turbulence grid, turning vanes, and a contraction used for controlling the
boundary-layer thickness. This upstream section was specifically
designed to simulate turbine exit flow conditions. The contraction
design was shaped to size the boundary layer thickness. Swirl
distributions were controlled with interchangeable swirl vanes, and
three custom turning vanes were designed to simulate the engine exit
Figure 1: Subsonic wind tunnel facility for the testing of 1/4 scale flow angle at design and off-design conditions. Operation without swirl
diffuser collector was also made possible by removing the swirl vanes and inserting at
both endwalls custom rings machined to allow for minimal change in
endwall profile tolerance. Tip leakage flow was not included in this
study. The turbulence grid was designed according to Roach et al. [20]
for an intensity level of 4% at the inlet plane to the diffuser. Hotwire
measurements validated these conditions.

A constant area duct was placed between the flow conditioning


section and the annular diffuser. Inlet measurements were gathered with
this constant area section. The test section consisted of an annular
diffuser of outlet-to-inlet cross-sectional area ratio of 2.45 and a
collector. The diffuser has eight integral struts that were defined by a
constant cross-section profile shown in Figure 4. The strut stagger angle
was adjusted along the center of the leading edge ellipse, and three
stagger angles of -12°, 0°, and +12° were investigated. The diffuser
struts were leaned in the rig as shown in Figure 5. Lean is incorporated

Figure 2: The diffuser-collector test rig hardware

Figure 4: Diffuser strut profiles and stagger angles

Figure 3: Section view of the diffuser-collector rig.


Figure 5: Forward looking aft view of the exhaust diffuser showing
strut lean

3 Copyright © 2017 Solar Turbines Incorporated


in gas turbine exhaust diffusers to reduce thermal stresses during relative to the maximum theoretical (ideal) pressure rise. The recovery
transient operation. A collector was used to turn the annular flow was calculated as shown below.
radially and exhaust the flow out of a vertical stack. A ramp just
downstream of the diffuser was positioned along the inner-wall to guide
the flow. The collector consisted of a flat-plate back wall, an angled
front wall, and round sidewalls. A top view portraying diffuser-collector
interface features was provided in Figure 8 and an isometric view Diffuser Pressure Measurements – Both the hub and shroud were
featuring the collector annular cross-section was provided in Figure 9. instrumented with axial rows of pressure taps at the 0o and 180o
locations, 0o being aligned with the collector exit and referred to as top-
dead-center. The axial locations of each tap were measured from the
2.3 Inlet Flow Conditions and Distributions start of the annular expansion at the inlet of the diffuser and normalized
by the axial length of the shroud as shown by the diagram in Figure 8.
The inlet flow conditions were aerodynamically scaled to match Additional taps measuring the circumferential variation of the pressure
turbine exit Mach number and Reynolds number of 0.42 and 950,000 distribution were located on the shroud at 55.1% and 96.5% of diffuser
respectively. The Reynolds number was based on diffuser inlet length and at 96.5% of the diffuser length on the hub.
hydraulic diameter. Diffuser inlet blockage of 5% was generated with
the custom-built contraction. A Pitot probe of 1/32 inch in diameter was
designed for boundary layer measurements and used to evaluate the inlet
blockage. The circumferential average diffuser inlet profile of total
pressure coefficient, Cpo, is shown in Figure 6 and calculated based on
the equation provide below. Similarly, distribution of diffuser inlet
Mach number were provided in Figure 7. Diffuser inlet distribution of
total pressure and Mach number were nearly uniform. Data from both
measurements and predictions were provided.

௉ ି௉
೟భ തതതത
೟భ
‫ ݋݌ܥ‬ൌ തതതത തതതത
௉೟భ ି௉ ೞభ

Uniform axial flow at a swirl angle of 0º and three non-uniform


swirl angle distributions with mass-averaged values of -14.6°, -6.6°, and
9.4° were considered. Swirl vanes were used to impose a radial gradient
in inlet angle. Both the -14.6° and -6.6° swirl vanes were designed with
greater turning at the hub. In contrast, the 9.4° swirl vane was designed
with greater turning at the shroud. The swirl angle distributions were
provided in Figure 14 located in Section 5.2.

Figure 6: Radial distribution of the inlet total pressure coefficient,


3. INSTRUMENTATION, DATA ACQUISITION, ANALYSIS Cpo - circumferentially averaged

Diffuser Inlet Measurements - A 3/16” United Sensor WAC-187 three-


hole wedge probe was used for measurement of total pressure (po1), and
static pressure (ps1) and this data was used to derive the inlet swirl angle.
These data were acquired at an axial location equivalent to 85% of the
annular height upstream of the inlet plane of the diffuser, as shown in
Figure 3. The measurements were taken using six radial traverses
equally spaced around the circumference. Data was gathered at span-
wise locations of 10%, 20%, 35%, 50%, 65%, 80% and 90% of the
annulus height. A smaller Pitot probe was used to acquire the boundary
layer profile on the hub and shroud at both the 0 and 180o locations.

Collector Exit Measurements - A row of 10 static pressure taps located


at the collector exit plane were used to record the exit pressure, Ps2. The
exit pressure taps shown in Figure 8 were located at a distance
equivalent to 1.05 diffuser exit diameters downstream of the rig
centerline. The back wall of the rig was not instrumented with pressure
taps since it was used for optical access for Particle Image Velocimetry
(PIV) measurements.

Performance Indices - The pressure recovery of the overall diffuser-


collector system was calculated based on mass averaged diffuser inlet Figure 7: Radial distribution of the inlet Mach number,
total and static pressures and average collector exit static pressure. The circumferentially averaged
definition of pressure recovery relates the actual static pressure rise

4 Copyright © 2017 Solar Turbines Incorporated


Flow Visualization Technique – Surface streak lines were visualized
using a silicon oil mixed with fluorescent pigment. The rig was then
run several times to allow the oil to setup completely. The rig was then
disassembled and illuminated with an ultraviolet light to enhance the
visualization of the fluorescent pigment in the oil. The results were
photographed using a high-resolution DSLR camera.

Stereoscopic PIV System - A commercial stereoscopic Particle Image


Velocimetry (PIV) system from LaVision was used to obtain the three-
dimensional flowfield within the collector at a frequency of 30Hz. An
oil-based seeding particle, atomized Di-Ethyl-Hexyl-Sebacat (DEHS),
was used to produce nominally 1µm diameter particles providing flow
response to approximately 100 kHz. The pair of flow imaging cameras
were calibrated in situ using a stereoscopic calibration plate and image
registration software provided by LaVision. Cross-correlation post-
processing was performed using LaVision DaVIS software.

The flow exiting the collector was captured at twelve two-


dimensional measurement planes oriented parallel to the back wall of
the collector. The data was re-created at a collector exit plane that was
normal to the flow and this data was used for comparison with the CFD
Figure 8: Location of the diffuser endwall and collector exit pressure results. A diagram of the PIV measurement volume was shown in
taps Figure 9.

DAQ and Instrumentation - Data were acquired using high-speed 16-


bit National Instruments PCI6259 M-Series A/D systems running
LabVIEW. Scanivalve ZOC17 analog pressure transducers were used
for all pressure measurements. Transducers with a +/- 5 psid range and
a +/- 0.08% full scale accuracy were used for data used to derive the
pressure recovery and a combination of +/- 5 and +/- 15 psid
transducers were used for measuring the diffuser pressure distributions.
Data reported at each of the 42-inlet measurement points were averaged
from 2,400 pressure measurements acquired over a three second period.
Due to the short test duration of the blow down facility, 12 separate tests
were needed to acquire all 42-measurement points. A reference inlet
total pressure probe and inlet wall static pressure taps were used to
monitor and account for any minor run-to-run variation in inlet
conditions that occurred.

Primary Uncertainties - All individual mean-square uncertainties were


combined to establish instrumentation uncertainty of a single
performance variable. Overall uncertainty was representative of
instrumentation and testing uncertainties. An account of experimental Figure 9: Locations of PIV measurement planes
uncertainties was evaluated and provided in Table 1. Uncertainty
analysis was completed in accordance with Boehm [21].

4. NUMERICAL APPROACH
Table 1: Table of Uncertainties of Performance Variables
Numerical analysis of the scaled rig geometry was completed for
all the combinations of inlet swirl flow distributions as well as strut
angles and geometries. CD-adapco STAR-CCM+ 10.06.009 was used
for all mesh generation and numerical computations. CFD predictions
made with an earlier version of this code (v5.04) were previously made
by Bernier et al. [22] for diffuser-collector geometries operating with
uniform and axial inlet flow conditions and with incompressible flow.
The authors made comparisons between test data and predictions made
with the realizable two-layer K-Epsilon turbulence model and reported
CFD under-predicted pressure losses by 5%. In the present study, the
computational domain included the full test rig as shown in Figure 3.
All numerical models included the bellmouth, the swirl vanes, the
contraction, and the scaled diffuser-collector. The swirl distribution at

5 Copyright © 2017 Solar Turbines Incorporated


Table 2: Mesh Sizing Summary two additional periodic oscillations. The steady-state simulations were
deemed converged when the exhaust recovery, Cp, was converged
K - Epsilon K - Epsilon EB SST (Menter) RSM (LPS & QPS) DES within ±0.0002.
Mesh Size (Millions) 62.1 71.0 71.9 68.6 87.8
Area Avg Y+ Hub 13.6 2.2 0.2 23.4 0.4
Area Avg Y+ Shroud 14.1 2.1 0.3 24.9 0.4
Area Avg Y+ Strut 9.8 2.6 0.5 9.8 0.7 5. RESULTS
Endwall Prism Sizing
Count 12 14 20 12 20 5.1 Turbulence Model Study
First Element Height [in] 0.0050 0.00070 0.00010 0.0080 0.00015
Overall Thickness [in] 0.10 0.10 0.10 0.10 0.10
Values of diffuser-collector recovery, as predicted with the
Strut Prism Sizing considered turbulence models were summarized in Figure 10. These
Count 10 12 15 10 16 predictions were provided in relative form with respect to the recovery
First Element Height [in] 0.0090 0.00050 0.00010 0.010 0.00010
Overall Thickness [in] 0.020 0.020 0.020 0.020 0.020 derived from measurements. The case studied involved an inlet swirl
angle of -6.6º and an axial strut geometry. The K-Epsilon two-layer
the inlet of the diffuser was determined by the solution and depended on model was found to be most accurate steady-state turbulence model, and
the swirl vane geometry that was modeled. The domain was extended diffuser recovery predictions were within 4 points (Cp of 0.04) of
downstream by three hydraulic diameters to reduce the impact of measured data. The diffuser-collector recovery as predicted with the
imposed exit boundary conditions and to add length for flow mixing. implicitly-unsteady models provided similar level of prediction
This additional length also enhanced the simulation numerical stability. capability with significantly higher computational requirements. The
Region based meshing technique was used and the mesh was sized in SST, K-Epsilon EB, and RSM-QPS models significantly under-
an iterative manner, with mesh generation followed by a review of the predicted the performance of this exhaust system.
solution’s mesh to capture the gradients in pressure and velocity fields.
This iterative approach ensured gradients within the growing diffuser Local pressure recovery, Cpi, was calculated by substituting
boundary layer were captured and gradual increase in mesh volume ratio endwall static pressure measurements for Ps ¯ 2 in the overall recovery
was incorporated between the boundary layer refinement and the equation. Local pressure recovery as a function of axial distance along
mainstream flow. This mesh sizing was completed for the K-Epsilon the inner and outer wall was provided in Figure 11. The predictions
model and the boundary layer thickness, distribution, and near wall were plotted against measured data and all plots were shown relative to
mesh sizing were adjusted to match the Y+ requirements of other the exhaust system inlet to outlet recovery derived from measurements.
models as per Table 2. The grids used in DES analyses were further The K-Epsilon model offered a good compromise between accuracy and
refined to a mesh size below the estimated large eddy size that was computational requirements. The increase in static pressure along the
calculated from the mean values of dissipation rate, η, velocity, U, and hub was initially delayed until flow passed the strut maximum thickness
viscosity, ν as predicted with the K-Epsilon model. For all cases, the location at mid chord. Good prediction accuracy was achieved along
total pressure inlet boundary condition of 16.89 psia was applied the top-dead-center (circumferential location of 0º) of the hub and
upstream of the bellmouth with flow direction specified as normal, shroud. Recovery predictions along the bottom-dead-center
turbulence intensity of 4% and viscosity ratio at 10%. An exit static (circumferential location of 180º) were lower than measurements.
pressure boundary condition was set to 15.87 psia with allowance to Recovery predictions just downstream of the strut (at normalized length
modulate and match an exit mass flow of 15.08 lb/s. of 40%) were lower than those derived from measurements indicating
that flow separation was over-predicted.
A turbulence model sensitivity study was completed for the
diffuser-collector geometry with an axial strut and an inlet swirl angle Along with the predictions made with the K-Epsilon model,
of -6.6°. Turbulence models evaluated with the steady-state numerical included in Figure 11 are predictions of endwall Cpi made with the SST,
scheme included two-layer K-Epsilon, K-Epsilon Elliptical Blending DES-EB and DES-SST models. Predictions with other models provided
(EB), Shear-Stress Transport (SST) - Menter, and Reynolds-Stress less accurate results and were therefore not included. The rate of static
Models (RSM) with the linear pressure strain (LPS) and quadratic pressure rise along the endwalls was different for all turbulence models
pressure strain (QPS) formulations. Implicit unsteady analysis with considered. The difference was especially notable within the strut
time advancement was also completed using the Delayed Detached region. The static pressure rise within the strut was significantly lower
Eddy Simulation model with Elliptical Blending (DES–EB) as well as for models that under-predict the overall performance (i.e. K-Epsilon
Improved Delayed Detached Eddy Simulation (DES – SST) model. EB, SST, and RSM –QPS), and in general, the rate of pressure recovery
Details of the model are provided in the software release notes [23] and along the length of the diffuser was lower than measurements.
the default model settings were used. These models use RANS Predictions within the struts were comparable with the RSM – LP and
equations with advanced modification of the turbulent dissipation rate K-Epsilon models, but the RSM – LP model predicted lower recovery
that attempt to capture LES-like flow. Details of DES-SST model is in the downstream portion of the diffuser. Endwall recovery predictions
summarized by Shur et al. [24]. The coupled solver was used made with the DES models were similar to those predicted with the
exclusively with the noted exception of the RSM models that were steady-state K-Epsilon model. Recovery predictions along the BDC
found to be numerically unstable, and the segregated solver was used to were more pessimistic than the K-Epsilon model. Along the TDC, the
solve the RSM models. The DES models used second order rate of static pressure rise was higher and a better match was offered at
discretization and were solved at a time-step of 2.5E-6s. The convective the diffuser exit. Prediction capability of the DES-SST model was
Courant number was below unity throughout the mesh. Each time-step comparatively worse within the strut region but static pressure was
was solved to ten inner iterations and the solution was deemed recovered along the aft portion of the diffuser to yield similar net
converged when four repeatable oscillations of the inlet and outlet mass performance as the DES-EB model.
flow and mass-flow averaged static pressure were calculated.
Ensemble-averaged flow velocities and pressures were gathered during

6 Copyright © 2017 Solar Turbines Incorporated


All models predicted an onset of separation along the bottom dead
center (BDC) of the diffuser, as seen in the contours of Mach number in
Figure 12. The contour plots generated from RSM predictions were
omitted as they were similar to those predicted with the K-Epsilon
model. Performance prediction capability of CFD was dependent on the
accurate prediction of the onset of separation, predicted size of the
separated zone, and the capability of the code to predict flow re-
attachment. The onset of separation was predicted along the shroud
with all models with the noted exception of the SST model. Unlike other
models, the SST model predicted the onset of separation along the hub
and not the shroud. Unsteady models predicted flow re-attachment that
was not predicted with steady-state methods. The onset of separation
was predicted further upstream with the DES-SST model than with the
DES-EB model, and re-attachment was predicted further upstream. The
difference in flow development within the diffuser likely affected the
vortex development within the collector, thus leading to differences in
the flow structure at the outlet. Figure 10: CFD prediction of Cp relative to experimental data,
evaluated for various turbulence models, Swirl 6.6°, Axial Strut
Whereas the pressure data was found to provide a good quantitative
assessment of CFD accuracy, the PIV measurements were used on a achieved with the K-Epsilon model with reduced computational
qualitative basis to select a turbulence model that provided a requirements. The K-Epsilon model was therefore exclusively used for
representative prediction of the flow structure within the collector. inlet swirl angle and strut stagger angle studies.
Normalized velocity contours at the collector exit plane previously
defined in Figure 9 were shown in Figure 13. The contours represent a
non-dimensional velocity defined by the normal component in the 5.2 Impact of swirl angle and strut incidence
direction of the exit plane normalized by the total velocity at the diffuser
inlet. The CFD predictions were shown against ensemble-averaged data The exhaust diffuser-collector recovery as a function of average
gathered with the PIV system. Optical imagery along the collector front inlet swirl angle was shown in Figure 14 for the three strut stagger
and back wall was not possible, and a region along these sections were angles considered. Radial distributions of the inlet swirl angle were also
shaded in blue. Higher velocities were measured along the back wall provided for direct comparison. The recovery values were related to a
and all turbulence models captured this trend. Both measurements and new baseline condition with the diffuser strut set axially at a stagger
predictions showed the development of two large counter-rotating angle of 0° and the swirl vanes removed to allow for axial inlet flow.
vortices within the collector. The white locations indicate flow reversal This later configuration was also modeled and the predicted value of
into the collector, and all turbulence models captured this reversal. In recovery was used as the baseline for relating CFD predictions. The
addition to affecting the predicted vortex size and distribution, the ability of CFD predictions to track performance differences with swirl
choice of turbulence model was found to affect the size of flow reversal angle changes was evaluated on a relative basis.
regions back into the collector. Among the steady-state models,
predictions made with the K-Epsilon and RSM-LPS models were most Contrary to findings reported for annular diffusers, the addition of
representative of measurements. The larger region of flow reversal inlet swirl did not positively impact the performance of this diffuser-
predicted with the RSM-LPS model was consistent with under- collector geometry. The diffuser-collector performance was similar
prediction of the overall exhaust system performance. The SST model with and without inlet swirl for cases where flow was aligned with the
was found to provide the least-accurate prediction of the overall flow strut angle to within 10º degrees. In this range, the CFD predictions of
structure, most notably the size of counter-rotating vortices and the exhaust recovery as shown on a relative basis were similar to measured
region where flow reversal occurred. Both the SST and RSM-QPS values. On an absolute basis, the difference between predicted and
models predicted higher velocities along the backwall that contributed measured recovery was of the same magnitude as quoted in the
to a pessimistic prediction of the overall pressure recovery. Improved turbulence model study.
resolution in the collector outlet flow structure was predicted with the
implicitly unsteady models, and the DES-SST model was most In the case of +12º strut stagger, a significant reduction in recovery
representative of the time-averaged PIV data. was measured for larger incidence angles (Figure 14a). CFD predictions
also showed a large reduction in recovery although at a greater rate. It
Time-resolved DES simulations provided greater resolution of the would be expected that the diffuser configuration with -12º strut stagger
flow field within a diffuser-collector exhaust system. The models would experience similar reduction in recovery at positive inlet swirl
offered improvements in accurately predicting endwall static pressure angles but the data in Figure 14b showed this not to be the case. A trend
distribution within the diffuser and resolving detailed flow features at was not captured with data and a weak trend was predicted with CFD
the collector exit. Most notably, DES models captured flow- analysis. A similar observation was made for the strut set at 0º, were
reattachment within the diffuser that were not captured with steady-state better recovery was measured at positive inlet swirl angles than at
analyses. These models require a significant computational effort with negative angles despite utilization of a symmetric strut profile. To
solver time in the order of one month with 200 processors. Major flow understand this discrepancy, attention was paid not only to the mean
features were adequately captured using the steady-state K-Epsilon incidence but also to the variation in incidence along the span. At a strut
model. Adequate accuracy in predicting both the static pressure rise stagger angle of +12º, the increase in incidence at the hub was higher
within the diffuser and the overall diffuser-collector performance was than at the shroud. With the strut stagger angle set at -12°, the incidence
at the shroud was more significantly increased. The distinction between

7 Copyright © 2017 Solar Turbines Incorporated


K-Epsilon Model SST Model

DES – EB Model DES – SST Model

Figure 11: CFD Prediction capability of endwall static pressure recovery for select turbulence model

a) b) c) d)

Figure 12: CFD predictions of the Mach number through the exhaust diffuser-collector cross-section
a) K-Epsilon Two-Layer, b) SST, c) DES-EB, d) DES-SST

8 Copyright © 2017 Solar Turbines Incorporated


DIFFUSER

BACK WALL

(a) PIV Measurements b) CFD Prediction – SST (Menter) Model

c) CFD Prediction – k – Epsilon Two - Layer Model d) CFD Prediction – K – Epsilon EB Model

e) CFD Prediction - RSM Model – Linear Pressure Strain f) CFD Prediction – RSM Model – Quad. Pressure Strain

g) CFD Prediction – DES-EB Model – Ensemble Average h) CFD Prediction – DES-SST Model – Ensemble Average

Figure 13: PIV measurements and CFD predictions of collector exit normal velocity component normalized to inlet velocity

9 Copyright © 2017 Solar Turbines Incorporated


a) Strut stagger angle +12° b) Strut stagger angle -12° c) Strut stagger angle 0°

Figure 14: Exhaust diffuser-collector recovery change as a function of inlet swirl and strut stagger angle (top),
compared along the radial distribution of inlet swirl angle (bottom)

the influence of hub and shroud incidence on diffuser-collector recovery i. Strut stagger of +12º and inlet swirl angle of -14.6º. The resulting
was apparent when comparing the swirl angle sensitivity on the axial incidence angle along the hub and shroud were -34° and -19º
strut. A performance degradation was measured with the increase in respectively,
flow angle at the hub. Shroud incidence of the same magnitude did not ii. Strut stagger of -12º and swirl angle of -6.6º. The resulting
impact the diffuser-collector performance. For flows with radial incidence angle along the hub and shroud were +2° and +10°
distributions of inlet swirl, the influence of swirl on the performance of respectively, and,
this diffuser-collector exhaust system was dependent on the radial iii. Strut stagger of -12º and swirl angle of 9.4º. The resulting
location associated with maximum strut incidence angle. incidence angle along the hub and shroud were +15° and +34°
respectively.
The diffuser-collector recovery as a function of strut incidence
angle at the hub was plotted in Figure 15a and the shroud in Figure 15b. Distributions of relative recovery along the hub and shroud were
The data was provided in relative form consistent with the baseline shown in Figure 16 and Figure 17 respectively. The plots were
configuration with axial flow aligned with the axial strut. A definitive separated for data gathered at the top-dead-center (TDC) and bottom-
correlation of recovery with hub incidence was observed. Exhaust dead-center (BDC). The measurements were not gathered upstream of
performance did not correlate with shroud incidence. The accuracy of 38% normalized axial distance and the complementary CFD analysis
CFD performance predictions became increasingly unreliable as the was used in this range to provide insight in flow development within
flow angle deviated from the strut by 12º and higher. The accuracy of this region.
CFD predictions was also dependent on the radial location of incidence.
Accurate predictions were not achieved when large incidence angles at For the case with least incidence, an initial reduction in recovery
the hub were present. was observed within the strut commensurate to the reduction of cross-
sectional flow area due to strut blockage. Along the hub, gradual
The influence of strut incidence angle on flow development was diffusion within the struts was observed once the flow passed through
further evaluated with the use of static pressure measurements gathered the strut region of maximum thickness (maximum reduction of cross-
at the endwalls. These measurements were converted to a local recovery section flow area). Along the shroud, initial recovery was low due to
value and shown in relative form. The data was related to the total local flow acceleration at the diffuser inlet, where the axial flow-
exhaust recovery measured for the baseline case with no swirl and no conditioning flow path connected with the diffuser. For this low
incidence. Three scenarios were considered: incidence case, continuous diffusion was observed throughout the
length of the diffuser downstream of the strut mid chord. CFD

10 Copyright © 2017 Solar Turbines Incorporated


predictions of endwall static pressure were in good agreement with The streamlines shown in Figure 19 and Figure 21 sub-labels c)
measurements. and d) correspond to the previously examined case i corresponding to
strut incidence at the hub of -34º, and case iii corresponding to strut
Pressure recovery, as derived from measurements, was incidence at the shroud of +34º respectively. These predictions were
significantly impacted for cases were the strut was subjected to large examined to assess possible reasons for reduced CFD prediction
levels of incidence. However, the rate of static pressure rise at the accuracy and establish a general trend regarding flow behavior at high
downstream 60% of the diffuser wall was similar, and the losses incidence angles along the hub and shroud. The diffuser-collector
occurred upstream of the locations were measurements were taken. geometry between cases b) and c) was identical and the average swirl
CFD predictions indicated that large pressure losses were present within angle was increased from -6.6° to -14.6°. Flow streamlines along the
the strut, and the magnitude of these losses was a function of the radial hub were similar for the two cases, but the region of flow recirculation
location of the greatest incidence i.e. the hub or the shroud. Losses due along the BDC has increased and also moved further upstream. Flow
to incidence effects on the strut were larger in the hub region than the separation along the shroud was predicted at increased inlet flow angles.
shroud, and the both measurements and predictions show this trend. At This additional separation along the shroud lead to further degradation
these high incidence angles, CFD predictions over-predicted the loss. in predicted exhaust recovery.

Flow visualization along both the hub and shroud walls was Similarly, the diffuser-collector geometry between cases a) and d)
completed to examine the impact of incidence on flow behavior near the were identical and the average swirl angle was increased from -6.6° to
endwalls. Photos of the streamlines along the hub and shroud were +9.4°. The increase in positive inlet swirl angle results in a stabilizing
shown in Figure 18 and Figure 20 respectively. The diffuser average effect on the flow along the hub despite an increase in strut incidence.
inlet swirl angle was set at -6.6°, and surface streak lines were visualized The flow interaction between the diffuser and collector were therefore
and compared for two cases of diffuser strut angle a) -12° (case ii above) sensitive to the diffuser inlet swirl angle. As seen in Figure 21 d), flow
and b) +12° (case i). As seen in Figure 18 a), separation along the BDC along the shroud separated at high incidence angles. The separated
of the hub was observed with the diffuser operating with little incidence. region shifted the general flow direction and resulted in an increase in
This separated region was at about 75% of axial length. The flow along the wetted area. The flow direction along the shroud followed the entire
the hub was completely changed by modifying the strut stagger angle, length of the diffuser. This was contrary to swirl angle sensitivity at the
and was characterized by recirculating flow along the entire diffuser hub, were significant strut incidence resulted in flow separation that
length at the BDC location. The transition from axial to radial flow that impeded effective utilization of the full diffuser length.
should occur within the collector commenced further upstream within
the diffuser. The reduction in exhaust diffuser-collector performance
resulted from the compounding effects of: 1) reduced diffuser length CONCLUSIONS
utilization that reduced the operating area ratio, 2) increased mixing
losses that resulted from flow separation and 3) increased blockage The impact of inlet swirl angle and associated radial distribution
caused by flow separation and increased through-flow velocities. The on diffuser-collector performance was examined along with the
streamlines along the shroud BDC as captured with flow visualization influence of strut stagger angle. Experimental measurements coupled
were shown in Figure 20. The flow along the shroud was mostly well with complimentary CFD analyses were completed on a quarter-scaled
behaved, and the flow direction was changed by the strut stagger angle. geometry representative of an industrial gas turbine exhaust. The
In Figure 20a) a region along the BDC existed were flow began to turn introduction of inlet swirl did not positively impact the performance of
well upstream of the diffuser-collector transition. The corresponding this diffuser-collector geometry. This finding was contrary to findings
strut incidence at the shroud was +10°. This flow phenomenon was not widely reported for annular diffusers without a downstream collector.
observed when the strut was re-set to +12º and the strut incidence at the The diffuser-collector was able to accommodate a significant amount of
shroud -14°. In summary, the flow along the shroud was changed with inlet swirl without a degradation in performance, so long as the inlet
a net increase in wetted area. The associated differences in flow flow direction did not significantly deviate from the strut stagger angle.
structure resulting from re-setting the diffuser strut were much larger A reduction in exhaust system performance was measured as the
along the hub. deviation between flow angle and strut stagger angle increased beyond
±12°. A symmetrical airfoil strut was used for this study and the impact
Comparative streamlines, as predicted with CFD analysis, were of strut geometry was not investigated.
shown in Figure 19 along the hub and Figure 21 along the shroud. These
velocity streamlines were plotted on a surface offset by 0.010 inches For flows with radial gradients of diffuser inlet swirl angle, the
from the endwalls. The CFD results in sub-labels a) and b) provided reduction in performance was dependent on the radial location
direct comparisons to streamlines shown using experimental techniques. corresponding to the highest strut incidence. Increases in incidence
Numerical predictions of streamlines were in good agreement with along the shroud had a weak effect on diffuser-collector performance.
experimental visualization in terms of capturing the overall flow Incidence effects on struts result in local flow separation that impact the
direction and associated strut impact. The flow along the hub was overall flow direction in the downstream section of the diffuser.
susceptible to separation at low strut incidence angles, and the full Changes in diffuser recovery were mainly attributed to differences in
length of the diffuser was not utilized along the BDC location. The wetted flow area and additional losses impacted by local flow
increase in strut incidence angle resulted in additional level of separation. A much stronger correlation between strut incidence at the
separation and a reduction in the utilized diffuser length. Predicted flow hub and diffuser-collector performance was identified. A reduction in
direction along the shroud was in good agreement with experimental diffuser-collector performance was measured as the swirl angle was
visualizations, and the impact of strut stagger angle on flow direction increased beyond 12º of strut incidence at the hub. Endwall static
was captured. CFD results showed an increase in flow velocities along pressure measurements, flow visualization, and numerical predictions
the shroud that resulted from flow separation induced blockage along were completed to identify differences in flow behavior at the hub and
the hub. shroud that impacted swirl angle sensitivities. Strut incidence at the hub

11 Copyright © 2017 Solar Turbines Incorporated


a) b)

Figure 15: Impact of incidence at the (a) hub and (b) shroud on exhaust diffuser-collector recovery

Figure 16: Pressure recovery along the hub (a) TDC, (b) BDC

Figure 17: Pressure recovery along the shroud (a) TDC, (b) BDC

12 Copyright © 2017 Solar Turbines Incorporated


a) Hub Incidence +2°, Shroud Incidence +10° b) Hub Incidence -22°, Shroud Incidence -14°

Figure 18: Surface flow visualization along the diffuser hub BDC location

a) Hub Incidence +2°, Shroud Incidence +10° b) Hub Incidence -22°, Shroud Incidence -14°

c) Hub Incidence -35°, Shroud Incidence -19° d) Hub Incidence +15°, Shroud Incidence +34°

Figure 19: Streamlines predicted with CFD along the diffuser-collector hub

13 Copyright © 2017 Solar Turbines Incorporated


a) Hub Incidence 2°, Shroud Incidence 10° b) Hub Incidence -22°, Shroud Incidence -14°

Figure 20: Surface flow visualization along the diffuser shroud BDC location

a) Hub Incidence +2°, Shroud Incidence +10° b) Hub Incidence -22°, Shroud Incidence -14°

c) Hub Incidence -35°, Shroud Incidence -19° d) Hub Incidence +15°, Shroud Incidence +34°

Figure 21: Streamlines predicted with CFD along the diffuser-collector shroud

14 Copyright © 2017 Solar Turbines Incorporated


resulted in formation of large regions of flow separation along the [7] Kumar D., Kumar K., 1980, Effect of Swirl on Pressure Recovery in
bottom-dead-center of the diffuser. The reduction in performance was Annular Diffusers, J. Mech. Eng. Sci., Vol 22 (6), pp. 305-313.
attributed to: [8] Vassiliev V., Irmisch S., Claridge M., Richardson D., 2003,
1) The location where flow direction changes from axial to radial “Experimental and Numerical Investigation of the Impact of Swirl on
was shifted upstream and was observed to be within the diffuser. The the Performance of Industrial Gas Turbines Exhaust Diffusers,”
diffuser area ratio and length were effectively reduced. ASME Turbo Expo, Atlanta, Georgia, USA., June 16-19, ASME
2) An increase in pressure losses resulted from the flow separation Paper No. GT2003-38424.
[9] Senoo Y., Kawaguchi N., Kojima T., Nishi M., 1981, “Optimum
3) An increase in blockage was caused by flow separation that
Strut-Configuration for Downstream Annular Diffusers with Variable
increased the flow velocities within the passage.
Swirling Inlet Flow”, J. Fluids Eng., 103, pp. 294-298.
[10] Vassiliev V., , Irmisch S., Florjancic S., 2002 “CFD Analysis of
Despite the presence of an adverse pressure gradient that is coupled Industrial Gas Turbine Exhaust Diffusers”, ASME Turbo Expo,
with highly unsteady three-dimensional chaotic vortical flow, steady- Amsterdam, The Netherlands, ASME Paper No. 2002-GT-30597.
state Reynolds Averaged Navier Stokes (RANS) with K-Epsilon two- [11] Japiske D., Pampreen R., 1979, “Annular Diffuser Performance for
layer turbulence closure model can be successfully applied for diffuser- an Automotive Gas Turbine”, ASME J. Eng. Power., 101, pp. 358-
collector performance evaluations at design-conditions. Predominant 372.
flow structures, static pressure distributions, and mass-averaged values [12] Xue S., Guillot S., Ng W., Fleming J., Lowe K, Samal N., Stang U.,
of diffuser recovery used for performance calculations were predicted 2016, “An Experimental Investigation of the Performance Impact of
within 5 points (Cp of 0.05). More advanced DES models are capable Swirl on Turbine Exhaust Diffuser/Collector for a Series of Diffuser
of capturing flow re-attachment that is not possible with steady-state Strut Geometries”, J. Eng. Gas Turbines Power 138(9), pp. 092603,
methods. At design conditions, the increase in flow resolution offered 1-8.
by the DES models does not warrant the increased computational effort. [13] Lui J. Cui Y., Jiang H., 2003, “Investigation of Flow in a Steam
The accuracy of steady-state CFD performance predictions become Turbine Exhaust Hood With/Without Turbine Exit Conditions
increasingly unreliable as the flow angle deviates from the strut stagger Simulated” ASME J. Eng. Gas Turbines Power, 125(1), pp. 292-299.
angle. The accuracy is especially impacted for simulations with high [14] Liu J., Hynes T., 2002, The Investigation of Turbine and Exhaust
Interaction in Asymmetric Flows Part 2 – Turbine-Diffuser-Collector
incidence at the hub, where flow behavior is significantly different than
Interactions, ASME Turbo Expo Amsterdam, The Netherlands, June
observed at design conditions. Simulations of these flow conditions
3-6, ASME Paper No. GT-2002-30343.
may warrant the application of DES models as the improved ability to [15] Fu J., Lui J., 2008, “Influences of Inflow Conditions on Non-
capture re-attachment may yield less pessimistic performance Axisymmetric Flows in Turbine Exhaust Hoods,” J. Thermal Sci.
predictions at off-design conditions. 17(4), pp. 305-313.
[16] Stastny M,. Tajc L., Kolar P., Ucek A., 2000 “Effects of Inlet Swirl on
the Flow in a Steam Turbine Exhaust Hood,” J. Thermal Sci., 9(4) pp.
ACKNOWLEDGMENTS 327-333.
The work was sponsored by Solar Turbines Incorporated. The [17] Sultanian B., Nagao S, Sakamoto T., 1999, “Experimental and Three-
authors would like to express their gratitude to Nihar Samal and Dimensional CFD Investigation in a Gas Turbine Exhaust System,”
Geoffrey Potts of Solar Turbines Incorporated for their contributions in ASME J. Eng. Gas Turbines Power, 121(1), pp 364-374.
designing two swirl vanes. The authors would like to thank Chao Zhang [18] Zhang W., Pail B., Jang Y., Lee S., Lee S., Kim J., 2007, “Particle
from CD-Adapco for providing technical support with numerical Image Velocimetry Measurement of the Three-Dimensional Flow in
simulations and contributing to the turbulence model study with an Exhaust Hood Model of a Low-Pressure Steam Turbine,” ASME J.
Eng. Gas Turbine Power, 129(2), pp. 411-419.
computational resources for the DES-SST case.
[19] Guillot S., Ng W., Hamm H., Stang U., Lowe K., 2015, “The
Experimental Studies of Improving the Aerodynamic Performance of
a Turbine Exhaust System,” J. Eng. Gas Turbines Power 137(1), pp.
REFERENCES 12601, 1-13.
[1] Rocha, G. and Etheridge, C. J., 1998 “Evolution of the Solar Turbines [20] Roach, P. E., 1987, “The generation of Nearly Isotropic Turbulence
Titan 130 Industrial Gas Turbine,” International Gas Turbine and by Means of Grids,” International Journal of Heat and Fluid Flow,
Aeroengine Congress and Exposition, Stockton, Sweden, June 2-5, Vol. 8, Issue: 2, Publisher: Elsevier, pp. 82-92.
ASME Paper No. 98-GT-590. [21] Boehm, B. P., 2012, “Performance Optimization of a Subsonic
[2] Waldhelm, C, 1992 “Evolution and Introduction of the Mars T-14,000 Diffuser-Collector Subsystem Using Interchangeable Geometries,”
Gas Turbine,” International Gas Turbine and Aeroengine Congress M. S. Thesis, Department of Mechanical Engineering, Virginia
and Exposition, Cologne, Germany, June 1-3, ASME Paper No. 92- Polytechnic Institute and State University, Blacksburg, VA.
GT-332. [22] Bernier, B., Ricklick, M., Kapat J. S., “Impact of a Collector Box on
[3] Jones, M., Pytanowski G., Files D., Wilson J., 1994 “Evolution and the Pressure Recovery of an Exhaust Diffuser System” ASME Turbo
Experience of the Mars 100LS Gas Turbine,” International Gas Expo, Vancouver, British Columbia, Canada, June 6-10, 2011, ASME
Turbine and Aeroengine Congress and Exposition, June 13-16, Paper No. GT2011-46455.
ASME Paper No. 94-GT-301. [23] Star-CCM, 2016, “Star-CCM+ Version 10.06, USA, CD-Adapco.
[4] Sovran,G., Klomp E., “Experimentally Determined Optimum [24] Shur, M.L., Spalart, P.R., Strelets, M.Kh., and Travin, A.K. 2008. “A
Geometries for Rectilinear Diffusers with Rectangular, Conical, or hybrid RANS-LES approach with delayed-DES and wall-modelled
Annular Cross-Section,” Fluid Dynamics of Internal Flow, Elsevier LES capabilities”, International J. Heat and Fluid Flow, 29(6), pp.
Publishing Co. 1967. 1638-1649.
[5] Japikse D., Baines, N., Turbomachinery Design Technology, Concepts
ETI, Inc. Norwich, Vermont, USA 05055, 1984.
[6] McDonald A., Fox R., Van Dewoestine R., 1971, “Effect of Swirling
Inlet Flow on Pressure Recovery in Conical Diffusers, AIAA J.,
9(10), pp. 2014-2018.

15 Copyright © 2017 Solar Turbines Incorporated

You might also like