anagnostis_g ( PDF ) - UFDC Image Array 2 - University of Florida

105
DESIGN POINT ANALYSIS OF THE HIGH PRESSURE REGENERATIVE TURBINE ENGINE CYCLE FOR HIGH-SPEED MARINE APPLICATIONS By GEORGE ANAGNOSTIS A THESIS PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE UNIVERSITY OF FLORIDA 2007

Transcript of anagnostis_g ( PDF ) - UFDC Image Array 2 - University of Florida

DESIGN POINT ANALYSIS OF THE HIGH PRESSURE REGENERATIVE TURBINE ENGINE CYCLE FOR HIGH-SPEED MARINE APPLICATIONS

By

GEORGE ANAGNOSTIS

A THESIS PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE

UNIVERSITY OF FLORIDA

2007

Copyright 2007

By

George Anagnostis

This thesis is dedicated to my parents, Victor and Linda Anagnostis. Without their emotional and financial encouragement this thesis would not exist.

iv

ACKNOWLEDGMENTS

I thank the members of my graduate committee members: Dr. William E. Lear, Jr.,

Dr. S. A. Sherif, and Dr. Herbert Ingley for their support on this thesis. Dr. Lear was

especially helpful, providing me with critical advice throughout this project. Next, I

would like to thank the Aeropropulsion Systems Analysis Office at the National

Aeronautics and Space Administration Glenn Research Center for their assistance on

Numerical Propulsion System Simulation program. Two members of that group provided

continued technical assistance—Scott Jones and Thomas Lavelle. Lastly, I thank two

special individuals that have provided me with insight and wisdom concerning matters of

engineering and life in general, John Crittenden and William Ellis.

v

TABLE OF CONTENTS page

ACKNOWLEDGMENTS ................................................................................................. iv

LIST OF TABLES............................................................................................................ vii

LIST OF FIGURES ......................................................................................................... viii

NOMENCLATURE ............................................................................................................x

CHAPTER

1 INTRODUCTION ........................................................................................................1

2 LITERATURE REVIEW .............................................................................................4

Brief History of Turbine Engine Development ............................................................4 Gas Turbine Engine Examples in Marine Applications ...............................................5 Advantages of Gas Turbine Engines in Marine Applications ......................................6 Recuperation and Inter-cooling ....................................................................................7 Semi-Closed Cycles......................................................................................................9 Computer Code Simulators.........................................................................................10 Previous Gas Turbine Research at the University of Florida .....................................12

3 NUMERICAL PROPULSION SYSTEM SIMULATION ARCHITECTURE.........16

Model..........................................................................................................................16 Elements .....................................................................................................................17 FlowStation.................................................................................................................18 FlowStartEnd ..............................................................................................................18 Thermodynamic Properties Package ..........................................................................20 Solver..........................................................................................................................21

4 CYCLE CONFIGURATIONS AND BASE POINT ASSUMPTIONS.....................25

Major Model Features.................................................................................................25 Flow Path Descriptions & Schematics .......................................................................26

Simple Cycle Gas Turbine Engine Model...........................................................26 High Pressure Regenerative Turbine Engine Efficiency Model .........................26

vi

High Pressure Regenerative Turbine Engine with Vapor Absorption Refrigeration System Efficiency Model ..........................................................27

Simple Cycle Gas Turbine Engine Design Assumptions and HPRTE Cycles Base Point Assumptions .................................................................................................28

5 THERMODYNAMIC MODELING AND ANALYSIS............................................33

Thermodynamic Elements ..........................................................................................33 Heat Exchangers..................................................................................................33 Mixers..................................................................................................................34 Splitter .................................................................................................................35 Water Extractor ...................................................................................................36 Compressors ........................................................................................................37 Turbines...............................................................................................................39 Burner ..................................................................................................................40

Sensitivity Analysis ....................................................................................................41

6 RESULTS AND DISCUSSION.................................................................................43

Cycle Code Comparison .............................................................................................43 Sensitivity Analysis ....................................................................................................44

Simple Cycle Gas Turbine Engine Model...........................................................44 Simple Cycle Gas Turbine Engine Model Sensitivity Analysis..........................48 High Pressure Regenerative Turbine Engine Efficiency Model .........................50 High Pressure Regenerative Turbine Engine Efficiency Model Sensitivity

Analysis............................................................................................................53 Cycle Comparison Analysis .......................................................................................61

Extreme Operating Conditions ............................................................................65 High Pressure Compressor Inlet Temperature Comparison for H-V Efficiency

Model ...............................................................................................................67 Final Design Point Parameter Comparison .........................................................68

7 CONCLUSIONS AND RECOMMENDATIONS.....................................................83

Conclusions.................................................................................................................83 Recommendations.......................................................................................................86

LIST OF REFERENCES...................................................................................................88

BIOGRAPHICAL SKETCH .............................................................................................91

vii

LIST OF TABLES

Table page 4-1 Comparison of major configuration features ...........................................................29

4-2 Simple Cycle Gas Turbine engine design point parameters.....................................32

4-3 Base case model assumptions for HPRTE cycles [3], [26], [27] .............................32

6-1 Cycle codes comparison: NPSS verses spreadsheet code for HPRTE Efficiency model data run. All temperatures are in °R. ............................................................70

6-2 Summary of the HPRTE Efficiency sensitivity analysis .........................................77

6-3 Comparison of the thermal efficiency maximums and their corresponding overall pressure ratios (OPRs)..................................................................................78

6-4 Comparison of the specific power maximum values and their corresponding OPRs.........................................................................................................................78

6-5 Comparison of exhaust temperature maximum values for the three engine configurations...........................................................................................................79

6-6 Engine cycles comparison for four extreme operating conditions ...........................81

6-7 High pressure compressor (HPC) inlet temperature comparison for the H-V Efficiency engine model...........................................................................................81

6-8 Final performance design point comparison for the engine configurations.............82

viii

LIST OF FIGURES

Figure page 3-1 Example NPSS engine model [19]...........................................................................23

3-2 State 7 of HPRTE engine cycle................................................................................24

4-1 Simple Cycle Gas Turbine (SCGT) engine model configuration ............................29

4-2 High Pressure Regenerative Turbine Engine model, both efficiency and power configurations represented .......................................................................................30

4-3 High Pressure Regenerative Turbine Engine-Vapor Absorption Refrigeration System, both efficiency and power model configurations represented....................31

4-4 Vapor Absorption Refrigeration Cycle with HPRTE flow connections ..................32

6-1 Thermal efficiency comparison is plotted with respect to OPR. NPSS results (with turbine inlet temperature (TIT) set to 2500°R) are compared to the derived and the ideal Brayton cycle expressions. .................................................................70

6-2 Thermal efficiency vs. OPR with sensitivity to TIT ................................................71

6-3 Specific power vs. OPR with TIT sensitivity...........................................................71

6-4 Thermal efficiency vs. ambient temperature with OPR sensitivity..........................72

6-5 Demonstrates agreement between NPSS and developed theory that describes the low pressure spool ....................................................................................................72

6-6 High pressure spool pressure ratio (HPPR) vs. ambient temperature with low pressure spool pressure ratio (LPPR) sensitivity......................................................73

6-7 Thermal efficiency vs. HPPR showing sensitivity to TIT .......................................73

6-8 Thermal efficiency vs. HPC inlet temperature for recirculation ratio sensitivity ....74

6-9 Thermal efficiency vs. turbine exit temperature (TET) with cooler pressure drop sensitivity .................................................................................................................74

6-10 Specific power vs. TET for HPC efficiency sensitivity ...........................................75

ix

6-11 Specific power vs. HPPR for HPT efficiency sensitivity.........................................75

6-12 Exhaust temperature vs. OPR for TIT sensitivity ....................................................76

6-13 Thermal efficiency vs. HPPR for turbocharger efficiency sensitivity .....................76

6-14 Thermal efficiency vs. LPPR for TIT sensitivity .....................................................77

6-15 Engine cycles comparison of thermal efficiency vs. OPR .......................................78

6-16 Engine cycles comparison of specific power vs. OPR.............................................79

6-17 Engine cycles comparison of exhaust temperature vs. OPR....................................80

6-18 Engine cycles comparison of thermal efficiency vs. ambient temperature..............80

x

NOMENCLATURE

DepV Dependent variable in a Jacobian matrix

IndV Independent variable in a Jacobian matrix

ε Heat exchanger effectiveness

inPP _0Δ Pressure drop as a percentage of the inlet stream pressure

Q& Heat flow rate (Btu/sec)

nm& Mass flow rate at station “n” (lbm/sec)

npC _ Specific heat at constant pressure at flow station “n” (Btu/lbm-°R)

inT _0 Stagnation temperature at the inlet to a physical cycle component (°R)

outT _0 Stagnation temperature at the exit to a physical cycle component (°R)

inP _0 Stagnation pressure at the inlet to a physical cycle component (psi)

outP _0 Stagnation pressure at the exit to a physical cycle component (psi)

inh _0 Mass specific stagnation enthalpy at the inlet to a physical cycle

component (Btu/sec-lbm)

outh _0 Mass specific stagnation enthalpy at the inlet to a physical cycle

component (Btu/sec-lbm)

nFAR Fuel-to-air ratio at state point “n”

xi

ntotm _& For splitters and separators, total mass flow rate at state point “n”

(lbm/sec)

BPR Flow bypass ratio for splitter elements

liquidOHm _2& Mass flow rate of liquid water being extracted in separator (lbm/sec)

liquidOHh _2 Mass specific enthalpy of liquid water being extracted (Btu/sec-lbm)

CompPR Pressure ratio any compressor

adComp _η Adiabatic efficiency of any compressor

ns _0 Mass specific stagnation entropy at flow station “n” (Btu/lbm-°R)

R Ideal gas constant (Btu/lbm-°R)

idh Mass specific enthalpy change for an isentropic process (Btu/sec-lbm)

inCompR _ Ideal gas constant at a compressor inlet state point (Btu/lbm-°R)

bη Burner efficiency

RQ Lower heating value of the fuel (Btu/lbm)

WAR Water to air ratio, mass basis

TIT Turbine inlet temperature

OPR Overall pressure ratio of a system

γ Ratio of specific heats, v

p

CC

SpPw Specific Power (HP-sec/lbm)

ambT Ambient Temperature (°R), also ambientT

LPPR Low pressure compressor pressure ratio

HPRR High pressure compressor pressure ratio

xii

adLPC _η Low pressure compressor adiabatic efficiency

adLPT _η Low pressure turbine adiabatic efficiency

xiii

Abstract of Thesis Presented to the Graduate School

of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Master of Science

DESIGN POINT ANALYSIS OF THE HIGH PRESSURE REGENERATIVE TURBINE ENGINE CYCLE FOR HIGH-SPEED MARINE APPLICATIONS

By

George Anagnostis

May 2007

Chair: William E. Lear, Jr. Major Department: Mechanical and Aerospace Engineering

A thermodynamic sensitivity and performance analysis was performed on the High

Pressure Regenerative Turbine Engine (HPRTE) and its combined cycle variation, the

HPRTE with a vapor absorption refrigeration system (VARS). The performance analysis

consisted of a comparison of three engine configurations, the two HPRTE variants and a

simple cycle gas turbine engine (SCGT), modeled after the production marine gas turbine

engine, ETF-40B. The engine cycles were optimized using a parametric analysis; a

sensitivities study was completed to establish which design parameters influence

individual engine model performance. The NASA gas turbine cycle code Numerical

Propulsion System Simulation (NPSS) was the software platform used to complete this

analysis.

The comparison was performed at sea level with an ambient temperature of 544°R.

The results for the SCGT predict a design-point optimized thermal efficiency of 33.4%

and an overall pressure ratio (OPR) of 10.4 with a specific power of 180 HP-sec/lbm.

xiv

The HPRTE engine, called HPRTE Efficiency for this thesis, had an expected design

thermal efficiency of 37.2% (OPR of 32.2) with a specific power rating of 593 HP-

sec/lbm—229% larger than the SCGT specific power. The combined-cycle HPRTE-

VARS, called H-V Efficiency in the analysis, had a predicted design thermal efficiency

of 45.0% (OPR of 32) with a specific power of 629 HP-sec/lbm. The H-V Efficiency

thermal efficiency was 34.7% higher than that of the SCGT designed for maximum

specific power. Exhaust gas temperatures varied significantly between the SCGT and the

HPRTE variants. The model engine exhaust for the SCGT was 1580°R while the exhaust

temperatures of the HPRTE Efficiency and H-V Efficiency were 801°R and 837°R,

respectively. On average, the HPRTE calculated exhaust temperature was 761°R less

than that of the SCGT. High pressure compressor (HPC) inlet temperature sensitivity

was considered for the H-V Efficiency. Two operating cases were considered—the HPC

inlet held constant at 499°R and 509°R. The 499°R case operated with a thermal

efficiency higher by 1.56% and a specific power higher by 1.62%.

The results of the analysis imply that HPRTE duct sizes will be smaller due to the

engine having significantly higher specific power. Since specific fuel consumption is

inversely proportional to thermal efficiency, the H-V Efficiency engine cycle will require

a smaller fuel tank to allow for additional cargo (or if the tank size is unchanged, the ship

range is increased). Future project considerations include an off-design performance

analysis using NPSS or another software package, additional NPSS model benchmarking

with a reputable cycle simulation code, and an analysis of the effects of moist ambient air

on evaporator water flow extraction rates.

1

CHAPTER 1 INTRODUCTION

Before the marine gas turbine, naval ships clipped through the water propelled by

sooty coal-fired steam turbines or diesel engines. The 1940s advent of the gas turbine jet

engine introduced a similar technology shift in the marine propulsion industry a decade

later. And now for the last 60 years marine gas turbine engine propulsion advancements

have derived mainly from aeronautical research and development programs. However,

there have been some instances where the marine propulsion industry has led the way in

development—most notably by the introduction of the Westinghouse-Rolls-Royce 21st

century (WR21) ICR program in the early 1990s. Inter-cooled compressors and exhaust

heat recuperation set the WR21 gas turbine engine apart. Ironically, the same ingenuity

that steered the Navy to develop the WR21 program was nowhere to be found during the

decision-making process time for the propulsion system for the 21st century speed ship-

to-shore transport.

The ETF-40B, a workhorse and variant of the original TF-40 that powered the

Navy landing craft air-cushion (LCAC) vessel for the last two decades will provide the

propulsion and lift thrust for the new J-MAC ship-to-shore transport. Despite interest in

new engine technologies, such as the High Power Regenerative Turbine Engine

(HPRTE), funding constraints prevented the Navy from further investigating novel

systems. This thesis will make the case for the HPRTE as an alternative engine concept

to the ETF-40B for the J-MAC program.

2

The motivation to compare the HPRTE to the ETF-40B is a result of previous

experimental and computational modeling efforts completed at the University of Florida

(UF) Energy and Gas Dynamics Laboratory to develop alternate engine technologies.

There other design considerations besides cost that drive engine development; the

HPRTE will outperform the ETF-40B, having a higher specific power ratio, improved

off-design performance, and a considerably lower infrared heat signature.

The HPRTE is a semi-closed, compressor inter-cooled, recuperative system. A

demonstration engine has been build and performance tested at UF, and the proof of

concept has been met. The laboratory demonstrator uses engine exhaust heat to power a

vapor absorption refrigeration system (VARS). This is representative of the combined

cycle system, one of the two HPRTE configurations, that is considered in this modeling

and analysis project. The base HPRTE is the other. The combined cycle variant is

expected to outperform the base HPRTE because the VARS unit provides additional

cooling to the high pressure compressor inlet of the engine.

The analysis in this thesis includes a parametric optimization and sensitivity studies

that determine design-critical parameters. There are three engine models total that are

considered—the two HPRTE variants (HPRTE Efficiency and H-V Efficiency) and a

simple cycle gas turbine engine (SCGT). The SCGT is modeled to represent the ETF-

40B engine configuration. Only two of the three engines examined are considered in the

sensitivity analysis; they are the SCGT and HPRTE Efficiency engine models. Sensitive

parameters for the HPRTE Efficiency are expected to be the similar for the H-V

Efficiency cycle, and therefore the exercise was deemed redundant.

3

The second part of the project is the cycle comparison analysis which will examine

the performance parameters such as thermal efficiency, specific power, exhaust gas

temperature, and high pressure compressor inlet temperature. Mission specifications and

material and component limitations provide the scope for many of engine variables that

are to be optimized. Being for a military application, the engine is expected to have

robust performance capabilities; therefore, run cases were analyzed representing a wide

range of ambient operating conditions for all cycle configurations.

The model processes were based on thermodynamics relationships. The complete

set of equations used to close the cycle model is discussed later. The flows were all

considered steady-state and incompressible, and the turbomachinery components and

ducting were all represented as adiabatic processes.

These considerations are built in to the cycle code called Numerical Propulsion

System Simulation (NPSS). This is a DOS driven, object-oriented program that has

design, off-design, and transient run operation capabilities. Technical support for this

program was provided by the ASAO group at the NASA Glenn Research Facility.

4

CHAPTER 2 LITERATURE REVIEW

Brief History of Turbine Engine Development

Between 150-50 B.C., a Greek named Hero, living in Alexandria, Egypt, boiled

water in a sealed container that had two spouts extending from the top and slightly curved

[1]. As the water boiled, steam billowed from the spouts, rotating the entire container.

At the time it was considered a toy, but today history remembers Hero as the inventor of

the steam turbine.

Despite this early application, the first documented use of the turbine engine for

propulsion purpose was not until 1791; John Barber, a British inventor designed a simple

steam engine with a chain-driven compressor to power an automobile [1]. Then in 1872,

nearly 100 years after Barber engine, steam-powered automobile was designed, Franz

Stolze designed the first axial gas turbine engine [2]. The practicality of the engine was

suspect and it never ran unassisted.

Interest in gas turbine engines continued to increase, and developmental

breakthroughs were made in the 1930s. Great Britain and Germany were the spearhead

of these efforts as tension between the European heavyweights mounted. Faster, more

agile aircraft were being conceived, and the air forces of both nations noticed the

advantages of the jet engine over conventional piston engines. Frank Whittle of Great

Britain worked out a concept for a turbojet engine and won a patent for it in 1930 [3].

Five years later in Germany Hans van Ohaim, working independently of Whittle,

patented his own gas turbine engine system [3]. Ohaim and his colleagues witnessed the

5

first flight of their turbojet engine on August 27, 1939, powering the He.S3B aircraft [3].

The Whittle concept was shelved until mid 1935 when finally with the help of two ex-

Royal Air Force pilots the engine was built and tested by Power Jets Ltd [3]. After

working through design setbacks, including fuel control issues, the first British—

designed turbojet-powered aircraft flew in May 1941 [3]. Even though the Germans

could claim the first turbojet powered flight, the British built the first production turbojet

engine, the Roll-Royce de Haviland [3]. Turbojet development sky-rocketed in the 1940s

and 1950s; a Whittle design provided the blueprints for the first American made turbojet

engine, the General Electric I-A [7].

Gas Turbine Engine Examples in Marine Applications

The British were using simple gas turbine engines to power gun boats as early as

1947 [4]. The HMS Grey Goose was the first marine vessel to be powered by a

turboshaft engine with an inter-cooled compressor and exhaust heat recuperation (ICR)

[5]. In 1956, the U.S. Navy contracted with Westinghouse to develop a gas turbine

engine for submersible operation [6]. They designed a two shaft semi-closed ICR engine;

a novel concept that but was limited by fuel-type availability. The use of heavy sulfur

fuels triggered sulfuric acid build-up in the intercoolers which degraded the metal

components in the heat exchanger. A direct effect heat exchanger was tried with sea

water, but this only succeeded in introducing salt into the engine which deposited on the

turbomachinery parts [6]. At the same time the Westinghouse engine was under

development, General Electric was looking to convert their profitable J79 engine into a

marine gas turbine. In 1959 they introduced the LM1500. It was a simple cycle gas

turbine that produced 12,500 SHP [7]. The General Electric LM2500, introduced in

1968, ushered in the second generation marine of marine turboshaft engines. Like the

6

LM1500, the LM2500 was a derivate of a proven aero engine that powered over 300 U.S.

Naval ships [8, 7]. Moreover, thermal efficiency was improved on the LM2500 to 37

percent [8].

Advantages of Gas Turbine Engines in Marine Applications

Gas turbine engines have overtaken diesels as the power plant of choice for ferries,

cruise liners and fast-attack military ships. This trend exists because gas turbines offer

higher power output-to-weight ratios, significantly higher compactness, higher

availability, and they produce fewer emissions than marine diesels [9, 4]. The power-to-

weight advantage is best realized with an example comparing a diesel engine to a gas

turbine engine of similar power rating. The 7FDM16 marine diesel offered from General

Electric produces 4100 BHP and weighs 48,800 lbs [10]. In comparison the Lycoming

TF-40 turboshaft marine engine, produces 4,000 BHP and weighs only 1,325 lbs [11].

The significant weight disparity favoring the TF-40 is a prime reason gas turbines are

being chosen to power marine vessels requiring agility and speed. Similarly, the

compactness that gas turbine engines offer greatly improves vessel versatility and crew

and cargo capacity optimization. As an example, the 7FDM16 diesel has a volume of

920 cubic feet, whereas the TF-40 has a volume of less than 43 cubic feet [10, 11].

Subsequently, the compact, light-weight gas turbines are easier to transport and switch-

out of ships. With skilled professionals available from the aviation industry trained on

gas turbines engines, there is an abundance of mechanics and support crew able to

maintain and operate these systems [4]. Moreover, the emission reductions achieved by

gas turbine engines over comparable diesels make them more attractive to commercial

and military forces needing to placate environmental agencies such as the EPA and other

7

international bodies. A simple open-cycle gas turbine engine produces 1/3 to ¼ the

emissions of a diesel engine of comparable technology [9].

Recuperation and Inter-cooling

Simple, open-cycle turbo-shaft engines exhaust hot gas products to the atmosphere

wasting high—quality heat energy; an increasingly common use of this available heat

energy in gas turbine engines is to pre-heat the compressed gas flow before the

combustion stage. This process is called exhaust heat recuperation. As a result of raising

the combustor inlet temperature, less fuel is required to achieve the desired turbine inlet

temperature and desired power output. This directly impacts the thermal efficiency and

specific power of the engine, raising thermal efficiency but dropping specific power in

most cases. Any instance in which fuel use can be decreased has a direct positive impact

on the cycle thermal efficiency. It is important to note that gas turbine engine

recuperators generally work better in engines with only moderate pressure ratios [12].

Qualitatively, one can see that as the engine pressure ratio rises, the compressor exit

temperature and turbine exit temperature approach each other. In practice this would

drop the capacity of the recuperator to pre-heat the compressed air before combustion,

thus rendering it ineffective.

A second improvement on the simple gas turbine engine is the addition of an inter-

cooler. Inter-coolers are placed between the low pressure and high pressure compressors

to reduce the air temperature exiting the last stage of the compressor. Assuming the

process is adiabatic and the air is a calorically perfect gas, the power required to drive the

compressor is written as TcmW pcomp Δ= && . This assumes a control volume analysis

around the entire compressor for all stages [3]. The inter-cooler delivers a lower

8

temperature fluid to the high pressure compressor stage. If the same pressure ratio is

applied to the high pressure stage, the exhausting fluid temperature would be lower than

if no inter-cooling had been performed. The outcome is that TΔ for the entire

compressor has been decreased, and subsequently, the total power requirement for the

compressor has also been decreased. The net effect on the cycle thermal efficiency is the

same as raising the adiabatic efficiency of the entire compressor. The outcome is a net

available power increase of 25 to 30% [5]. Coolants exist for both sea and air

applications. Jet aircraft have -50°C ambient air available and naval ships have the

abundant salt water reserves of the oceans.

Additionally, combining both compressor inter-cooling and exhaust gas

recuperation provides a further improvement to cycle thermal efficiency. Engines that

employ this technology are referred to as inter-cooling recuperation (ICR) engines. With

the inter-cooler cooling the compressor discharge, the temperature difference between it

and the turbine discharge increases—the outcome is an improved recuperator

performance [12]. In 1953 Rolls Royce introduced the RM60 ICR engine which powered

the gunboat HMS Grey Goose [5]. Though innovative and more efficient than the steam

engine it replaced, the RM60 was too complex to operate using existing controls

technology. A further example reviewed for this project compares two gas turbine

engines, a simple open-cycle and an ICR, for a marine destroyer application. The study

noted that fuel use is reduced by 30% with the ICR engine [5, 13].

In 1990, General Electric began retrofitting their mid-size turboshaft engine, the

LM2500, in hopes of improving its thermal efficiency by 30% [13]. This project was

sidelined in 1991 when a team led by Northrop Grumman won a $400 million, 9-year

9

development contract to develop and build a replacement for the LM2500 marine gas

turbine [14]. Program leaders Northrop Grumman and Rolls-Royce chose an ICR engine

design, called the WR-21, for the navies of the United States, Canada, Great Britain, and

France [14]. John Chiprich, who managed the ICR development program, noted that the

new engine will reduce the fuel consumption for the entire marine turbine powered fleet

of the United States by 27 to 30% [14].

One negative aspect to the ICR concept is that it has a lower power limit for it to be

considered effective. Blade tip leakage for gas turbine engines that have a nominal

power rating below 1.5MW overrides any efficiency gained from the implementation of

ICR technology [15].

Semi-Closed Cycles

A semi-closed gas turbine cycle is one in which hot exhaust products are

recirculated, combined with fresh air, and then burned again in the combustion chamber.

Example configurations can include inter-cooling and recuperation, and some are

turbocharged to boost core engine pressures. Despite the added complication of engine

components and weight addition; many semi-closed cycle configurations have significant

performance related benefits. For instance, semi-closed cycles that are turbocharged,

have higher specific power, reduced recuperator size (if a recuperator is present) which

improves heat transfer coefficients, and higher part-load performance characteristics [13].

All semi-closed cycles benefit from reduced emissions since reduced oxygen

concentrations reduce flame temperatures [13].

Some of the earliest semi-closed gas turbine engine configurations were proposed

by the Sulzer Brothers in the late 1940s [16]. Their 20 MW gas turbine system for the

Weinfelden Station was a complex system that achieved a cycle thermal efficiency of

10

32% for full load capacity and 28 % for half load capacity [16]. The earliest example of

a semi-closed gas turbine system for naval propulsion was the Wolverine engine

developed by Westinghouse [6]. The submarine engine program which began in 1956

called for a two-shaft, semi-closed, ICR turboshaft engine [6]. It was never a production

engine because of sulfuric acid buildup that degraded the metallic intercooler

components. This was attributed to the high concentration of sulfur in early diesel fuels.

More recent research projects on semi-closed gas turbine cycles conducted by the

University of Florida, Energy and Gas Dynamics Laboratory will be highlighted in the

final section of this chapter.

Computer Code Simulators

Because of the complexity of the cycles that need to be simulated and the iterative

nature of semi-closed cycle modeling, it is convenient to employ the use of a

computational code to perform the numerous calculations. There were several

computational thermodynamic cycle programs that were potential platforms for this

project. Below is a brief overview of the programs surveyed.

Gas turbine Simulation Program (GSP) is a product of the National Aerospace

Laboratory—The Netherlands (NLR) [17]. The GSP website boasts of a user friendly

platform with drag-and-drop components ready for building engines models. The code

can be used for steady-state as well as transient simulation. Material specifications and

life-cycle information can be incorporated for failure and deterioration analysis.

Unknown, however, is whether or not GSP can model semi-closed engine cycles. A

second code called GASCAN was reviewed by Joseph Landon. This code models fluid

movement as well as thermodynamic state variables for engine simulations. Semi-closed

11

operation is not explicitly discussed but simple and complex cycles are apparently easily

modeled.

A third modeling program reviewed was Navy/NASA Engine Program (NEPP); it

was developed to perform gas turbine cycle performance analysis for jet aircraft engines.

NEPP is an older component-based engine modeling program that has design and off-

design modeling capabilities with performance map integration. User instantiated

variables can be controlled to hold specific parameters constant while the program

converges to its solution. This program was eliminated because it can not model

recirculated flows [13]. NEPP was only the first of three NASA programs evaluated for

this modeling project. The second NASA code was ROCket Engine Transient

Simulation (ROCETS) developed at Marshall Space Flight Center. This program

provides a suite of engine component modules to assist users in building their models; it

also allows users to create their own modules to model more exotic engine cycles [18].

Like NEPP, ROCETS gives the developer the ability to vary certain parameters until

other constraints are satisfied and a converged solution is determined [18]. Users have

the option of operating in design or off-design mode as the program has the capability of

reading performance maps for compressors and turbines. ROCETS was used in

modeling efforts at the University of Florida in the 1990s. The program is capable of

modeling recirculation in gas turbines and water particulate extraction. Being somewhat

antiquated, the program was dismissed as a possible platform for the project considering

the unlikely availability of user support.

A commercial software package option was the versatile ASPEN PLUS. The

ASPEN PLUS engineering suite is a robust package of software programs that can handle

12

all of the modeling requirements for this project. Once again, here is a program that

provides users with the option of running their cycle in design, off-design, or transient

modes. Their website displays screen shots of a pleasant graphic user interface with

drag-n-drop engine components [19].

The third software program from NASA, Numerical Propulsion System Simulation

(NPSS) is a product of the Aeropropulsion Systems Analysis Office (ASAO) at the Glenn

Research Center. NPSS is set up to operate similar to the earlier programs NEPP and

ROCETS. Accordingly, NPSS offers users the convenience of object-oriented engine

components for building cycle models [20]. Off-design and transient modeling are

options in addition to running in the design point mode [20]. The model developer has

control of convergence through constraint handling. Since this program became the

platform of choice for this project, its capabilities will be discussed in further detail in

Chapter 2.

Previous Gas Turbine Research at the University of Florida

In 1995 Todd Nemec performed a thermodynamic design point analysis on a semi-

closed ICR gas turbine engine with a Rankine bottoming cycle [21]. Nemec developed

his model using the ROCETS program discussed earlier—his analysis concluded that the

combined cycle with superheated steam in the bottoming cycle resulted in an overall

efficiency of 54.5% [21]. The next body of work on semi-closed cycles was performed

by Joseph Landon. Landon performed design and off-design point analysis of two

separate regenerative feedback turbine engines (RFTE) [13]. The turbocharger

configuration resembled the topping cycle that Nemec modeled. The other configuration

sent the combustion products through a power turbine before the recuperation heat

exchanger. The analysis predicted that the power turbine configuration produced the

13

highest thermal efficiency, 48.2%, compared to 46% for the turbocharger case [13]. Off-

design analysis revealed that the turbocharger model was the most efficient between 20%

and 80% power capacity [13].

Russell MacFarlane used the ROCETS program to model water extraction and

injection on the RFTE engine [12]. MacFarlane found that water removal caused a

decrease in specific fuel consumption and a slight increase of specific power [12]. He

surmised that water removal was particularly influenced by “recirculation ratio, cooler

effectiveness, and first stage pressure ratio” [12]. George Danias extended the study of

the RFTE cycle and investigated design and off-design performance of three separate

configurations for a helicopter engine application [18]. His conclusions stated that the

three RFTE configurations were 30 to 35% more efficient than the T700-701C, baseline

engine [18].

Currently, a research project is underway to design and develop a combined cycle,

power-refrigeration cycle called the HPRTE-VARS. The High Power Regenerative

Turbine Engine (HPRTE) uses exhaust gas heat to power the vapor absorption

refrigeration system (VARS). A design point performance study was carried out by

Joseph Boza analyzing two HPRTE-VARS engine sizes, a small 100 kW engine and a

larger 40 MW engine. Boza calculated the performance parameters based on a constant

high pressure compressor (HPC) inlet temperature of 5 ° C. Excess refrigeration

capacity (that capacity not used to cool the HPC inlet stream) was considered in the

combined cycle efficiency value. The larger engine analysis predicted a combined cycle

efficiency of 63% while the small engine efficiency was determined to be 43% [22]. He

determined that increasing ambient temperature limits the excess refrigeration capacity,

14

and at an ambient temperature of 45 ° C the combined-cycle system has no excess

refrigeration. For his analysis, Boza used a spreadsheet cycle code to predict the

performance of the HPRTE; this was in conjunction with a VARS model that he created.

In Chapter 6 the spreadsheet model has been used to benchmark the NPSS program used

in this project. The spreadsheet HPRTE model is not configured to consider the low

pressure spool of the engine as a turbocharger—in the comparison in Chapter 6, the

spreadsheet cycle model will be constrained manually for the turbocharger configuration.

Life cycle cost analyses of the HPRTE-VARS was performed and compared to a

microturbine engine by Viahbav Malhatra. Using a standard life cycle cost analysis

procedure, Malhatra determined that the HPRTE-VARS system exhibited a life cycle cost

savings of 7% over the competing microturbine system [23]. One primary reason for the

cost savings was associated with the HPRTE being turbocharged—this enabled smaller

and less expensive engine components to be considered. The other reason for the cost

savings was directly related to fuel consumption. HPRTE fuel costs were partially

compensated by the proceeds from available refrigeration capacity of the VARS unit

[23]. To obtain his results Malhatra used a Fortran model of the HPRTE-VARS created

by Jameel Khan. Khan performed his dissertation study on the design and optimization

of the HPRTE-VARS combined cycle developing a high fidelity, thermodynamic model

for both the engine and the refrigeration systems. He used the optimization package

LSGRG2 to determine the best design-point engine parameters considering such outputs

as power, refrigeration, and water. His results for the combined cycle with the

OHNH 23 / refrigeration system predicted a cycle thermal efficiency of 40.5% with a

ratio of water production to fuel (propane) consumption of 1.5 [24]. Including the excess

15

refrigeration produced by the cycle, a combined cycle thermal efficiency was evaluated

as 44%.

16

CHAPTER 3 NUMERICAL PROPULSION SYSTEM SIMULATION ARCHITECTURE

Numerical Propulsion System Simulation (NPSS) was developed by

Aeropropulsion Systems Analysis Office (ASAO) at the National Aeronautics and Space

Administration (NASA) Glenn Research Center, Cleveland, OH in conjunction with the

Department of Defense and leaders in the aeropropulsion industry. The purpose of the

code was to speed the development process of new gas turbine engine concepts for

military and civilian applications. It is a component-based engine cycle simulation

program that can model design and off-design point operation in steady-state or transient

mode [20]. The code can be used as a stand-alone analysis program or it can be coupled

in conjunction with other codes to produce higher fidelity models.

Model

Engine models are created using any standard text editor such as Microsoft

Wordpad. The model file contains the instructions and commands required by NPSS to

build an engine model. The engine model file combines the engine components

(elements) in a systematic manner that is consistent with the engine cycle the user is

modeling. Here, elements are connected to create the flow stations of the engine; these

flow stations are created by linking the flow ports between elements. In the model the

thermodynamic package, solver solution method, and model constraints should also be

specified if different than the defaults. These subjects will be discussed in further detail

later in the Chapter 3.

17

Figure 3-1 is a schematic representation of an example engine modeled using

NPSS. The elements are plainly listed; there is an inlet, compressor, burner, turbine,

shaft, duct, and exhaust. The working fluid properties are passed through flow ports from

one element to the next. Shaft ports connect the compressor and turbine with the shaft

element in order to perform the power balance for the engine. The interaction of a

subelement, CompressorMap with its parent element, Compressor, is shown with its

socket link. This particular model has an assembly for the major engine components.

The assembly compartmentalizes any processes or calculations performed by these

components from the rest of the model.

Elements

Elements are the corner stones of the engine model. Although NPSS comes with a

full suite of engine component modules, users are encouraged to create their own

elements to model their unique circumstances using the C++ type syntax of NPSS. As

mentioned above, elements are responsible for performing the individual thermodynamic

processes that simulate the physical engine components. The modules use standard

thermodynamic relationships to simulate these processes. The level of modeling

sophistication is entirely user driven as loss coefficients and scalars may be applied to

variables. Mach number effects are calculable. For higher fidelity models heat and

frictional energy dissipation may be considered. For the purpose of this analysis the

cycle models were kept as simple as possible to shorten computing run-times.

Nevertheless, even simple models require a certain level of complexity—for those

cases there are supplemental routines added to elements called subelements and

functions. Subelements are subroutines that can be called by elements to perform

calculations or performance table look-ups. For instance, the turbine element for a model

18

that is operating in off-design mode would use a subelement to determine the efficiency

value from data tables. Functions are a type of subroutine that is user instantiated in a

particular element that requests particular calculations be performed. Function

calculations take precedence over the solver driven calculations. They may be performed

before, after, or during solver run-time depending on the desire of the user.

FlowStation

For an element to perform its calculations, properties and state information must be

known as initial conditions. These initial conditions are set by the user or the computer

and passed to the element through a flow port. When flow ports are used to link two

elements, this bridge is called a FlowStation. There is a main FlowStation subroutine and

then there are the specific FlowStation subroutines unique to each thermodynamic model.

The main FlowStation subroutine is responsible for linking the model to the appropriate

subroutines that handle the subroutine look-ups. When NPSS uses the Chemical

Equilibrium with Applications (CEA) thermodynamic software, the main FlowStation

subroutine links the model file/files with the CEA program allowing the passage of

species and state information between the two programs.

FlowStartEnd

There are elements in NPSS specifically designed to either begin or end a fluid

flow path. Semi-closed gas turbine engine modeling in NPSS makes use of these flow

start/end elements to obtain converged solutions. The solution solver in NPSS requires a

single initial pass through the model elements to create the flow path and flow stations—

and essentially build the engine model. For open cycle gas turbine engines this task

requires no extra consideration by the modeler. The solution solver can logically step

through the engine from the inlet element to the exhaust element for the preprocess pass.

19

However, all of the HPRTE configurations have mixing junctions upstream of the core

engine components adding a further level of complexity that the solution solver must

negotiate.

The solution requires added components, FlowStart and FlowEnd elements, and

additional constraints added to the solution solver. For convenience and brevity the

ASAO developed the element FowStartEnd to replace the FlowStart/FlowEnd

elements—this element also contains the additional constraints required, eliminating the

necessity to initialize these in the main model file.

To be complete it is best to describe the coding required to gain convergence of a

regenerative gas turbine model using FlowStart, FlowEnd, and FlowStartEnd elements.

When the solver is stepping through the HPRTE it expects to have a hot-side flow station

already instantiated when it reaches the recuperator inlet after the high-pressure

compressor exit. Therefore, a FlowStart element is created and added to the solver

sequence (responsible for the order of element preprocess loading) before the high-

pressure recuperator flow station is created. Initial conditions are given to the stream

including temperature, pressure, mass flow rate, fuel-to-air ratio, water-air-ratio, and fuel

type. This flow station is 7a.

Now the solver can continue to load the model to the point of the high-pressure

turbine exit flow station. This is the point where the ‘bridge’ is made with the FlowStart

element instantiated earlier. Here, a FlowEnd element is created and the state of the flow

exiting the high-pressure turbine is stored in this element. The flow station here is 7b.

Since the flow conditions cannot be directly passed from the FlowEnd element to the

FlowStart element, the solver is given the task of iterating on all the flow station 7

20

parameters until the conditions match in both elements. To make this happen the user

sets up five variables, which NPSS considers ‘independents’, to iterate on until their five

counterpart constraints, which NPSS deems the ‘dependents’, are satisfied. These five

independent variables are listed as: stagnation temperature and pressure, mass flow rate,

fuel-air-ratio, and water-air-ratio.

The constraints are generally written as equations that must be satisfied for solver

convergence to be recognized. One example of a dependent constraint from the

FlowStartEnd element is given below in NPSS syntax.

Dependent dep_P{

eq_lhs = "Fl_I.Pt";

eq_rhs = "Fl_O.Pt";

autoSetup = TRUE;

}

The constraint variable is ‘dep_P’. The left hand side of the equation is set equal to

the stagnation pressure of the flow entering FlowStartEnd, and the right hand side is set

equal to the exiting stagnation pressure. This constraint is added to the solver along with

four others corresponding to the variables listed above. Figure 3-2 shows the schematic

representation of the procedure that was just described.

Thermodynamic Properties Package

Chemical Equilibrium with Applications (CEA), obtains chemical equilibrium

compositions for pre-defined thermodynamic states. Two thermodynamic state

properties must be known for the rest to be calculated or obtained from table subroutines.

This requires two input files:

21

1. Thermo.inp—Contains thermodynamic property data in least squares coefficients. These data can be used to calculate reference-state molar heat capacity, enthalpy, and entropy at a given temperature.

2. Trans.inp—Contains the transport property coefficients for the species CEA uses the Gibbs free-energy minimization method to calculate chemical

equilibrium at each state point. Chemical reaction equations are unnecessary when using

the free-energy minimization method and chemical species can be treated individually.

For a detailed description of the theory and methods used in CEA please see reference

[25].

CEAFlowstations are responsible for passing constituent and state point

temperature and pressure from NPSS to CEA.

Solver

The NPSS solver is responsible for bringing the model to a converged solution. In

order to accomplish this task the user must choose which engine parameters to constrain.

Constrained parameters are called model “dependent variables”. To satisfy the dependent

variables a set of “independent variables” must be defined and iterated. This iterative

approach to find a solution begins with an initial state guess, and that is subsequently

refined until a satisfactory solution is found.

The solver solution method is a quasi-Newton method. For a simple description

assume there is only one constraint on the model, and as a result only one variable to

iterate to meet it. The initial value of the independent variable is user specified, and with

that the initial value of the variable desired to be constrained can be found. Then the

independent variable is perturbed a certain amount chosen by the solver and a new value

for the dependent variable is found. The solver now must decide if this new value of the

variable to be constrained is a satisfactory one. A partial derivative error term is

calculated,

22

( )( )II

II

tValueIndependentValueIndependenalueDependentValueDependentVErrorTerm

−−

= +

+

1

1

, (3.1)

where I denotes the iteration number. If it is outside the acceptable tolerance region, the

process is begun again. With a system of constraints a Jacobian matrix would be created

to hold all the error terms. The new perturbation terms would be calculated from the

previous Jacobian matrix:

[ ]

( )( )

( )( )

( )( ) ⎥

⎥⎥⎥⎥⎥⎥⎥⎥

⎢⎢⎢⎢⎢⎢⎢⎢⎢

−•••

−••

=

+

+

+

+

+

+

11

11

1

111

11

11

11

11

1

IndVIndVDepVDepV

IndVIndVDepVDepV

IndVIndVDepVDepV

J

I

Im

Im

nI

n

II

I

II

I (3.2)

Here there are “n” number of independents and “m” number of dependents.

The Jacobian can be related to the independent variables with the expression

[ ] [ ] ( )[ ]III xFxJ −=Δ⋅ , (3.3)

where [ ]IxΔ is the matrix composed of the independent perturbation values. The ( )[ ]IxF

matrix holds the values of the dependent constraints at the thI ' iteration. The new

independent values may now be calculated with the following:

[ ] [ ] [ ] ( )[ ]IIII xFJxx ⋅−=−+ 11 . (3.4)

With [ ]1+Ix now determined, [ ]1+Δ Ix and ( )[ ]1+IxF can be found and a new Jacobian

matrix created. The process continues until the Jacobian error values are within the

acceptable tolerance limits of the solver.

23

Figure 3-1 Example NPSS engine model [19]

24

HPT Recuperator

NPSS Code/Element Representation of Above Engine State

HPT FlowStart

Recuperator

RecuperatorFlowEnd

HPT FlowStartEn

Simplified Code Representation

7a 7b

7a 7b

7temp

Figure 3-2 State 7 of HPRTE engine cycle

25

CHAPTER 4 CYCLE CONFIGURATIONS AND BASE POINT ASSUMPTIONS

Before discussing the thermodynamics relationships used in the analysis, it is

necessary to give an overview of the cycles from a systems standpoint. This analysis

compares the design point performance of three engine configurations. The first engine

is a simple cycle gas turbine engine (SCGT). It has been modeled to predict the

performance of the production engine, ETF-40B, which powers the military LCAC for

the United States Navy. The SCGT will be compared to two variations of the HPRTE

engine, the base HPRTE and a variant that uses refrigeration capacity to cool the high

pressure compressor inlet stream.

Major Model Features

When comparing engine systems, it is convenient to understand the major features

of each model. Listed in Table 4.1 is a breakdown of the features that distinguish the

engine configurations from one another. The HPRTE cycles are two spool engines with

exhaust gas product heat recuperation. Both are semi-closed and have compressor inter-

cooling. The H-V Efficiency has additional cooling capacity provided by a vapor

absorption refrigeration system (VARS). The additional cooling enables exhausted water

vapor to be condensed and collected for use elsewhere or for injection after the high

pressure compressor.

26

Flow Path Descriptions & Schematics

Simple Cycle Gas Turbine Engine Model

As mentioned earlier, the SCGT is a simple, open cycle gas turbine engine. For

this analysis the model with have a total of five flow stations (Figure 4-1). State 1 is the

inlet stream. From State 1 to 2 the flow undergoes an adiabatic compression process in

compressor, C1. From State 2 to 3 there is a constant area, premixed burner, B. The

process from State 3 to 4 is an adiabatic expansion process through the turbine, T1.

Mechanical work generated by the turbine drives the compressor and supplies power for

the ship propellers or lift fans. State 5 is the fuel flow station. JP-4 was the fuel of

choice for this analysis because it is widely used in industry and has a high availability.

High Pressure Regenerative Turbine Engine Efficiency Model

Figure 4-2 is a schematic representation for the Efficiency and Power modes of the

HPRTE cycle. The Power mode concept incorporates a flow splitter to bypass some

exhaust from the high pressure turbine and send it directly to the low pressure turbine.

Initially, the Power mode had been considered for this project to give additional boost

capabilities to the low pressure spool. However, while completing the analysis it was

determined that the Efficiency mode predicts sufficient boost for the system and any

additional boost pressure would result in a turbocharger design outside of modern

technology limits.

There are 14 states for the basic HPRTE (the Power mode has 16). Air enters at

State 1 and undergoes an adiabatic compression process in the low pressure compressor,

LPC, before reaching State 2. Next, the fresh air from State 2 is combined with the

recirculated exhaust gas products from State 10 in an isobaric, adiabatic mixing process.

The resultant State is 2.9. Now the combined flow passes through a sea water cooled

27

heat exchanger called the main gas cooler (MGC). The effectiveness, pressure drop, and

process fluid temperature are all given. The resulting State is 3.0. After the gas has been

cooled it goes through another adiabatic compression process in the high pressure

compressor, HPC. The resultant State 4 has the maximum system pressure. Following

the HPC there is a heat recuperation process (RHX) in which high-temperature exhaust

gas product stream preheats the State 4 flow resulting is State 5. From state 5 to 6 the gas

is mixed with fuel and ignited in the combustion chamber, B. A small pressure drop is

applied before State 6 to simulate friction losses in the combustor. The high pressure

turbine inlet temperature, or TIT, was chosen to be 2500°R—an acceptable value for a

medium size engine.

The expansion across the high pressure turbine, HPT, produces the power to drive

the HPC and the net BHP is available power for the vessel. State 7 is State 7.11 in the

Efficiency mode, and that flow passes through the RHX, rejecting heat to State 4. The

only flow splitter for the Efficiency cycle comes at State 9. Here, a user defined

recirculation ratio determines the mass flow rates at State 7.15 and 10. State 10

recombines with fresh air flow from the LPC exit. State 7.2 is also State 7.3 in Efficiency

mode. The final expansion process across the low pressure turbine, or LPT, exhausts to

the environment at State 8.

High Pressure Regenerative Turbine Engine with Vapor Absorption Refrigeration System Efficiency Model

Figure 4-3 is a schematic representation of the H-V Efficiency. The HPRTE-

VARS modes differs from the HPRTE modes only by the addition of two heat

exchangers in the flow path after the recirculated gas products combine with the fresh

inlet air at State 2.9. The generator (GEN) and the evaporator (EVP) are two of the heat

28

exchangers that make up part of the VARS. A schematic of the VARS is also included as

Figure 4-4 for clarification. It was not modeled since the scope of this analysis only

included modeling the gas path side of the combined cycle system. The point of water

collection is shown on the figure, as well. The computational model of this cycle

required the addition of a separator element to perform the water extraction. The

separator is discussed in Chapter 5, Thermodynamic Modeling and Analysis.

Notice that the HPRTE cycles require an iterative solution method to obtain model

convergence because of the semi-closed operation. For the first iteration of the engine

cycle an initial guess for the temperature at State10 is given.

Simple Cycle Gas Turbine Engine Design Assumptions and HPRTE Cycles Base Point Assumptions

The SCGT is a medium size, open-cycle gas turbine engine modeled after the ETF-

40B. The ETF-40B has a seven stage axial compressor followed by a single stage

centrifugal compressor yielding an overall pressure ratio of 10.4 [Robert Cole]. The

nominal output shaft horsepower is 4000 SHP. Turbine inlet temperature was assumed to

be 2500°R. Turbomachinery efficiency information was provided by Dan Brown of

Brown Turbine Technologies. All other engine design parameters were chosen based on

conservative current technology limits. See Table 4-2 for complete details.

The base point HPRTE component parameters are listed in Table 4-3. The same

methodology used to determine the design parameters for SCGT was considered when

deciding base-line design values for the HPRTE engine cycle configurations—size and

technology limitations were applied.

There were material and computational limitations that existed and needed to be

accounted for to preserve the fidelity of the engine model. They are as follows: TIT

29

maximum was 2500°R, hot side recuperator inlet temperature maximum was 2059°R,

turbocharger pressure ratio maximum was 7.5, and HPC inlet temperature minimum was

491°R (NPSS limitations).

Table 4-1 Comparison of major configuration features

SCGTHPRTE Efficiency * * * *H-V Efficiency * * * * * *

Model FeaturesIntercooled

CompressorsVARS cooling

Water ExtractionModel Semi-Closd Turbocharger

Pressurized Recuperatored

1 2 3 4B

C1 T1

5Fuel

Air

Figure 4-1 Simple Cycle Gas Turbine (SCGT) engine model configuration

30

Figure 4-2 High Pressure Regenerative Turbine Engine model, both efficiency and power configurations represented

77.

12

7.2

7.3

86

54

32.

92

1

109

7.11

LPC

RH

X

HPC

MG

CH

PTB

LPT

Fuel

Air

Sea

Wat

er

31

Figure 4-3 High Pressure Regenerative Turbine Engine-Vapor Absorption Refrigeration System, both efficiency and power model configurations represented

77.

12

7.2

7.3

86

54

32

1

109

7.11

2.91

2.92

2.9

LPC

MG

CH

PC

RH

X

BH

PTLP

TG

ENEV

P

Fuel

Air

Wat

e

Sea

Wat

erH

P R

efrig

eran

tLP

Ref

riger

ant

32

GEN

EVP

CND

CND

2.9 2.91

2.923

PumpExpander

Figure 4-4 Vapor Absorption Refrigeration Cycle with HPRTE flow connections

Table 4-2 Simple Cycle Gas Turbine engine design point parameters Parameter ValueC1 Adiabatic Efficiency 0.858Burner Efficiency 0.99Burner Pressure Drop 0.03Turbine Inlet Temperature 2500°RT1 Adiabatic Efficiency 0.873

Table 4-3 Base case model assumptions for HPRTE cycles [3], [26], [27] Parameter Value

Ambient Temperature 544.67°RSea Water Temperature 544.67°R

Ambient Pressure 14.7 PSILPC Adiabatic Efficiency 0.83

GEN Effectiveness 0.85GEN Pressure Drop 0.03MGC Effectiveness 0.85MGC Pressure Drop 0.03EVP Effectiveness 0.85EVP Pressure Drop 0.03

HPC Adiabatic Efficiency 0.858RHXEffectiveness 0.85

RHX Pressure Drop State 4-5 0.04RHX Pressure Drop State 7.11-9 0.04

B Efficiency 0.99B Pressure Drop 0.03

HPT Inlet Temperature 2500°RHPT Adiabatic Efficiency 0.873

Recirculation Ratio 3LPT Adiabatic Efficiency 0.87

Fuel Hydrogen to Carbon Ratio 1.93:1

33

CHAPTER 5 THERMODYNAMIC MODELING AND ANALYSIS

Chapters 3 and 4 addressed the computational structure of NPSS and the cycle

configurations of the models including the design point assumptions. While top level

NPSS calculations are performed by the solution solver, the intermediate operations

performed during every iterative pass to calculate the thermodynamic states are discussed

next. Chapter 5 develops the theory for these auxiliary thermodynamic relations that

drive the model elements (subroutines). These relations are developed using fundamental

thermodynamic concepts.

Thermodynamic Elements

Heat Exchangers

Heat exchangers are an important component in HPRTE cycles. The base HPRTE

Efficiency model mode has two heat exchangers, MGC and RHX; and the combined

cycle, H-V Efficiency, has four heat exchanger elements including three for compressor

inter-cooling. Those for the inter-cooling have defined process inlet flow states. Mass is

conserved by setting the exit mass flow rate equal to the entrance mass flow rate.

User defined inputs include effectiveness, ε , and 1_0 inPPΔ . Let effectiveness be

defined as

( ) ( )( ) ( )

( ) ( )( ) ( )2_01_011

1_01_011

__0__0min_min

__0__0_

max ininpin

outinpin

incoldinhotp

outhotinhothotphot

TTCmTTCm

TTCmTTCm

QQ

−⋅

−⋅=

−⋅

−⋅==

&

&

&

&

&

&ε (5.1)

hotpC _ is the hot side specific heat at constant pressure, and min_pC is the specific heat of

the minimum capacity flow stream.

34

Therefore, ( )( )2_01_0

1_01_0

inin

outin

TTTT

−=ε . (5.2)

The only unknown in Equation 5.2 is 1_0 outT .

The capacity of the process fluid is set such that it is always the maximum capacity

stream. This ensures that it is not used in the calculation above.

The exit pressure is determined using the following equation:

( )1_01_01_0 1 ininout PPPP Δ−⋅= . (5.3)

Know known are the parameters 1_0 outT , 1_0 outP , and 1outm& . The exit state is set.

Mixers

The mixer is modeled as an adiabatic, constant static pressure process. Because

there is no consideration given for Mach number effects, the stagnation pressures of the

two flows entering the mixer must be identical. This requires a model constraint be set

up by the user for each HPRTE model and satisfied by the solution solver. All HPRTE

models have recirculation mixers which are tasked with combing the recirculated exhaust

gas products with fresh air discharged from the low pressure compressor.

A mass balance requires:

21 ininout mmm &&& += (5.4)

Assuming adiabatic mixing, the energy balance is as follows:

out

ininininout m

hmhmh

&

&& 2_021_01_0

+= . (5.5)

Constant pressure mixing implies:

outinin PPP _02_01_0 == . (5.6)

35

Other parameters such as the outFAR and the mass fractions are mass averaged. For

example:

out

ininininout m

FARmFARmFAR

&

&& 2211 += . (5.7)

With outh _0 , outP _0 , outm& , and the exit state mass fractions all known, all other

thermodynamic properties can be found.

Splitter

In Chapter 4 the cycle schematics for the HPRTE cycle models showed flow

splitting occurring at State 9. To accomplish this feature with a computer model a splitter

component must be defined to separates flow into two streams before exhausting to the

environment. The recirculation splitter is tasked with the job of splitting the flow stream

on a mass basis after the high temperature recuperation process (State 9). A portion of

the flow is reconstituted with fresh air before heading back through the core engine

components while the rest is directed to the low pressure turbine (LPT) to power the

turbocharger. A bypass ratio, BPR, is user defined to represent the mass basis split of the

flow streams. The recirculation splitter inlet state is defined by the following know

parameters: inT _0 , inP _0 , intotm _& , inFAR , inh _0 , and mass fractions for all species.

In general BPR is defined as

1_

2_

outtot

outtot

mm

BPR&

&= . (5.8)

For this application the bypass ratio is defined as

exhaustedtot

recirctotcirc m

mBPR

_

_Re &

&= . (5.9)

36

recirctotm _& is the mass flow rate recirculated and mixed with fresh air. exhaustedtotm _& is

the mass flow rate that passes directly to the low pressure turbine and be exhausted from

the system at State 8.

The user also reserves the option of applying flow pressure drops to either or both

of the split streams, but for this analysis the splitter is modeled as an isobaric process.

Similarly, the process is adiabatic, as there is no heat transfer. The mass fractions are

unchanged; therefore, the exit state of each flow is defined.

inoutout PPP _02_01_0 == (5.10)

inoutout TTT _02_01_0 == (5.11)

Water Extractor

The water extraction component is only present in the H-V Efficiency

configuration. Because water vapor is present in the recirculated mixed gases and the

cooling capacity of the three heat exchangers is significant to cause condensation to occur

in the flow stream, it is desirable to separate the liquid water from the gas flow before the

inlet to the high pressure compressor. The separation of liquid water from the flow

stream is modeled as an isentropic process. The inlet state is completely defined;

therefore, liquidOHm _2& and liquidOHh _2 are readily available from CEA. The exit state is

defined by first setting the inlet and exit temperatures and pressures equal.

inout PP _0_0 = (5.12)

inout TT _0_0 = (5.13)

Then the exit mass flow rate and enthalpy are set.

liquidOHintotouttot mmm _2__ &&& −= (5.14)

37

liquidOHinout hhh _2_0_0 −= (5.15)

The exit state of the water extractor is now defined.

Compressors

Compressors inlet states are defined with the following parameters passed to the

element: inT _0 , inP _0 , intotm _& , inFAR , inh _0 , and mass fractions for all species. The

performance of the compressor is determined by the following parameters: pressure ratio

( CompPR ) and adiabatic efficiency ( adComp _η ).

Exit pressure is determined first with the equation

inCompout PPRP _0_0 ⋅= . (5.16)

The other thermodynamic parameter, the adiabatic efficiency, is used to calculate

the exit state point parameters in the NPSS Compressor module. Define the adiabatic

compressor efficiency as

inout

inidealoutadComp hh

hhworkcompressoradiabatic

workcompressorideal

_0_0

_0__0_ __

__−

−==η . (5.17)

Determining the ideal exit state enthalpy is straight forward knowing outP _0 and

idealouts __0 if idealoutin ss __0_0 = . Since entropy and enthalpy are only functions of

temperature; the exit state ideal temperature is quickly found along with enthalpy. Now,

rearrange and directly solve Equation 5.17 for outh _0 . With the exit pressure and

enthalpy know known, all exit state thermodynamic parameters are readily calculated by

CEA.

The power required by the compressor is also calculated.

outoutininComp hmhmW _0_0_0_0 ⋅−⋅= &&& (5.18)

38

The power is converted from Btu/sec to HP:

HP

lbfftBTU

lbfftWComp

1sec

5501

778

⋅⋅& . (5.19)

Polytropic efficiency, polyComp _η , is an output parameter calculated from the

entrance and exit entropies and pressures. The derivation is as follows:

The definition of the polytropic efficiency is

dhdhi

polyComp =_η . (5.20)

To arrive at this equation, first consider a reversible form of the energy equation.

Since PdvvdPdudh ++= , (5.21)

vdPdhPdvvdPPdvdhPdvduTds −=+−−=+= )( (5.22)

Therefore, P

dPRTdhdP

Tv

Tdhds −=−= . (5.23)

Solving Equation 5.23 for Tdh yields

PdPRds

Tdh

+= . (5.24)

For an isentropic process 0=ds . Therefore,

PdPR

Tdhi = . (5.25)

Combining Equations 5.24 and 5.25 results in the following:

PdPRds

PdPR

Tdh

Tdh

dhdh i

ipolyComp

+===_η . (5.26)

Integrating Equation 5.26 from the inlet state to the exit state yields:

39

( )( )CompinCompinout

CompinComppolyComp PRRss

PRRlog

log

__0_0

__ ⋅+−

⋅=η . (5.27)

Turbines

Turbines provide the power to drive the compressors as well as the net power for

the ship propellers and lift fans (if LCAC is the mission). The NPSS model Turbine

element requires a defined entrance state to include such parameters as inT _0 , inP _0 ,

intotm _& , inFAR , inh _0 , and mass fractions for all species present. As was the case with the

compressors, the performance of the turbine components is determined by the defined

parameters: pressure ratio ( TurbPR ) and adiabatic efficiency ( adTurb _η ). NPSS defines

TurbPR differently than most turbomachinery reference texts. Here it is defined as:

out

inTurb P

PPR

_0

_0= . (5.28)

The exit state can be determined by first applying the turbine pressure ratio.

Turb

inout PR

PP _0

_0 = . Turbinout PRPP _0_0 = (5.29)

As was the case for the compressor, outh _0 is the other thermodynamic parameter

necessary to in order to define the exit state. The turbine adiabatic efficiency is defined

as:

idealoutin

outinadTurb hh

hhworkturbineideal

workturbineadiabatic

__0_0

_0_0_ __

__−

−==η . (5.30)

The power generated by the turbine is also calculated.

outoutininTurb hmhmW _0_0_0_0 ⋅−⋅= &&& (5.31)

40

This power is converted to horsepower as it is in the compressor. The polytropic

efficiency is an output parameter calculated using the same approach described in the

compressor section. The final equation is given below.

( )( )TurbinTurb

TurbinTurbinoutpolyTurb PRR

PRRss/1log

/1log

_

__0_0_ ⋅

⋅+−=η (5.32)

Burner

The Burner element is a constant volume burner. The entrance state is completely

defined; those parameters include: inT _0 , inP _0 , intotm _& , inFAR , inh _0 , and mass fractions

for all species. Also specified are the bη and burner inPP _0Δ . The exit stagnation

pressure, outP _0 , is found with the equation:

( )ininout PPPP _0_0_0 1 Δ−⋅= . (5.33)

outT _0 must be specified by the user in order to determine the incoming fuel flow

rate, fuelm& . In order to determine the exit state, the burner subroutine makes an initial

guess for the fuel flow rate, 1fuelm& , using a straightforward energy balance.

(5.34)

The model assumes a lower heating value, RQ , of 18400 Btu / lbm. It also assumes

a constant specific heat, pC , of 0.285 Btu / lbm-R. The inlet conditions and the first fuel

flow rate iteration, 1fuelm& , are then passed to CEA from the NPSS subroutine calcBurn.

CEA calculates the burner exit state point including: equilibrium composition and the

new burner exit temperature iteration, 1_0 outT . The burner exit conditions ( 1

_0 outT , outP _0 ,

⎟⎟⎠

⎞⎜⎜⎝

−=

out

inoutinairfuel T

TTmm

_0

_0_0_

1

285.0/18400&

41

outFAR , outWAR , outh _0 , and mass fractions for all species) are then passed back to

calcBurn where the burner efficiency is applied to determine the actual burner exit

temperature, actoutT _1

_0 .

( ) ininoutbactout TTTT _0_0

1_0

_1_0 +−=η (5.35)

Then, actoutT _1

_0 is used to determine the next fuel flow rate iteration, 2fuelm& , with the

energy balance described above (Equation 5.34). An error check is performed on the fuel

flow rate values every iteration to determine when the loop can be exited.

ToleranceErrormmm fuelfuelerrorfuel _12_ ≤−= &&& (5.36)

Once fuelm& is determined, the exit state point is completely defined.

Sensitivity Analysis

No formal optimization program was used for this project; instead, each engine

cycle model was roughly optimized manually starting from base case assumptions listed

in Chapter 4. The sensitivity analyses were performed on the SCGT and HPRTE

Efficiency models to determine the influences of particular design parameters. The H-V

Efficiency model was not included in these studies because the results would mirror those

for the HPRTE Efficiency model analysis. Two primary dependent parameters

investigated in the sensitivity analysis include thermal efficiency and specific power.

The thermal efficiency is defined as:

Rfuelth Qm

W⋅

=&

&η , (5.37)

where RQ is the lower heating value of the fuel and W& is the net power.

42

The specific power is defined as: inairm

WSP_&

&= . (5.38)

Influence coefficients are use to quantify the sensitivity of resultant parameters as

they relate to perturbed input parameters. The dimensional influence coefficient is

defined as

( )( )ParameterInput

Resultant∂

∂ . (5.39)

To relate the magnitudes of influence coefficients to one another, they must be non-

dimensionalized is required. This is accomplished by dividing the perturbed value by its

base case quantity:

( )

( )Value BaseParameter Input

ParameterInput Value BaseResultant

Resultant

∂. (5.40)

Such an example of a non-dimensional influence coefficient is given below. Here,

the HPC inlet temperature is perturbed from its base value and the resultant change to thη

is expressed in the following form. A value of 1 suggests that a 1% perturbation in HPC

inlet temperature results in a 1% change in thη . In this way the sensitivity of input

parameters is determined.

( )

base

baseth

th

TinHPCTinHPC

_)_(

_

∂η

η

(5.41)

43

CHAPTER 6 RESULTS AND DISCUSSION

The results and discussion of the analysis performed using the cycle code NPSS are

presented in Chapter 6. The first section in this chapter, Cycle Code Comparison,

compares results from the spreadsheet code (used by Boza [22]) and the NPSS program

for one operating point of the HPRTE Efficiency model. Next, sensitivity studies were

performed on the SCGT and HPRTE Efficiency cycles and influence coefficients were

calculated. Engine model results are given and compared to derived thermodynamic

expressions. Finally, plots and tables are presented that compare the performance

parameters of the three engine configurations.

Cycle Code Comparison

Before initiating the sensitivities studies, it is important to benchmark the NPSS

program and compare results of one model configuration to those results obtained from

running a proven cycle analysis program. One operating point for the HPRTE Efficiency

model was chosen for the comparison, and the results from the two cycle codes are

presented in Table 6-1. The third column in the table lists the absolute differences of the

two data sets parameters in percentages. Agreement of the data between the two codes is

high; values for thη , OPR, HPT exit temperature (TET), inHPCT _ , and exhaustT are all

within acceptable limits. The SpPw calculated by the spreadsheet model was 12.5%

higher than that calculated in NPSS. There are three possible reasons for the disparity in

the output values. First, it is impossible to implicitly balance the low pressure spool

44

specific work; therefore the turbine specific work is never properly matched to the

specific power of the compressor. This could very easily result in a different airm& . Two,

different fuels are used in the codes. The hydrogen-to-carbon ratio is 1.93 in NPSS and

2.03 in the spreadsheet code. Because the fuels are different, the curve-fit coefficients

used to calculate the enthalpies for the spreadsheet code could be different than the ones

used by NPSS.

Sensitivity Analysis

Simple Cycle Gas Turbine Engine Model

Of particular interest for this project is the sensitivity of the open-cycle engine

thermal efficiency and specific power to variations in turbine inlet temperature and

ambient temperature. Figures 6-2 through 6-4 show the results of the analysis. Unless

otherwise specified the following parameters remained constant throughout the analysis:

ambient temperature is 544°R, TIT is 2500°R, and nominal power output is 4000 BHP.

Figure 6-2 displays the thermal efficiency as a function of the overall cycle pressure ratio

(OPR). The TIT variations have a strong influence on the outcome of the thermal

efficiency value. Raising TIT has a positive effect on thη which also implies that the

total heat added to the system has been reduced since the power output remains steady.

Let thη be defined by the following relationship:

Rfuelth Qm

W⋅

=&

&η . (6.1)

Since, RQ , the fuel lower heating value, is constant; fuelm& has to decrease in order

to reduce the total heat added to the system in order to raise the thη . For all run cases in

the SCGT analysis, variations in thη are the direct result of variations in fuelm& .

45

To better understand the relationship between thη and the input parameters in this

sensitivity study, the following derivation has been included. The State numbers in the

equations correspond to those for the SCGT cycle schematic given in the Configurations

chapter (Figure 4-1). Figure 6-1 shows the comparison between the theoretical

expression below, the run data from NPSS, and ideal thermal efficiency expression for an

air standard Brayton cycle available in any thermodynamic text. The derived expression

predicts a curve with higher efficiency than the NPSS output; this is expected because the

pressure losses in the combustor are not accounted for, as it was assumed that the turbine

pressure ratio is equal to the compressor pressure ratio. That and the fact that the specific

heat ratio (γ ) was averaged for the entire cycle may explain the discrepancy between the

results from NPSS and the derived curve-fit expression.

There are several assumptions exercised in this derivation to develop the final

expression in terms of only the following parameters: 01

03

TT

, γ , PR , adTurb _η , and

adComp _η . They are as follows:

fuelm& is much less than airm&

( )Thh =

constant≅pC

ratio of specific heats, constant≅γ

PRPP

PP

=≅04

03

01

02

Beginning with the definition of thermal efficiency:

46

( ) ( )[ ]Rfuel

airp

Rfuelth Qm

TTTTmCQm

W⋅

−−−=

⋅=

&

&

&

& 01020403η . (6.2)

Factoring out 01T gives:

Rfuel

airp

th QmTT

TT

TT

TmC

⎥⎦

⎤⎢⎣

⎡+−−

=&

& 101

02

01

04

01

0301

η . (6.3)

Now the adiabatic turbomachinery efficiencies derived in the previous chapter may be

rewritten in the form:

⎟⎟⎠

⎞⎜⎜⎝

⎛−

⎟⎠

⎞⎜⎝

⎛ −=

1

1

01

02

1

_

TT

PR

adComp

γγ

η . (6.4)

⎟⎠

⎞⎜⎝

⎛ −

⎟⎟⎠

⎞⎜⎜⎝

⎛−

=−

1

1

104

03

γη

PR

TT

adTurb (6.5)

Solving the compressor Equation 6.4 for 01

02

TT

and Equation 6.5 for 04T yields the

following two expressions.

⎟⎠

⎞⎜⎝

⎛ −+=−

1111

_01

02 γγ

ηPR

TT

adComp

(6.6)

⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜

+⎟⎠

⎞⎜⎝

⎛ −=

11 _

103

04

adTurbPR

TT

ηγγ

(6.7)

Substituting Equations 6.6 and 6.7 into Equation 6.3 gives the following expression:

47

Rfuel

adCompadTurb

airp

th Qm

PRPRT

TTT

TmC

⎥⎥⎥⎥

⎢⎢⎢⎢

⎟⎠

⎞⎜⎝

⎛ −−

⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜

+⎟⎠

⎞⎜⎝

⎛ −−

=

&

& 11

11

1 1

__

101

03

01

0301

γγ

γγ ηη

η (6.8)

The denominator can be rewritten in terms of 01

03

TT

, γ , PR , and adComp _η .

The energy balance of the combustor is as follows:

( ) 030302 TCmTCmmmQTCm pairpfuelairfuelRpair &&&&& ≅+=+ . (6.9)

Now, solving for RfuelQm& gives:

( ) ⎟⎟⎠

⎞⎜⎜⎝

⎛−=−=

01

02

01

03010203 T

TTT

TCmTTCmQm pairpairRfuel &&& . (6.10)

Substituting in the expression for 01

02

TT in Equation 6.6 results in the following:

⎥⎥⎦

⎢⎢⎣

⎡−⎟

⎞⎜⎝

⎛ −−=−

111 1

_01

0301

γγ

ηPR

TT

TCmQmadComp

pairRfuel && . (6.11)

The final expression for thη is

⎥⎥⎦

⎢⎢⎣

⎡−⎟

⎞⎜⎝

⎛ −−

⎥⎥⎥⎥

⎢⎢⎢⎢

⎟⎠

⎞⎜⎝

⎛ −−

⎟⎟⎟⎟⎟

⎜⎜⎜⎜⎜

+⎟⎠

⎞⎜⎝

⎛ −−

=−

111

11

11

1

1

_01

03

1

__

101

03

01

03

γγ

γγ

γγ

η

ηηη

PRTT

PRPRT

TTT

adComp

adCompadTurb

th . (6.12)

As mentioned above, the ideal thermal efficiency for an air standard Brayton is also

plotted on Figure 6.1.

48

γγη 1_11 −−=

PRidealth (6.13)

Simple Cycle Gas Turbine Engine Model Sensitivity Analysis

Figure 6.2 is a plot of the thermal efficiency as a function of OPR with separate

curves for TITs. These curves peak at certain OPR values; continuing to increase

pressure ratio will continue to decrease the heat added per unit mass to the cycle,

however, the thermal efficiency will drop because the recuperator capacity to exchange

heat is being neutralized. When this happens the thermal efficiency begins to drop again.

The influence coefficients were determined for Figure 6.2. The base case engine

chosen had a TIT of 2500°R with an OPR of 24. These numbers imply that changes in

TIT effect greater change in thη than changes in OPR. One other point to note is the fact

that no matter what the TIT is for the engine model, the maximum thermal efficiency

always occurs when the compressor power is about twice as large as the net BHP.

( )898.0

)(_ =

base

baseth

th

TITTIT

ηη

( )0322.0

)(_ =

base

baseth

th

OPROPR

ηη

Figure 6-3 displays specific power sensitivity to changes in TIT over a pressure

range from 8 to 18. The drop in specific power for increasing OPRs can be explained as

follows. Consider the engine as a control volume that produces a constant power output.

Ignoring the small effects to thermal efficiency that increasing OPR produces, the heat

energy added to the engine must be constant. However, as OPR increases the heat rate

per unit mass added in the combustor drops. Therefore, more air must be brought into the

combustor to maintain the total heat energy input required to produce a constant power

engine output. The other trend visible has to do with increasing TIT and its influence on

49

specific power. Drawing two cycles, with different TITs, on the same T-S diagram

clearly demonstrates this phenomenon.

Influence coefficients were calculated using the base case parameters from the last

section. Specific power is more affected by changes in TIT than changes in OPR. In

fact, there is 1 order of magnitude difference between the two parameters. The negative

sign quantifies the drop in specific power with increasing OPR that was discussed earlier.

( )36.3

)(=

base

base

TITTIT

SpPwSpPw

( )339.0

)(−=

base

base

OPROPR

SpPwSpPw

Figure, 6-4 examines operating temperature variation as it influences cycle thermal

efficiency and OPR. The OPR curves are nearly linear and the slopes become more

negative as OPR increases. Notice that the thermal efficiencies approach 40% as ambient

temperature falls to 509°R. Cold day operation translates to good thermal efficiency for

the SCGT. Notice the effect of changing OPR for the case when ambientT is 572°R. The

higher OPR curves collapse on each other implying that thermal efficiency maximums

are at or near their peak values here, and any further boost in OPR drops the thermal

efficiency off the other side of the curve that would appear if this was a three dimensional

figure. For example, at ambientT 554°R, the maximum thermal efficiency corresponds to

an OPR of 24. A further boost to an OPR of 28 results in a reduced thermal efficiency.

The influence coefficients corresponding to Figure 6-4 for thermal efficiency and

specific power are calculated assuming the base case cycle where ambient temperature is

518°R and OPR is 24.

50

( )669.0

)(_ −=

base

baseth

th

TambTamb

ηη

( )0682.0

)(_ =

base

baseth

th

OPROPR

ηη

The thermal efficiency coefficients suggest that drops in ambient operating

temperature impact thermal efficiency more so than variations in pressure ratio. This

implies that OPR is a secondary issue if the engine is being designed with the intent to

optimize thermal efficiency.

( )76.1

)(−=

base

base

TambTamb

SpPwSpPw

( )2676.0

)(−=

base

base

OPROPR

SpPwSpPw

The specific power influence coefficients quantify the negative impact on specific

power when either ambient temperature or OPR is increased. Of course, the effect of

ambient temperature is much more significant—almost an order of magnitude greater.

High Pressure Regenerative Turbine Engine Efficiency Model

Before beginning the sensitivity analysis of the HPRTE Efficiency, it is important

to examine the validity of the data output from NPSS. To accomplish this task an

expression has been derived to test the validity of NPSS output. After the derivation is

complete, the resulting expression is normalized; and it is only a function of the

following parameters: 01

2.07

TT

, LPPR , γ , adLPC _η , and adLPT _η . The plotted NPSS data

should agree with the derived expression. The ten points in Figure 6-5 represent ten

distinct converged operating points from NPSS runs. The standard deviation of the set of

points is 0.00398, signifying close conformity with the derived thermodynamic

expression. The development uses the following assumptions:

51

fuelm& << airm&

( )Thh =

constant≅pC

Ratio of specific heats, constant≅γ

LPPRPP

PP

=≅08

2.07

01

02

The development begins with the adiabatic efficiency expressions for the low pressure

compressor and turbine:

⎟⎟⎠

⎞⎜⎜⎝

⎛−

⎟⎠

⎞⎜⎝

⎛ −=

1

1

01

02

1

_

TT

LPPR

adLPC

γγ

η (6.14)

⎟⎠

⎞⎜⎝

⎛ −

⎟⎟⎠

⎞⎜⎜⎝

⎛−

=−

1

1

108

2.07

γη

LPPR

TT

adLPT . (6.15)

Solving Equation 6.14 for 01

02T

T yields

⎟⎠

⎞⎜⎝

⎛ −⎟⎟⎠

⎞⎜⎜⎝

⎛=−

1111

_01

02 γγ

ηLPPR

TT

adLPC

. (6.16)

Now, introducing the power balance for the turbocharger gives

( ) ( )082.070102 )( TTCmmTTCm pfuelinpin −+=− &&& . (6.17)

And simplifying Equation 6.17 using the assumptions provided above results in

( ) ( )082.070102 TTTT −=− . (6.18)

52

Rearranging Equation 6.18 and solving for 101

02 −TT produces

⎟⎟⎠

⎞⎜⎜⎝

⎛−=−

01

08

01

2.07

01

02 1TT

TT

TT

. (6.19)

Setting Equation 6.16 equal to Equation 6.19 gives

⎟⎟⎠

⎞⎜⎜⎝

⎛−=⎟

⎞⎜⎝

⎛ −⎟⎟⎠

⎞⎜⎜⎝

⎛ −

01

08

01

2.071

_

11TT

TT

LPPRadLPC

γγ

η. (6.20)

Solving Equation 6.20 for 08T produces

⎥⎥⎦

⎢⎢⎣

⎡⎟⎠

⎞⎜⎝

⎛ −⎟⎟⎠

⎞⎜⎜⎝

⎛−⎟⎟

⎞⎜⎜⎝

⎛=

11 1

_01

2.070108

γγ

ηLPPR

TT

TTadLPC

. (6.21)

Next, rearranging Equation 6.15 and solving for 08T yields

111

_

2.0708

+⎟⎠

⎞⎜⎝

⎛ −=

−γ

γ

η LPPR

TT

adLPT

. (6.22)

Setting Equation 6.21 equal to Equation 6.22 results in the following expression:

⎥⎥⎦

⎢⎢⎣

⎡⎟⎠

⎞⎜⎝

⎛ −⎟⎟⎠

⎞⎜⎜⎝

⎛−⎟⎟

⎞⎜⎜⎝

⎛=

+⎟⎠

⎞⎜⎝

⎛ −

−11

11

1

_01

2.07011

_

2.07 γγ

γγ ηη

LPPRTT

TLPPR

T

adLPCadLPT

. (6.23)

Rearranging,

⎥⎦

⎤⎢⎣

⎡+⎟

⎞⎜⎝

⎛ −⋅⎥⎥⎦

⎢⎢⎣

⎡⎟⎠

⎞⎜⎝

⎛ −⎟⎟⎠

⎞⎜⎜⎝

⎛−⎟⎟

⎞⎜⎜⎝

⎛=

−−

1111 1

_

1

_01

2.07

01

2.07 γγ

γγ

ηη

LPPRLPPRTT

TT

adLPTadLPC

. (6.24)

Then dividing both sides by 01

2.07

TT

yields the final expression plotted in Figure 6-5.

53

⎟⎟⎠

⎞⎜⎜⎝

⎥⎦

⎤⎢⎣

⎡+⎟

⎞⎜⎝

⎛ −⋅⎥⎥⎦

⎢⎢⎣

⎡⎟⎠

⎞⎜⎝

⎛ −⎟⎟⎠

⎞⎜⎜⎝

⎛−⎟⎟

⎞⎜⎜⎝

=

−−

01

2.07

1

_

1

_01

2.07 1111

1

TT

LPPRLPPRTT

adLPTadLPC

γγ

γγ

ηη

. 6.25

High Pressure Regenerative Turbine Engine Efficiency Model Sensitivity Analysis

Figure 6-6 shows low pressure spool pressure ratio (LPPR) variation influences on

the high pressure compressor pressure ratio (HPPR) as ambient temperature varies. For

these run cases the TIT and the nominal shaft power output were held constant at 2500°R

and 4000 BHP, respectively. There are three important features represented by Figure 6-

6. First, there is the interaction between LPPR and HPPR. Monotonically increasing

LPPR results in a similarly increasing HPPR. Examining the raw data shows that that

mass flow rate drops with increasing LPPR; this drop in mass flow rate requires a greater

expansion on the high pressure turbine to produce the nominal power output. The second

trend to notice is that raising ambient temperature results in raising HPPR. Raising

ambient temperature decreases air density which in turn causes a decrease in mass flow

rate. With less mass flow rate to the core components, the heat added per unit mass to the

combustor must be increased in order to maintain the constant power output. The way to

accomplish this task with an HPRTE engine is to increase the HPPR. Increasing HPPR

drops the hot side recuperator inlet temperature and as a result the combustor inlet

temperature drops, too. Thermal efficiency increases with increasing HPPR until HPPR

is about 5.1. Then, any further increasing of HPPR results in a drop in thermal

efficiency.

Figure 6-6 can be used to find the operating lines for a properly designed HPRTE.

An HPRTE with waste-gating capabilities would have operating curves of constant

54

HPPR; therefore, the horizontal grid lines on the Figure 6-6 could be called operating

curves as well. Moreover, the line corresponding to a HPPR of 5.1 would represent the

highest thermal efficiency operating curve.

The influence coefficients for Figure 6-6 are listed below. The base case has an ambient

temperature of 528°R and a LPPR of 6.0. ambientT and LPPR variations both have

significant resultant effects on HPPR. Operationally speaking, ambientT is related to airm&

which can cause large changes to the specific power of the system.

( )63.1

)(=

base

base

LPPRLPPR

HPPRHPPR

( )

15.6)(

=∂

base

base

TambTamb

HPPRHPPR

Figure 6-7 shows cycle thermal efficiency sensitivity to TIT variation for a range of

HPPRs. The analysis holds R, ambient temperature, output shaft horse power, and all

component efficiencies constant. NPSS convergence is difficult to achieve if HPPR is

changed manually by the user; instead, to effect change in HPPR, the LPPR is controlled

by the modeler. As LPPR was increased, the model solver reduced fresh air flow rate to

the engine. The mass flow rate reduction caused the HPPR to rise for the same reason

discussed in the previous section. Figure 6-6 shows that the thermal efficiency peaks

very close to an HPPR of 5.1. Along a curve, TET and the stoichiometry change because

R is held constant.

The influence coefficients for thη and SpPw were produced by making

perturbations around the base run case where TIT was 2500°R and HPPR was 5.12.

( )847.0

)(_ =

base

baseth

th

TITTIT

ηη

( )0106.0

)(_ −=

base

baseth

th

HPPRHPPR

ηη

55

The interpretation of these influence coefficients suggests that changing TIT by 1%

causes a 0.847% change in thη . The second coefficient was calculated with data taken

from the right side of the plot where HPPR is monotonically increasing and thη continues

to decrease. The small coefficient value suggests that a large change in HPPR has only

limited effect on thη . This is expected when the influence coefficient is calculated near

an optimum thη point on the curve. Below, notice that TIT perturbations cause

significant resultant changes to the value of specific power.

( )59.2

)(=

base

base

TITTIT

SpPwSpPw

( )

318.0)( =∂

base

base

HPPRHPPR

SpPwSpPw

Figure, 6-8 shows the importance with respect to thermal efficiency of reducing the

HPC inlet temperature for HPRTE engine cycles. Sensitivity to circBPRRe is shown and

its value is varied from 2.0 to 3.5. Those parameters held constant for the analysis are as

follows: output BHP, ambientT , TIT, LPPR, all turbomachinery efficiencies, and the sea

water temperature. The HPC inlet temperature was varied by changing the effectiveness

of the main cooler. The trend toward higher thermal efficiencies for lower HPC inlet

temperatures is the same phenomenon seen in Figure 6-4. A lower inlet temperature to

the high pressure core results in an increase in predicted thermal efficiency. Examination

of the model data used to produce Figure 6-8 shows that as HPC inlet temperature drops,

HPPR increases. As a result the temperature change across the HPT increases and the

mass flow requirements drop. This in turn means that the fuel flow requirement is less.

As shown by Equation 6.1, the drop in fuel flow directly affects thermal efficiency since

56

power output is constant. Lower circBPRRe helps thermal efficiency because TIT must be

maintained at 2500°R. The more inert combustion products that are mixed with fresh air

act to drive down TIT making it necessary to burn more fuel to keep TIT constant.

However, higher recirculation boosts specific power as less fresh air is required to

produce the desired power output. The ultimate choice is the designer’s—if weight and

compactness are important, circBPRRe would be maximized. However, on a naval vessel,

weight might not be the paramount consideration.

The influence coefficients for thη and SpPw are calculated using the base case

circBPRRe of 3 and a HPC inlet temperature of 622°R.

( )58.1

_)_(

_ −=∂

base

baseth

th

TinHPCTinHPC

ηη

( )0610.0

)(_ −=

base

baseth

th

RRη

η

( )66.3

_)_(

−=∂

base

base

TinHPCTinHPC

SpPwSpPw

( )624.0

)(=

base

base

RR

SpPwSpPw

The influence coefficients based on variations to circBPRRe help to quantify the

effects on thermal efficiency and specific power that were discussed above. Moreover,

reducing the HPC inlet temperature boosts both thermal efficiency and specific power.

This is an observed behavior in recuperated gas turbine engines.

Figure 6-9 displays the thermal efficiency sensitivity to pressure drops in the

coolers. To obtain the curves below, HPT exit temperature variations were created by the

modeler. This was done in the same manner as it was for the Figure 6-7. LPPR was

manually controlled to effect change in HPC pressure ratio which caused the HPT exit

temperature to change.

57

The general trends are consistent with the results of Figure 6-7. The curve of

Figure 6-7 corresponding to that of a TIT of 2500°R is the same as the curve in Figure 6-

9 for a cooler PPΔ of 3%. There the HPT exit temperature is 1884°R. The propensity

to see a lowered thermal efficiency when the cooler PPΔ is raised has a straightforward

explanation. Increasing cooler PPΔ results in a drop in the OPR of the cycle. Since the

net output power remains constant, mass flow must be increased. This in turn causes a

rise in fuel flow rate and a drop in thermal efficiency.

Influence coefficients for cooler PPΔ and HPT exit temperature are listed below

for the base cycle case where cooler PPΔ is 3% and HPT exit temperature is 1884°R.

Specific power influence coefficients are included. The results show that neither cooler

PPΔ nor HPT exit temperature affect significant change in thermal efficiency.

However, the specific power is positively influenced by decreasing HPT exit

temperature.

( )0258.0

/)/(

_ −=

ΔΔ∂

base

baseth

th

PPPPη

η

( )0617.0

_)_(

_ =∂

base

baseth

th

TexitHPTTexitHPT

ηη

( )0407.0

/)/(

−=

ΔΔ∂

base

base

PPPP

SpPwSpPw

( )86.1

_)_(

−=∂

base

base

TexitHPTTexitHPT

SpPwSpPw

Figure 6-10 considers specific power sensitivity to HPC adiabatic efficiency. For

an engine with constant power output and TIT, raising TET also results in core mass flow

increasing. That is because a higher low pressure recuperator inlet temperature drives the

combustor inlet temperature up. The effect of driving that temperature up is similar to

what happens for SCGT when OPR is increased. While the total heat added to the engine

58

may be remain almost constant, the heat added per unit mass drops off as combustor inlet

temperature rises. Consequently, the mass flow coming into the combustor must go up.

The result is a drop in specific power. The other noteworthy trend here is the positive

effect on specific power that comes from raising HPC adiabatic efficiency. Raising HPC

adiabatic efficiency decreases the power requirement of the HPC thereby increasing the

specific power of the cycle.

Influence coefficients are considered for a base engine with HPC adiabatic

efficiency of 86 % and HPT exit temperature of 1883°R. Both thermal efficiency and

SpPw influence coefficient are given. Compare these results with the influence

coefficients for the cooler PPΔ discussed above. The adComp _η has a stronger effect on

the performance of the cycle than does PPΔ .

( )06.1

)(_

_

_ =∂

baseadComp

adComp

baseth

th

ηη

ηη

( )74.1

)(_

_=

baseadComp

adComp

baseSpPwSpPw

ηη

Figure 6-11 is a sister plot to Figure 6-10. It describes the same specific power

trends, however, now they are in terms of HPC pressure ratio (HPPR). HPT adiabatic

efficiency ( adTurb _η ) is varied to show specific power sensitivity to this parameter. The

expectation for high adTurb _η to result in an improved specific power is met. From a

thermodynamic standpoint, higher adTurb _η means a greater temperature drop across the

turbine for the same HPPR. Therefore, less mass flow is required to produce the same

power—resulting in a boosted specific power.

59

Influence coefficients for SpPw and thη were calculated from the base cycle case

where adTurb _η was 87% and HPPR was roughly 5.1. Note: since HPPR is an output, it

can not be explicitly set. The coefficient values imply significant influence of adTurb _η on

thη and SpPw . Similarly, notice that the coefficients calculated based on HPPR

variation are very close to those same coefficients calculated for Figure 6-7 when TIT

was the sensitivity parameter. In fact, there is less than a 1% difference between the

values.

( )0105.0

)(_ −=

base

baseth

th

HPPRHPPR

ηη

( )46.1

)(_

_

_ =∂

baseadTurb

adTurb

baseth

th

ηη

ηη

( )316.0

)(=

base

base

HPPRHPPR

SpPwSpPw

( )36.2

)(_

_=

baseadTurb

adTurb

baseSpPwSpPw

ηη

Figure 6-12 displays exhaust gas temperature sensitivity to OPR with TIT

variations included. Cycle constants included: output BHP, LPPR, ambient temperature,

R, and all component efficiencies except for MCG effectiveness. Since LPPR was held

constant, MCG effectiveness was varied to effect HPPR change. The results indicate that

exhaust gas temperature is not sensitive to either OPR or TIT variation. This should be

expected in an inter-cooled recuperated system. The heat exchangers act to damp exhaust

temperature variations that may result from parametric tweaking. The influence

coefficients further illustrate this phenomenon. The base case cycle had an OPR of 24

and a TIT of 2500°R. The computed values being small and negative imply that

60

significant manipulation of TIT or OPR is necessary before any change in exhaust

temperature is noticed.

( )0683.0

)(−=

base

base

TITTIT

TexhaustTexhaust

( )

0201.0)(

−=∂

base

base

OPROPR

TexhaustTexhaust

Figure 6-13 examines the Turboη impact on thη as it relates to HPPR. Here, the

parameter driving convergence will be LPPR. This means that LPPR is controlled by the

modeler and both, LPPR and HPPR, will be varying. Most noteworthy about Figure 6-13

is that Turboη alone does not significantly affect thη . The influence coefficient affirms this

assertion. Figure 6-13 likewise gives a clear indication that an HPPR of 5.1 produces the

maximum thη —this optimum HPPR value agrees with the earlier optimum predicted in

the plot of thermal efficiency verses HPPR with TIT sensitivity.

( )0228.0

)(_ −=

baseTurbo

Turbo

baseth

th

ηη

ηη

Figure 6-14 is the last figure of the sensitivity analysis for the HPRTE Efficiency

cycle, and it examines LPPR variations and their effect on thermal efficiency. Thermal

efficiency sensitivity to TIT perturbations has previously been analyzed for Figure 6-7.

However, this plot is useful in that it gives optimum LPPRs for particular TITs.

Particularly interesting is the curve for a TIT of 2500°R since this is the base and design

optimum TIT. A quick regard of the curve reveals that the best LPPR is about 6.2. That

value is high but within the limits of single stage centrifugal compressor technology.

Taking a look at the influence coefficient suggests that perturbing LPPR does not

61

significantly cause change to the resultant, thη . Again, this is most likely because the

influence coefficient was evaluated near a design optimum and the curve was flat.

( )0645.0

)(_ =

base

baseth

th

LPPRLPPR

ηη

Table 6-2 summarizes the results of the sensitivity analysis. The perturbed

parameters are listed in the left hand column and a value between 1 and 3 (1 being the

most important) was assigned to indicate the degree of importance that a particular

parameter had on the resultant in the top row.

Below are the definitions for the values assigned to the sensitivity parameters listed

in summary (Table 6-2).

• 1: 1% fluctuation from parameter produces >1% fluctuation in resultant) • 2: 1% fluctuation of parameter produces between 0.1%<resultant change<1% • 3: 1% change of parameter causes <0.1% change in resultant

Cycle Comparison Analysis

Three engine configurations are included in the comparison analysis: the SCGT,

the HPRTE Efficiency, and the third engine configuration, H-V Efficiency. The rough

optimization of the SCGT and the HPRTE Efficiency cycles has been completed in the

sensitivity analysis, and now those optimized results will be compared to the predicted

performance results from H-V Efficiency analysis. Not explicitly shown in this analysis

was the iterative process used to determine the best low and high spool pressure ratios for

the H-V Efficiency cycle. For Figures 6.16 through 6.18, LPPR was held at 3.5. This

value proved to be the best LPPR value for the largest number of H-V Efficiency data

runs. As was the case with HPRTE Efficiency, LPPR was controlled by the modeler and

HPPR was controlled by the solution solver. Since heat signature is a serious design

62

consideration for military vessels, an exhaust gas temperature comparison will be

included and discussed. Next, performance comparisons will be made considering

different air and sea water temperatures. The last table will list the performance details

for the optimized engine configurations side by side.

Figure 6-15 provides a clear indication of the thermodynamic advantage that the H-

V Efficiency cycle enjoys over the other two cycles in the comparison. The HPC inlet

temperature of the H-V Efficiency engine was maintained at 509°R by the refrigeration

system. Contrast that against the HPC inlet temperatures of the HPRTE Efficiency

engine. The results reveal that the H-V Efficiency HPC inlet temperature was between

99-107°R below that of the HPRTE Efficiency.

Table 6-3 highlights key features of Figure 6-15. Considering the results, it is easy

to make a case for the H-V Efficiency engine configuration. It has the highest max_thη by

17.1% over the SCGT and 17.2% over the standard HPRTE without refrigerator. Its

meanth _η is 44.2% which indicates that the curve is very flat. This implies that the H-V

Efficiency design point is not significantly sensitive to the OPR choice; therefore,

existing, off the shelf turbomachinery components can be used to save on capital

investment costs. Moreover, the meanth _η for HPRTE Efficiency is 3.08% higher than that

of the SCGT configuration, indicating that the HPRTE Efficiency configuration is also

less sensitive to OPR choice than SCGT. The second data point for SCGT on Table 6-3

is the case where the engine is designed for optimum specific power (SCGT SpPw cycle

mode). Here, the thη is only 33.4%—a full 10.2% less than the maximum design thη for

the HPRTE Efficiency cycle.

63

Figure 6-16 compares SpPw performance characteristics for the three cycle

configurations. Here it can be shown that a sizable advantage is enjoyed by the HPRTE

cycles over the SCGT. Since the three cycles have the same output BHP requirements,

this implies that the HPRTE Efficiency and H-V Efficiency engine cycles operate at

much reduced airm& levels. This disparity can be attributed to the three main differences

between the HPRTE and SCGT cycles: exhaust gas recirculation, inter-cooling of the

compressors, and recuperative heating before the combustor.

Closer examination of the two HPRTE engine configurations reveals that

compressor inter-cooling boosts specific power significantly by itself, especially at lower

OPRs. This can be shown from an examination of the run data for an OPR of 14.5. The

H-V Efficiency cycle has a calculated specific power of 560 lbmHP sec⋅ (units are

industry standard) compared to the HPRTE Efficiency specific power of 458

lbmHP sec⋅ for the same OPR. That computes to a specific power increase of 22.2%

for the H-V Efficiency over the standard HPRTE Efficiency. Table 6-4 further illustrates

that gap in specific power between the two HPRTE engine cycles and the standard

SCGT.

Now notice the divergence of the SCGT curve from the HPRTE curve. The SCGT

mass flow rate requirement increases as OPR increases in order to maintain constant

power BHP output. This is because even though the heat added per unit mass drops the

total heat input must remain nearly unchanged to produce the same power output.

Meanwhile, increasing the OPR for either HPRTE cycle improves specific power because

the recuperator has less available heat to drive up the combustor inlet temperature.

Reviewing the analysis for Figures 6-10 and 6-11 may provide additional clarity.

64

Figure 6-17 shows the operating exhaust gas temperatures for various OPR design

points for the three engine configurations. There are three features to the plot that are

worth noting. First, as expected, the SCGT exhaust gas temperature drops in a weak

exponential manner as the design OPR increases. The ETF40B engine was designed to

optimized specific power, and it had an OPR of 10.4. Assuming the SCGT model is a

good representation of the ETF-40B engine, it implies that its exhaust gas temperature is

about 1580°R. The second point brought to light by Figure 6-17 is the fact that the

HPRTE cycles have constant exhaust temperatures for a broad range of design OPRs.

Third, the HPRTE Efficiency cycle has lower exhaust gas temperatures than the

combined cycle H-V Efficiency. The reason for this is that for the same OPR, the H-V

Efficiency has a lower LPPR than the HPRTE Efficiency, and less expansion across the

LPT means the exhaust gas temperature will be higher. Following that logic suggests

that the presence of a VARS unit actually raises the exhaust gas temperature slightly.

Table 6-5 is a list of the maximum exhaust gas temperature cases for each cycle

with their corresponding OPR values from Figure 6-17. The mean exhaust temperatures

are also given to indicate the flatness of the plotted curves for the HPRTE cycles.

Figure 6-18 compares thη values of the three engine configurations under different

ambientT operating conditions. While LPPRs were held constant, OPRs could not be held

constant because of NPSS operational limitations on the HPRTE models discussed above

for Figure 6-7. The parabolic shape of the HPRTE engine curves resembles the thermal

efficiency plots produced in the sensitivity analysis for the HPRTE Efficiency model.

Similarly, notice the linear trend of the SCGT as ambient temperature drops. Eventually,

the H-V Efficiency and SCGT curves will intersect at an ambient temperature of 491°R

65

(32 ° F). The most likely operational scenario for the H-V Efficiency will be in a desert

environment where the environment temperatures are above 544°R (85°F). The engines

will also be performance rated at or above 518°R (59°F). At 518°R the H-V Efficiency

thη is 14.0% higher than the SCGTs thη . Moreover, above 547°R (88°F), the thη of the

HPRTE Efficiency surpasses that of the SCGT. Above 547°R meanth _η for the HPRTE

Efficiency engine is 36.9% compared to 36.1% for the SCGT.

Extreme Operating Conditions

Four extreme operating cases were chosen for this examination of how the engine

configurations would perform in the severest of environments. Ambient temperature and

sea water temperature were the dependent inputs. Two extreme cases were chosen for

each dependent input resulting in a total of four operating cases—cold day/cold water,

cold day/warm water, hot day/cold water, and hot day/warm water. NPSS limited the

ability to compare the engine cycles with constant OPRs. While the LPPRs for the

engine configurations were chosen from approximated design points maximizing thermal

efficiency from the analysis in the last section, it is necessary to let mass flow rates and

HPPRs float to obtain convergence from the solution solver. The OPR for the SCGT

remained constant at 24. OPRs are listed in the Table 6-6. Notice that for the 569°R day,

the OPRs for the HPRTE engine cycles are very high. High temperature days decrease

the air density and NPSS drops the mass flow rates as a result. This in turn requires the

HPRTE cycles to increase their pressure ratios to maintain constant power output.

Case 1 is the cold day with warm sea water condition. These conditions loosely

represent night operations in a dry desert climate. A key comparison for this case

includes the thη values for the H-V Efficiency and SCGT cycles. Table 6-6 shows that

66

the thη for the SCGT model is higher than that of the H-V Efficiency cycle. This

operating point represents a case described during the discussion of Figure 6-18 when

ambient temperature drops to the point where the thη values of H-V Efficiency and

SCGT converge. That analysis showed that if the ambient temperature curves were

extended down to 489°R, the curves of the SCGT and H-V Efficiency would eventually

meet. Another key observation from case 1 data is the wide bridge between the specific

power values for the HPRTE cycles in comparison to the SCGT. The HPRTE Efficiency

configuration enjoys a 104% improvement in specific power over the SCGT. This theme

runs through the whole analysis, and as the operating conditions get warmer the disparity

becomes more pronounced.

Case 2 is the cold day/cold water temperature condition. In a North Atlantic

mission, conditions similar to these might exist. Cold water operating points improve the

thη and specific power of the HPRTE engine cycles by improving compressor inter-

cooling. The most significant example is in the change of the thη of the HPRTE

Efficiency engine between case 1 and 2. It increases by 13.2% and achieves its

maximum value for any of the four operating points examined. From a specific power

standpoint, the HPRTE Efficiency cycle bests the other two configurations in case 2—

beating the H-V Efficiency by 17.3% and the SCGT by 168%. Unaffected by the drop in

water temperature, the thη of the SCGT continues to hold steady at a sporty 40.1%. The

argument for choosing the H-V Efficiency cycle strengthens under case 2 operating

conditions because it displays the highest thη of the three configurations at 40.9%.

Case 3 is the hot day/hot sea water temperature condition. This operating point is

characteristic of a desert day scenario—the most likely mission conditions for the ETF-

67

40B and its replacement. Hot day design points drive up predicted specific power

performance values for the HPRTE engine models. The cause of this is described in the

introduction to this section.

From a thermodynamic standpoint the H-V Efficiency outperforms the other two

configurations. Here, the thη of the H-V Efficiency cycle is 20.4% better than H-V

Efficiency and 23.5% better than SCGT. The specific power of the H-V Efficiency cycle

is 12.1% higher than HPRTE Efficiency and 358% higher than the SCGTs. The rise in

thη from case 1 to 3 for the H-V Efficiency is directly indicative of the ambient

temperature rise which caused a noteworthy rise in OPR from 10.2 to 30.7. The thη rise

constitutes an 11.3% jump between the case points 1 and 3.

Case 4 is an unlikely operating point—when ambient temperature is high and water

temperature is very cold. A summer day in the North Atlantic is the closest example of

this condition. Here the thη of the H-V Efficiency cycle bests the HPRTE Efficiency by

15.3% and the SCGT by 24.1%. If the engines were design based on specific power

alone, the HPRTE Efficiency is a considerable threat to the H-V Efficiency engine mode.

Its specific power of 675 HP-sec/lbm, respectively, is the highest of the three engines for

case 4. However, this is an improbable engine design with an OPR of 55.4.

High Pressure Compressor Inlet Temperature Comparison for H-V Efficiency Model

Up until this point in the analysis the HPC inlet temperature for the H-V Efficiency

cycle has been held at a conservative 509°R. This section compares the same engine at

two different HPC inlet temperatures, 509°R and 499°R, respectively (Table 6-7). In

other words, careful consideration was taken to ensure that the HPPRs for both model

68

cases were the same. This was a time intensive consideration that could not be used for

other parts of the comparison studies. Other parameters held constant include: nominal

output BHP, TIT, ambientT , all turbomachinery η values, and R.

The results of the analysis conclude that there are performance improvements that

result from decreasing HPC inlet temperature but they are not stunning. The thη

increases by 1.56% and the specific power increases by 1.62% for the lower HPC inlet

temperature case. LPPR is also 5.14% lower when HPC inlet temperature is 499°R.

Essentially, from a computer model handling standpoint, the lower HPC inlet temperature

is achieved by lowering the LPPR slightly. This causes OPR to be 5.45% less, as well.

Additionally, there is very little influence on exhaustT which drops only 2°R when HPC

inlet is lowered 10°R. In summary, the performance parameters are positively affected

by decreasing HPC inlet temperatures, but it is unclear whether or not the difference is

significant enough to warrant implication. Moreover, the refrigeration capacity used to

cool the HPC inlet could be used elsewhere in applications not examined in this analysis.

Final Design Point Parameter Comparison

While the extreme operating conditions section provides key insight into off-design

point performance of the three engine cycles, it is necessary to compare the cycles with

their optimized design parameters at the most likely operating condition. Two versions

of the SCGT cycle are compared—the SCGT Efficiency is the open cycle optimized for

maximum thermal efficiency and the SCGT SpPw is the open cycle optimized for

maximum specific power. For this analysis the ambient temperature and the sea water

temperature were both set to 544°R. The engine output requirement was unchanged, at

4000 BHP. TIT was held to a maximum of 2500°R. TET was restricted below 2059°R.

69

The turbocharger pressure ratio was limited to 7.5. Table 6-8 results are consistent with

the comparative analysis plots.

• From a thermodynamic stance, the H-V Efficiency engine cycle has a higher design point thermal efficiency besting the SCGT Efficiency by 20.6%, the SCGT SpPw cycle version by 34.7%, and the HPRTE Efficiency engine cycle by 21.0%. The SCGT Efficiency thermal efficiency has an expected thermal efficiency that is 10.2% higher than the SCGT SpPw engine configuration. When comparing the specific power results of the HPRTE cycles the performance gap isn’t quite as large. For that parameter, the H-V Efficiency has a predicted specific power that is only 6.08% better than the HPRTE Efficiency. The SCGT SpPw cycle has a predicted specific power that is 13.2% greater than the SCGT Efficiency configuration.

• TET for the H-V Efficiency is 110°R less than the HPRTE Efficiency providing some leeway in material selection.

• Equivalence ratios are all within the reasonable limit (0.9 was maximum allowable). As the equivalence ratio value approaches a value of 1, the oxygen concentration in the gas is being limited. Limiting the excess oxygen helps to reduce the soot and harmful emissions production.

• Both HPRTE engine cycles have R values of 3 or above. This directly affects equivalence ratio. Increasing R limits the fresh air dilution into the combustion chamber—the net effect is an increase in the equivalence ratio value.

• The extra cooling capacity of the H-V Efficiency cycle causes water vapor from the combustion products to condense to liquid. That mass flow rate has been included in the table, as well. It is also convenient to relate that value to the mass flow rate of fuel used by the engine. The mass basis ratio of fuel burned to water extracted from the flow path was 1.13. Note: for simplicity, the ambient air was considered dry with a %RH of 0.

• Exhaust gas temperatures for both HPRTE cycles are approximately 500°R less than the exhaust temperature of the SCGT Efficiency configuration and nearly 800°R less than the exhaust temperature of the SCGT SpPw configuration . Naval forces are concerned with IR signatures produced by engine exhaust coming from their ships, the 800°R difference represents a significant stealth advantage over the SCGT. Moreover, reduced exhaust temperatures suggests that air density is higher; and as a result, the exhaust duct size can be smaller.

70

Table 6-1 Cycle codes comparison: NPSS verses spreadsheet code for HPRTE Efficiency model data run. All temperatures are in °R.

SpreadSheet HPRTE Eff. NPSS HPRTE Eff. ABS % Difference

37.0% 37.2% 0.54

(HP-sec/lbm) 519 593 12.5OPR 32.0 32.2 0.621TET 1880 1880 0.00R 3.30 3.30 0.00Equivalence Ratio 0.827 0.894 7.51

544 544 0.00LPPR 6.25 6.25 0.00HPPR 5.29 5.31 0.377

632 614 2.93TIT 2500 2500 0.00

790 801 1.37All Temperatures are in °R

thη

SpPw

ambientT

inHPCT _

exhaustT

0.35

0.4

0.45

0.5

0.55

0.6

0.65

15 20 25 30 35

OPR

Ther

mal

Effi

cien

cy

Ideal Eta

Derived Eta

NPSS

Figure 6-1 Thermal efficiency comparison is plotted with respect to OPR. NPSS results (with turbine inlet temperature (TIT) set to 2500°R) are compared to the derived and the ideal Brayton cycle expressions.

71

0.3

0.32

0.34

0.36

0.38

7 10 13 16 19 22 25 28 31 34 37

OPR

Ther

mal

Effi

cien

cy

TIT = 2500°RTIT = 2400°RTIT = 2300°RTIT = 2200°R

SCGT

Figure 6-2 Thermal efficiency vs. OPR with sensitivity to TIT

100

120

140

160

180

200

6 8 10 12 14 16 18 20

OPR

Spec

ific

Pow

er (H

P-se

c/lb

m)

TIT = 2500°RTIT = 2400°RTIT = 2300°RTIT = 2200°R

SCGT

Figure 6-3 Specific power vs. OPR with TIT sensitivity

72

0.33

0.34

0.35

0.36

0.37

0.38

0.39

0.4

500 520 540 560 580Ambient Temperature (°R)

Ther

mal

Effi

cien

cy

OPR = 28OPR = 24OPR = 20OPR = 16OPR = 12

SCGT

Figure 6-4 Thermal efficiency vs. ambient temperature with OPR sensitivity

0.90

0.95

1.00

1.05

1.10

4 5 6 7

LPPR

RH

S of

Equ

atio

n 6.

25

Model Data Points

HPRTE Efficiency

Figure 6-5 Demonstrates agreement between NPSS and developed theory that describes the low pressure spool

73

0

1

2

3

4

5

6

7

8

500 520 540 560 580

Ambient Temperature (°R)

HPP

R

LPPR = 6.4LPPR = 6.0LPPR = 5.5LPPR = 5.0LPPR = 4.5

HPRTE Efficiency

Figure 6-6 High pressure spool pressure ratio (HPPR) vs. ambient temperature with low pressure spool pressure ratio (LPPR) sensitivity

0.33

0.335

0.34

0.345

0.35

0.355

0.36

0.365

0.37

0.375

3 3.5 4 4.5 5 5.5 6 6.5 7 7.5 8

HPPR

Ther

mal

Effi

cien

cy TIT = 2500°RTIT = 2450°RTIT = 2400°RTIT = 2350°R

HPRTE Efficiency

Figure 6-7 Thermal efficiency vs. HPPR showing sensitivity to TIT

74

0.26

0.28

0.3

0.32

0.34

0.36

0.38

600 650 700

HPC Inlet Temperature (°R )

Ther

mal

Effi

cien

cy RecircRatio = 2.0RecircRatio = 2.5RecircRatio = 3.0RecircRatio = 3.5

HPRTE Efficiency

Figure 6-8 Thermal efficiency vs. HPC inlet temperature for recirculation ratio sensitivity

0.35

0.355

0.36

0.365

0.37

0.375

1750 1825 1900 1975 2050

TET (°R)

Ther

mal

Effi

cien

cy Cooler dP/P = 3%Cooler dP/P = 4%Cooler dP/P = 5%Cooler dP/P = 6%

HPRTE Efficiency

Figure 6-9 Thermal efficiency vs. turbine exit temperature (TET) with cooler pressure drop sensitivity

75

380

420

460

500

540

580

620

1750 1825 1900 1975 2050

TET (°R)

Spec

ific

Pow

er (H

P-se

c/lb

m)

HPC_eff = 0.86HPC_eff = 0.85HPC_eff = 0.84HPC_eff = 0.83

HPRTE Efficiency

Figure 6-10 Specific power vs. TET for HPC efficiency sensitivity

425

475

525

575

625

3 4 5 6 7 8

HPPR

Spec

ific

Pow

er (H

P-se

c/lb

m)

HPT_eff = 0.88HPT_eff = 0.87HPT_eff = 0.86HPT_eff = 0.85

HPRTE Efficiency

Figure 6-11 Specific power vs. HPPR for HPT efficiency sensitivity

76

800

802

804

806

808

810

812

17 22 27 32 37

OPR

Exha

ust T

empe

ratu

re (°

R)

TIT = 2350°R

TIT = 2400°R

TIT = 2450°R

TIT = 2500°R

HPRTE Efficiency

Figure 6-12 Exhaust temperature vs. OPR for TIT sensitivity

0.335

0.34

0.345

0.35

0.355

0.36

0.365

0.37

0.375

2.5 4.5 6.5 8.5

HPPR

Ther

mal

Eff

icie

ncy

Turbo Eff = 0.7138Turbo Eff = 0.6970Turbo Eff = 0.6804

HPRTE Efficiency

Figure 6-13 Thermal efficiency vs. HPPR for turbocharger efficiency sensitivity

77

0.33

0.34

0.35

0.36

0.37

0.38

4 4.5 5 5.5 6 6.5 7 7.5

LPPR

Ther

mal

Eff

icie

ncy

TIT = 2500°RTIT = 2450°RTIT = 2400°RTIT = 2350°R

HPRTE Efficiency

Figure 6-14 Thermal efficiency vs. LPPR for TIT sensitivity

Table 6-2 Summary of the HPRTE Efficiency sensitivity analysis

Perturbed Parameter Specific Power HPPR

TIT 2 1 3

TET 3 1

1

1 1

1 1

1 1

3 1

HPPR 3 2

LPPR 3 1

OPR 3

R 3 2

Cooler 3 3

Resultant Parameter

ambientT

inHPCT _

adComp _η

adTurb _η

Turboη

thη exhaustT

78

Table 6-3 Comparison of the thermal efficiency maximums and their corresponding overall pressure ratios (OPRs)

ConfigurationOPR

H-V Eff. 45.0% 23.6 0.442HPRTE Eff. 37.2 32.2 0.368

SCGT 37.3 24.0 0.361SCGT SpPw 33.4 10.4

Parameter

max_thη meanth _η

0.3

0.32

0.34

0.36

0.38

0.4

0.42

0.44

0.46

5 10 15 20 25 30 35 40 45 50

OPR

Ther

mal

Eff

icie

ncy

H-V Eff.HPRTE Eff.SCGT

Figure 6-15 Engine cycles comparison of thermal efficiency vs. OPR

Table 6-4 Comparison of the specific power maximum values and their corresponding OPRs

ConfigurationOPR

H-V Eff. 665 35.0 586HPRTE Eff. 624 42.6 565

SCGT 180 10.5 155

Parameter

SpPw units are industry standard (HP-sec/lbm)

maxSpPw meanSpPw

79

100

200

300

400

500

600

700

5 10 15 20 25 30 35 40 45

OPR

Spec

ific

Pow

er (H

P-se

c/lb

m)

H-V Eff.HPRTE Eff.SCGT

Figure 6-16 Engine cycles comparison of specific power vs. OPR

Table 6-5 Comparison of exhaust temperature maximum values for the three engine configurations

Configuration OPR

H-V Eff. 855 49.0 832HPRTE Eff. 805 42.6 799

SCGT 1660 8.00 1360

Parameter

Temperatures have units of °R

max_exhaustT meanexhaustT _

80

600

800

1000

1200

1400

1600

1800

5 10 15 20 25 30 35 40 45 50

OPR

Exha

ust T

empe

ratu

re (°

R)

SCGTH-V Eff.HPRTE Eff.

Figure 6-17 Engine cycles comparison of exhaust temperature vs. OPR

0.34

0.36

0.38

0.4

0.42

0.44

0.46

500 520 540 560 580 600

Ambient Temperature (°R )

Ther

mal

Effi

cien

cy

H-V Eff.HPRTE Eff.SCGT

Figure 6-18 Engine cycles comparison of thermal efficiency vs. ambient temperature

81

Table 6-6 Engine cycles comparison for four extreme operating conditions

Case SpPw OPR SpPw OPR SpPw OPR1 489°R 544°R 0.397 411 10.2 0.341 390 17.8 0.401 191 24.02 489 499 0.409 436 10.8 0.386 512 23.7 0.401 191 24.03 569 544 0.442 659 30.7 0.367 588 42.0 0.358 144 24.04 569 499 0.444 670 32.3 0.385 675 55.4 0.358 144 24.0

Operating Point H-V Efficiency HPRTE Efficiency SCGTEngine Type

All Specific Power calculations have industry standard units of HP-sec/lbm

thη thη thηambientT waterT

Table 6-7 High pressure compressor (HPC) inlet temperature comparison for the H-V Efficiency engine model

Design Point Parameter 509°R 499°R

45.0% 45.7%

(HP-sec/lbm) 629 639OPR 23.6 22.3TET 1768 1769R 3.00 3.00Equivalence Ratio 0.799 0.799

(lbm/sec) 6.36 6.26

(lbm/sec) 0.306 0.306

545 545LPPR 3.5 3.32HPPR 7.37 7.35

509 499TIT 2500 2500

837 835Temperatures are in °R

airm&

liquidOHm _2&

thη

SpPw

ambientT

inHPCT _

exhaustT

inHPCT _

82

Table 6-8 Final performance design point comparison for the engine configurations Design Parameter

SCGT Eff. SCGT SpPw HPRTE Eff. H-V Eff.

37.3% 33.4% 37.2% 45.0%

(HP-sec/lbm) 159 180 593 629OPR 24 10.4 32.2 23.6TET 1330 1580 1880 1770

R N/A N/A 3.30 3.00Equivalence Ratio 0.240 0.303 0.894 0.799

(lbm/sec) 25.2 22.2 6.75 6.36

(lbm/sec) N/A N/A N/A 0.306

544 544 544 544

LPPR N/A N/A 6.25 3.50HPPR N/A N/A 5.31 7.37

544 544 614 509

TIT 2500 2500 2500 2500

1330 1580 801 837

Temperatures are in °R

Engine Configuration

airm&

liquidOHm _2&

thη

SpPw

ambientT

inHPCT _

exhaustT

83

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS

Conclusions

The analysis performed for this thesis project consisted of parametric studies to

establish design point parameters, sensitivity studies to examine specific

parameter/resultant interactions, and design point comparisons of the performance

characteristics of the three gas turbine engine configurations. The engine models were

developed using a steady-state, incompressible thermodynamic approach with the engine

cycle code NPSS developed by NASA Glenn Research Center. The mission requirement

for the engine was produce continuous nominal power output of 4000 BHP. To refrain

from investing capital in exotic material development for the engine components, the TIT

maximum was limited to 2500°R and the hot side recuperator inlet temperature was

constrained to not exceed 2059°R. The turbocharger pressure ratio was designed to not

exceed a value of 7.5—this was a design limitation determined for single-stage

centrifugal compressors. Moreover, NPSS and CEA do not have thermodynamic

properties for solids; therefore, to prevent icing before the in the HPRTE engines, the

high pressure compressor (HPC) inlet temperature minimum value was 492°R (33°F).

The dependent variables optimized during the parametric performance comparison of the

engine cycles are listed in their order of significance: thη , specific power, and exhaust gas

temperature.

84

The conclusions are as follows:

• Comparison output from a NPSS run case of the HPRTE Efficiency matched well with output from a similarly configured engine cycle using the spreadsheet code. Those parameters whose values from NPSS agreed with the counterpart values from the spreadsheet code included: thη , OPR, TET, inHPCT _ , and exhaustT . The specific power outputs from the two codes did not match well. Their values differed by 12.5%. The difference in these output values can be associated to three causes. The operational handling of the spreadsheet code was such that it was impossible to model the HPRTE in the proper turbocharger configuration. Moreover, the codes modeled the engine using different fuels and thermodynamic curve-fit equations.

• For the SCGT sensitivity analysis, the results showed that the cycle thermal efficiency and specific power were both particularly sensitive to TIT and ambientT variations but were not very sensitive to OPR changes. This is not surprising when considering the dominance of the temperature ratio in the development of the theoretical thermal efficiency expression in the first section in Chapter 6. The three most noteworthy influence coefficients from this section are

( )36.3

)(=

base

base

TITTIT

SpPwSpPw

( )76.1

)(−=

base

base

TambTamb

SpPwSpPw

( )898.0

)(_ =

base

baseth

th

TITTIT

ηη

.

• Two design points were considered for the SCGT engine configuration. One was optimized for maximum thη (SCGT Efficiency ) and the other was optimized for maximum specific power (SCGT SpPw). The SCGT SpPw best predicts the ETF-40B design point. The SCGT Efficiency had a predicted thη value of 37.3%—10.2% higher than the thη predicted for the SCGT SpPw engine configuration. The SCGT SpPw had a predicted specific power of 180 HP-sec/lbm—13.2% greater than the SCGT Efficiency configuration.

• HPRTE Efficiency sensitivity analysis was performed next. Variations in the following parameters affected the thη resultant by a proportional amount: inHPCT _ ,

adTurb _η , adComp _η , and TIT to a lesser extent. Specific power was decidedly sensitive to these input parameters: TIT, TET, inHPCT _ , adTurb _η , adComp _η , and

Turboη . The HPPR was sensitive to the following parameter inputs: ambientT and LPPR.

85

• A byproduct of the sensitivity analysis for the HPRTE Efficiency was that the optimized pressure ratios for the two spools were determined. The optimization was based on maximizing thη rather than specific power. The turbocharger pressure ratio was chosen to 6.25, and the HPPR was chosen to be 5.31.

• Exhaust gas temperature ( exhaustT ) was an important consideration in the engine cycles comparison studies. exhaustT values for the HPRTE cycles were an average of 550°R less than the exhaustT values of the SCGT Efficiency design point. When compared to the SCGT SpPw design point, the exhaustT values for the HPRTE cycles were almost 800°R less. Cooler exhaust temperatures directly impact the survivability of the ship. Naval ships powered by HPRTE engines instead of SCGT engines would have a greatly reduced infrared detection signature. Moreover, an 800°R reduction in temperature would increase the density of the exhausted gases implying that the exhaust ducting would be smaller in diameter for the HPRTE system.

• The thη values for both HPRTE cycles remain consistently high through a wide operating range of pressure ratio designs. The meanth _η for H-V Efficiency is 44.2% and for HPRTE Efficiency it is 36.8%. The meanth _η of the SCGT was 36.1%. The HPRTE cycle curves for this part of the analysis were very flat meaning that the

thη value is not greatly affected by OPR variations. The implication here is that existing turbomachinery components could most likely be used to design a production HPRTE system.

• The four extreme operating cases analyze the performance characteristics of the competing engine cycles head-to-head. Many conclusions can be drawn from this section of work. First, the H-V Efficiency has better thermal performance in hot weather than in cold weather. Second, thermal performance of the H-V Efficiency is not significantly affected by water temperature. Third, raising air temperature positively impacts the thermal performance and specific power of both HPRTE cycles while negatively affecting both performance characteristics of the open cycle SCGT. Under hot conditions (cases 3 and 4) the H-V Efficiency performed with an average thermal efficiency of 44.3%. That is an average of 17.8% higher than the HPRTE Efficiency cycle and 23.7% higher than the SCGT.

• The effects of decreasing HPC inlet temperature on a H-V Efficiency configured engine were analyzed. The data reveals that there are minor increases (less than 2%) in the performance variables thη and specific power when HPC inlet temperature is dropped from 509°R to 499°R; however, those increases are not significant enough to make a recommendation for this concept.

• For their optimized design point parameters, the H-V Efficiency has a thη of 45.0%—besting the SCGT Efficiency by 20.6%, the SCGT SpPw cycle version by

86

34.7%, and the HPRTE Efficiency engine cycle by 21.0%. Thermal efficiency is inversely related to the specific fuel consumption. For the same ship platform, the H-V Efficiency engines would allow for increased cargo if the mission range is unchanged. The HPRTE cycles should also be considered if the important design issue is mission range. Having lower specific fuel consumption than the SCGT SpPw design point suggests that the HPRTE cycles would exhibit an increased mission range capability if the fuel tank size is a constant parameter.

• There are significant specific power differences between the HPRTE engine cycles and the SCGT. The mean specific power of the H-V Efficiency cycle is nearly four times larger than that of the SCGT. Since specific power is directly proportional to the area of the ducting in the engine, the core HPRTE engine would be almost four times smaller than a SCGT of the same power capacity (of course some additional space would be needed to house the VARS unit). Currently, it is unknown if the size and performance trade-offs would cancel each other.

Recommendations

• An off-design point analysis should be completed on the engine cycle configurations reported on in this analysis because over 93% of naval ship operation time is spent operating at or below 35% engine power [Landon]. NPSS would be the obvious software choice for this next step since it has off-design point modeling capabilities. Moreover, the engine models have already been created in NPSS and the output has been benchmarked to a certain degree. Performance map integration and scaling is a critical competency that must be addressed before this next step is taken; and it is unknown how the performance map look-ups would affect model computing times. Of course, any added sophistication to a model increases the expectation that it will lengthen the model run-time. Currently, for the H-V Efficiency model, the run-times range from 4 to 12 minutes, depending on the model its constraints. A comprehensive off-design study should include range and propeller analysis, as well.

• Further benchmarking of the results of this analysis is a necessary next step to guarantee the accuracy of the work. The accuracy of NPSS is not in question; it was developed by NASA in conjunction with leading United States aero-propulsion companies. However, checking the fidelity of the HPRTE models against other thermodynamic cycle codes, such as the code created by Jameel Kahn, is a useful next step to give confidence to this work. Furthermore, comparing NPSS model results against experimental test data from the laboratory demonstrator should be considered.

• A more robust model would include the VARS unit, at least to the extent that the HPC inlet temperature is set by calculations representative of having the VARS unit in the model. Thus far, excess cooling capacity of the VARS has not been addressed, but there are obvious uses for additional refrigeration on a naval ship.

87

• Other software modeling programs should be considered for future performance analysis of the power and refrigeration cycles being examined by the University of Florida (UF). One commercial software package currently being explored by the Energy and Gas Dynamics Laboratory at UF is ASPEN Plus. ASPEN Plus is an industry leader in process flow modeling and has proven capabilities modeling gas turbine power and refrigeration cycles. Not only can it model design point performance, but off-design analysis is available.

• There are other HPRTE layout designs being explored by UF. Even the current test rig in the Energy and Gas Dynamics Laboratory has a slightly different gas flow path than the HPRTEs in this analysis. Instead of introducing the fresh air before the evaporator, the current test engine rig has the air inlet ducted into the entrance to generator heat exchanger of the VARS unit.

• Since the analysis was done assuming dry ambient air, perhaps more cases should be run on the H-V Efficiency to determine how the added moisture of humid air would affect the water extraction rate.

• A variant design of the H-V Efficiency cycle would have a lower nominal power output rating. Then, during the infrequent instances of operation that full power (4000 BHP) is needed, the water being extracted after the evaporator could be re-injected downstream of the HPC exit to provide the needed boost in horsepower.

88

LIST OF REFERENCES

[1] About Inc., “The History of Engines-How Engines Work Part 2: A Short History and Timeline of Gas Turbine Engines,” The New York Time Company, [online], http://inventors.about.com/library/inventors/blenginegasturbine.htm, [retrieved 18 December 2006].

[2] Shaw, R. J., “A Short History of Turbine Engines,” Ultra-Efficient Engine

Technology Program, NASA, John H. Glenn Research Center at Lewis Field, Cleveland, OH, http://www.ueet.nasa.gov/engines101.php#history, [retrieved 18 December 2006].

[3] Mattingly, J. D., Elements of Gas Turbine Propulsion, McGraw Hill, Inc., New York,

1996, pp. 232-233. [4] Brady, C.O., & Luck, D.L., “The Increased Use of Gas Turbines as Commercial

Marine Engines,” Journal of Engineering for Gas Turbines & Power, ASME, April 1994, pp. 428-433.

[5] Harmon, R., “Marine Gas Turbine: A New Generation,” Mechanical Engineering,

May 1990, pp. 48-51. [6] Davis, R. C., “Final Test Report of Wolverine Gas Turbine Plant,” Research and

Development Report #030076 Ns-6220210, U.S. Naval Engineering Experiment Station, Annapolis, 14 August 1956.

[7] GE Aviation, “The History of Aircraft Engine,” General Electric Company, [online],

http://www.geae.com/aboutgeae/history.html, [retrieved 18 December 2006]. [8] Valenti, M., “Propelling Jet Turbines to New Uses,” Mechanical Engineering, March

1993, pp. 68-72. [9] Ness, J.C., Franks, C.B., Sadala, R.L., “Economic Considerations for a New Gas

Turbine System in the U. S. Navy,” Journal of Engineering for Gas Turbines & Power, ASME, April 1988, pp. 271-272.

[10] GE Infrastructure Marine, “Making Waves: The Rugged, Reliable 7FDM Diesel

Engine,” The General Electric Company, [online brochure], http://www.mshs.com/pdf/GE7FDMBrochure.pdf, [retrieved 18 December 2006].

89

[11] Vericor, “TF Series Marine Gas Turbine Engines – 4000 to 5000 Hp,” MTU Aero Engines Company, [online brochure], p http://www.vericor.com/products/downloads/Marine_tfseries_datasheet.pdf, [retrieved 18 December 2006].

[12] MacFarlane, R.S., Lear, W.E., “System Impact of H2O Production and Injection of a

Novel Semi-Closed Cycle Gas Turbine,” AIAA, 35th Aerospace Sciences Meeting and Expo, Reno, NV, USA, Jan. 1997.

[13] Landon, J.C., “Design and Off-Design Point Study of Two Regenerative Feedback

Turbine Engines for Marine Applications,” Master’s Thesis, University of Florida, Gainesville, Dept. of Mechanical Engineering, 1996.

[14] “Fuel-Saving Warship Drives,” Mechanical Engineering Magazine Online, Aug.

1998: n. page, Online, ASME, 5 Aug. 2006. [15] Korakianitis, T., Beler, K.J., “Investigation of the Part-Load Performance of Two

1.12 MW Regenerative Marine Gas Turbines,” ASME, Journal of Engineering for Gas Turbines & Power, April 1994, pp. 418-423.

[16] Gasparovic, N., “The Advantage of Semi-Closed Cycle Gas Turbines for Naval Ship

Propulsion,” Naval Engineers Journal, April 1968, pp. 275-281. [17] GSP Team, “GSP-Gas turbine Simulation Program,” National Aerospace

Laboratory—The Netherlands, [online], http://www.gspteam.com/main/main.shtml, [retrieved 18 December 2006].

[18] Danias, G.A., “Design and Off-Design Point Study of Two Regenerative Feedback

Turbine Engines for Marine Applications,” Master’s Thesis, University of Florida, Gainesville, Dept. of Mechanical Engineering, 1997.

[19] “Aspen Plus”, Aspen Technology, Inc., [online] July 2006,

http://www.aspentech.com/brochures/aspenplus.pdf, [retrieved 18 December 2006].

[20] United States, NASA, John H. Glenn Research Center at Lewis Field, NPSS User

Guide, Cleveland, OH, 2006. [21] Nemec, T., Lear, W. E., “Thermodynamic Performance of a Semi-Closed Gas

Turbine Combined with a Rankine Bottoming Cycle,” AIAA, 36th Aerospace Sciences Meeting and Expo, Reno, NV, USA, Jan. 1998.

[22] Boza, J. J., 2003, “Performance of a Novel Semi-Closed Gas Turbine Refrigeration

Combined Cycle,” ASME, Turbo Expo: Power for Land, Sea, & Air, Atlanta, GA, USA, June 2003.

90

[23] Malhatra, V., “Life-Cycle Cost Analysis of a Novel Cooling and Power Gas Turbine Engine,” Master’s Thesis, University of Florida, Gainesville, Dept. of Mechanical and Aerospace Engineering, 2005.

[24] Khan, J., Lear, W., Sherif, S., “Performance of a Novel Combined Cooling and

Power Gas Turbine with Water Harvesting,” ASME, Turbo Expo: Power for Land, Sea, & Air, Vienna, Austria, April 2004.

[25] Gordon, S., and McBride, B. J., ( 1994, October), “Computer Program for

Calculation of Complex Chemical Equilibrium Compositions and Applications,” Cleveland, OH, NASA Reference Publication 1311.

[26] Hill, P., and Peterson, C., Mechanics and Thermodynamics of Propulsion, Addison-

Wesley Publishing Company, Reading, MA, 1992, pp. 173. [27] Turns, S. R., An Introduction to Combustion: Concepts and Applications, McGraw-

Hill Book Company, Boston, MA, 2000.

91

BIOGRAPHICAL SKETCH

The author is a native of Flagler Beach, Florida. He earned a bachelor of science

degree in aerospace engineering from the University of Florida in 2004. For the past 2 ½

years he has been a member of the Energy and Gas Dynamics Laboratory at the

University of Florida. His graduate emphasis was firmly focused in the thermal sciences

with selective coursework including gas turbine propulsion systems, combustion, and

entrepreneurship for engineers. His future plans include working in nuclear power

generation in Crystal River, FL.