You are on page 1of 24

Simulation Modelling Practice and Theory 44 (2014) 95–118

Contents lists available at ScienceDirect

Simulation Modelling Practice and Theory


journal homepage: www.elsevier.com/locate/simpat

Multi-field coupling dynamic modeling and simulation


of turbine test rig gas system
Yang Chen a,⇑, Guobiao Cai a, Zhenpeng Zhang a, Yulong Huang b
a
School of Astronautics, Beihang University, Beijing 100191, PR China
b
Beijing Aerospace Measurement & Control Corp., Beijing 100041, PR China

a r t i c l e i n f o a b s t r a c t

Article history: On the basis of early research on the 38-component system, this paper extends the estab-
Received 23 June 2013 lished finite volume model to the multi-field coupling numerical system which can
Received in revised form 28 October 2013 describe flow, heat transfer, combustion and rotation, and conducts, based on the experi-
Accepted 12 March 2014
ment, dynamic modeling and simulation research on the turbine test rig gas system which
Available online 17 April 2014
includes turbine and load. The comparison among the simulation results, test data and
early simulation results indicates the 42-component system established by this paper
Keywords:
has made improvements against many weaknesses of the Previous simulation in an all-
Turbine test rig gas system
Modularization modeling
round way. Accordingly, it can be concluded that the case setting and algorithm improve-
Multi-field coupling ment are effective. It is also found that the modeling of module with chemical reaction, the
Dynamics spool throttling modeling of various regulator valves, the modeling of the turbine charac-
Numerical simulation teristics and the modeling of wall heat transfer are four key factors which affect simulation
accuracy, and that the coupling modeling of flow and combustion is the key factor which
affects simulation stability.
Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction

The idea of modularization [1] is gradually formed to meet the demand of versatile system modeling. Its basic idea is
firstly viewing the system as an assembly of some typical components which are called modules, secondly establishing
numerical model of every module and encapsulating it as an independent function module, finally establishing numerical
model of the whole system by combination of relevant modules according to special rules. As all components of the same
type are described by one module, the modeling and simulation problem of all kinds of systems with different structures
can be solved conveniently by computer. In the liquid propulsion system (LPS) simulation field, this idea is deemed to be
put forward by A.P. Tishi and L.P. Gurova [2]. Since late 1980s, relevant researches have been evolving toward maturity
and put into application in the form of simulation software [3–25]. Ref. [26] summarizes the relevant researches and points
out that future improving direction in the aspect of algorithm will focus on more detailed and accurate modular modeling of
typical component, and integration of simulation and optimization function for optimization application.
At the level of component research, the high-fidelity full-scale three-dimensional (3D) or two-dimensional (2D) modeling
and simulation can be conducted by employing the leading CFD (Computational Fluid Dynamics) software tools (e.g. CFX,
FLUENT) for many components of LPS. However, at the level of system research, that has not become the mainstream prac-
tice mainly because: for a transient complex system consisting of tens of or even hundreds of components, the CFD softwares

⇑ Corresponding author. Tel.: +86 010 82316540.


E-mail address: yangchen@buaa.edu.cn (Y. Chen).

http://dx.doi.org/10.1016/j.simpat.2014.03.004
1569-190X/Ó 2014 Elsevier B.V. All rights reserved.
96 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

Nomenclature

A cross-sectional area (m2)


Abmin the smallest flow section area of turbine stationary blade nozzle (m2)
acr critical sound velocity (m/s)
Cd flow coefficient
Csb flow coefficient of turbine stationary blade nozzle
cp specific heat at constant pressure (J/(kg K))
DTM turbine cascade average diameter (m), (Dhub + Dtip)/2
E total energy per unit volume (J/m3)
fmix mixture fraction
h specific enthalpy (J/kg) or spool opening (m)
Ma mach number
MLoad load torque (N m)
MT turbine drive torque (N m)
m,_ Qm mass flow rate (kg/s)
mr mean mass mixture ratio
nrot rotate speed of rotor (r/min)
nT rotate speed of turbine (r/min)
P power (W)
p,pt static pressure and total pressure (Pa)
Qm_np turbine flow capacity parameter
Qv volume flow rate (m3/s)
q_ heat flux density per unit area (W/m2)
qrho heat flux density per unit area from in-tube fluid to pipe wall (W/m2)
qrho2 heat flux density per unit area from wall to out-tube fluid (ambient atmosphere) (W/m2)
R specific gas constant (J/(kg K))
Re Reynolds numbers
r pipe-wall radial coordinate or radius (m)
rQ m the mass-flow-rate ratio of oxidant to fuel flowing into combustion zone
S surface area (m2)
T, Tt static temperature and total temperature, respectively (K)
T f2 fluid static temperature outside the pipe (ambient temperature) (K)
t time (s)
u fluid velocity in x direction (m/s)
uTM median-location circular velocity of turbine blade (m/s)
V volume (m3)
W work per unit mass working medium (J/kg)
x axial direction of pipe (one-dimensional flow direction)
a2 coefficient of convective heat transfer from pipe wall to out-tube fluid (W/(m2 K))
c ratio of specific heat capacities of gas
e1, e2 blackness of pipe wall and ambient environment, respectively
es system blackness
g efficiency
k thermal conductivity (W/(m K))
ku median-location circular velocity coefficient of turbine blade
p circular constant
ppt turbine total pressure ratio
q density (kg/m3)
r0 blackbody radiation constant (W/(m2 K4))
s valve relative opening
XT turbine reaction degree
x rotational angular velocity (radians/s)

Subscripts
ad adiabatic
c combustion zone
ektexine external surface of pipe wall
f fluid
f fuel
i serial number of a axial-direction grid
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 97

n total number of flow-field grids


nw total number of wall temperature-field radial-direction grids
o oxidant
rb rotating blade
sb stationary blade
T turbine
TM median location of turbine blade
w wall

like CFX and FLUENT, with current technological conditions, are not able to undertake the system-level modeling and sim-
ulation in terms of complexity and task volume because it involves multi-component and multi-disciplinary combined tran-
sient modeling of components with different structures and functions, flow of different fluids and various non-pure flow
phenomena; even if the modeling was successfully made with the CFD tools, the computing cycles would not be acceptable
due to the huge computation.
Therefore, a reasonable balance must be made between the accuracy of component models and the complexity of the
whole system for the system-level research. The general approach is to use 1D or even 0D simplified models to conduct mod-
eling and simulation, with focus on the overall performance of the system and the specific role of a single component in the
dynamic change process of the system. Currently, the relatively mature system-level simulation software tools, such as RO-
CETS [3] and GFSSP [6,7] (USA), AMESim [10] (France) and Flowmaster [12] (UK), are developed in this way and put into
commercial use. Moreover, according to the researches [13–25] of recent years, the modeling method in the system-level
simulation field, represented by such software tools as ROCETS [13,14], GFSSP [15,16], AMESim [17], CARINS [18], Flowmas-
ter [19], Matlab/Simulink [20,21] and so on [22–25], does not change intrinsically. These software tools just have more func-
tions and more extensive application scope by constantly adding new modules and improving algorithm, and are applied to
modeling, simulation and optimization of some specific complex systems.
It is worth noting that NASA (National Aeronautics and Space Administration) has realized multidisciplinary analysis on
the whole engine system at a variety of resolution levels (from 1D steady state to 3D transient state) in the NPSS (Numerical
Propulsion System Simulation) project [27–47] which has been carried out for more than a decade, but the 1D framework of
the flow field simulation, NCP [27,28] (National Cycle Program, released by the Lewis Research Center, the predecessor of
NASA Glenn Research Center, during 1997–1998), was still developed by the foregoing method. Furthermore, the high-
dimensional frameworks [29–31] are coupling systems consisting of many single-discipline software tools such as SINDA
(Heat transfer), ENG10 and ADPAC (Fluids), TETRA and NASTRAN (Structures). Particularly emphasizing on 2D or 3D qua-
si-steady simulation, the high-dimensional frameworks do not give a solution to high-dimensional multi-disciplinary tran-
sient combined simulation of flow, heat transfer and combustion.
The NPSS project was proposed by NASA Glenn Research Center in late 1990s [32] with the primary purpose of achiev-
ing high-fidelity system analysis ability by the integrated computer simulation of many disciplines including fluid
mechanics, heat transfer, combustion, structural mechanics, materials, controls, manufacturing and so on in order to im-
prove the confidence level of design and decrease the cost of test and related hardware facilities. The ultimate goal of NPSS
is modeling the entire propulsion system at the highest level of fidelity (3D transient and multidisciplinary) for revealing
the physical mechanisms that affect system characteristics, while two problems prevent this from being a viable option in
most cases [33,34]: first, the level of detailed information needed as boundary and initial conditions to get a converged,
validated solution would be extremely difficult to collect; second, the computational time and cost would be prohibitively
high for effective use in a design environment. Therefore, a kind of important method called zooming [29–37] is presented
for multi-components coupling simulation in NPSS framework. Zooming is defined as the ‘‘automated, seamless integra-
tion of one or more high resolution analyses with a full system simulation’’ [35,36]. Its basic idea is integrating the high-
dimensional models of some key components (such as fan, compressor, combustor and turbo) in an aerospace system with
the low-dimensional full engine model by the directly linking approach (fully integrated zooming) or the characteristic-
map approach (partially integrated zooming) [34,38,39] to conduct variable-fidelity simulation of the entire system.
The zooming methods have been extensively used in the simulation and optimization design of aerospace propulsion sys-
tems and components [29–45]. However, there is a lack of explicit application report on full-system high-dimensional
transient simulation.
As a continuation of the numerical study on the turbine test rig system, this paper connects the turbine component to
main test system [26] by the characteristic-map approach in zooming methods, and carries out the modeling and simulation
of full gas system of the test rig by expanding the algorithm system and improving the regulating time sequence.
98 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

Fig. 1. Schematic diagram of the turbine test rig system.

2. Physical model of the system

Fig. 1 shows the schematic diagram of the turbine test rig system which can be approximately divided into three parts:
main test system, fuel–oil supply and ignition system, turbine and load system. The function of main test system is to pro-
vide driving gas for rotation of the test turbine. The function of fuel–oil supply and ignition system is to provide fuel, which
then mixes with air to generate high-temperature and high-pressure gas after combusting in the BK-1 type heater as the
driving source of turbine rotation, for main test system. Turbine and load system is test object of the test rig. The turbine
test rig gas system in this paper comprises two major parts, the main test system and the turbine and load system.
It can be seen from Fig. 1 that the test rig contains a large number of fluid pipes and regulating apparatuses, the gas tem-
perature at turbine inlet is adjusted by controlling the combustion mixture ratio and the mixing of hot combustion gas and
normal air, the turbine inlet and outlet pressures are adjusted by regulating the openings of many valves. Since different test
conditions of turbine have different requirements on such parameters as air flow rate, fuel flow rate, turbine pressure ratio
and output power, there are great differences among corresponding regulating schemes of valves. In addition, as only several
tests have been conducted since the test rig was established, and each valve is regulated by manual control and there are no
detailed records on the regulating time sequence, only limited adjusting experience has be accumulated. The combination of
the two aspects makes present adjusting and testing of the test rig to be a very heavy task for engineering test personnel. In
order to provide engineers with advice on the rig regulating test, Ref. [26] establishes a 38-component numerical system
(hereinafter referred to as the Previous Case or Previous) for the main test system which excludes the turbine and the seg-
ments behind it by employing computer numerical simulation method, then conducts the simulation study based on a rig
regulating test. Comparison between simulation results and experimental data indicates that the errors between 20–260 s
dynamic simulation results and test measurement results at various measuring points are below 20% on the whole (except
that the errors of the total temperature curve at main-line measurement section and total temperature curve at turbine inlet
exceed 20% around 60–160 s period), and the errors at several measuring points are below 6%. However, the following prob-
lems still exist:

(1) Given that the simulation program has a problem of stability, the numerical system which is finally used keeps the
combustion gas velocity far away from the low velocity scope by narrowing the inner diameter of the heater outlet
pipe. The measure improves numerical stability by avoiding simulation instability arising out of instantaneous back-
flow caused by oscillation when the system starts up and shifts from a stage to another. Although the method is very
effective and would not deviate the simulation results at other parts from the original numerical system, manually
changing the structure size makes the simulation results represented by the flow velocity of the heater outlet pipe
disagree with the actual situation.
(2) The simulation framework fails to include core parts of the test rig – turbine and its load, and the outlet boundaries of
the 38-component numerical system depend too much on test data. In the case of high measuring accuracy, using
actual measurement data as boundary conditions may ensure the consistency between the boundaries of the numer-
ical system and those of the test system, but it, from the perspective of simulation verification, results in less test
curves to verify simulation results.
(3) It is very obvious in the rig regulating test that there is a phenomenon of temperature gradually reducing along the
flow direction due to pipe wall absorbing heat when high-temperature combustion gas flows through the pipeline,
that is so-called ‘‘thermal stratification’’ phenomenon. Given the long distance of the pipeline in which the phenom-
enon occurs, the obvious temperature stratification and the large number of sensor measuring points involved, the
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 99

phenomenon is a good indicator for evaluating the simulation accuracy. Although the errors of the Previous simulation
results are less than 20% on the whole, great improvements need to be made in the simulation algorithm and the case
setting due to numerous sensor measuring curves.

In view of the abovementioned weaknesses, this paper firstly further extends the established finite volume model system
[26] to the multi-field coupling field which can describe flow, heat transfer, combustion and rotation by theoretical deepen-
ing and extending, numerical expansion and integration, and practical application and evaluation, secondly conducts, based
on the experiment, a comprehensive numerical modeling and simulation study on the gas system which includes the turbine
and its load, finally makes assessment and analysis on the simulation results based on the test data.

3. Modeling approach

Based on the finite volume model system [48,49] developed by us which attributes to the improvement and evolution of
the finite element state-variable model system [1], Ref. [26] has given an introduction to the finite volume model used for
calculating the transient flow field of quasi-one-dimensional compressible variable-section pipe flow, the axisymmetric two-
dimensional finite volume model used for calculating pipe-wall transient heat transfer, the valve spool and orifice throttling
models, the thermodynamic calculation model, and the flowchart of modularization modeling and simulation. Code optimi-
zation has been made on some less efficient sentences in the follow-up work, which not only simplifies the program but also
reduces unnecessary program operation, improves readability of the program and helps increase calculation speed of sim-
ulation. Emphases are placed on introducing greater model improvements and numerical models of newly added modules as
follows.

3.1. Improvement in pipe-wall temperature field model

Given the high wall temperature in the pipe network area through which the combustion gas flows, improvement is made
in the pipe-wall external boundary heat transfer model in the case of single-tube pipeline structure and the radiation heat
transfer effect is taken into consideration based on the convective heat transfer model [26] in the Previous Case.
The main assumptions include:

(1) The inner tube of pipeline is a long cylinder pipe and the layer number of its wall is no more than two. The heat con-
duction within the wall is axisymmetric.
(2) In case of a single-tube pipeline structure, there is only convective heat transfer or simple superposition of convection
and radiation heat transfer between the external surface of pipe wall and the external environment, and the atmo-
sphere temperature outside the pipe is constant and equal to local environmental temperature. The radiation heat
transfer between the external surface of pipe wall and the external environment is considered as the radiation heat
exchange between the external surface of finite-volume gray-body grid and the infinite-surface gray body.

In the case of single-tube pipeline structure, the external boundary conditions of convective heat transfer are calculated
by:
@T
q_ i;nw ¼ ki;nw ¼ a2 ðT ektexine  T f2 Þ; i ¼ 0; 1; . . . ; n  1 ð1Þ
@r
In the case of single-tube pipeline structure, the external boundary conditions of combined action of convective and radi-
ation heat transfer are calculated by:
@T
q_ i;nw ¼ ki;nw ¼ a2 ðT ektexine  T f2 Þ þ es r0 ðT 4ektexine  T 4f2 Þ ð2Þ
@r
.h i
S1

As S2  S1ektexine , the system blackness is calculated by: es ¼ 1 e11 þ ektexine
S2
1
e2  1  e1 .
In above equations, i means the serial number of a grid in axial direction, nw is the total number of wall radial-direction
grids, r and T are radial coordinate and temperature of pipe-wall control volume, respectively, k is the thermal conductivity
of control volume material; Tektexine is the temperature of pipe-wall outer layer, T f2 is the static temperature of fluid outside
the pipe, a2 and q_ i;nw mean the coefficient of convection heat transfer from pipe wall to fluid outside the pipe and the heat
flux density per unit area, respectively; S1ektexine , e1, S2 and e2 are the external surface area and the blackness of pipe wall, the
internal surface area and the blackness of ambient environment, respectively, r0 means the blackbody radiation constant
and es is named the system blackness.

3.2. Development of valve spool and orifice throttling models

Throttling phenomenon occurs when the fluid flows through a valve spool or an orifice (a suddenly-narrowed cross sec-
tion). Along with generation and interaction of vortexes, the physical discontinuity cross section will definitely bring out
mathematical discontinuity. Meanwhile, the opening degree of spool can range from complete shutdown to full opening.
100 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

Therefore, it would be a rather difficult proposition to accurately describe the dynamic throttling effect of the spool with
one-dimensional flow model. The valve spool and orifice throttling models which have been built include the pressure ra-
tio-based injector orifice model and the pressure difference-based injector orifice model [26]. The paper proposes a dual-
model solution of calculating spool throttling to the flow-rate distortion problem of traditional injector orifice models in
the case of large opening, high pressure level and slight differential pressure, with basic idea lying in the practice that the
injector orifice models are used in the case of large pressure loss while the local loss model similar to the model of pipe mod-
ule is used in the case of small pressure loss. Furthermore, an injector orifice model based on the flow coefficient obtained by
converting from turbine flow capacity parameter is established for throttling of turbine stationary blade nozzle, details of
which will be given in the introduction to the numerical model of gas turbine module. The priority is given to introduce
the built calculation methods of flow coefficient in the following.
The flow-rate formulas for spool or orifice throttling are semi-theoretical semi-empirical formulas. For compressible gas,
one of common forms is the spool mass flow rate equation [26] in the pressure ratio-based injector orifice model. It can be
seen from the equation that when static pressures and densities at upstream and downstream of a valve spool or orifice are
known, the calculated value of flow rate at throttling position is directly decided by values of flow section area and flow coef-
ficient Cd which are functions of spool opening h. The flow section area is a univariate function of the opening, so the relation
formula between it and the opening can be obtained via mathematical derivation when the shapes of valve spool and seat
are determined. The flow coefficient is a function of such factors as the opening, thermophysical properties of fluid and the
spool shape, so it is quite complicated to determine the relation formula between it and the opening. The usual practice is to
conduct interpolation or fitting with test data of spool characteristics.
Eight calculation schemes of flow coefficient are set up upon relevant references and a variable, n_Type_Cd, for controlling
the flow coefficient scheme is set up for users to choose an appropriate solution. Specifically, the variable equals to 0 for con-
stant coefficient scheme (Then users can further set the constant value of flow coefficient.), 1 for the scheme of variable-coef-
ficient default formula (a relational expression between flow coefficient and relative opening, Cd  s), 2 for the
PolynomialPerry formula scheme (a common relational expression between flow coefficient and downstream-to-upstream
static pressure ratio, Cd  p2/p1), 3 for the equal percentage flow characteristic fitting formula scheme of HCB cage-type dou-
ble-seat high-capacity sleeve regulator valve (Cd  s), 4, 5 and 6 for the flow characteristic fitting formula schemes of DN200,
DN250 and DN400 VBS soft-seat butterfly valve, respectively (Cd  s), 7 for the flow characteristic fitting formula scheme of
turbine flow capacity parameter (Cd  s, ppt, where Cd stands for Qm_np and s means ku). When there is a new spool or orifice
throttling model, users, thanks to this design, can obtain a new model by expanding the value of n_Type_Cd (8, 9, . . .) and
adding a corresponding model formula (Cd  s, p2/p1, etc.) into the module of flow coefficient calculation schemes. The pur-
pose of this design is to facilitate expanding the experience library used for calculating flow coefficient so as to facilitate
application of various throttling characteristic test data and finally to get the simulation value of mass flow rate at throttling
position which can meet the requirement of accuracy. Under the circumstances where the upstream and downstream static
pressures, densities and spool opening are known, it is a key for the simulation value of flow rate at the regulator valve spool
to agree with the test value. The other key is the univariate function relationship between flow section area and spool open-
ing. Unlike the condition that the flow coefficient is affected by various uncertain factors (due to the zero-dimensional
description mechanism highly depending on experience), the relation formula between flow section area and spool opening
can be uniquely determined as long as the spool structure is fixed.

3.2.1. Variable-coefficient default formula scheme


A relation formula between flow coefficient and relative opening is fitted according to the characteristic test data of a type
of valve [50]. It is developed into the form in this paper as follows:
8
< C d ðs ¼ 0:30Þ
> 0 6 s < 0:30
C d ðsÞ ¼ 1:30131 þ 9:9684s  26:1845s2 þ 28:3646s3  9:8375s4 0:30 6 s 6 0:93 ð3Þ
>
:
C d ðs ¼ 0:93Þ 0:93 < s 6 1:0

3.2.2. PolynomialPerry formula scheme


The flow coefficient is considered as the function of downstream-to-upstream static pressure ratio at spool throttling po-
sition in the scheme [10], that is:
   2  3  4  5
p p p p p
C d ¼ 0:8414  0:1002 2 þ 0:8415 2  3:9 2 þ 4:6001 2  1:6827 2 ð4Þ
p1 p1 p1 p1 p1

3.2.3. Four flow characteristic fitting formulas of regulator valves developed by Wuzhong Instrument Co., Ltd.
Wuzhong Instrument Co., Ltd. provides the flow characteristic curves [51] of flow coefficient and spool stroke percentage
(relative opening) for various regulator valves it develops. The characteristic data can be extracted from these characteristic
curves in line with the type of the regulator valve which is used in the gas system, and then, four characteristic formulas of
the flow coefficient Cd and the relative opening s can be fitted based on these characteristic data in the form of Y = f(X). In
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 101

P  point b i Þ2 and the standard error of the estimate (SEE) is


these formulas, the error sum of squares is defined as SSE ¼ ni¼1 ðY i  Y
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
defined as r ^ ¼ SSE=ðnpoint  ncoefficient Þ; npoint and ncoefficient mean the number of data points and the number of coefficients
in the fitting formula, respectively; Yi and Y b i are the actual value of the given data point and the predicted value calculated by
corresponding formula, respectively.

(1) Equal percentage flow characteristic fitting formula scheme of HCB cage-type double-seat high-capacity sleeve regu-
lator valve

If ssubcr 6 s 6 0:12; C d ðsÞ ¼ C dRating  ð0:00153 þ 9:64283s3 Þ


2
ð5Þ
If 0:12 6 s 6 1:0; C d ðsÞ ¼ C dRating  e5:05068þ9:20303s4:16797s
According to the above formula, the SEE between the calculated results of the flow coefficient percentage and the char-
acteristic data is 1.2503E02.
Where the flow coefficient percentage is defined as Cd% = Cd/CdRating; CdRating is the flow coefficient rating, that is, the value
of Cd in the case of s = 1.

(2) Flow characteristic fitting formula scheme of DN200 VBS soft-seat butterfly valve

C d ðsÞ ¼ 508:178 þ 1549:48s2  ln s þ 549:800es ð6Þ


The SEE of this formula is 11.941.

(3) Flow characteristic fitting formula scheme of DN250 VBS soft-seat butterfly valve

C d ðsÞ ¼ 804:109 þ 2258:69s2  ln s þ 854:150es ð7Þ


The SEE of this formula is 15.574.

(4) Flow characteristic fitting formula scheme of DN400 VBS soft-seat butterfly valve

C d ðsÞ ¼ 1997:29 þ 6175:65s2  ln s þ 2174:00es ð8Þ


The SEE of this formula is 49.920.

(5) Processing methods of various flow characteristic fitting formulas in the case of small opening

The flow characteristic curves provided by Wuzhong Instrument Co., Ltd. need to be used together with the pressure dif-
ference-based injector orifice model, in which the mass flow rate equation of valve has nothing to do with the flow section
area. Therefore, on the one hand, the flow coefficient Cd = 0 must be ensured when the relative opening s = 0, only based on
which the calculated flow rate Qm = 0 can be ensured when s = 0. On the other hand, the respective fitting formulas are ob-
tained according to the characteristic curves of three kinds of VBS soft-seat butterfly valves which have different nominal
diameters. Although the data of the regulator valves with different nominal diameters seem to be independent, a principle
can be discovered by analysis and comparison that the ratios among the flow coefficients of regulator valves with different
nominal diameters are approximately equal to the ratios among their flow sectional areas at the same point of relative open-
ing. Hence, the principle should be followed when special processing is conducted in small-opening region which the fitting
formulas cannot cover.
For the above two reasons, the aforementioned four characteristic fitting formulas are only allowed to be used in the case
of ssubcr 6 s 6 1.0. When 0 6 s < ssubcr, linear interpolation of two points, C d ðsÞ ¼ s  C d ðssubcr Þ=ssubcr , is adopted, which not
only supplements the formula near zero-opening area but also guarantees Qm = 0 when s = 0.
Thereinto, ssubcr = 0.01 is for the HCB cage-type double-seat high-capacity sleeve regulator valve; for VBS soft-seat but-
terfly valve, ssubcr = 0.24 is set as the flow coefficients calculated by DN200, DN250 and DN400 fitted formulas at this point
follow the principle that the ratios among flow coefficients approximate the ratios among their flow sectional areas.

3.3. Improvement in mixture ratio model

The following three concepts shall be taken into consideration:

Mass-flow-rate mixture ratio: the mass-flow-rate ratio of oxidant to fuel injection, i.e., rQ m ¼ m_ in _ in
o =mf .
Mean mass mixture ratio (mass mixture ratio for short): the ratio of the mass of material originating in oxidant to the mass
of material originating in fuel, i.e., mr = mo/mf.
Mixture fraction: the ratio of the mass of material originating in fuel to the total mass of mixture, i.e., fmix = mf/(mo + mf).
102 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

For the modeling of the combustor device (the module of gas generator pipe), Ref. [26] uses the scheme based on the
mass-flow-rate mixture ratio which assumes that the mixture ratio of the combustion zone is only determined by the
mass-flow-rate ratio of oxidant to fuel injection, without consideration of residual gas taking part in combustion, and sup-
poses that the capacity of combustion gas generated in the combustion zone to do work, (RT)g, is related to r Q m and the static
pressure in the combustion zone, pc. Therefore, the thermodynamic calculation works out:
ðRTÞg ¼ f ðr Q m ; pc Þ ð9Þ

In order to get stable combustion and flow coupling calculation, the improved combustion-zone thermodynamic calcu-
lation adopts the mean mass mixture ratio of the zone, mrc, considering both the instantaneous mass flow rates of oxidant
and fuel injection and the equivalent mixture ratio effect of residual gas. In addition, the combustion-zone mixture fraction,
fc,mix, is used as the variable of mixture ratio in the calculation:
ðRTÞg ¼ f ðfc;mix ; pc Þ ð10Þ

It can be seen from the relational expression, fc,mix = 1/(1 + mrc), that the problem of infinite mean mass mixture ratio is
avoided and the range of mixture-ratio variable is converted into 01 of fc,mix from 0  þ1 of mrc.
By derivation at both sides of the definition of the mixture fraction to time, the equation of the combustion-zone mixture
fraction can be obtained:
dfc;mix 1  
¼ _ in
ð1  fc;mix Þ  m _ in
f  fc;mix mo ð11Þ
dt qc V c
The mixture fraction-based model can solve the problem of transient simulation stability in the Previous research. Fig. 2
shows the stability comparison of two kinds of thermodynamic calculation schemes when structural size parameters of the
38-component system are in strict accordance with the assembly and detail drawings. It can be seen from Fig. 2 that there is
sharp oscillation in the simulation curves of the mass-flow-rate mixture ratio-based model while the simulation curves of
the mixture fraction-based model have good stability. The small fluctuations in various curves of Fig. 2b result from the small
fluctuations of test data including fuel–oil flow rate which are used as the boundary conditions of the numerical system.

3.4. Gas turbine model

Fig. 3 shows the finite control volume grids of the gas turbine module which has a gas pipe inlet and a gas pipe outlet. At
its boundaries, there are boundary grids of connected components. The inlet and outlet passages of gas turbine are consid-
ered as two lumped-parameter volumes. Such state parameters as static pressure, density, total energy per unit volume
within each grid are instantaneously consistent and homogeneous. The middle passages of stationary and rotating blades
are uniformly regarded as a throttle resistance element.
Assumptions:

(1) The combustion gas is a completely frozen flow with invariable species and thermophysical properties.
(2) The combustion gas is adiabatic when flowing through the passages of stationary and rotating blades, without consid-
ering the heat transfer between gas and wall.
(3) The frictional and local losses, and the effects of axial heat conduction and gravitational field are ignored.

900 800 320 600 0.01

800 Tg 300
700
700 rQm Tg 500
0.008
280
600 fmix(0)
600
260 mr(0) 400
500 500
0.006
fmix (0)

240
mr (0)

400
Tg / K
Tg / K
rQm

400 300
300 220
0.004
200 300
200 200
100
200
180
0 0.002
100
100 160
-100

-200 0 140 0 0
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t /s t /s
(a) The scheme based on the mass-flow-rate mixture ratio (b) The scheme based on the mixture fraction

Fig. 2. Simulation stability comparison of two kinds of thermodynamic calculation schemes.


Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 103

The last state element The first state element


u n-1 of gas pipe u n AT1 u 1 A sb2 Arb1 u2 AT2 u 0 of gas pipe u 1
Asb1 psb2 Arb2
Gas pipe ρ n-1 En-1 ρ1 p 1 T1 p sb1 Tsb2 p rb2 ρ 2 p 2 T2 ρ 0 E0 Gas pipe
prb1
inlet p n-1 V1 Tsb1 Trb2 V2 p0 outlet
Trb1
x0 x1 Qm x4 x5
An-1 A n-1 An Stationary x2 x3 Rotating blade A0 A0 A1
blade MT nrot
Rotor outlet

Fig. 3. Finite control volume grids of gas turbine.

(4) Three major turbine characteristic parameters including turbine flow capacity parameter, efficiency and reaction
degree are obtained by fitting the characteristic data provided by Beijing Aerospace Measurement & Control Corp.
and a approximate method is used for missing data processing at the starting and stalling stages.
(5) The wall is rigid, with ignorance of the elastic deformation of the wall.

3.4.1. Inlet and outlet cavities

8
< dq1 ¼ 1 ðqin
1 u n An  Q m Þ
dt V1
Continuity equations : ð12Þ
: dq2 ¼ 1 ðQ m  qout u0 A0 Þ
dt V2 2

8  
< dp1 ¼ c pin out Q m c1 _
1 un An  p1 qout  V 1 q1 S1
dt V1
Energy equations : 1
ð13Þ
: dp2 ¼ c ðRT in Q  pout u A Þ  c1 q_ S
dt V2 2 m 2 0 0 V2 2 2


Varn1 if un P 0
The upwind scheme is used for boundary parameters of the state elements: Varin 1 ¼ ,
Var1 if un < 0
out Var1 if Q m P 0 in Varrb2 if Q m P 0 out Var2 if u0 P 0
Var1 ¼ , Var2 ¼ , Var2 ¼ , Var e {q, p, T}
Varsb2 if Q m < 0 Var2 if Q m < 0 Var0 if u0 < 0
Flow velocity calculation formula : u1 ¼ Q m =q1 AT1 ; u2 ¼ Q m =q2 AT2 ð14Þ

Static temperature calculation formula : T 1 ¼ p1 =ðq1 RÞ; T 2 ¼ p2 =ðq2 RÞ ð15Þ

 c  c1
c
c  1 2 c1 c1
Total pressure calculation formula : pt1 ¼ p1 1 þ Ma1 ; pt2 ¼ p2 1 þ Ma22 ð16Þ
2 2
   
c1 2 c1 2
Total temperature calculation formula : T t1 ¼ T 1 1 þ Ma1 ; T t2 ¼ T 2 1 þ Ma2 ð17Þ
2 2
pffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffi
Mach number calculation formula : Ma1 ¼ u1 = cRT 1 ; Ma2 ¼ u2 = cRT 2 ð18Þ

3.4.2. Turbine stationary blade nozzle throttling point


The mass flow rate equation of turbine stationary blade nozzle is given by:
8 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

 2  cþ1
>
>   c
>
> C sb Abmin psb1 RT1 c21 c psb2 c
 ppsb2
c psb2
> 2 c1
< sb1 psb1 sb1 psb1 cþ1
Qm ¼ rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi psb1 P psb2 P 0 ð19Þ
>
>  ccþ1   c
>
> c1
: C sb Abmin p 1
c 2 1 psb2
6 2
sb1 RT sb1 cþ1 psb1 cþ1

In the case of reverse flow, that is, when Eq. (19) meets 0 6 psb1 < psb2, the upstream and downstream parameters of the
stationary blade (e.g. psb1 and psb2) should exchange positions, and a minus sign should be added in the front of the mass flow
rate result.
In the equation, Csb is the flow coefficient of turbine stationary blade nozzle, standing for nozzle throttling characteristic;
Abmin is the smallest flow section area of turbine nozzle (Abmin = Asb2). In addition, Varsb1, Varsb2, Varrb1 and Varrb2 are the state
parameters of turbine stationary blade inlet and outlet, rotating blade inlet and outlet, respectively, Var e {q, p, T}.
The flow coefficient calculation scheme is given by:
104 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u  ccþ1
Q m np u
tR 2 1
C sb ¼ ð20Þ
Abmin c c þ 1

Turbine flow capacity parameter is obtained by fitting based on the characteristic data:
Q m np ¼ f ðku ; ppt Þ ð21Þ

3.4.3. Turbine stationary and rotating blade sections

(1) The state parameters of stationary blade inlet and outlet, rotating blade inlet and outlet
Stationary blade inlet parameters : psb1 ¼ p1 ; T sb1 ¼ T 1 ð22Þ

"  c1 #c1


c
 c1
p c psb2 c
Stationary blade outlet parameters : psb2 ¼ psb1 XT þ ð1  XT Þ rb2 ; T sb2 ¼ T sb1 ð23Þ
psb1 psb1

Rotating blade inlet parameters : prb1 ¼ psb2 ; T rb1 ¼ T sb2 ð24Þ

Dhb ðc  1ÞDhb
Rotating blade outlet parameters : prb2 ¼ p2 ; T rb2 ¼ T sb1  ¼ T sb1  ð25Þ
cp cR

(2) Without consideration of combustion gas heat transfer, the actual enthalpy drop of working medium flowing through
the turbine is:
1
Dhb ¼ ðu2rb2  u2sb1 Þ þ W T ð26Þ
2

In the equation; urb2 ¼ u2 ; usb1 ¼ u1 ð27Þ

(3) Turbine power equation


Turbine isentropic expansion work per unit mass working medium:
cRT t1 h c1 i
W ad ¼ cp ðT t1  T t2 Þ ¼ 1  ðpt2 =pt1 Þ c
c1
So the turbine work is calculated by: WT = gTWad.
The turbine shaft power is calculated by:
PT ¼ gT W ad Q m ð28Þ

(4) Turbine rotate speed:


nT ¼ nrot ð29Þ

(5) Turbine drive torque equation:


PT 30PT
MT ¼ ¼ ð30Þ
xT nT p

(6) Turbine reaction degree equation


The turbine reaction degree should be obtained by fitting based on corresponding turbine characteristic data:
XT ¼ fXT ðku ; ppt Þ ð31Þ
Due to lack of dynamic data, it only can be set as a constant according to its steady-state value: XT = 0.232

(7) The turbine efficiency is obtained by fitting based on the characteristic data provided:

gT ¼ fgT ðku ; ppt Þ ð32Þ

(8) Turbine total pressure ratio:


ppt ¼ pt1 =pt2 ð33Þ
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 105

(9) Median-location circular velocity coefficient of turbine blade:


ku ¼ uTM =acr ð34Þ
Given the turbine rotational angular velocity xT = 2pnT/60, so the turbine median-location circular velocity is calculated
by:
uTM ¼ xT r TM ¼ nT pDTM =60 ð35Þ
The critical sound velocity is calculated by:
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2c
acr ¼ RT ð36Þ
c þ 1 t1
where nT and nrot are the rotate speeds of turbine and rotor, respectively (unit: r/min, rpm); DTM is the turbine cascade aver-
age diameter, c is the ratio of specific heat capacities of gas, and R is the specific gas constant.

3.4.4. Wall heat transfer model


For the walls of the inlet passage (inlet cavity) and the outlet passage (outlet cavity), one of the radial one-dimensional
heat transfer model, the zero-dimensional heat transfer model and the adiabatic model can be chosen in line with the actual
condition of the temperature field. The flow through the passages of stationary and rotating blades is supposed to be
adiabatic.

3.5. Modularization modeling approach

Table 1 gives the module information on ten typical components arisen after modularization disassembly of the turbine
test rig gas system. Among them, fluid source, gas pipe, gas volume, gas valve, gas generator pipe, gas turbine and rotor are
seven basic modules, based on which other components are born through combination and development.

4. Numerical modeling of gas system

4.1. System model

The turbine test rig is a complex system, where there are numerous physical and chemical phenomena such as atomiza-
tion, ignition, combustion, mixing, rotation, flow and heat transfer, as well as flows of various media like fuel oil, water, nor-
mal-temperature air, high-temperature combustion gas and mixed gas. Furthermore, each flow is controlled by many
regulator valves of corresponding types. Considering the great difficulty in the entire system modeling, Ref. [26] only con-
ducts modeling and simulation on the main test system. In this paper, however, the turbine and the trailing pipelines are
also considered. The fuel–oil supply system, with experimental measuring data (time-parameters) of fuel volume flow rate
Qvf (converted to mass flow rate Qmf when used) and oil-source static temperature Tf as the inlet boundary conditions of the
system, can be simplified into a fluid source of flow rate and static temperature inlet. The ignition system can be simplified
into an event-triggered variable by a real-time control variable of ignition time established in corresponding module of the
BK-1 type heater. Although the rotor module has set up a scheme of rotate-speed dynamics model, the nrot value of the rotor,
considering that the rotate-speed test data are more accurate, is determined by directly adopting the time-varying test data
of the rotate speed in the case research in order to place emphasis on verifying the coincidence between the simulation
curves of flow-field and wall-temperature-field parameters and the test curves. As the gas system outlet of the rest rig ends
at the atmospheric environment, the test data of the atmospheric pressure p0 are used as boundary condition of the gas sys-
tem outlet, that is, static pressure outlet fluid source. F1-7 seven regulator valves are considered as gas valve module com-

Table 1
Modularization disassembly of turbine test rig gas system.

Module type name Type code Identifier Module attribute and status
Fluid source 00 FS Basis, improved
Gas pipe 02 GP Basis, improved
Gas volume 12 GVol Basis, improved
Gas mixing apparatus 17 GMA Combination, old
Gas steady apparatus 20 GSA Combination, improved
Gas valve 22 GV Basis, improved
Gas injector 32 GI Combination, improved
Gas generator pipe 35 GGP Basis and combination, improved
Gas turbine 41 GTurbo Basis, newly added
Rotor 42 Rotor Basis, newly added
106 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

ponents, and the spool throttling model and flow coefficient calculation scheme can be decided in accordance with the valve
type by the data and curves from Wuzhong Instrument Co. Ltd.
Fig. 4 shows the numerical simulation model of the turbine test rig gas system with structural size parameters in strict
accordance with the assembly and detail drawings. During the modular modeling, the system is divided into 2 fluid sources
(FS1-2), 20 gas pipes (GP1-20), 5 gas volumes (GVol1-5), 1 gas mixing apparatus (GMA1), 1 gas steady apparatus (GSA1), 9
gas valves (GV1-9), 1 gas injector (GI1), 1 gas generator pipe (GGP1), 1 gas turbine (GTurbo1) and 1 rotor (Rotor1), by which a
42-component numerical system is established.
The following demonstrates that the numerical system is applied to simulating the first 260 s rig regulating processing of
a test. The length (unit: m), outer diameter and thickness (unit: mm) of every pipe are as shown in Fig. 4. The flow field grids
of pipe-type components are divided by 100 mm/grid. The length along flow direction of the lumped parameter components
(such as gas valve and gas volume) should be two standard grid units, i.e. 200 mm. The total volume of the gas source (com-
posed of a number of high-pressure air bottles) is designated to be 2000 m3. When the simulation starts, a approximate dis-
tribution of initial flow field and wall temperature field is given, and the ignition is made after 10 s cold flow simulation. The
emphasis should be placed on comparing the 20–260 s results of simulation and experiment. In this way, the initial field is
not necessarily designed to be exactly the same as that in experiment. The test data are used for boundary conditions of two
fluid sources (FS1-2), and the scheme of time-speed data input file is adopted for the rotor-type component (Rotor1), that is,
directly using the test data in the form of time-parameters point list of Qmf (converting with a fuel–oil density of 780 kg/m3,
slightly regulated based on 830 kg/m3 in the Previous Case), p0 and nrot for the boundary conditions of the 42-component
numerical system.
The pipe-wall axisymmetric two-dimensional heat transfer model is used for the gas-pipe-module components and the
flow zone of gas-generator-pipe-module component; the pipe-wall radial one-dimensional heat transfer model for the mod-
ule components of gas valve, gas steady apparatus and gas injector, the inlet and outlet passages of gas-turbine-module com-
ponent; the pipe-wall zero-dimensional heat transfer model for the gas-volume-module components; the adiabatic model
for the gas-mixing-apparatus-module component, the combustion zone of gas-generator-pipe-module component, and the

Fig. 4. Numerical simulation model of the turbine test rig gas system.

Table 2
Throttling calculation schemes of GV1, GV4, GV5 and GI1 in 42-component gas system.

Simulation System Relative Valve spool throttling Flow coefficient Rated Flow coefficient Cd Remarks
case component opening model scheme flow or local loss
s n_Type_Cd coefficient coefficient
42-Component GV1 0.65 Pressure ratio-based 1 0.149 (0.0969) The scheme in
system case (1.0) injector orifice model parentheses is an
GV4 0.75 1 0.3 (0.225) equivalent scheme
(1.0)
GV5 1.0 Pressure difference- 3 780 767.91 or 0.05 The flow coefficient is
based dual-model proportional to the
solution rating
GI1 1.0 Pressure difference- 3 449.185 442.23
based injector orifice
model

Note: There is Qm = f(Cd  s) for the pressure ratio-based injector orifice model, so regulating the flow coefficient can be equivalent to regulating the relative
opening.
The flow coefficient is for the injector orifice model; the local loss coefficient is for the local loss model.
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 107

Table 3
Improved regulating time sequence design of seven regulator valves in 0–260 s.

Simulation Manually-recorded values (bold-faced and italic) and analyzed Sensor-measured values in the test (gauge Test time
time assumptive values (normal) on relative openings of seven pressure conversion with p0 as standard
regulator valves in numerical model atmospheric pressure)
GV3 GV2 GV5 GV6 GV7 GV8 GV9 FS1 FS2 Rotor1 10:16:00
DN100 DN200 DN150 DN200 DN200 DN250 DN400 780 kg/m3 Starting point
Relative opening Fuel–oil Static Rotate Load
flow rate pressure speed torque
t/s F1 F2 F3 F4 F5 F6 F7 Qmf kg/s p0/Pa nrot r/ MLoad/ Time/s
min Nm
0s 0.184 0 1.0 0 (0.25) 0 (0.341) 0s
0.38 0.90
10 s 0.013577 102362 14.063 1.6222 10
Ignition
62.9 0.01 0.012983 102376 23.907 1.4656 62.9
65 0.015 0.013425 102368 225.01 1.9579 65
67 0.018 0.013540 102373 393.76 2.2935 67
70 0.02 0.013304 102373 570.96 2.8809 70
71 0.03 0.013415 102363 702.21 3.2445 71
73 0.04 0.013100 102376 1074.4 4.0836 73
75 0.05 0.013726 102372 1486.9 4.0836 75
78 0.06 0.013388 102375 2101.9 5.2863 78
80 0.068 0.013385 102373 2339.1 5.9575 80
120 0.07 0.013476 102363 2329.8 5.4820 120
123 0.08 0.013369 102372 2661.6 6.4890 123
125 0.085 0.013956 102363 3264.5 7.2721 125
127 0.09 0.013129 102371 3827.9 8.6146 127
130 0.095 0.012856 102365 4693.3 10.573 130
133 0.10 0.013190 102380 5487.4 12.139 133
135 0.103 0.013820 102366 5569.9 14.516 135
138 0.106 0.012978 102372 5073.9 13.929 138
150 0.190 0.014285 102375 5043.0 13.789 150
155 0.200 0.013031 102373 5029.8 14.824 155
160 0.205 0.014078 102372 5031.7 15.495 160
162 0.208 0.013437 102373 5028.0 15.663 162
165 0.210 0.013044 102360 5014.8 16.027 165
170 0.215 0.013574 102372 5040.2 16.894 170
175 0.220 0.013706 102377 5209.8 18.404 175
180 0.230 0.014292 102376 5040.2 20.194 180
185 0.235 0.013848 102365 5016.7 20.809 185
190 0.240 0.014312 102373 5038.3 21.872 190
195 0.245 0.014354 102380 5052.3 23.271 195
202.3 0.250 0.015477 102371 5083.2 26.157 202.3
208.5 0.255 0.015872 102371 5023.7 28.487 208.5
210 0.260 0.016643 102365 4991.4 28.977 210
216 0.270 0.017418 102377 5018.6 31.326 216
220 0.272 0.018090 102370 5021.4 34.291 220
221 0.280 0.017809 102370 5009.2 34.962 221
222.5 0.285 0.017898 102374 5025.2 35.997 222.5
225 0.290 0.017658 102372 5043.0 37.983 225
227 0.300 0.017953 102373 5013.0 39.437 227
232 0.305 0.020933 102365 5008.3 42.402 232
233.6 0.310 0.020726 102374 5032.1 43.437 233.6
236 (0.317) 0.020288 102375 4998.0 44.528 236
239 0.305 0.11 0.022087 102373 5151.7 46.989 239
240 0.12 0.022802 102376 5095.5 55.184 240
241 0.13 0.023411 102373 5097.3 59.519 241
242.3 0.300 0.135 0.022700 102372 5078.3 66.249 242.3
244 0.140 0.022809 102362 5058.0 74.147 244
246 0.145 0.022741 102378 5038.3 81.084 246
254 0.159 0.022319 102373 5029.8 83.070 254
260 s No manually-recorded values on valve opening available at the     
following stage
Simulation        2494 s
end
10:57:34
Test end

stationary and rotating blades passages of gas-turbine-module component. The pipe-wall radial grid sequence numbers of
the components adopting two-dimensional or one-dimensional heat transfer model total 4. The heat transfer between outer
108 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

wall of various components and the external environment is considered as the simple superposition of convection heat
transfer and radiation heat transfer. The convection heat transfer coefficient is set as 6 W/(m2 K) in the sections of nor-
mal-temperature air and 8 W/(m2 K) in the sections of high-temperature combustion gas, and the ambient temperature is
279.4 K.
GV2 (F2) and GV3 (F1) adopt equal percentage flow characteristic fitting formula scheme of HCB cage-type double-seat
high-capacity sleeve regulator valve. GV7 (F5), GV8 (F6) and GV9 (F7) adopt flow characteristic fitting formula schemes of
DN200, DN250 and DN400 VBS soft-seat butterfly valves, respectively. The five regulator valves correspondingly employ
pressure difference-based injector orifice model. GV6 (F4) remains fully closed during the rig regulating test, so it is not
important to choose which formula to be used and it would be workable if only the flow value, worked out through the for-
mula under zero opening, is zero, where the same scheme is adopted as GV2 and GV3. The flow coefficients at the throttling
places of two orifice plates in the gas steady apparatus GSA1 are set to be 0.3 and 0.4, respectively.
GV1 (F0), GV4 (safety valve), and GV5 (F3) remain fully open in the test, so there is no change in the opening degree. Con-
sequently, the throttling conditions of these valves would not affect the variation trends of simulation curves. However, the
three fully opened valves are under high pressure level and slight differential pressure, where the traditional injector orifice
models often do not work, so the flow-rate simulation results at the throttling points tend to be distorted. Furthermore,
although the size of injection orifice in the gas injector GI1 is unchangeable, the coupling throttling effect of GI1, GV5
and GV6 directly determines the flow distribution relationship between the heater main line and the air mixing bypass
(As GV6 and GV5 remain fully closed and fully open in this rig regulating test, respectively, the effect does not exist here.
However, the openings of GV6 and GV5 change in other rig regulating tests.). Therefore, the throttling models of GI1 and
GV5 should match the throttling model of GV6. Based on the above considerations, the Previous debugging experience
and tremendous simulation debugging during this stage, the throttling calculation schemes shown in Table 2 are designed
for the three full-open valves and the gas injector.

4.2. Regulating time sequence design

By comprehensive analyses on various stages of 0–260 s test curves and based on numerous simulation pilot calculations,
the improved regulating time sequence design of seven regulator valves in 0–260 s is made as shown in Table 3. Similar to
the Previous design [26], the data in the table are divided into two categories: the relative opening data of F1-7 seven reg-
ulator valves in the left part and the test data of fuel–oil flow rate, atmospheric pressure, rotor speed and load torque in the
right part. Among the left data, the data in bold and italic are opening data of several state points from manual records based
on pressure gauge observation; the data in round brackets are from manual records of the test but not adopted for the time
sequence design in the simulation; the data in normal form (non-bold and upright) are opening variation values speculated
by analysis. All data in the right part are values measured by corresponding sensors in the test.

5. Results and discussion

Fig. 5 shows the experimental curves of static pressures, total pressures, static temperatures, total temperatures, wall
temperatures, fuel–oil flow rate and turbine rotate speed in the turbine test rig system which are measured by sensors in
this rig regulating test. The data acquisition frequency stands at 1 Hz, i.e. a data point per second. The experimental curves
of 2500 s have recorded the variations of state parameters at various measuring points in the speed-up process of the tur-
bine from starting rotation to rotate speed of 27,500 r/min by stages after ignition and in the slowdown process from the
highest speed by stages. This paper has conducted numerical simulation on the first 260 s of the rig regulating process as
shown in Fig. 6. Locations of the measuring points are shown by the schematic diagram of the turbine test rig system in
Fig. 1. Some measuring points are designed for measuring the parameters of the same location, but only typical measuring
curves are picked up in charting. For example, Tt13 and Tt14 are total temperatures measured at the third and the fourth Tt1
measuring points. It can be seen that the curves measured by different sensor measuring points at the same location are not
certainly identical.
Fig. 7a–p shows comparison between the simulation results and the experimental measurements. The simulation data
output frequency is 10 Hz, namely, a data point per 0.1 s. The subheads of figures are named after the respective measuring
points. The locations of measuring points are shown in Fig. 1, and the locations of the simulation data points are shown by
the numerical simulation model of the gas system in Fig. 4. In Fig. 7a–k, m and o, the parameter curves marked with symbols
like triangle, circle, etc. and the turbine speed nT curve are experimental data while the parameter curves without symbols or
with the component name (such as GP1 and GGP1) as prefix are simulation data of various component grids in Fig. 4 cor-
responding to the sensor-measuring points in Fig. 1. In names of simulation curves, 0, n/2 and n  1 in flow direction stand
for the entrance grid, middle grid and exit grid of pipe flow field, respectively while 0 and nw in radial direction stand for the
internal pipe-wall grid and external pipe-wall grid, respectively; qrho(i) is the convective heat flux density from the fluid in
the in-pipe grid with i as serial number to the interior wall; qrho2(i, 1) and qrho2(i, 2) are the heat flux densities of convective
and radiation heat transfer from the exterior wall grid with i as axial serial number to the ambient atmosphere, respectively.
The simulation on the 38-component numerical system obtained results, as shown in Ref. [26] Fig. 8 (hereinafter referred
to as Previous Fig. 8), relatively accordant with the test results as a whole, but there are still five obvious weaknesses: (1)
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 109

1.8 0.03 30000 580 140 30000

1.6 540 120


1.4
p91 500
100
1.2 p31 0.02 20000 20000
p51 460

Q mf / (kg/s)
pt7

nT / (r/min)

nT / (r/min)
p,p t /MPa

1 80
pt8

Tt / K

M / (Nm)
Q mf 420
0.8 nT Tt91
60
ptI1 Tt14
ptIII9* 380 Tt33
0.6 0.01 10000 10000
pII3 Tt51
Tt7 40
340
0.4 Tt8
nT
300 MLoad 20
0.2
TtI1
TtIII1
0 0 0 260 0 0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500 0 250 500 750 1000 1250 1500 1750 2000 2250 2500
t/s t/s
(a) Static pressures, total pressures and fuel-oil mass flow rate (b) Total temperatures, load torque and turbine rotate speed
(converting with a fuel-oil density of 780kg/m3)

460 30000 580 30000

440
540

420
500
400
20000 20000
Tt91 460
380 Tw10

nT / (r/min)
n T / (r/min)
T , Tw / K

Tt33
Tt1 / K

360 Tt36 420


Tw4
340 Tt51
380
Tw6 10000
10000 Tt11
320 Tf1
Tt12
T0 340 Tt14
300 nT
Tt15
300
nT
280

260 0 260 0
0 250 500 750 1000 1250 1500 1750 2000 2250 2500 0 250 500 750 1000 1250 1500 1750 2000 2250 2500
t/s t/s
(c) Static temperatures, total temperatures and wall temperatures (d) Heater-downstream total temperatures by different
sensor measuring points at the same location

Fig. 5. Experimental measured results of the turbine test rig.

Previous Fig. 8b: although the simulation curve of heater-downstream total temperature is well consistent with the test
curves on the whole, it cannot accord well with the test curve in some local change details (especially in 140–240 s); (2) Pre-
vious Fig. 8c, e, and f: although there is evident thermal stratification phenomenon along combustion-gas flow direction
among the simulation results of various pipeline components behind the heater, the simulation curves of total temperatures
at other three measuring points are obviously higher than the test curves (Tt51, Tt52, Tt33, Tt36 and Tt7) except that the sim-
ulation curve of heater-downstream total temperature falls in the scope of two sensor measuring curves (Tt13 and Tt14);
(3) Previous Fig. 8d: there are big magnitude differences in the period of 20–60 s between the static pressure simulation
curves of GP17 grids and the test curves of main-line measurement nozzle upstream and downstream static pressures.
Around 240–260 s, the test curves begin to drop while the simulation curves fail to drop obviously. Furthermore, the sim-
ulation curves are not as smooth as the test curves in some local change details at the regulator valve action stages; (4) Pre-
vious Fig. 8e and f: the total-temperature simulation curves of GP17 and GP18 grids climb fast in 60–80 s as GV8 opens,
while Tt3 and Tt7 test curves climb slowly. Meanwhile, there are magnitude differences between them; and (5) Previous
Fig. 8i and k: there are magnitude differences between the temperature simulation curves of GP16 and GP17 wall grids
and the test curves.
It can be seen from Fig. 7 that many improvements have been achieved in the simulation results of 42-component gas
system (hereinafter referred to as this Case) by comparison with the Previous simulation results, which are described one
by one as follows against abovementioned five weaknesses:

(1) Fig. 7b shows that the simulation curve of heater-downstream (GGP1 middle section) total temperature, after elabo-
ration and improvement of regulating time sequence and slight adjustment of fuel–oil density as shown in Table 3,
especially after elaboration and improvement of the total air flow-rate regulator valve GV3 (F1), not only falls com-
pletely between the two test curves of total temperatures (Tt13 and Tt14) measured by two sensors at this position
on the whole but also accords well with the test curves in some local change details (especially in 140–240 s).
110 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

0.55 0.03 90 580


Tt91
6000
0.5 Tt13
540
75 Tt14
0.45
p31 500 Tt33
0.4
p51 0.02 60 Tt36
4000 460
p , pt / MPa

0.35 pt7 Tt51

Qmf / (kg/s)

M / (Nm)
nT / (r/min)

Tt / K
pt8 Tt7
0.3
Qmf
45 420
Tt8
0.25 nT
380 TtI1
MLoad 0.01 30
2000 TtIII1
0.2 ptI1
340
0.15 ptIII9*
pII13 15
300
0.1

0.05 0 0 0 260
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t/s t/s
(a) Static pressures, total pressures, fuel-oil mass flow rate, (b) Total temperatures
turbine rotate speed and load torque

460 Tt91 580 6000


Tw10 6000
440 Tt33
560
Tt36 5000
420
Tw4 5000 Tt11
Tt51 540 Tt12
400
Tw6 Tt13 4000
Tf1 4000 Tt14

n T / (r/min)
380
n T /(r/min)

520 Tt15
T0
T,Tw /K

Tt1 / K nT
360 nT 3000
3000
500
340
2000
320 2000
480

300
1000 1000
460
280

260 0 440 0
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t/s t/s
(c) Static temperatures, total temperatures and wall temperatures (d) Heater-downstream total temperatures by different
sensor measuring points at the same location

Fig. 6. Experimental measurements in 0–260 s.

(2) According to Fig. 7b, d, f and g, this Case fails to make up for the deficiency that the simulation curves of total tem-
peratures at three measuring points behind GVo15 are obviously above the test curves, but the total temperature sim-
ulation curve magnitudes of GP16, GP17 and GP18 evidently decline compared with the Previous simulation results
and thereby are closer to the test curves of Tt5, Tt3 and Tt7, respectively. It proves that the description of thermal strat-
ification in flow direction among various pipeline components behind the heater by the 42-component gas system
case becomes more accurate. Table 4 shows comparison of total temperature test and simulation values at four mea-
suring points in 30 s and 260 s.
(3) According to Fig. 7e, by improvement and elaboration of regulating time sequence, especially elaboration of the tur-
bine-inlet regulator valve GV8 (F6), adjustment of opening of exhaust-bypass regulator valve GV7 (F5) and elaboration
of the total air flow-rate regulator valve GV3 in 150–250 s, the variation trends of static pressure simulation curves of
GP17 grids are well in line with the test curves in the whole period of 20–260 s after the heater igniting and the mag-
nitudes in 20–60 s are also close to the test curves, giving a sound solution to the problem in the Previous GP17 sim-
ulation results. Table 5 shows comparison of static pressure test and simulation values at p31 measuring point in 50 s,
100 s and 235 s.
(4) As shown in Fig. 7f and g, although this Case fails to cover the shortage that the simulation curves of total tempera-
tures at Tt3 and Tt7, with GV8 from fully closed status to opening gradually during 60–80 s, rise more quickly than the
test curves, the magnitudes of total temperature simulation curves at various locations in the turbine main line obvi-
ously drop and then are closer to the test curves after elaboration and improvement of regulating time sequence and
slight adjustment of fuel–oil density. In addition, after elaboration of GV8 in 60–70 s, the total temperature simulation
curves climb sharply exactly at the time when the test curves do.
(5) According to Fig. 7m–p, as the relative opening of GV7 is adjusted to 38% in this Case from 25% in the Previous Case,
the combustion-gas flow rate through the exhaust bypass increases while the combustion-gas flow rate through the
turbine main line decreases, and the scouring effect of high-temperature gas on the exhaust bypass gets stronger
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 111

1.64
p91 275
p101 580 10000
Tt91
Tt92
GP1-p(n-1) 560
1.63 GP1-Tt(n-1) 274
8000
540

1.62 273
520
6000

n T / (r/min)
p / MPa

Tt / K

Tt / K
1.61 272 500

4000
480
1.6 271

460
Tt13 2000
1.59 270 Tt14
440
nT
GGP1-Tt(n/2)
1.58 269 420 0
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t/s t/s
(a) Static pressures and total temperatures of inflow (b) Heater-downstream total temperatures
measurement nozzle upstream and downstream

0.11 10000 500 10000

490

8000 480 8000


0.105
470

6000 460 6000


nT / (r/min)

n T / (r/min)
p / MPa

Tt / K

0.1 450

4000 440 4000


p51 Tt51
430
p61
0.095 Tt52
nT 2000 420 2000
GP16-p(0) nT
GP16-p(n/2) 410 GP16-Tt(n/2)
GP16-p(n-1) GP16-Tt(n-1)
0.09 0 400 0
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t/s t/s
(c) Static pressures of bypass measurement nozzle (d) Total temperatures at bypass measurement section
upstream and downstream

0.6 10000 500 10000

480
0.55
p31
p41 8000 460 8000
nT
0.5 GP17-p(n/2) 440
GP17-p(n-1)
6000 420 6000
n T / (r/min)
n T / (r/min)

0.45
p / MPa

Tt / K

400
0.4
4000 380 4000

0.35 360 Tt33


Tt36
2000 340 nT 2000
0.3 GP17-Tt(n/2)
320
GP17-Tt(n-1)
0.25 0 300 0
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t/s t/s
(e) Static pressures of main-line measurement nozzle (f) Total temperatures at main-line measurement section
upstream and downstream

Fig. 7. Comparison of simulation results with experimental measurements.


112 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

Fig. 7 (continued)
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 113

420 900000 GP16-Re(0) 26000


Tw6
GP16-Tw(0.0) GP16-qrho(0) 24000
410 800000 GP16-Re(n-1)
GP16-Tw(n/2.0) 22000
GP16-qrho(n-1)
400 GP16-Tw(n-1.0) GP16-qrho2(0.1)
700000 20000
GP16-Tw(n-1.nw) GP16-qrho2(n-1.1)
390 18000
GP16-qrho2(0.2)

qrho, qrho2 / (W/m 2)


600000 GP16-qrho2(n-1.2)
380 16000

500000 14000
Tw / K

370

Re
12000
360 400000
10000
350 8000
300000
340 6000
200000 4000
330
2000
100000
320 0
310 0 -2000
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t/s t/s
(m) Wall temperatures at bypass measurement section
(n) Simulation results of Reynolds numbers, convective heat flux
densities of interior wall, convective and radiation heat flux
densities of exterior wall at GP16 entrance and exit grids

GP17-Re(0)
GP17-qrho(0)
365 700000 GP17-Re(n-1) 10000
Tw4 GP17-qrho(n-1)
GP17-qrho2(0.1) 9000
360 GP17-Tw(0.0) GP17-qrho2(n-1.1)
GP17-Tw(n/2.0) 600000 GP17-qrho2(0.2)
GP17-qrho2(n-1.2) 8000
355
GP17-Tw(n-1.0)
GP17-Tw(n-1.nw) 7000
500000

qrho, qrho2 / (W/m2)


350
6000
400000
345 5000
Tw / K

Re

340 300000 4000

3000
335
200000
2000
330
1000
100000
325 0

320 0 -1000
0 20 40 60 80 100 120 140 160 180 200 220 240 260 0 20 40 60 80 100 120 140 160 180 200 220 240 260
t/s t/s
(o) Wall temperatures at main-line measurement section (p) Simulation results of Reynolds numbers, convective heat flux
densities of interior wall, convective and radiation heat flux
densities of exterior wall at GP17 entrance and exit grids
Fig. 7 (continued)

Table 4
Comparison of total temperature test and simulation values at four measuring points (Unit: K).

Test Time Tt13/Tt14 Tt51/Tt52 Tt33/Tt36 Tt7


30 s 501.86/556.99 434.47/411.16 354.13/336.21 326.07
260 s 465.90/477.77 439.38/420.72 435.84/411.79 401.13
Previous Time GGP1 middle-grid total GP16 exit-grid total GP17 exit-grid total GP18 exit-grid total
Case temperature temperature temperature temperature
30 s 536.21 478.15 342.87 325.80
260 s 482.89 454.17 451.49 446.74
This Case Time GGP1 middle-grid total GP16 middle-grid total GP17 exit-grid total GP18 exit-grid total
temperature temperature temperature temperature
30 s 523.90 472.20 334.29 318.25
260 s 479.67 454.64 439.82 433.18

Note: As the 42-component system expands in downstream boundary compared with the 38-component system, the spatial location of GP16 middle grid in
this Case is close to that of the Previous GP16 exit grid.
114 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

Table 5
Comparison of static pressure test and simulation values at p31 measuring point (Unit: MPa).

Location/time 50 s 100 s 235 s


Test p31 0.465 0.370 0.540
Previous Case GP17 middle and rear grids 0.554 0.378 0.468
This Case 0.472 0.376 0.546

while the scouring effect on the turbine main line gets weaker. Meanwhile, the combustion temperature in the heater
declines somewhat compared with the Previous Case, which weakens the wall temperature rise caused by the scour-
ing effect. Therefore, by comparison with the Previous simulation curves, the temperature-rise rates of the simulation
curves of GP16 wall axial initial and middle grids (Please note that as the 42-component system expands in down-
stream boundary compared with the 38-component system, the spatial locations of GP16 middle and exit grids move
toward downstream compared with the Previous Case and the spatial location of GP16 middle grid in this Case is close
to that of GP16 exit grid in the Previous Case.) only increase slightly while the temperature-rise rates of the simulation
curves of GP17 wall grids decrease evidently and Fig. 7o shows that the wall-temperature simulation curves of GP17
middle and rear sections are well in line with the test curves.

With regard to the five obvious weaknesses of the Previous simulation, the above gives analysis on the obvious improve-
ments in the simulation results of the 42-component system compared with those of the 38-component system. The below
analyzes the coincidences one by one between the simulation and the test in each figure of Fig. 7.
The comparison in Fig. 7a, b, d, e and f has been described and analyzed in Ref. [26], and the major differences between
the simulation results of this Case and the Previous simulation results are analyzed thereinbefore, which will not be de-
scribed again here.
As shown in Fig. 7c, as the bypass measurement nozzle downstream is in the atmospheric environment with constant
pressure, both simulation curves and test curves show that the static pressures of bypass measurement nozzle upstream
and downstream are equivalent to the local atmospheric pressure, around 0.10 MPa in 0–260 s.
According to Fig. 7g, the simulation curves of total pressures and total temperatures at GP18 grids nearly coincide with
the test curves of total pressure and total temperature at the turbine inlet respectively and basically remain stable (slightly
fall in fact) in 0–60 s. Around 65–70 s when GV8 gradually opens to a certain degree, the total temperature simulation curves
fast climb but the test curve slowly climbs, and thereafter there is a big difference in magnitude between them. In addition,
in the Previous Case, the turbine inlet total pressure test data are directly used for simulation as outlet boundary condition,
so the total pressure simulation curve is coincident with the test curve. However, in this Case, due to the boundary expansion
of the 42-component numerical system, turbine inlet total pressure measurement point becomes a internal point of the sys-
tem, and then there is no longer direct relation between simulation data and test data. Fig. 7g shows the total pressure sim-
ulation curves basically coincide with the test curve in terms of the variation trend, only with magnitude much higher than
the test curve. Similar to the reasons for the magnitude difference between simulation and test in Fig. 7e, there are still dif-
ferences between the opening-flux relationships calculated by throttling models of various regulator valves and the relation-
ships in the actual rig regulating experiment. Meanwhile, the incompleteness of the manual recording test data for time
sequence, with which the regulating time sequence is designed in accordance, is another factor.
According to Fig. 7g, h and j, the turbine I-I and III-III sections are located at turbine inlet section and outlet section,
respectively, so their total temperature test curves are between the total temperature test curves of turbine inlet and outlet,
and their total pressure test curves, due to small pressure losses, are equivalent to the total pressure test curves of turbine
inlet and outlet, respectively. Similar to the situation in Fig. 7g, in Fig. 7h, the simulation curves and test curves of total pres-
sures and total temperatures basically remain stable (slightly fall in fact) in 0–60 s. Around 65–70 s when GV8 gradually
opens to a certain degree, the total temperature simulation curves of turbine inlet-cavity grid and outlet-cavity grid fast
climb but the total temperature test curves of turbine I-I and III-III sections slowly rise, and thereafter there is difference
in magnitude between them. The total pressure simulation curve of turbine inlet cavity basically coincides with the total
pressure test curve of turbine I-I section in terms of the variation trend, only with magnitude much higher than the test curve
for the same reasons as Fig. 7e and g. The turbine outlet-cavity total pressure simulation curve completely falls among the
three total pressure (ptIII2 ; ptIII3 ; ptIII4 ) test curves measured at this location by three sensors of III-III section. Furthermore, as
also indicated by Fig. 7h, the total temperature or total pressure curves of turbine I-I and III-III sections separate gradually
after the turbine starts rotation, and the gap increases with time, which proves that the turbine drives the load device to
rotate by getting energy from the high-temperature combustion-gas flow field, i.e. the role of the turbine is converting com-
bustion-gas energy into mechanical work for driving load rotation.
According to Fig. 7i, as this Case directly uses time-speed test data in the form of point list as the boundary condition, the
turbine rotate speed curve is coincident with the test curve. The turbine drive torque curve calculated by turbine torque
equation basically accords with the test curve of load torque in terms of variation trend, but the magnitude of the drive tor-
que curve is much higher than that of the load torque curve. However, there should be a small difference between them in
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 115

fact. It can be concluded that the gas turbine numerical model has acceptable qualitative accuracy but inadequate quanti-
tative accuracy.
According to Fig. 7j, similar to the situation in Fig. 7g, the total pressure and total temperature simulation curves of GP19
grids nearly coincide with the test curves of total pressure and total temperature at the turbine outlet respectively and basi-
cally remain stable in 0–60 s. Around 65–70 s when GV8 gradually opens to a certain degree, the total temperature simula-
tion curves fast climb but the test curve slowly climbs, and thereafter there is a big difference in magnitude between them.
As the opening of GV9 (F7) in Table 3 is adjusted to 90% in this Case from 34.1% of manual record based on a large number of
pilot calculation results, the total pressures at various grids of GP19, the pipe before GV9, are close to the atmospheric pres-
sure and are consistent with the total pressure test curve of turbine outlet. However, if the opening of GV9 is set as 34.1% of
manual record, the total pressure simulation curves of all GP19 grids would be much higher than the test curve.
Fig. 7k–p shows the heat transfer conditions between in-pipe flow field or external environment and pipe wall at three
pipe-wall-temperature measuring points as well as the comparison between the pipe-wall-temperature simulation results
and the experimental measurements.
According to Fig. 7k and l, different from the pipes which the combustion gas flows through, the densities of heat transfer
between GP1 flow-field grids and the interior wall, the exterior wall and the environment are negative in 20–260 s, which
shows that the in-pipe air absorbs heat from the pipe wall and the wall absorbs heat from the external environment. Com-
pared with the Previous simulation results, the heat transfer effects of the wall to the gradually cooling in-pipe air and the
external environment to the wall are enhanced slightly. Fig. 7l shows that the heat flux densities of convection heat transfer
and radiation heat transfer between GP1 exterior-wall grids and the external environment stand at the level of 51 W/m2 and
8 W/m2 in the simulation process of 20–260 s, respectively, which means that the total heat flux density between the exte-
rior wall and the external environment is about 59 W/m2, higher than the convection heat flux density of 43 W/m2 in the
Previous Case which does not take the radiation heat transfer effect into consideration.
As shown in Fig. 7m and n, the densities of heat transfer between the flow-field grids of such pipeline components as
GP16 which the combustion gas flows through and the interior wall, the exterior wall and the environment are positive
in 20–260 s, which shows that the combustion gas releases heat to the wall and the wall releases heat to the external envi-
ronment. Compared with the Previous simulation results, the heat transfer effects of the combustion gas to the wall and the
wall to the external environment are enhanced. Fig. 7n shows that the heat flux densities of convection heat transfer and
radiation heat transfer between GP16 exterior-wall grids and the external environment gradually increase as the pipe-wall
temperatures get higher in the simulation process of 20–260 s. Table 6 shows comparison of heat flux densities of convection
and radiation heat transfer between GP16 exterior-wall axial initial/last grids and the external environment in Previous Case
and this Case in 30 s and 260 s.
According to Fig. 7o and p, GP17 flow-field grids, similar to the situation in GP16, release heat to the wall and the wall
releases heat to the external environment. Different from GP16, by comparison with the Previous simulation results, the heat
transfer effect of combustion gas in GP17 to the wall evidently weakens while the heat transfer effect of the wall to the exter-
nal environment obviously intensifies. Fig. 7p shows that the heat flux densities of convection heat transfer and radiation
heat transfer between GP17 exterior-wall grids and the external environment do not change too much during 20–60 s when

Table 6
Comparison of heat flux densities of heat transfer between GP16 exterior-wall axial initial/last grids and the external environment in Previous Case and this
Case (Unit: W/m2).

Time (s) Simulation case GP16 exterior-wall axial initial grid and the external GP16 exterior-wall axial last grid and the external
environment environment
Convection heat flux density Radiation heat flux density Convection heat flux density Radiation heat flux density
30 Previous Case 320.30 Ignored 281.57 Ignored
This Case 409.44 66.325 371.96 58.809
260 Previous Case 779.62 Ignored 539.60 Ignored
This Case 1046.4 251.94 747.97 150.16

Table 7
Comparison of heat flux densities of heat transfer between GP17 exterior-wall axial initial/last grids and the external environment in Previous Case and this
Case (Unit: W/m2).

Time (s) Simulation case GP17 exterior-wall axial initial grid and the external GP17 exterior-wall axial last grid and the external
environment environment
Convection heat flux density Radiation heat flux density Convection heat flux density Radiation heat flux density
30 Previous Case 278.45 Ignored 271.87 Ignored
This Case 358.09 56.109 349.57 54.471
260 Previous Case 592.78 Ignored 500.48 Ignored
This Case 630.01 117.48 524.65 91.511
116 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

GV8 is unopened, but increase as the wall temperatures rise in the simulation process of 80–260 s when GV8 is opened grad-
ually. Table 7 shows comparison of heat flux densities of convection and radiation heat transfer between GP17 exterior-wall
axial initial/last grids and the external environment in Previous Case and this Case in 30 s and 260 s. The weaker heat absorp-
tion of the interior wall and the stronger heat release of the exterior wall contribute to better coincidence of the wall-tem-
perature simulation curves of GP17 middle and rear sections than the Previous simulation curves with the test curve.

6. Conclusions

Compared with the Previous research [26], the research in this paper adds two brand-new modules – gas turbine and
rotor, improves the algorithm of module with chemical reaction to solve the problem in the stability of simulation, adds
the radiation heat transfer algorithm based on the algorithm of natural convection heat transfer between wall and environ-
ment to enhance wall heat transfer, adds the regulator valve F7 and improves the spool throttling models, modifies the mod-
el size of the heater inlet segment according to its part drawing, improves the way of data conversion from gauge pressure to
absolute pressure and slightly regulates the density and initial static temperature of the fuel oil (RP-3 kerosene) in line with
engineering data, improves the regulating time sequence according to manually-recorded data through a large number of
simulation pilot calculations, and finally obtains the 42-component numerical system which can describe the gas system.
The simulation on this system can conclude:

(1) The 42-component system comprehensively makes up the five weaknesses of the Previous simulation and has
achieved the following progresses: the heater-downstream total temperature simulation curve coincides well with
the test curves both on the whole and on local change details; the description on the thermal stratification phenom-
enon in the flow direction among the pipeline components after the heater is more accurate; the static pressure sim-
ulation curves of GP17 grids coincide better with the test curves of main-line measurement nozzle upstream and
downstream static pressures in terms of magnitude and variation trend during 20–260 s after ignition of the heater;
the simulation curves of total temperatures at various locations of the turbine main line are closer to the test curves in
terms of magnitude and the time of initial sharp climbing; the simulation curves of various pipe-wall temperatures
coincide well with the test curves. In the scope of measuring points which the 42-component system and the 38-com-
ponent system both have, there are still two major differences between the simulation data and the test data: the total
temperature simulation curves of three measuring points behind GVol5 are obviously higher than the test curves; Tt3
and Tt7 total temperature simulation curves climb quickly with GV8 from fully closed status to opening gradually
while the test curves climb slowly. In the follow-up work, the idea of fully integrated zooming will be adopted to inte-
grate high-dimensional models of the key components such as the BK-1 type heater with low-dimensional test rig
model so as to further improve the description accuracy of the thermal stratification phenomenon.
(2) The 42-component system contains the numerical description on such test rig downstream parts as turbine and rotor,
with outlet ending at constant-pressure atmospheric environment. From the macro perspective, therefore, the biggest
achievement of this paper lies in the point that the 42-component system is a relatively self-consistent and compre-
hensive gas numerical system which is less dependent on test data. Secondly, this paper has solved the problem of
simulation stability and reduced unnecessary iterative parts by algorithm improvement. Hence, another biggish
achievement of this paper is that, under the circumstances of the structure and size of the gas system model strictly
in line with the drawings, the stability of multi-field coupling simulation is quite sound and the computation speed
does not decline but increases with structure expansion. It is on the basis of comprehensive and accurate structure
as well as stable and efficient computation that evident improvement has been made on coincidence between simu-
lation and test results by such means as elaboration and improvement of regulating time sequence, optimization and
improvement of spool throttling and turbine characteristic models, deepening and expansion of wall heat transfer
model, rational utilization of test data and so on.
(3) According to experience of debugging and improvement, the modeling of module with chemical reaction, the spool
throttling modeling of various regulator valves (Meanwhile, the throttling models in corresponding branches need
to match with each other.), the modeling of the turbine characteristics and the modeling of wall heat transfer are four
key factors which affect simulation accuracy, and the coupling modeling of flow and combustion is the key factor
which affects simulation stability.
(4) Although there are many uncertainties in the simulation process of the 42-component system, such as the problem of
spool opening-flux relationships of various regulator valves, the matching problem of branch throttling models and
the processing problem of turbine characteristic models, the variation trends of various state parameters in the sim-
ulation case which is finally obtained are coincident with the test curves as a whole, and the simulation curves at most
measuring points are well coincident with the test curves during 20–260 s after the heater ignites, which proves that
the analyses on test data and simulation pilot-calculation results are reasonable, the case setting and the algorithm
improvement are effective, and the established numerical system are accurate and effective in some degree.
(5) The simulation curves which are greatly different from the test curves are centralized around the turbine. For example,
under the circumstances of rotate speed test data being boundary condition, the simulation curve of turbine rotate
speed overlaps with the test curve, and the simulation curve of drive torque is in accord with the test curve of load
Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118 117

torque in terms of variation trend while the magnitude is 6–1.6 times (the latter the timeline gets, the smaller the time
becomes) as high as the test curve after GV8 opens. In fact, the gap between them should be very narrow. The big dif-
ference attributes to the data deficiency of three turbine characteristic parameters including turbine flow capacity
parameter, reaction degree and efficiency as well as the shortcoming of the gas turbine numerical model. More accu-
rate simulation needs to, based on a more precise and comprehensive gas turbine model, make characteristic tests
with combustion gas as medium on working conditions of different turbine speeds and loads to get corresponding
transient characteristic curves.
(6) The test curves measured by different sensors at the same location are not completely consistent, so the reliability
assessment needs to be made on dynamic measurement of experimental data.

Acknowledgments

This work is financially supported by the National Natural Science Foundation of China (No. 11101023) and the China
Scholarship Council (No. 201203070237). The authors would like to thank Dr. Peng Xu from the research team for his guid-
ance in software interface, Professor Haixing Wang from BUAA and the anonymous referees for their revision advice, Dr.
Huasheng Wang from Queen Mary, University of London and Ms. Jihong Zhao for their language support.

References

[1] K. Liu, Y.L. Zhang, A study on versatile simulation of liquid propellant rocket engine systems transients, in: AIAA 2000-3771, 2000.
[2] H.J. Liu, Investigation on Static Characteristics and Dynamic Response Characteristics of Staged Combustion Cycle Rocket Engine, Ph.D. Thesis, Shanxi
Engine Design Institute, Xi’an, China, 1998 (in Chinese).
[3] J.R. Mason, R.D. Southwick, Large Liquid Rocket Engine Transient Performance Simulation System (Final Report), NASA CR-184099, 1990.
[4] D. Nguyen, A. Martinez, Versatile engine design software, in: AIAA 93-2164, 1993.
[5] C. Goertz, A modular method for the analysis of liquid rocket engine cycles, in: AIAA 95-2966, 1995.
[6] A.K. Majumdar, J.W. Bailey, P. Schallhorn, et al., A generalized fluid system simulation program to model flow distribution in fluid networks, in: AIAA
98-3682, 1998.
[7] M. Sozen, A.K. Majumdar, A novel approach for modeling chemical reactions in generalized fluid system simulation program, in: AIAA 2003-4467,
2003.
[8] A. Tarafder, S. Sarangi, CRESP-LP: a dynamic simulator for liquid-propellant rocket engines, in: AIAA 2000-3768, 2000.
[9] N. Yamanishi, T. Kimura, M. Takahashi, et al., Transient analysis of the LE-7A rocket engine using the rocket engine dynamic simulator (REDS), in: AIAA
2004-3850, 2004.
[10] IMAGINE S.A., AMESim4.2 User Manual, 2004.
[11] V. Leudiere, G. Albano, G. Ordonneau, et al., CARINS: a versatile and flexible tool for engine transient prediction development status, ISTS 2004-t-14,
24th International Symposium on Space Technology and Science, Miyazaki, Japan, 2004.
[12] Flowmaster International Ltd., Flowmaster version 6.2 user guide, Flowmaster International Ltd., Towcester, 2006. <www.flowmaster.com>.
[13] J.E. Crowley, J.R. Olds, Co-OPT: a constrained optimizer for propulsion tools, in: AIAA 2005-4126, 2005.
[14] K.W. Nelson, S.P. Simpson, Engine system model development for nuclear thermal propulsion, in: AIAA 2006-5087, 2006.
[15] R.D. Salvo, S. Deaconu, A.K. Majumdar, Development and implementation of non-Newtonian rheology into the generalized fluid system simulation
program (GFSSP), in: AIAA 2006-4869, 2006.
[16] A.C. LeClair, A.K. Majumdar, Computational model of the chilldown and propellant loading of the space shuttle external tank, in: AIAA 2010-6561,
2010.
[17] X.Y. Ren, J.L. Wang, X.L. Zhao, Simulation of turbofan engine main fuel control system based on AMESim, J Aerosp Power 25 (12) (2010) 2816–2820 (in
Chinese).
[18] V. Leudiere, P. Supie, M. Villa, KVD1 engine in LOX/CH4, in: AIAA 2007-5446, 2007.
[19] H.W. Ng, F.L. Tan, Simulation of fuel behaviour during aircraft in-flight refueling, Aircraft Eng. Aerosp. Technol. 81 (2) (2009) 99–105.
[20] B.T. Burchett, Simulink model of the Ares I upper stage main propulsion system, in: AIAA 2008-6544, 2008.
[21] D.K. Frederick, A new method for constructing fast models of jet engines in Simulink, in: AIAA 2009-5419, 2009.
[22] H. Karimi, R. Mohammadi, Modeling and simulation of a two combustion chambers liquid propellant engine, Aircraft Eng. Aerosp. Technol. 79 (4)
(2007) 390–397.
[23] H. Karimi, A. Nassirharand, M. Mohseni, Modeling and simulation of a class of liquid propellant engine pressurization systems, Acta Astronaut. 66 (3)
(2010) 539–549.
[24] H. Karimi, A. Nassirharand, A. Zanj, Integration of modeling and simulation of warm pressurization and feed systems of liquid propulsion systems, Acta
Astronaut. 69 (5–6) (2011) 258–265.
[25] Y. Matsutomi, S.D. Heister, Studies on valveless pulse detonation engines, in: AIAA 2009–859, 2009.
[26] Y. Chen, G.B. Cai, Z. Wu, Modularization modeling and simulation of turbine test rig main test system, Appl. Math. Model. 35 (11) (2011) 5382–5399.
[27] A.L. Evans, G. Follen, C. Naiman, et al., Numerical propulsion system simulation’s national cycle program, in: AIAA 98-3113, 1998.
[28] R.H. Ashleman, T. Lavelle, F. Parsons, The national cycle program – a flexible system modeling architecture, in: AIAA 98-3114, 1998.
[29] R.W. Claus, S. Townsend, T. Lavelle, et al., A case study of high fidelity engine system simulation, in: AIAA 2006-4971, 2006.
[30] M. Turner, J. Reed, Variable fidelity analysis of complete engine systems, in: AIAA 2007-5042, 2007.
[31] R.W. Claus, T. Lavelle, S. Townsend, et al., Challenges in the development of a multi-fidelity, coupled component simulation of complex systems, in:
AIAA 2008-4651, 2008.
[32] NPSS User Guide, Software Release: NPSS_2.3.0, Rev. 2, Numerical Propulsion System Simulation Consortium, July 7, 2010.
[33] J.K. Lytle, The numerical propulsion system simulation: an overview, in: NASA TM-2000-209915, 2000.
[34] V. Pachidis, P. Pilidis, T. Alexander, et al, Advanced performance simulation of a turbofan engine intake, J. Propul. Power 22 (1) (2006) 201–204.
[35] J. Lytle, G. Follen, C. Naiman, et al., Numerical propulsion system simulation (NPSS) 1999 industry review, in: NASA TM-2000-209795, 2000.
[36] J.F. Dannenhoffer, N.D. Varano, A new look at zooming, in: AIAA 2005-496, 2005.
[37] M.G. Turner, Lessons learned from the GE90 3D full engine simulations, in: AIAA 2010-1606, 2010.
[38] R. Sampath, R. Irani, M. Balasubramaniam, et al., High fidelity system simulation using NPSS, in: AIAA 2004-371, 2004.
[39] E.L. Butzin, P.K. Johnson, R.E. Creekmore, Airframe thermal management system modeling in NPSS, in: AIAA 2007–5048, 2007.
[40] S.K. Rallabhandi, D.N. Mavris, Simultaneous airframe and propulsion cycle optimization for supersonic aircraft design, in: AIAA 2008-143, 2008.
[41] P. Osterbeck, E.L. Butzin, P. Johnson, Air vehicle sizing and performance modeling in NPSS, in: AIAA 2009-5417, 2009.
118 Y. Chen et al. / Simulation Modelling Practice and Theory 44 (2014) 95–118

[42] K. Lee, T. Nam, C. Perullo, et al, Reduced-order modeling of a high-fidelity propulsion system simulation, AIAA J. 49 (8) (2011) 1665–1682.
[43] B.C. Huffman, T.M. Lavelle, A.K. Owen, An NPSS model of a proposed altitude test facility, in: AIAA 2011-312, 2011.
[44] H.J. Kim, T. Kumano, M.S. Liou, et al, Flow simulation of supersonic inlet with bypass annular duct, J. Propul. Power 27 (1) (2011) 29–39.
[45] R.H. Nichols, A.G. Denny, J.A. Calahan, et al., Firebolt v1.0 – coupling of transient and steady engine performance models with a high-fidelity Navier-
Stokes code, in: AIAA 2011-3190, 2011.
[46] T.L. Benyo, Flow matching results of an MHD energy bypass system on a supersonic turbojet engine using the numerical propulsion system simulation
(NPSS), environment, in: AIAA2011-3591, 2011.
[47] J.A. Clough, R.P. Starkey, M.J. Lewis, et al., Reactor modeling for bimodal nuclear thermal rockets, in: AIAA 2006-4556, 2006.
[48] Y. Chen, F. Gao, Z.P. Zhang, et al, Finite volume model for quasi one-dimensional compressible transient pipe flow (I) finite volume model of flow field,
J. Aerosp. Power 23 (2) (2008) 311–316 (in Chinese).
[49] Y. Chen, F. Gao, Z.P. Zhang, et al, Finite volume model for quasi one-dimensional compressible transient pipe flow (II) finite volume model of
temperature field, J. Aerosp. Power 23 (2) (2008) 317–322 (in Chinese).
[50] M.S. Cheng, Research on the Dynamic Characteristics of Prepressurized and Evacuated Spacecrafts Propellant Feed Systems Priming Process, M.E.
Thesis, National University of Defense Technology, Changsha, China, 1997 (in Chinese).
[51] Wuzhong Instrument Co., Ltd, Selection Method of Valves, Related Information about Products of Wuzhong Instrument Co., Ltd. (in Chinese).

You might also like