A Sequential Model-Based Approach For Gas Turbine Performance Diagnostics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 41

University of Birmingham

A sequential model-based approach for gas turbine


performance diagnostics
Chen, Yu-Zhi ; Zhao, Xu-Dong ; Xiang, Heng-Chao ; Tsoutsanis, Elias

DOI:
10.1016/j.energy.2020.119657
License:
Creative Commons: Attribution-NonCommercial-NoDerivs (CC BY-NC-ND)

Document Version
Peer reviewed version

Citation for published version (Harvard):


Chen, Y-Z, Zhao, X-D, Xiang, H-C & Tsoutsanis, E 2021, 'A sequential model-based approach for gas turbine
performance diagnostics', Energy, vol. 220, 119657. https://doi.org/10.1016/j.energy.2020.119657

Link to publication on Research at Birmingham portal

General rights
Unless a licence is specified above, all rights (including copyright and moral rights) in this document are retained by the authors and/or the
copyright holders. The express permission of the copyright holder must be obtained for any use of this material other than for purposes
permitted by law.

•Users may freely distribute the URL that is used to identify this publication.
•Users may download and/or print one copy of the publication from the University of Birmingham research portal for the purpose of private
study or non-commercial research.
•User may use extracts from the document in line with the concept of ‘fair dealing’ under the Copyright, Designs and Patents Act 1988 (?)
•Users may not further distribute the material nor use it for the purposes of commercial gain.

Where a licence is displayed above, please note the terms and conditions of the licence govern your use of this document.

When citing, please reference the published version.


Take down policy
While the University of Birmingham exercises care and attention in making items available there are rare occasions when an item has been
uploaded in error or has been deemed to be commercially or otherwise sensitive.

If you believe that this is the case for this document, please contact UBIRA@lists.bham.ac.uk providing details and we will remove access to
the work immediately and investigate.

Download date: 02. ... 2566


Revised Manuscript with changes Marked

1
2
3
4
5 1 A sequential model-based approach for gas turbine
6
7 2 performance diagnostics
8
9
10 3 Yu-Zhi Chen1, Xu-Dong Zhao2,*, Heng-Chao Xiang3, Elias Tsoutsanis4
*
11 4 Corresponding author (xdzhaohit@gmail.com)
12 5
13
14 6 ABSTRACT
15
7
16
17
8 The gradual degradation of gas turbine components is an inevitable result of engine
18
19
20 9 operation, impacting engine availability, reliability, and operating cost. Gas path analysis plays
21
22 10 an essential role in engine fault diagnosis. Accurate and fast diagnosis of multiple simultaneously
23
24 11 degraded components has always posed a challenge, especially when the number of available
25
26 12 measurements is limited. This paper proposes a novel performance diagnostic method that
27
28 13 partitions the engine diagnosis into a series of steps to remove the “smearing effect” and reduce
29
30 14 the matrix dimensions in the iterative diagnostic algorithm. An engine performance model of a
31
32 15 triple-shaft gas turbine has been developed and validated against commercial software, in order
33
34 16 to assess the accuracy and computational performance of the proposed method. The advantage
35
36 17 of the proposed method lies in its capability to detect the severity of engine component
37
38 18 degradation, such as compressor fouling and turbine erosion, with greater accuracy and
39
40 19 computational efficiency than other model-based methods that use the same number of
41
42 20 measurements. The newly developed method provides an accurate diagnosis with a reduced set
43
44 21 of measurements. The method can deal effectively with the presence of random noise in the
45
46 22 measurements and carries a significantly lower computation burden in comparison to existing
47
48 23 methods. The proposed method could be used as a tool for supporting condition monitoring
49
50 24 systems for improved gas turbine reliability and energy efficiency.
51
52
53
54
55 1
Key Laboratory of Intelligent Control and Optimization for Industrial Equipment (Dalian University of Technology),
56 Ministry of Education, Dalian 116024, China
57 2
Key Laboratory of Intelligent Control and Optimization for Industrial Equipment (Dalian University of Technology),
58 Ministry of Education, Dalian 116024, China
59 3
China Gas Turbine Establishment, Aero Engine Corporation of China, Chengdu, 610500, China.
60 4
Department of Mechanical Engineering, University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK
61
62
63 1
64
65
1
2
3
4 25 Key Words: Gas Turbine Diagnostics; Gas Path Analysis; Model-Based Diagnostics; Gas Turbine
5
26 Performance.
6
7 27
8
9
10 28 Nomenclature
11 29
12
13 30 𝐶𝑃 = Characteristic parameter of component
14
15 31 𝑓(∙) = Nonlinear vector-value function
16
17 32 𝐹𝐴𝑅 = Fuel-air ratio
18
19
33 𝐹𝑃𝑇 = Free power turbine
20
21
34 𝐻 = Enthalpy [kJ/kg]
22
23
24 35 𝐻𝑃𝐶 = High-pressure compressor
25
26 36 𝐻𝑃𝑇 = High-pressure turbine
27
28 37 𝐿𝑃𝐶 = Low-pressure compressor
29
30 38 𝐿𝑃𝑇 = Low-pressure turbine
31
32 39 n = Number of operating points
33
34 40 𝑃 = Total pressure [atm]
35
36 41 𝑅𝐻 = Relative humidity [%]
37
38 42 𝑅𝑀𝑆𝐸 = Root mean square error
39
40 43 𝑆 = Entropy [kJ/(kg∙K)]
41
42 44 𝑇 = Total temperature [K]
43
44 45 𝑊 = Mass flow rate [kg/s]
45
46 46 𝑊𝐴𝑅 = Water-air ratio
47
48 47 𝑋 = Iteration variables, covers degradation factor variables
49
50
48 𝑍 = Measurement parameter
51
52
49
53
54
55
56 50 Greek Letters
57 51
58
59 52 𝜆 = Relative error of degradation factor as a percentage
60
61
62
63 2
64
65
1
2
3
4 53 𝜌 = Density [kg/𝑚3 ]
5
6 54
7
8
9
55 Subscripts
10
11 56
12
13 57 ac = Actual
14
15 58 CW = Compressor work
16
17 59 𝐸 = Efficiency
18
19 60 𝐹 = Flow capacity
20
21 61 L = Number of degradation factors
22
23 62 M = Number of measurement parameters
24
25 63 Mea = Measurement
26
27 64 𝑝𝑑 = Predicted
28
29 65 𝑃𝑅 = Pressure ratio
30
31 66 TW = Turbine work
32
33 67
34
35
36
68 1. Introduction
37
38 69
39
40 70 The pursuit of high reliability, availability, and efficiency in gas turbines has governed the evolution of engine
41
42 71 maintenance methods [1]. Currently, the maintenance cost of the gas turbine is an important aspect of engine lifecycle
43
44 72 expenditure. For instance, the lifecycle expenditure of the Siemens V94.3A gas turbine is expected to be 51.34 million
45
46 73 Euros, which is 17.9 times the initial purchase cost of 2.86 million Euros, according to its 40-year life maintenance
47
48 74 plan [2]. It is suggested that a more cost-efficient way of operating gas turbines could be achieved by enhanced engine
49
50 75 condition monitoring and appropriate repairs [3,4]. Talebi and Tousi (2017) [5] demonstrated that gas path analysis
51
52 76 (GPA), introduced by Urban (1969) [6], remains one of the soundest technologies for engine health monitoring and is
53
54 77 widely used for gas turbine condition monitoring to detect, identify, and assess component degradation. This, in turn,
55
56 78 affects the maintenance of gas turbine assets [7].
57
58 79 The degradation of gas turbine components has a great impact on the engine’s loss of performance from both a
59
60 80 thermodynamic and an economic perspective [8]. Some of the most common types of gas turbine degradation are
61
62
63 3
64
65
1
2
3
4 81 fouling, erosion, corrosion, rubbing wear, hot section damage, seal damage, and object damage [9]. The types of
5
6 82 deterioration fall into two classes: recoverable and unrecoverable [10]. Diagnostic methods are also classified into
7
8 83 three categories: model-based, data-driven, and hybrid methods [11,12].
9
10 84 The diagnostic accuracy of model-based methods relies heavily on the gas turbine model, which requires extensive
11
12 85 expert knowledge related to the model’s development and presents a great challenge. On the other hand, data-driven
13
14 86 approaches such as artificial neural networks [13], and deep learning [14], have excellent accuracy, subject to an
15
16 87 extensive training phase. The latter methods are limited by identifying new sets of data that are not used in their
17
18
88 training phase. A family of object-oriented Artificial Intelligence methods is also gaining significant ground in the
19
20
89 engine diagnostics arena [10,15]. The hybrid approaches can address some, but definitely not all, of the above
21
22
90 limitations by combining two or more methods. Thus, there are trade-offs in accuracy, computational performance,
23
24
25 91 and measurement noise, to name only a few considerations when selecting a diagnostic method. However, real-time
26
27 92 diagnosis is crucial for decision-making to ensure optimum, safe, and reliable engine operation. This study will focus
28
29 93 on model-based approaches, which present greater challenges in terms of accuracy and computational speed,
30
31 94 especially when such solutions are to be deployed in a real-time condition monitoring system.
32
33 95 The number of simultaneous fault components can profoundly affect the performance of the diagnosis [16,17].
34
35 96 When there are more than two degraded components, the complexity of nonlinear diagnostic systems is significantly
36
37 97 increased [17]. Traditionally, the number of engine measurements should be larger than the number of health
38
39 98 parameters to produce a unique diagnostic solution [1]. Hence, an increase in the number of engine components that
40
41 99 can degrade will not only increase the number of health parameters but also increase the number of measurement
42
43 100 parameters. In such a condition, two issues are raised for engine diagnosis: limited availability of engine measurements
44
45 101 for predicting every health parameter correctly and an increase of the matrix dimensions, which reduces computational
46
47 102 efficiency.
48
49 103 Regarding diagnostic accuracy, it has been pointed out that monitoring more degraded components using a limited
50
51 104 number of measurements could cause a severe “smearing effect” and lead to low precision [16]. The “smearing effect”
52
53 105 is the result of combinations of different degradation footprints in measured parameters. Hanachi et al. (2018) [18]
54
55 106 emphasized that accurate diagnosis of engine faults through limited measurements has always posed a challenge.
56
57 107 However, increasing the number of engine sensors will improve the precision of the diagnosis, but life cycle cost will
58
59 108 be increased substantially [19]. The improvement of the diagnostic accuracy of gas turbines engines with limited
60
61
62
63 4
64
65
1
2
3
4 109 measurement sets has attracted the attention of researchers in both academia and industry. Jasmani et al. (2011) [20]
5
6 110 reported a measurement selection method for triple-shaft engine diagnostics that showed improved accuracy when
7
8 111 compared with measurements on-site. However, the prediction error increased when there were more than three
9
10 112 degraded components. Pinelli et al. (2012) [21] considered keeping some of the health parameters fixed by an a-priori
11
12 113 optimized selection for engine diagnosis and used the concept of multiple operating points for addressing limited
13
14 114 measurements on-site. Hanachi et al. (2014) [22] introduced a diagnostic technique for gas turbines, in which the
15
16 115 degradation magnitudes are quantified by heat loss and power deficit indices, rather than the health index of each
17
18
116 rotating component. Lu et al. (2016) [23] proposed an improved and extended Kalman filter to address the shortage
19
20
117 of available measurements by the linear combination of the health parameters, but the diagnostic accuracy was affected
21
22
118 by the transformation matrices. Mohammadi and Montazeri-Gh (2016) [24] developed a global optimization-based
23
24
25 119 engine diagnostic method to overcome the lack of measurement instrumentation. Qingcai et al. (2016) [25] conducted
26
27 120 a series of sensitivity analyses, in which they chose different degradation levels to quantify the measurement deviation
28
29 121 of a triple-shaft engine. However, actual engine component degradation and ambient conditions could fall outside the
30
31 122 range of the case studies examined. Besides, the estimation of the correct degradation level through real measurement
32
33 123 deviation and chart of sensitivity analysis remains questionable. Sun et al. (2016) [26] proposed a GPA method to
34
35 124 overcome the lack of measurement parameters by fusing information from other sources. Simon and Rinehart (2016)
36
37 125 [19] suggested a sensor selection for aero-engines based on the Kalman filter and a maximum a posteriori estimator,
38
39 126 but they assumed the faults occurred in isolation. Yang et al. (2018) [27] proposed multiple interacting models for
40
41 127 fault detection and isolation. Then, they applied a generalized likelihood ratio approach for fault quantification. As
42
43 128 the number of multiple models was limited, their scheme assumed that the failures did not occur simultaneously. In
44
45 129 2019, Yang et al. [28] developed a new multiple model-based engine fault diagnosis algorithm, but the assumption of
46
47 130 multiple models remained. Despite the recent progress in engine diagnostics, the limited set of engine measurements
48
49 131 is still one of the most significant challenges for fault diagnosis [17].
50
51 132 From a computation perspective, increasing the matrix dimensions for the iterative diagnostic algorithm may lead
52
53 133 to the dimensionality problem [29]. It is worth emphasizing that the computation will increase exponentially under
54
55 134 these conditions [30]. Daroogheh et al. (2017) [31] pointed out that the number of required samples increases
56
57 135 exponentially for particle filters when the dimensionality of the health parameters increases. To date, studies have
58
59 136 investigated the demand for improving the computation speed for engine diagnostics. Tsoutsanis et al. (2014) [32]
60
61
62
63 5
64
65
1
2
3
4 137 proposed an adaptive diagnostics method for detecting compressor degradation through map tuning based on a
5
6 138 heuristic optimization technique. However, increasing the number of degraded components may lead to local and not
7
8 139 globally optimal solutions. Ying et al. (2016) [16] conducted fault detection before fault diagnosis, which could reduce
9
10 140 the dimension of the fault coefficient matrix. However, the scheme is not applicable when all components degrade
11
12 141 simultaneously. Yang et al. (2018) [33] suggested that the computation burden could be reduced by the generalized
13
14 142 expression of the Jacobian matrix, although the dimension of the matrix remained the same. Lu et al. (2018) [34]
15
16 143 proposed a fusion unscented Kalman filter to reduce the computation time of fault diagnosis by improving the
17
18
144 convergence speed. However, the diagnosis only considered efficiency degradation, and the flow capacity was
19
20
145 excluded during fault diagnosis.
21
22
146 Overall, the above studies highlight the need for accurate diagnosis of engine degradation with a limited number
23
24
25 147 of measurements and improved computational performance under the simultaneous deterioration of multiple
26
27 148 components. In this study, a sequential diagnostic method for improving the precision and computation efficiency is
28
29 149 proposed and applied to a triple-shaft industrial gas turbine with all five rotating components degraded simultaneously.
30
31 150 The procedure of sequential diagnosis allows the partition of the diagnostic algorithm into several serial mechanisms
32
33 151 to remove the smearing effect and reduce the matrix dimension in the iterative diagnostic algorithm. Furthermore, the
34
35 152 problem of limited measurements is addressed by feeding multiple operating points into the diagnostic process. The
36
37 153 novel contributions of this work are as follows:
38
39 154 1) A gas turbine engine model with object-oriented and modularized architecture has been developed in the
40
41 155 Microsoft Visual Studio C# environment [35], which is validated against GasTurb. The model’s architecture
42
43 156 is suited to the sequential diagnostic algorithm, which is evaluated through a well-used diagnostic method.
44
45 157 2) The new method improves diagnostic accuracy by isolating the fault components and eliminating the
46
47 158 smearing effect via sequential analysis.
48
49 159 3) The novel algorithm decreases the computation time by reducing both the matrix’s dimensions in the iteration
50
51 160 algorithm and the number of calls to engine sub-models (compressor model, burner model, turbine model,
52
53 161 etc.).
54
55 162 4) The proposed scheme can ensure the required diagnostic accuracy, under a limited number of measurements,
56
57 163 by multiple operating point analysis. Reducing the number of measuring sensors can potentially bring
58
59 164 economic benefits to the engine operator and decrease sensor related problems.
60
61
62
63 6
64
65
1
2
3
4 165 5) The negative impact of measurement noise on fault diagnosis accuracy is addressed by an averaging filter for
5
6 166 data filtration to demonstrate the suitability of the method for real-world applications.
7
8 167 The remainder of this paper is organized as follows. Section 2 is concerned with the methodology used for this
9
10 168 study. Section 3 demonstrates the validation of the developed engine model using commercial software. Section 4
11
12 169 analyses the results of the fault diagnosis. The final section presents the conclusions of the research.
13
14
15
16
170 2. Methodology
17
18 171 2.1 Engine Performance Model
19
20 172 The triple-shaft industrial gas turbine engine used in this study (Fig. 1) is similar to the Rolls-Royce RB211-24G
21
22 173 operated at the China Petroleum Pipeline Langfang compressor group [36]. The power output is selected as the control
23
24 174 variable in this study, but this could be any other control parameter, such as rotation speed, fuel flow rate, etc. The
25
26 175 existing measurement parameters on-site are shown in Table 1 [36]. It is worth noting that the measurements at 𝐻𝑃𝑇
27
28 176 outlet are not available due to the high gas temperature.
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49 177
50
51 178 Fig. 1 Schematic layout of the triple-shaft engine configuration, with station numbering.
52
179
53
180 In general, the performance of a gas turbine engine is a function of its components’ performance [37]. Hence, the
54
55
56 181 degradation factor (𝑋) of each component is defined as the ratio of the degraded state over the healthy state of each
57
58 182 characteristic parameter (𝐶𝑃), as Eq. (1) [38]. 𝑋 equals unity means a healthy/clean state.
59
60
61
62
63 7
64
65
1
2
3
4
5 𝐶𝑃𝑎𝑐 (1)
6 𝑋=
𝐶𝑃𝑖𝑑𝑒𝑎𝑙
7 183
8
9 184 where the subscript “ac” represents the actual characteristic parameter of component, while “ideal” represents the
10
11 185 healthy/clean characteristic parameter obtained from the component map.
12
13
14 186 Table 1 Available engine gas path measurements on-site [36].
15
16 No Available Measurement Parameters Symbol
17 1 Ambient pressure 𝑃1
18
19 2 Ambient temperature 𝑇1
20
3 Ambient relative humidity 𝑅𝐻1
21
22 4 Free power turbine shaft rotational speed 𝑁𝐹𝑃𝑇
23
24 5 Free power turbine output 𝑇𝑊𝐹𝑃𝑇
25 6 Low-pressure compressor (LPC) exit pressure 𝑃3
26
27 7 LPC exit temperature 𝑇3
28 8 High-pressure compressor (HPC) exit pressure 𝑃4
29
30 9 HPC exit temperature 𝑇4
31 10 Low-pressure turbine (LPT) exit pressure 𝑃9
32
33 11 LPT exit temperature 𝑇9
34 12 Free power turbine (FPT) exit pressure 𝑃10
35
36 13 FPT exit temperature 𝑇10
37 14 LP shaft rotational speed 𝑁𝐿𝑃
38
39 15 HP shaft rotational speed 𝑁𝐻𝑃
40 16 Burner fuel flow rate 𝑊𝐹𝑢𝑒𝑙
41
42 187
43
44 188 The engine model is crucial for model-based diagnostics in order to assess the performance state of a gas turbine
45
46 189 engine [39]. A thermodynamic model of the engine has been developed in Microsoft Visual Studio C#. A detailed
47
48 190 description of the model and its governing equations are presented in Appendix B. The balancing process for off-
49
50 191 design simulation of the triple-shaft industrial gas turbine is based on [40]. The assumptions that have been made for
51
52 192 the developed model are as follows:
53
54 193 1) It is assumed that the compressors are of fixed geometry.
55
56 194 2) Pressure losses in the burner and duct models are accounted for by assuming that the losses are proportional
57
58 195 to the inlet conditions.
59
60 196 3) Isentropic expansion is assumed in the model of the exhaust nozzle.
61
62
63 8
64
65
1
2
3
4 197 4) The mixture model includes two inlet flows, namely the main flow and the cooling flow. The mixture's total
5
6 198 outlet pressure is assumed to be equal to the absolute inlet pressure of the main flow.
7
8 199
9
10 200 The effect of relative humidity is considered in the developed engine model. The gas property of the engine model
11
12 201 is generated via the NASA CEA program [41], where both fuel-air ratio (𝐹𝐴𝑅) and water-to-air-ratio (𝑊𝐴𝑅) are
13
14
202 considered to define the mixtures. When two of the gas properties and mixtures are defined/known, the remaining gas
15
16
203 properties can be calculated. For example, when the temperature (𝑇), pressure (𝑃), 𝐹𝐴𝑅, and 𝑊𝐴𝑅 are known, the
17
18
204 enthalpy (𝐻), entropy (𝑆), and density (𝜌) etc. could be obtained using Eq. (2).
19
20
21 (2)
22 [𝐻, 𝑆, 𝜌, … ] = 𝐺𝑎𝑠𝑃𝑟𝑜𝑝[𝑇,𝑃] (𝑇, 𝑃, 𝐹𝐴𝑅, 𝑊𝐴𝑅)
23
24 205 It is worth mentioning that the gas turbine of interest does not include water injection at any station of the engine.
25
26 206 Hence, the 𝑊𝐴𝑅 is a constant throughout the engine and varies only with a change in ambient conditions.
27
28 207
29
30
31 208 2.2 Conventional Diagnostic Method
32
33 209 The scheme of the conventional model-based diagnostic system is shown in Fig. 2 [42–44]. Irrespective of the
34
35 210 number of degradation components considered, the diagnostic system should run the performance simulation of the
36
37 211 entire engine model (call all engine sub models each time). In such a condition, the algorithm will consume significant
38
39 212 computational power and may suffer from smearing. The nonlinear gas path analysis (NLGPA) combined with
40
41 213 multiple operating point analysis has been widely used for model-based fault diagnosis, as shown in Eq. (3) [45]:
42
43
44 𝑍𝑖∙𝑚 = 𝑓(𝑋𝑙 ) (3)
45
46 214 where 𝑓(∙) is the nonlinear vector-valued function of gas turbine performance, 𝑍 denotes the measurement parameters,
47
48 215 ∀ 𝑖 = 1, … , 𝑛, ∀ 𝑚 = 1, … , 𝑀, and ∀ 𝑙 = 1, … , 𝐿 where “𝑛”, “𝑀”, and “𝐿” denote the number of operating points, the
49
50 216 number of measurements, and the number of degradation factors, respectively. The NLGPA solver could be of any
51
52 217 type, but the most popular for gas turbine engines are: Newton-Raphson [46], Kalman filter [47], Particle filter [31],
53
54 218 and Genetic Algorithms [25].
55
56 219 Remark 1. It is worth noting that the first five measurements in Table 1 are used to establish the engine operating
57
58 220 condition for conventional model-based diagnostics, and as such, the gas path measurements for fault diagnosis in
59
60 221 Eq. (3) are the remaining 11 parameters.
61
62
63 9
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
222
34
35
36 223 Fig. 2 Schematic layout of the conventional gas path analysis process [42–44].
37 224
38 225 The component health parameters considered in this study are summarized in Table 2.
39
40
41 226 Table 2 Degradation factors of the individual rotating component.
42
43 Component Symbols Component Health Parameters
44 𝑋𝐿𝑃𝐶,𝐸 LPC efficiency degradation factor
45 𝐿𝑃𝐶 𝑋𝐿𝑃𝐶
46 𝑋𝐿𝑃𝐶,𝐹 LPC flow capacity degradation factor
47 𝑋𝐻𝑃𝐶,𝐸 HPC efficiency degradation factor
48 𝐻𝑃𝐶 𝑋𝐻𝑃𝐶
49 𝑋𝐻𝑃𝐶,𝐹 HPC flow capacity degradation factor
50 𝑋𝐻𝑃𝑇,𝐸 HPT efficiency degradation factor
51 𝐻𝑃𝑇 𝑋𝐻𝑃𝑇
52 𝑋𝐻𝑃𝑇,𝐹 HPT flow capacity degradation factor
53 𝑋𝐿𝑃𝑇,𝐸 LPT efficiency degradation factor
54 𝐿𝑃𝑇 𝑋𝐿𝑃𝑇
55 𝑋𝐿𝑃𝑇,𝐹 LPT flow capacity degradation factor
56 𝑋𝐹𝑃𝑇,𝐸 FPT efficiency degradation factor
57 𝐹𝑃𝑇 𝑋𝐹𝑃𝑇
58 𝑋𝐹𝑃𝑇,𝐹 FPT flow capacity degradation factor
59 227
60
61
62
63 10
64
65
1
2
3
4 228 The root mean square error of the measurement parameters (𝑅𝑀𝑆𝐸𝑀𝑒𝑎 ) is calculated using Eq. (4) to check the
5
6 229 convergence of NLGPA iterations. When the maximum allowed iteration step (21 in this study) is achieved, then the
7
8 230 calculation will stop without convergence. In this situation, the 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 is larger than the threshold (1E-5).
9
10
11 (4)
12 𝑛∙𝑀 2
𝑍𝑖,𝑎𝑐 − 𝑍𝑖,𝑝𝑑
13 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 = /(𝑛 ∙ 𝑀)
14 𝑍𝑖,𝑎𝑐
𝑖=1
15
16
17 231 where the subscript “𝑝𝑑” and “ac” referred to the predicted and actual values, respectively.
18
19 232 The root mean square error of the degradation factor (𝑅𝑀𝑆𝐸𝑋 ) is defined by Eq. (5) to evaluate the diagnosis of
20
21 233 the degradation factor.
22
23
24 (5)
𝐿 2
25 𝑋𝑖,𝑎𝑐 − 𝑋𝑖,𝑝𝑑
26 𝑅𝑀𝑆𝐸𝑋 = /𝐿
27
𝑋𝑖,𝑎𝑐
𝑖=1
28
29 234 Remark 2. It should be noted that the actual degradation is not available, and 𝑅𝑀𝑆𝐸𝑋 is used in this study only
30
31 235 to assess the performance of the developed method.
32
33
236 The relative error (𝜆𝑖 ) is defined by Eq. (6) and represents the percentage ratio of the absolute difference between
34
35
237 the predicted and actual/implanted degradation factors, to the actual degradation.
36
37
38 (6)
39
𝑋𝑖,𝑎𝑐 − 𝑋𝑖,𝑝𝑑
𝜆𝑖 = × 100%
40 𝑋𝑖,𝑎𝑐
41
42 238
43
44
45 239 2.3 Novel Sequential Diagnostic Method
46
47 240 2.3.1 Novel Sequential Diagnostic Method with All Available Measurements
48
49 241 The architecture of sequential diagnosis is shown in Fig. 3, where the dotted lines and the solid lines in the graph
50
51 242 indicate the flow of information of target parameters and to-be adapted parameters, respectively. The diagnostic
52
53 243 scheme partitions the engine diagnosis into four sequential steps for the triple-shaft engine of interest. The 𝐹𝑃𝑇
54
55 244 diagnostic is carried out first. The diagnosis then resumes in the 𝐿𝑃𝐶, then in the 𝐻𝑃𝐶, and finally, the 𝐻𝑃𝑇 and 𝐿𝑃𝑇
56
57 245 conclude the diagnosis. The dotted boxes in Fig.3 indicate the available gas path measurements for each step.
58
59 246 The subsequent sections describe the sequential procedure in more detail.
60
61
62
63 11
64
65
1
2
3
4 247 ➢ Step 1: Free Power Turbine Diagnostics
5
6 248 Step 1 in this process involves the tuning of the 𝐹𝑃𝑇 map, through the scaling factors 𝑋𝐹𝑃𝑇 (𝑋𝐹𝑃𝑇,𝐸 , 𝑋𝐹𝑃𝑇,𝐹 ), which
7
8 249 will eventually enable the model to match the available measurements at outlet temperature (𝑇10 ) and turbine power
9
10
250 output (𝑇𝑊𝐹𝑃𝑇 ). This step involves only the turbine model during iteration, which can potentially save a lot of
11
12
251 computation time.
13
14
15 252 ➢ Step 2: Low-Pressure Compressor Diagnostics
16
17 253 Similar to the previous step, the 𝑋𝐿𝑃𝐶 (𝑋𝐿𝑃𝐶,𝐸 , 𝑋𝐿𝑃𝐶,𝐹 ) is estimated through an iterative process to scale the
18
19 254 compressor map according to the compressor model based on the 𝐿𝑃𝐶 outlet temperature (𝑇3 ) and 𝑊2 .
20
21 255 ➢ Step 3: High-Pressure Compressor Diagnostics
22
23 256 At the 𝐻𝑃𝐶 stage, the 𝑋𝐻𝑃𝐶 (𝑋𝐻𝑃𝐶,𝐸 , 𝑋𝐻𝑃𝐶,𝐹 ) is also tuned to scale the compressor map for the simulation of the
24
25 257 compressor model. The two measurements to be satisfied are 𝑇4 and 𝑊3 , with the latter having been calculated from
26
27 258 Step 2. By iteration, the corrected 𝑋𝐻𝑃𝐶 can be determined, and the 𝐻𝑃𝐶 diagnostic needs to utilize only the
28
29 259 compressor model during iteration.
30
31 260 ➢ Step 4: High-pressure and Low-pressure Turbine Diagnostics
32
33 261 The pressure ratio of 𝐻𝑃𝑇 (𝑃𝑅𝐻𝑃𝑇 ), 𝑋𝐻𝑃𝑇 and 𝑋𝐿𝑃𝑇 are the iteration variables for Step 4, where 𝑋𝐻𝑃𝑇 (𝑋𝐻𝑃𝑇,𝐸 ,
34
35 262 𝑋𝐻𝑃𝑇,𝐹 ) and 𝑋𝐿𝑃𝑇 (𝑋𝐿𝑃𝑇,𝐸 , 𝑋𝐿𝑃𝑇,𝐹 ) are tuned to scale the 𝐻𝑃𝑇 and 𝐿𝑃𝑇 maps respectively during iteration. The work
36
37
263 compatibility of the HP and LP shafts, and 𝐿𝑃𝑇 outlet temperature (𝑇9 ) are available as convergence criteria to tune
38
39
264 the iteration variables.
40
41
42 265 Remark 3. There are three convergence criteria for a single operating point of the gas turbine in Step 4. Hence,
43
44 266 the number of convergence criteria is (𝑛 × 3) for n operating points. The total iteration variables are (4 + 𝑛) for n
45
46 267 different operating points, where the number “4” denotes the four degradation factors for 𝐻𝑃𝑇 and 𝐿𝑃𝑇, and “n”
47
48 268 denotes the required number of 𝑃𝑅𝐻𝑃𝑇 for each operating point of the gas turbine. The 𝑛 is assigned to be three,
49
50 269 which is the minimum number of operating points to satisfy the requirement that the number of convergence criteria
51
52 270 (nine) should be more than the iteration variables (seven). A more detailed description of the iterative matrix
53
54 271 computation for the proposed diagnostic method is provided in Appendix A.
55
56
57
58
59
60
61
62
63 12
64
65
1
2
3
4
5 FPT Diagnostic (Step 1)
6 LPC Diagnostic (Step 2)
7
8 HPC Diagnostic (Step 3)
9 HPT & LPT Diagnostic (Step 4)
10
11
12 Data: Gas Property
13 T1, P1, RH1 Model Data:
14 T10, TWFPT
15 WAR 1
16 T10, TW FPT
Data:
17 Turbine Model T10, TWFPT
WFuel, NFPT, XFPT
18 (FPT) compatibility?
T9, P9, P10
19
20 T10, TW FPT
21 Data:
22 W2
WAR 1 T3
23
T3, W2
24
25 Data: Compressor T3 , W2
XLPC
26 NLP , T2, P2, P3 Model (LPC) compatibility?
27
28 T3, W2
29
30
Data:
31 W3
T4
32 WAR 1
33 T4, W3
34 Data: Compressor T4 , W3
35 XHPC
NHP, T3, P3, P4 Model (HPC) compatibility?
36
37 WAR 1, W4, T4,
WAR 1, W42, P4, W41, T41, P41
38 T4, W3 n×CWLPC
T42, P42
39
40 Burner &
Data: Data:
41 Mixture n×CWHPC
WFuel n×T9
42 Model
43 WAR 1, FAR6, T6, P6
44 n×(T9, CWLPC, CWHPC)
45 Data: Turbine Model
46 n×TW HPT
NHP, WFuel (HPT) XHPT,
47
48 WAR 1, FAR7, T7, P7 n×PRHPT
49 n×(T9,
50 Mixture LP, HP power
51 Model balance)
52 WAR 1, T8, P8 compatibility
53 ?
54
55 Data: Turbine Model
XLPT
56 NLP , P9, WFuel (LPT)
Note: n=3
57 n×(T9, TWLPT)
58 272
59
60 273 Fig. 3 Sequential diagnostic scheme of gas turbine with all available measurements.
61
62
63 13
64
65
1
2
3
4 274 2.3.2 Novel Sequential Diagnostic Method with Reduced Number of Measurements
5
6 275 Shaft power of aero engines is a parameter which is more challenging to obtain accurately in comparison to land-
7
8 276 based or marine engines, where it can be found through engine load. Hence, it is necessary to test the 𝐹𝑃𝑇 diagnostic
9
10 277 without the measurement of power output (𝑇𝑊𝐹𝑃𝑇 ). Additionally, the gas path measurements at the 𝐿𝑃𝐶 outlet (𝑇3
11
12
278 and 𝑃3 ) will be assumed unavailable.
13
14
279 Now, the sequential diagnostic divides the entire triple-shaft engine into three diagnostic steps: 𝐹𝑃𝑇 diagnostics,
15
16
17 280 𝐿𝑃𝐶 and 𝐻𝑃𝐶 diagnostics, and 𝐻𝑃𝑇 and 𝐿𝑃𝑇 diagnostics. The diagnostic scheme with reduced measurements is
18
19 281 shown in Fig. 4, where the dotted lines and solid lines indicate the flow of information of target parameters and to-be
20
21 282 aligned parameters, respectively. The 𝐹𝑃𝑇 diagnostic step is now modified, in comparison to the previous approach,
22
23 283 in order to assess the suitability of this method for aero engine applications. Meanwhile, 𝐻𝑃𝐶 and 𝐿𝑃𝐶 diagnostics
24
25 284 (Step 2-3) are also modified to reduce the number of gas path measurements further. For 𝐻𝑃𝑇 and 𝐿𝑃𝑇, the diagnosis
26
27 285 is the same as presented in Fig. 3, and the calculation process will not be further discussed here. The dotted boxes
28
29 286 indicate the available gas path measurements for each step.
30
31 287 ➢ Step 1: Free Power Turbine Diagnostics
32
33 288 The 𝑋𝐹𝑃𝑇 (𝑋𝐹𝑃𝑇,𝐸 , 𝑋𝐹𝑃𝑇,𝐹 ) is estimated through an iterative process to scale the component map for the turbine
34
35 289 model. There is only one targeted parameter available for a single operating point (n) of the gas turbine, and that is
36
37 290 the 𝐹𝑃𝑇 outlet temperature (𝑇10 ).
38
39 291
40
41 292 ➢ Step 2: High-pressure and Low-pressure Compressors Diagnostics
42
43 293 In the second step of this process, the pressure ratio of 𝐿𝑃𝐶 (𝑃𝑅𝐿𝑃𝐶 ), 𝑋𝐿𝑃𝐶 and 𝑋𝐻𝑃𝐶 are utilized to satisfy three
44
45 294 measurement parameters, namely 𝑊2 , 𝑊3 and 𝑇4 . Once again 𝑋𝐿𝑃𝐶 (𝑋𝐿𝑃𝐶,𝐸 , 𝑋𝐿𝑃𝐶,𝐹 ) and 𝑋𝐻𝑃𝐶 (𝑋𝐻𝑃𝐶,𝐸 , 𝑋𝐻𝑃𝐶,𝐹 ) are
46
47
295 tuned to scale the 𝐿𝑃𝐶 and 𝐻𝑃𝐶 maps, respectively, during iteration.
48
49
50 296 Remark 4. The number of operating points, n, is three for every step, which is the least number of operating points
51
52 297 capable of satisfying the convergence requirements.
53
54
55
56
57
58
59
60
61
62
63 14
64
65
1
2
3
4
5 FPT Diagnostic (Step 1)
6
7 LPC & HPC Diagnostic (Step 2)
8
9
10 HPT & LPT Diagnostic (Step 3)
11
12 Data: Gas Property
13
T1, P1, RH1 Model Data:
14
15 WAR 1 n×T10
16 n×T 10
17 Data:
18 Turbine Model n×T10
WFuel, NFPT, XFPT
19 (FPT) compatibility?
T9, P9, P10
20
21 n×T 10
22 Data:
23 n×W 2
WAR 1 T4
24 n×W 3
25
26 Data: Compressor
27 NLP , T2, P2 Model (LPC) n×W 2
28
29 n×(W2, W3, T4)
XLPC,
30 WAR 1, T3, P3 n×PRLPC
31
32 n×
33 Data: Compressor (W2, W3, T4 )
XHPC
34 NHP, P4 Model (HPC) compatibility
35 ?
36 WAR 1, W4, T4,
WAR 1, W42,
P4, W41, T41, P41 n×(W3, T4)
37 T42, P42
n×CWLPC
38
39 Burner &
Data: Data:
40 Mixture n×CWHPC
WFuel n×T9
41 Model
42 WAR 1, FAR6, T6, P6
43 n×(T9, CWLPC, CWHPC)
44
Data: Turbine Model
45 n×TW HPT
46 NHP, WFuel (HPT) XHPT,
47 WAR 1, FAR7, T7, P7 n×PR HPT
48 n×(T9,
49 Mixture LP, HP power
50
51
Model balance)
52 WAR 1, T8, P8 compatibility
53 ?
54 Data: Turbine Model
55 XLPT
NLP , P9, WFuel (LPT)
56 Note: n=3
57 n×(T9, TWLPT)
58 298
59
60 299 Fig. 4 Sequential diagnostic scheme of gas turbine with reduced measurements.
61
62
63 15
64
65
1
2
3
4 300 2.4 Method Overview
5
6 301 Essentially, the proposed sequential diagnostic method partitions the engine model into several sequential steps to
7
8 302 reduce the dimensions of the matrices and the number of calls to each engine component model during diagnosis. A
9
10 303 detailed analysis of the computational burden of both methods is given in Appendix A. Another key feature of the
11
12 304 sequential diagnosis method is its capability of eliminating the smearing effect by isolating components with the aid
13
14 305 of multiple operating point analysis. The sequential diagnostic method is tested with reduced engine gas path
15
16 306 measurements, which will not only reduce the cost to engine operators but also reduce flow disturbances caused by
17
18
307 the installation of sensors.
19
20
21
22 308 3. Engine Model Validation
23
24 309 The developed engine model has been validated against the commercial gas turbine software GasTurb [44]. The
25
26 310 engine operating conditions at the design point are shown in Table 3. The design point simulation algorithm is
27
28 311 validated first, and the relative errors are shown in Table 4, which indicates that the maximum relative error is less
29
30 312 than 0.31%. It follows that the developed engine performance model is satisfactory at the design point.
31
32
33 313 Table 3 Engine specification.
34 Symbols
Parameters Unit Value
35
36 Ambient Pressure 𝑃1 atm 1.000
37 Ambient Temperature 𝑇1 Kelvin 288.15
38 Ambient Relative Humidity 𝑅𝐻1 % 60.00
39
Inlet Air Flow Rate 𝑊1 kg/s 83.40
40
41 LPC Rotational Speed 𝑁𝐿𝑃 rpm 6611
42 HPC Rotational Speed 𝑁𝐻𝑃 rpm 9305
43 FPT Rotational Speed 𝑁𝐹𝑃𝑇 rpm 4800
44 FPT Power Output 𝑇𝑊𝐹𝑃𝑇 MW 28.31
45
46 314
47
48 315 The power output of the gas turbine varies with respect to the actual load demand. Hence, it is essential to check
49
50 316 whether the developed algorithm can provide satisfactory results at steady-state off-design conditions. The same
51
52 317 component maps are used for both the GasTurb and the developed engine model for off-design validation. Four
53
54 318 measured parameters are shown in Fig. 5, where a comparison is made between the GasTurb and the developed model.
55
56 319 Specifically, the 𝐻𝑃𝐶 outlet temperature (𝑇4 ), 𝐹𝑃𝑇 inlet pressure (𝑃9 ), 𝐿𝑃𝐶 rotational speed (𝑁𝐿𝑃 ), and fuel flow rate
57
58 320 (𝑊𝐹𝑢𝑒𝑙 ) are considered at different power settings, varying from 65% to 100% with a step of 5%.
59
60
61
62
63 16
64
65
1
2
3
4 321 Table 4 Engine model validation at design point [44].
5
6 Parameters Units GasTurb[44] Developed Model Relative Error [%]
7 𝑃3 atm 4.850 4.850 0.000
8
9 𝑇3 Kelvin 468.70 468.79 0.019
10 𝑃4 atm 20.749 20.749 0.000
11 𝑇4 Kelvin 737.11 737.10 0.001
12 𝑃5 atm 20.127 20.126 0.005
13
14 𝑇5 Kelvin 1494.00 1494.20 0.013
15 𝑃6 atm 20.127 20.126 0.005
16 𝑇6 Kelvin 1459.49 1459.77 0.019
17 𝑃7 atm 8.605 8.588 0.198
18
𝑇7 Kelvin 1233.58 1234.41 0.067
19
20 𝑃8 atm 8.605 8.588 0.198
21 𝑇8 Kelvin 1216.03 1216.84 0.067
22 𝑃9 atm 4.647 4.633 0.301
23 𝑇9 Kelvin 1071.23 1071.84 0.057
24
25 𝑃10 atm 1.139 1.136 0.263
26 𝑇10 Kelvin 793.26 793.52 0.033
27 𝑊𝐹𝑢𝑒𝑙 kg/s 1.866 1.867 0.006
28 322
29
30 323 The results demonstrate that the developed model is capable of predicting the gas turbine performance at different
31
32 324 operating conditions with a high degree of precision relative to GasTurb. The maximum relative error increases
33
34
325 slightly as the power decreases, which is reasonable since we are moving further away from the design point, and the
35
36
326 maximum relative error for all parameters listed in Table 4 is less than 0.71% at 65% power setting for 𝑊𝐹𝑢𝑒𝑙 , as
37
38
327 shown in Fig. 5. Another cause for the increasing error may be attributed to the different methods of reading the
39
40
41 328 component map in the calculation procedure. Although the errors increase at lower power settings, the developed
42
43 329 model retains a good agreement with GasTurb. Therefore, the developed engine model will be implemented to
44
45 330 quantify the level of degradation for both conventional and newly proposed staged diagnostic methods.
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63 17
64
65
1
2
3
4 1.0 4.8 1.0
740 P9 GasTurb
5 T4 GasTurb
P9 Developed Model
6 730 T4 Developed Model
0.8
4.6
Relative Error 0.8
7 Relative Error
720
8

Relative Error [%]

Relative Error [%]


4.4
710
9 0.6 0.6
T4 [Kelvin]

P9 [atm]
10 700 4.2
11 690 0.4 0.4
12 4.0
13 680

14 670
0.2
3.8
0.2

15
660
16 0.0 3.6 0.0
17 60 65 70 75 80 85 90 95 100 105 60 65 70 75 80 85 90 95 100 105

18 Power Settings [%] Power Settings [%]


19
20 (a) Relative error of 𝑇4 . (b) Relative error of 𝑃9 .
21
1.0 1.0
22 NLP GasTurb 1.9 W Fuel GasTurb
23 6600
NLP Developed Model W Fuel Developed Model
24 Relative Error 0.8 1.8 Relative Error 0.8
25
Relative Error [%]

Relative Error [%]


6400 1.7
26
0.6 0.6
WFuel [kg/s]

27
NLP [rpm]

1.6
28 6200
29 0.4 1.5 0.4

30
1.4
31 6000
0.2 0.2
32 1.3
33
34 5800 0.0 1.2 0.0
60 65 70 75 80 85 90 95 100 105 60 65 70 75 80 85 90 95 100 105
35 Power Settings [%] Power Settings [%]
36
37
(c) Relative error of 𝑁𝐿𝑃 . (d) Relative error of 𝑊𝐹𝑢𝑒𝑙 .
38
39
40 331 Fig. 5 Engine model validation at off-design points [44].
41
42 332
43
44 333 4. Application and Analysis
45
46
47 334 4.1 Case Study Description
48
49 335 Four case studies are conducted to assess the accuracy and the computational performance of the proposed
50
51 336 diagnostic method. Moreover, the developed method will be compared with the conventional NLGPA method [45].
52
53 337 The case studies are as follows:
54
55 338 Case 1: The objective of this case study is to test the conventional diagnostic method NLGPA [45] in order to
56
57 339 establish a benchmark against which comparisons will be made.
58
59
60
61
62
63 18
64
65
1
2
3
4 340 Case 2: The objective of this case study is to test the proposed sequential method in terms of diagnostic accuracy
5
6 341 and computational speed.
7
8 342 Case 3: This case study is based on the sequential method, similar to Case 2, with the only difference being the
9
10 343 reduced number of available engine measurements. The goal is to assess the suitability of the proposed method for
11
12 344 application to gas turbine engines that have relatively fewer measurement sensors.
13
14 345 Case 4: The objective of this case study is to test the effectiveness of the proposed method in providing an accurate
15
16 346 diagnosis in the presence of measurement noise.
17
18
347 With the accumulation of running time, all components will degrade. In this paper, the typical degradation
19
20
348 implanted into all rotating components, to represent compressor fouling and turbine erosion, of the engine model is
21
22
349 shown in Table 5 [46]. The typical degradation is injected into the reference engine state using Eq. (1) to obtain
23
24
25 350 component characteristics under deterioration. The degradation factor, 𝑋, consists of isentropic efficiency and flow
26
27 351 capacity.
28
29 352 Table 5 Typical degradation level for engine rotating components [46].
30
31 Component Degradation Type Parameter Degradation Level [%]
32 𝑋𝐿𝑃𝐶,𝐸 -1.0
33 LPC Fouling
𝑋𝐿𝑃𝐶,𝐹 -4.0
34
35 𝑋𝐻𝑃𝐶,𝐸 -1.0
HPC Fouling
36 𝑋𝐻𝑃𝐶,𝐹 -4.0
37 𝑋𝐻𝑃𝑇,𝐸 -1.0
38 HPT Erosion
𝑋𝐻𝑃𝑇,𝐹 +2.0
39
40 𝑋𝐿𝑃𝑇,𝐸 -1.0
LPT Erosion
41 𝑋𝐿𝑃𝑇,𝐹 +2.0
42 𝑋𝐹𝑃𝑇,𝐸 -1.0
43 FPT Erosion
𝑋𝐹𝑃𝑇,𝐹 +2.0
44
353
45
46
47 354 The case studies have been conducted in a PC with Intel® Core™ i7, 2.9 GHz, and 16 GB RAM. The software
48
49 355 environment in which the model is developed in Visual Studio C# and the iterative algorithm used in all case studies
50
51 356 is the Newton-Raphson [48]. The above features remain constant for all case studies in order to demonstrate and
52
53 357 illustrate the advancement of the sequential diagnostic method.
54
55
56 358 4.2 Case 1: Conventional Diagnostic
57
58 359 Case 1 involves five simultaneously degraded components, with up to three operating points (𝑛), and it follows
59
60 360 the process is as schematically represented in Fig. 2. In the first diagnostic attempt we use a single operating point in
61
62
63 19
64
65
1
2
3
4 361 the NLGPA method [45], and the calculation converges in 7 steps, shown in Fig. 6 (a), with the 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 less than
5
6 362 1E-5, see Table 6. However, when checking the 𝑅𝑀𝑆𝐸𝑋 , it becomes evident that the predicted degradation factor is
7
8 363 not identical to the implanted fault. This highlights the presence of the smearing effect, where a different combination
9
10 364 of degradation factors could match the engine measurements.
11
12 0.030 7.00 Case 1 - Attempt 1
13 Single Operating Point 6.50 Case 1 - Attempt 2
Dual Operating Points 6.00
14 0.025
Triple Operating Points
Case 1 - Attempt 3
5.50
15 5.00
16 0.020 4.50

Relative Error [%]


17 4.00
RMSEMea

3.50
18 0.015
3.00
19
0.05
20 0.010
0.04
21
22 0.005 0.03

23 0.02

24 0.000 0.01

25 0.00
0 5 10 15 20 25 XLPC,E XHPC,E XHPT,E XLPT,E XFPT,E
26 XLPC,F XHPC,F XHPT,F XLPT,F XFPT,F
Number of Iterations
27
28 (a) Convergence of diagnosis. (b) Relative diagnostic error.
29 365 Fig. 6 Convergence performance and relative diagnostic error for Case 1.
30
31 366
32 367 Table 6 Diagnostic results of Case 1.
33
34 Items Symbols Units Attempt 1 Attempt 2 Attempt 3
35 Computation Time 𝐶𝑇 Second 22.659 129.682 197.368
36
37 Measurement Error 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 - 4.3E-6 2.1E-4 1.6E-4
38 Degradation Factor Error 𝑅𝑀𝑆𝐸𝑋 - 2.9E-2 1.5E-2 1.2E-2
39 Operating Point 𝑛 - 1 2 3
40 No. of Iteration Steps 𝑁𝐼𝑆 - 7 21 21
41
No. of total calls to engine sub-models 𝑁𝑂𝑀𝑜𝑑𝑒𝑙 - 94644 566775 841797
42
43 368
44
45 369 In the second and third attempts, we consider two and three operating points, respectively. In both attempts the
46
47 370 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 condition is not satisfied, and the iterations terminate at the maximum allowed step, as shown in Table 6.
48
49 371 Although the 𝑅𝑀𝑆𝐸𝑋 decreases as the operating points increase, it is still noticeable (Table 6). The results reveal that
50
51 372 the multiple operating point analysis did not eliminate the smearing effect in these two attempts. The computation
52
53 373 time increases significantly when increasing the operating points from one to two to three; 22.659, 129.682, and
54
55 374 197.368 seconds, respectively. The comparison between implanted and predicted degradation factors is shown in Fig.
56
57 375 6 (b), where all three cases could not achieve high diagnostic accuracy. Table 6 illustrates the computation burden of
58
59 376 three attempts by conventional diagnosis methods with an increasing number of operating points.
60
61
62
63 20
64
65
1
2
3
4 377 4.3 Case 2: Sequential Diagnostic with All Measurements
5
6 378 In Case 2, the sequential diagnostic method is employed, where the diagnosis is partitioned into four steps, as
7
8 379 shown in Fig. 3. Convergence processes for these four sequential steps are demonstrated in Fig. 7 (a). It is clear from
9
10 380 Fig. 7 (b) that the sequential diagnostic has the capability to estimate the degradation factor with greater accuracy than
11
12 381 the conventional diagnostic method (Case 1). It is worth noting that the last step of the proposed sequential method
13
14 382 requires three operating points to determine the degradation factor and eliminate the smearing effect, something that
15
16 383 the conventional diagnostic could not do.
17
18 0.0016
19 0.030
Step 1 Case 2
20 Step 2 0.0014

21 0.025 Step 3
Step 4 0.0012
22 Relative Error [%]
0.020
23 0.0010
RMSEMea

24
0.015 0.0008
25
26 0.010 0.0006
27
28 0.005
0.0004

29
0.0002
30 0.000
31 0.0000
32 1 2 3 4 5 6 XLPC,E XHPC,E XHPT,E XLPT,E XFPT,E
Number of Iterations XLPC,F XHPC,F XHPT,F XLPT,F XFPT,F
33
34 (a) Convergence of diagnosis. (b) Relative diagnostic error.
35 384 Fig. 7 Convergence performance and relative diagnostic error for Case 2.
36
37 385
38
39 386 The four sequential diagnostic steps converge in 0.014, 0.011, 0.013, and 0.032 seconds, respectively, as shown
40
41 387 in Table 7. The proposed diagnostic process converges in 0.070 seconds (sum of all four sequential steps) for the
42
43 388 simultaneous deterioration of all five rotational components. The reason for this fast convergence lies in the reduction
44
45 389 of matrix dimensions in the iteration algorithm and the reduced total number of calls to the engine sub-models (Table
46
47 390 7). Comparing the total number of calls to engine sub-models in Table 6 and Table 7 illustrates that the sequential
48
49 391 diagnostic requires less computation than the conventional diagnosis. Moreover, it is evident that the novel sequential
50
51 392 diagnostic process is also superior to the traditional diagnostic method in terms of accuracy, as shown in Fig. 6 (b)
52
53 393 and Fig. 7 (b). It should be pointed out that the level of diagnostic accuracy achieved by this method remains the same,
54
55 394 even for smaller levels of engine component deterioration.
56
57 395
58
59 396
60
61
62
63 21
64
65
1
2
3
4 397 Table 7 Diagnostic results of Case 2.
5
6 Items Symbols Units Step 1 Step 2 Step 3 Step 4
7 Computation Time 𝐶𝑇 Second 0.014 0.011 0.013 0.032
8
Measurement Error 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 - 2.4E-6 7.6E-6 7.6E-6 3.6E-6
9
10 Degradation Factor Error 𝑅𝑀𝑆𝐸𝑋 - 5.2E-6 3.0E-6 3.4E-6 5.7E-6
11 Operating Point 𝑛 - 1 1 1 3
12 No. of Iteration Steps 𝑁𝐼𝑆 - 6 6 6 6
13 No. of total calls to engine sub-models 𝑁𝑂𝑀𝑜𝑑𝑒𝑙 - 18 18 18 438
14
15 398
16
399 4.4 Case 3: Sequential Diagnostic with Reduced Number of Measurements
17
18
400 For aero engines, the indirect measurement of turbine power output may not be possible or lead to poor precision.
19
20
401 Hence, it is worth testing the sequential diagnostic method, without the power output measurement of the last turbine,
21
22
23 402 in order to assess the suitability of the proposed approach for aero engine applications. It follows that the 𝐹𝑃𝑇 fault
24
25 403 diagnosis step, shown in Fig. 3, will be adapted so that it does not require the shaft power of the 𝐹𝑃𝑇, shown in Fig.
26
27 404 4.
28
29 405 As can be seen in Table 1, there is no measurement available at the outlet of the 𝐻𝑃𝑇 due to the high gas
30
31 406 temperature. Moreover, the sequential diagnosis with multiple operating points can estimate the correct degradation
32
33 407 factor for 𝐻𝑃𝑇 and 𝐿𝑃𝑇 in Case 2. Hence, it is worth testing the possibility of eliminating the measurements 𝑇3 and
34
35 408 𝑃3 at the 𝐿𝑃𝐶 outlet.
36
37 0.05 0.025
Step 1 Case 3
38 Step 2
39 0.04 Step 3
0.020
40
41
Relative Error [%]

0.03
42 0.015
RMSEMea

43
44 0.02
45 0.010

46
0.01
47
0.005
48
49 0.00
50 0.000
51 0 2 4 6 8 10 XLPC,E XHPC,E XHPT,E XLPT,E XFPT,E
52 Number of Iterations XLPC,F XHPC,F XHPT,F XLPT,F XFPT,F
53 (a) Convergence of diagnosis. (b) Relative diagnostic error.
54
55 409 Fig. 8 Convergence performance and relative diagnostic error for Case 3.
56 410
57
58
59
60
61
62
63 22
64
65
1
2
3
4 411 In Case 3, the NLGPA includes three sequential steps, as shown in Fig. 4. The convergence performance of the
5
6 412 three steps is shown in Fig. 8 (a). It is clear from Fig. 8 (b) that the sequential diagnostic with reduced number of
7
8 413 measurements could still predict the degradation factor with excellent precision.
9
10 414 The degradation diagnosis for the three steps converges in 0.012, 0.065, and 0.031 seconds, respectively, as shown
11
12 415 in Table 8. In total, the sequential diagnostic takes 0.108 seconds for the simultaneous deterioration of all five
13
14 416 rotational components. Although the computation time has increased by 54.29% when compared with Case 2 (0.070
15
16 417 seconds) due to increased matrix dimensions and computation burden attributed to the multiple operating point
17
18
418 analysis (Table 8), the measurement of power output and sensors at 𝐿𝑃𝐶 exit can be removed in this situation. It is
19
20
419 noted that all three steps of the proposed sequential method require three operating points to determine the degradation
21
22
23 420 factor and eliminate the smearing effect under the reduced measurement condition.
24
25 421 Table 8 Diagnostic results of Case 3.
26
27 Symbol
Items Units Step 1 Step 2 Step 3
28 s
29 Computation Time 𝐶𝑇 Second 0.012 0.065 0.031
30 Measurement Error 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 - 1.0E-6 3.7E-6 3.6E-6
31 Degradation Factor Error 𝑅𝑀𝑆𝐸𝑋 - 5.5E-6 8.8E-6 8.5E-5
32
33 Operating point 𝑛 - 3 3 3
34 No. of Iteration Steps 𝑁𝐼𝑆 - 6 9 6
35 No. of total calls to engine sub-models 𝑁𝑂𝑀𝑜𝑑𝑒𝑙 - 54 432 438
36 422
37
38 423 Comparison of the total number of calls to the engine sub-models as shown in Table 6 and Table 8, highlights that
39
40 424 sequential diagnosis with a reduced number of measurements requires less computation than the conventional method
41
42
425 [45]. Additionally, it is clear that the novel sequential diagnosis with reduced number of measurements still has a
43
44
426 precision advantage over the traditional diagnostic method [45], as shown in Fig. 6 (b) and Fig. 8 (b). Furthermore,
45
46
427 the 𝑅𝑀𝑆𝐸𝑋 of the 𝐻𝑃𝑇 and 𝐿𝑃𝑇 diagnoses in Case 3 is increased by an order of magnitude in comparison with Case
47
48
49 428 2. Nevertheless, the 𝑅𝑀𝑆𝐸𝑋 is still quite small and less than 1E-4. Despite a slight sacrifice in computation efficiency
50
51 429 and diagnostic precision, reducing the number of sensors can reduce both the initial and operating cost of the gas
52
53 430 turbine, but a compromise is always necessary between diagnostic accuracy and number of sensors installed.
54
55 431 The relative error of 10 degradation factors defined in Eq. (6) for Cases 1-3 are summarized in Fig. 9 in order to
56
57 432 demonstrate the merits of the proposed sequential diagnosis. It is clear that the diagnostic precision of the proposed
58
59 433 method is far more accurate than the conventional method [45]. The computation time, the maximum relative error of
60
61
62
63 23
64
65
1
2
3
4 434 each case and the number of required sensors for Cases 1-3 are summarized in Table 9. There is a significant
5
6 435 improvement when comparing the novel sequential and conventional methods in both diagnostic precision and
7
8 436 computation speed. In Case 2, the maximum relative error is 1.4E-3 %, which is substantially smaller than that of the
9
10 437 traditional method, since the sequential diagnostic method can resolve the smearing effect. The calculation speed of
11
12 438 the method is over 300 times faster than the conventional method, and this is attributed to the reduction of the matrix’s
13
14 439 dimensions in the iterative diagnostic algorithm. In Case 3, the method demonstrates a maximum relative error of
15
16 440 2.2E-2 %, and the computation time of the proposed approach is more than 200 times faster than the traditional method.
17
18
19 441 Table 9 Comparison between three diagnostics cases.
20
21 Case 1 Case 1 Case 1
Items Symbols Units Case 2 Case 3
22 Attempt 1 Attempt 2 Attempt 3
23 Computation Time 𝐶𝑇 Second 22.659 129.682 197.368 0.070 0.108
24 Max Relative Error 𝜆𝑚𝑎𝑥 % 6.52 4.53 4.45 1.4E-3 2.2E-2
25 No. of Sensors Required 𝑁𝑂𝑆 - 16 16 16 16 13
26
27 442
28
7.00
29 Case 1 - Attempt 1
30 6.00 Case 1 - Attempt 2
31
Case 1 - Attempt 3
Relative Error [%]

32 5.00
33 Case 2
34 4.00
Case 3
35
36 3.00
37
38 0.03
39 0.02
40
41 0.01
42
43 0.00
44 XLPC,E XHPC,E X X X
45 XLPC,F XHPC,F HPT,E XHPT,F LPT,E XLPT,F FPT,E XFPT,F
46 443
47
48 444 Fig. 9 Comparison of relative diagnostic error for three diagnostic cases.
49
50 445 4.5 Case 4: Effect of the Measurement Noise
51
52 446 To analyze the influence of noise on diagnostic accuracy, measurement noise generated via the engine
53
54 447 thermodynamic model is imposed on the measured values. We have assumed that the noise is subject to a Gaussian
55
56 448 distribution with zero mean and standard deviation one, and the maximum deviation is shown in Table 10 [49,50]. We
57
58 449 assume that the engine measurements are recorded every 15 seconds, and for 5 minutes of steady-state operation, this
59
60 450 means that there are 20 operating points available. Then, 20 sets of measurements with random noise added are
61
62
63 24
64
65
1
2
3
4 451 generated for 𝐹𝑃𝑇 fault diagnosis (Step 1) in Case 2. Two examples of such noisy measurements, namely 𝑁𝐹𝑃𝑇 and
5
6 452 𝑇9 , are shown in Fig. 10. The 21st point is the post-filtered value of the measurement obtained from the previous 20
7
8 453 noisy data by an averaging filter [51], which is an averaging process of each measurement. The 22nd point represents
9
10 454 the actual value of measurement, with no noise added. The preprocessed measurements (21st point) are used in Case
11
12
455 4, which is essentially a repetition of 𝐹𝑃𝑇 fault diagnosis (Step 1) in Case 2.
13
14
15 456 Table 10 Maximum measurement noise [49,50].
16
17 Measurement Range Typical Error
18 0.204-3.06 0.50%
19 Pressure [atm]
0.544-31.30 0.5% or 0.125 atm whichever is greater
20
-65-290 ±3.3
21
22 Temperature [℃] 290-1000 ±√2.52 + (0.0075 × 𝑇)2
23 1000-1300 ±√3.52 + (0.0075 × 𝑇)2
24 Shaft Power - 0.10%
25
26 Rotational Speed - 0.10%
27 Relative Humidity [%] - 0.10%
28 Up to 5450 63.4
29 Fuel Flow [kg/h]
Up to 12260 142.7
30
457
31
32 4806
33 Without Noise Filter
34 4804 With Noise Filter
35 4802 True Value
NFPT

36 4800
37
38 4798
39 4796
40 1094
41 1092
42
T9 [Kelvin]

1090
43
1088
44
45 1086
46 1084
47 1082
48 0 5 10 15 20
49 Sample Points
50 458
51
52 459 Fig. 10 Measured values of 𝑵𝑭𝑷𝑻 and 𝑻𝟗 with added Gaussian measurement noise, post-filtered measurement,
53 460 and true value.
54
55 461 The diagnostic performance of Case 4 is shown in Fig. 11 (a). The relative error of 𝐿𝑃𝑇 degradation parameters,
56
57 462 predicted from the sequential diagnostic with respect to the implanted faults, is shown in Fig. 11 (b). Table 11 indicates
58
59 463 that the noisy measurements impact the prediction of the degradation factors when comparing with 𝐹𝑃𝑇 fault
60
61
62
63 25
64
65
1
2
3
4 464 diagnosis (Step 1) in Case 2. Nevertheless, the estimation errors of the degradation factors are still relatively small
5
6 465 and acceptable even when the effect of measurement noise has been considered, with the relative errors of efficiency
7
8 466 and flow capacity being 0.11% and 0.16%, respectively.
9
10
11 467 Table 11 Diagnostic results of Case 4.
12
13 Items Symbols Units Step 1
14 Computation Time 𝐶𝑇 Second 0.011
15 Measurement Error 𝑅𝑀𝑆𝐸𝑀𝑒𝑎 - 2.6E-6
16 Degradation Factor Error 𝑅𝑀𝑆𝐸𝑋 - 1.4E-3
17
18 Operating point 𝑛 - 1
19 No. of Iteration Steps 𝑁𝐼𝑆 - 6
20 No of total calls to engine sub-models 𝑁𝑂𝑀𝑜𝑑𝑒𝑙 - 18
21 468
22 0.008
Step 1
0.18
Case 4
23 0.007 0.16
24 0.006 0.14
25
0.005
26
Relative Error [%]

0.12
RMSEMea

27 0.004
0.10
28 0.003
0.08
29 0.002
30 0.06
0.001
31 0.04
32 0.000
0.02
33 -0.001
1 2 3 4 5 6
34 0.00
XFPT,F
Number of Iterations XFPT,E
35
36 (a) Convergence of diagnosis. (b) Relative diagnostic error.
37 469 Fig. 11 Convergence performance and relative diagnostic error for Case 4.
38
39 470
40
41 471 From a practical point of view, several aspects should be taken into consideration in the application of the proposed
42
43 472 diagnostic method. Ensuring appropriate systems and support of data filtration [52] and sensor validation [53] should
44
45 473 be a priority for fault diagnosis. There is, therefore, a definite need for map adaptation [8,54] of the engine model to
46
47 474 align with actual engine measurements based on healthy data and generic maps from the open literature. This
48
49 475 adaptation process should be carried out every time maintenance is carried out. This is crucial in refining and updating
50
51 476 the engine model that establishes the benchmark upon which any further diagnostic analysis is going to be based. The
52
53 477 proposed diagnostic method is not only adequate for off-line steady-state diagnosis but can also be applied in real-
54
55 478 time since the fast convergence of the algorithms provides flexibility in its implementation. The comparison between
56
57 479 Case 2 and Case 3 reveals that the sequential diagnostic scheme should be reconfigured when there are changes in the
58
59 480 type of engine, location of sensors and the number of available sensors. This is a typical limitation from a modeling
60
61
62
63 26
64
65
1
2
3
4 481 perspective and something that characterizes the model-based diagnostics methods, which rely heavily on the engine
5
6 482 model. However, the substantial accuracy and efficiency improvements of this method, in comparison to existing
7
8 483 model-based techniques [45], trade-off the previous limitation.
9
10 484 The findings of this study have several practical implications. More broadly, to develop a full picture of condition-
11
12 485 based maintenance, additional studies such as fault prognosis [11,37,55], maintenance optimization [56], economic
13
14 486 analysis [57], engine emissions modeling [58] could be implemented to complement the developed model. Further
15
16 487 studies, which take these aspects into account, should enable a prognostic health management solution for the gas
17
18
488 turbine engine. It will not only improve the reliability and availability of gas turbines but also economically benefit
19
20
489 engine stakeholders. The desirable features and excellent performance capabilities of the proposed method motivate
21
22
490 the inclusion of transient operating conditions in the diagnosis and variable geometry of compressors; tasks that the
23
24
25 491 authors are currently engaged in.
26
27
28 492 5. Conclusions
29
30 493 This study proposes a novel sequential diagnostic method for gas turbines with the primary aim of improving the
31
32 494 accuracy and computation speed when compared with the conventional model-based GPA. The engine performance
33
34 495 model is validated against commercial software for both the design point and off-design steady-state conditions. The
35
36 496 novel sequential diagnosis approach is evaluated via a well-used GPA method.
37
38 497 The conclusions drawn from this study are summarized as follows:
39
40 498 ⚫ The maximum relative errors between the developed engine performance model and GasTurb is less than
41
42 499 0.71% for the given test conditions.
43
44 500 ⚫ The developed sequential diagnostic algorithm is superior in both diagnostic accuracy and computation speed
45
46 501 to the conventional GPA, with a maximum relative error less than 1.4E-3 % and convergence in 0.070
47
48 502 seconds.
49
50
503 ⚫ The sequential diagnostic algorithm with reduced number of engine measurements outperforms the
51
52
504 conventional method. In such a case, the maximum relative error is only 2.2E-2 %, which is significantly
53
54
505 lower than the one achieved by the existing NLGPA method. Additionally, the proposed method converges
55
56
506 more than 200 times faster than the NLGPA.
57
58
59
60
61
62
63 27
64
65
1
2
3
4 507 ⚫ With the aid of a noise filter, the impact of measurement uncertainty can be reduced to an acceptable level
5
6 508 for engineering applications. In this situation, the relative errors of two degradation factors for the free power
7
8 509 turbine are 0.11% and 0.16% for efficiency and mass flow capacity, respectively.
9
10 510 The present study establishes a sequential framework for engine performance diagnosis with better precision and
11
12 511 computation efficiency than the existing model-based method, by eliminating the smearing effect and reducing the
13
14 512 matrix dimensions in the iterative diagnostic algorithm. Additionally, the novel diagnostic method with reduced
15
16 513 measurement could potentially decrease the cost of engine operations and the flow disturbances caused by sensors in
17
18
514 the engine gas path.
19
20
515 Overall, the results of the case studies demonstrate the superiority of this method in terms of diagnostic accuracy
21
22
516 and computation time, even with a reduced number of measurements, in comparison with the existing NLGPA method.
23
24
25 517 This new approach is sufficiently modular to be applied in all types of gas turbine engines with the potential to support
26
27 518 the operation and maintenance of gas turbine assets more cost-effectively and accurately than existing GPA methods.
28
29 519
30
31
32 520 Acknowledgments
33
34 521 This work is supported by the Fundamental Research Funds for the Central Universities under Grant No.
35
36 522 DUT19ZD218, the National Natural Science Foundation of China under Grant Nos. 61722302 and 61573069.
37
38 523
39
40
41 524 Appendix A
42
43
44
45 525 A.1 Computation Efficiency
46
47 526 This appendix provides detailed information about the iterative diagnostic algorithm from a computational
48
49 527 perspective. The sub-model of engine performance simulation consumes most of the computational resources during
50
51 528 GPA. The conventional diagnostics covers the “outer loop” (NLGPA: iteration of the degradation factor) and “inner
52
53 529 loop” (iteration of the engine performance model). The size of the Jacobian matrix, involved in the iteration, influences
54
55 530 the computation speed.
56
57 531 Fig. A. 1 demonstrates the dimensions of the Jacobian matrix for the outer and inner loops of the conventional
58
59 532 method [45]. The traditional diagnostic method includes 11 gas path measurements and 10 health parameters (Table
60
61
62
63 28
64
65
1
2
3
4 533 2), which establish the dimensions of the matrix in Fig. A. 1 (pictured at right). The subscript “𝑛” denotes the number
5
6 534 of operating points that determine the number of calls to the engine model. In this study, “𝑛” is set to one, two, and
7
8 535 three, which correspond to a maximum of 3 operating points. As far as off-design performance is concerned, the
9
10 536 developed model needs 8 iteration variables (Appendix B), and the dimension of the matrix for engine simulation are
11
12
537 shown in Fig. A. 1. Such a nested iteration reduces the computational efficiency of the engine diagnosis dramatically
13
14
538 and could be affected by the smearing effect. The number of calls to the engine model in the diagnostic algorithm and
15
16
539 the number of calls to the engine sub-models in engine simulation for one step is also demonstrated in Fig. A. 1.
17
18
540
19
20
21
22
23
24
25
26
27
28
29
30 541
31
32 542 Fig. A. 1 Matrix dimensional analysis of conventional fault diagnosis.
33
34 543
35
36
37
38
39
40
41
42 544
43
44 545 Fig. A. 2 Matrix dimensional analysis of sequential diagnostic with all available measurements.
45 546
46
47 547 Contrary to the conventional method, the sequential diagnostic method needs to call only the specific sub-model
48
49 548 directly in the engine performance model, which can reduce computing complexity. Furthermore, the sequential
50
51 549 diagnosis has a lower number of matrix dimensions during iteration, which could diminish issues of dimensionality.
52
53 550 The size of the Jacobian matrix for the novel sequential approach is also highlighted, and the number of calls to each
54
55
551 engine sub-model during one iteration is demonstrated. For sequential diagnostics with all available measurements,
56
57
552 the number of dependent variables, independent variables, and calls to engine sub-models are summarized in Fig. A.
58
59
553 2 based on Fig. 3. The number of iteration variables is the longitudinal length of four matrices, which are 2, 2, 2, and
60
61
62
63 29
64
65
1
2
3
4 554 7 for each sequential step in order, respectively. In contrast, the convergence criteria are the transverse length of four
5
6 555 matrices, which are 2, 2, 2, and 9, respectively.
7
8 556 For sequential diagnostics with reduced measurements, the number of dependent variables, independent variables,
9
10 557 and calls to engine sub-models are summarized in Fig. A. 3, based on Fig. 4. The number of iteration variables is the
11
12 558 longitudinal length of three matrices, which are 2, 7, and 7 for each sequential step in order, respectively. In contrast,
13
14 559 the convergence criteria are the transverse length of three matrices, which are 3, 9, and 9, respectively. The comparison
15
16 560 between the conventional method and novel method on the computation burden is illustrated in Table A. 1. As the
17
18
561 actual computation burden relies on the convergence steps, all results presented in Table A. 1 are referred to a single
19
20
562 iteration step for obtaining the Jacobian matrix.
21
22
563
23
24
25
26
27
28
29
30
31
32
33 564
34 565 Fig. A. 3 Matrix dimensional analysis of sequential diagnostic with reduced number of measurements.
35
36 566
37
38
567 Table A. 1 Comparison of computation burden to each engine sub-model.
39
40 Item Symbol CD SDA SDR
41
42 No. of Calls to Intake Model 𝑁𝑂𝐼𝑁𝑇 𝑛 ×80 0 0
43 No. of Calls to Compressor Model 𝑁𝑂𝐶𝑂𝑃 𝑛 ×160 4 42
44 No. of Calls to Burner Model 𝑁𝑂𝐵𝑢𝑟𝑛 𝑛 ×80 0 0
45
46 No. of Calls Mixture Model 𝑁𝑂𝑀𝐼𝑋 𝑛 ×160 21 21
47 No. of Calls to Turbine Model 𝑁𝑂𝑇𝑈𝑅𝐵 𝑛 ×240 44 48
48 No. of Calls to Duct Model 𝑁𝑂𝐷𝑈𝐶𝑇 𝑛 ×80 0 0
49
50 No. of Calls to Nozzle Model 𝑁𝑂𝑁𝑜𝑧𝑧𝑙𝑒 𝑛 ×80 0 0
51 568 CD: Conventional Diagnostic.
52 569 SDA: Sequential Diagnostic with All Available Measurements.
53 570 SDR: Sequential Diagnostic with Reduced Number of Measurements.
54 571
55 572
56 573
57 574
58
59
60
61
62
63 30
64
65
1
2
3
4 575 Appendix B
5
6
7
576 B.1 Modular Modelling of Gas Turbine Components
8
9
10 577 For an industrial gas turbine, the engine consists of six major components: intake, compressor, burner, mixture,
11
12 578 turbine, duct, and exhaust nozzle. The algorithm of engine performance simulation was adapted from the method
13
14 579 described by [59–63].
15
16 580 ➢ Intake Model
17
18 581 For a stationary gas turbine engine, the engine Mach Number is zero, and it is assumed that the engine is installed
19
20 582 at sea level with no pressure loss at the intake. Hence, the modeling of intake only needs to decide the 𝑊𝐴𝑅 for
21
22 583 calculating the gas properties at the following simulation. The 𝑊𝐴𝑅 is calculated by Eq. (B.1) [63].
23
24
25 𝑊𝑊𝐴 0.622𝑃𝑠𝑎𝑡 (B.1)
𝑊𝐴𝑅 = =
26 𝑊𝐷𝐴 𝑃𝑎𝑚𝑏
− 𝑃𝑠𝑎𝑡
27 0.01𝑅𝐻
28
584 where 𝑊𝑊𝐴 is mass flow rate of water vapor, 𝑊𝐷𝐴 denotes the mass flow rate of dry air and 𝑃𝑠𝑎𝑡 denotes the saturation
29
30
31 585 pressure of water vapor.
32
33 586 𝑃𝑠𝑎𝑡 is obtained by Eq. (B.2), which is related to the ambient pressure and temperature [64].
34
35
17.502(𝑇𝑎𝑚𝑏 − 273.15) (B.2)
36 𝑃𝑠𝑎𝑡 = (1.0007 + 3.46 ∙ 10−6 ∙ 101.325𝑃𝑎𝑚𝑏 ) ∙ 0.61121 ∙ 𝑒𝑥𝑝
37 𝑇𝑎𝑚𝑏 − 32.18
38
39 587
40
41 588 ➢ Compressor Model
42
43 589 For calculating compressor performance, the inlet temperature (𝑇𝑖𝑛 ), pressure (𝑃𝑖𝑛 ), shaft rotational speed (𝑁),
44
45 590 compressor flow capacity degradation factor (𝑋𝐶,𝐹 ), and the efficiency degradation factor (𝑋𝐶,𝐸 ) should be known so
46
47 591 that the 𝑋𝐶,𝐹 and 𝑋𝐶,𝐸 are applied to scale the healthy compressor map by Eq. (1), where 𝑋𝐶,𝐹 and 𝑋𝐶,𝐸 are equal to
48
49 592 one at a healthy/clean state.
50
51 593 When shaft speed 𝑁 and inlet conditions are known, the corrected shaft rotational speed (𝐶𝑁) is expressed by Eq.
52
53
594 (B.3) [60,61].
54
55
56 (B.3)
57 (𝑁 √𝑇𝑖𝑛 )𝑂𝐷
𝐶𝑁 =
58 (𝑁 √𝑇𝑖𝑛 )𝐷𝑃
59
60 595 where the subscript “𝐷𝑃” and “𝑂𝐷” represents the design point and off-design point, respectively.
61
62
63 31
64
65
1
2
3
4 596 If the compressor pressure ratio (𝑃𝑅) is known, then the outlet pressure (𝑃𝑜𝑢𝑡 ) is determined by:
5
6
7 𝑃𝑜𝑢 𝑡 = 𝑃𝑖𝑛 ∙ 𝑃𝑅 (B.4)
8
9 597 As 𝐶𝑁 and 𝑃𝑅 are known, the compressor efficiency (𝐸𝑓𝑓𝐶 ) and corrected mass flow (𝐶𝑀) could be obtained
10
11 598 from the scaled component map. The inlet temperature, pressure, and flow capacity are known, hence the inlet mass
12
13 599 flow (𝑊𝑖𝑛 ) is given by:
14
15
16 𝑃𝑖𝑛 𝑃𝑆𝐿𝑆 (B.5)
17 𝑊𝑖𝑛 = 𝐶𝑀 ∙
18 √𝑇𝑖𝑛 𝑇𝑆𝐿𝑆
19
600 where the subscript “𝑆𝐿𝑆” represents the sea level static conditions.
20
21
601 The inlet entropy and enthalpy of the compressor are obtained as follows:
22
23
24 [𝑆𝑖𝑛 , 𝐻𝑖𝑛 ] = 𝐺𝑎𝑠𝑃𝑟𝑜𝑝[𝑇,𝑃] (𝑇𝑖𝑛 , 𝑃𝑖𝑛 , 𝐹𝐴𝑅, 𝑊𝐴𝑅) (B.6)
25
26 602 Hence, the enthalpy at isentropic compression (𝐻𝑖𝑠 ) is given by:
27
28
29 𝐻𝑖𝑠 = 𝐺𝑎𝑠𝑃𝑟𝑜𝑝[𝑆,𝑃] (𝑆𝑖𝑛 , 𝑃𝑜𝑢 𝑡 , 𝐹𝐴𝑅, 𝑊𝐴𝑅) (B.7)
30
31 603 And the outlet enthalpy is given by Eq. (B.8) [59] as follows:
32
33
34 𝐻𝑜𝑢 𝑡 = 𝐻𝑖𝑛 − (𝐻𝑖𝑛 − 𝐻𝑖𝑠 ) ∙ 𝐸𝑓𝑓𝐶 (B.8)
35
36 604 Then, the outlet temperature (𝑇𝑜𝑢𝑡 ) can be determined by the following relationship:
37
38
39 𝑇𝑜𝑢 𝑡 = 𝐺𝑎𝑠𝑃𝑟𝑜𝑝[𝐻,𝑃] (𝐻𝑜𝑢𝑡 , 𝑃𝑜𝑢 𝑡 , 𝐹𝐴𝑅, 𝑊𝐴𝑅) (B.9)
40
41 606 The calculation of bleeding in a compressor is based on modularizing computations; as demonstrated below for
42
43 607 one bleeding path. The required inputs for bleeding include the bleed pressure ratio fraction (𝐹𝑟𝑎𝑐𝑃𝑅 ) and bleed mass
44
45 608 flow rate fraction (𝐹𝑟𝑎𝑐𝑊 ).
46
47 609 Then, the bleed outlet pressure (𝑃𝑅𝑏𝑙𝑒𝑒𝑑 ) and mass flow rate (𝑊𝑏𝑙𝑒𝑒𝑑 ) are computed by Eqs. (B.10), (B.4) and
48
49 610 (B.11):
50
51
52 𝑃𝑅𝑏𝑙𝑒𝑒𝑑 = 𝑃𝑅 ∙ 𝐹𝑟𝑎𝑐𝑃𝑅 (B.10)
53
54 611
55
56
57 𝑊𝑏𝑙𝑒𝑒𝑑 = 𝑊𝑖𝑛 ∙ 𝐹𝑟𝑎𝑐𝑊 (B.11)
58
59
60
61
62
63 32
64
65
1
2
3
4 612 Hence, the bleeding outlet enthalpy (𝐻𝑏𝑙𝑒𝑒𝑑 ) and temperature (𝑇𝑏𝑙𝑒𝑒𝑑 ) are determined by Eqs. (B.7), (B.8) and
5
6 613 (B.9), respectively.
7
8 614 The outlet mass flow (𝑊𝑜𝑢𝑡 ) is calculated by:
9
10
11 𝑊𝑜𝑢 𝑡 = 𝑊𝑖𝑛 − 𝑊𝑏𝑙𝑒𝑒𝑑 (B.12)
12
13 615 Finally, the compressor work (𝐶𝑊) is:
14
15
16 𝐶𝑊 = 𝑊𝑜𝑢 𝑡 ∙ (𝐻𝑜𝑢 𝑡 − 𝐻𝑖𝑛 ) + 𝑊𝑏𝑙𝑒𝑒𝑑 ∙ (𝐻𝑏𝑙𝑒𝑒𝑑 − 𝐻𝑖𝑛 ) (B.13)
17
18 616
19
20 617 ➢ Burner Model
21
22 618 The burner pressure drop (∆𝑃) can be obtained by Eq. (B.14), which relates the specified design point pressure
23
24 619 loss and kinetic head (𝐾𝐻) of burner inlet at both design point and off-design point conditions [62]:
25
26
27 ∆𝑃𝐷𝑃 𝐾𝐻𝐷𝑃 (B.14)
28 =
∆𝑃𝑂𝐷 𝐾𝐻𝑂𝐷
29
30 620 where the kinetic head is referred to Eq. (B.15) as follows:
31
32
33 𝑊𝑖𝑛2 ∙ 𝑇𝑖𝑛 (B.15)
34 𝐾𝐻 =
35
𝑃𝑖𝑛
36 621 The burner exit pressure is computed by:
37
38
39 𝑃𝑜𝑢 𝑡 = 𝑃𝑖𝑛 ∙ ∆𝑃𝑂𝐷 (B.16)
40
41 622 The enthalpy released (∆𝐻) by fuel combustion is given by:
42
43
44 ∆𝐻 = 𝑊𝐹𝑢𝑒𝑙 ∙ 𝐿𝐻𝑉 ∙ 𝐸𝑓𝑓𝐵 (B.17)
45
46 623 where 𝑊𝐹𝑢𝑒𝑙 is the burner fuel flow rate, 𝐿𝐻𝑉 is low heating value, and 𝐸𝑓𝑓𝐵 is burner efficiency.
47
48 624 The computation of the exit enthalpy of combustion gas is based on mass and energy conservation:
49
50
51 𝐻𝑖𝑛 ∙ 𝑊𝑖𝑛 + ∆𝐻 (B.18)
52
𝐻𝑜𝑢 𝑡 =
𝑊𝑜𝑢 𝑡
53
54 625 The 𝐹𝐴𝑅 of the combustor gas is determined through 𝑊𝐴𝑅, 𝑊𝐹𝑢𝑒𝑙 and 𝑊𝑜𝑢𝑡 as follows:
55
56
57 𝑊𝐹𝑢𝑒𝑙 (B.19)
58
𝐹𝐴𝑅 =
(𝑊𝑜𝑢𝑡 − 𝑊𝐹𝑢𝑒𝑙 ) ∙ (1 − 𝑊𝐴𝑅)
59
60 626 Finally, the burner exit temperature is obtained by Eq. (B.9).
61
62
63 33
64
65
1
2
3
4 627 ➢ Mixture Model
5
6 628 The mixture model is applied for mixing of core flow, and cooling flows in a constant area, when one inlet flow
7
8 629 is much smaller than the other. It is assumed that the total outlet pressure of mixed flow is equal to the total inlet
9
10 630 pressure of core flow. The mass flow of mixed flow can be determined based on mass conservation. Then, the enthalpy
11
12 631 of exit is:
13
14
15 𝑊𝑚𝑎𝑖𝑛 ∙ 𝐻𝑚𝑎𝑖𝑛 + 𝑊𝑏𝑙𝑒𝑒𝑑 ∙ 𝐻𝑏𝑙𝑒𝑒𝑑 (B.20)
16 𝐻𝑜𝑢 𝑡 =
𝑊𝑜𝑢 𝑡
17
18 632 Finally, the 𝐹𝐴𝑅 of mixed gas is referred to Eq. (B.19) and the outlet temperature of mixed flow is determined by
19
20 633 Eq. (B.9).
21
22 634
23
24 635 ➢ Turbine Model
25
26
636 For calculating the turbine performance model, the 𝑇𝑖𝑛 , 𝑃𝑖𝑛 , 𝑁, 𝑊𝐹𝑢𝑒𝑙 , turbine flow capacity degradation factor
27
28
637 𝑋𝑇,𝐹 , and turbine efficiency degradation factor 𝑋𝑇,𝐸 should be known and 𝑋𝑇,𝐹 are 𝑋𝑇,𝐸 are applied to scale the original
29
30
31 638 health map by Eq. (1), where 𝑋𝑇,𝐹 and 𝑋𝑇,𝐸 are equal to one at a healthy/clean state.
32
33 639 When shaft speed 𝑁 and the inlet conditions are known, the 𝐶𝑁 could be obtained by Eq. (B.3). If the turbine
34
35 640 expansion ratio of pressure (𝑃𝑅) is known, the turbine outlet pressure (𝑃𝑜𝑢𝑡 ) is given by:
36
37
38 𝑃𝑜𝑢𝑡 = 𝑃𝑖𝑛 𝑃𝑅 (B.21)
39
40 641 As 𝐶𝑁 and 𝑃𝑅 are known, the actual turbine efficiency (𝐸𝑓𝑓𝑇 ) and corrected mass flow (𝐶𝑀) could be obtained
41
42 642 from the scaled turbine map. Due to the turbine inlet temperature, pressure, and corrected mass flow are known, the
43
44 643 turbine inlet mass flow could be obtained by Eq. (B.5).
45
46 644 Since the 𝐸𝑓𝑓𝑇 is known, the turbine outlet enthalpy is computed by:
47
48
49 𝐻𝑜𝑢 𝑡 = 𝐻𝑖𝑛 − (𝐻𝑖𝑛 − 𝐻𝑖𝑠 ) ∙ 𝐸𝑓𝑓𝑇 (B.22)
50
645 The turbine outlet temperature could be calculated by Eq. (B.9), and turbine work (𝑇𝑊) is as follows:
51
52
53 𝑇𝑊 = 𝑊𝑖𝑛 ∙ (𝐻𝑖𝑛 − 𝐻𝑜𝑢 𝑡 ) (B.23)
54
55 646 ➢ Duct Model
56
57 647 In the Duct model, we considered a total pressure loss that could be obtained by Eq. (B.14). The outlet mass flow
58
59 648 and total enthalpy are the same as the inlet condition. Hence, the outlet temperature is calculated by Eq. (B.9).
60
61
62
63 34
64
65
1
2
3
4 649 ➢ Nozzle Model
5
6 650 For industrial gas turbines, the Mach number of nozzle exit flow is less than one since the nozzle is under sub-
7
8 651 critical condition. Hence, the following calculation will only discuss the calculation process for subsonic flow where
9
10
652 the nozzle exit static pressure (𝑝𝑜𝑢𝑡 ) is equal to the ambient pressure. Moreover, the assumed isentropic expansion
11
12
13 653 means that the static entropy (𝑠𝑜𝑢𝑡 ) is equal to total entropy (𝑆𝑜𝑢𝑡 ) and obtained by Eq. (B.6). The nozzle static
14
15 654 temperature (𝑡𝑜𝑢𝑡 ), static density (𝜌𝑜𝑢𝑡 ), heat capacity ratio (𝛾), and gas constant (𝑅𝑔 ) are given by:
16
17
18 [𝑡𝑜𝑢 𝑡 , 𝜌𝑜𝑢 𝑡 , 𝛾, 𝑅𝑔 ] = 𝐺𝑎𝑠𝑃𝑟𝑜𝑝[𝑆,𝑃] (𝑠𝑜𝑢 𝑡 , 𝑝𝑜𝑢 𝑡 , 𝐹𝐴𝑅, 𝑊𝐴𝑅) (B.24)
19
20 655 Then, the nozzle outlet velocity (𝑉𝑜𝑢𝑡 ) is calculated by:
21
22
23 (B.25)
24 𝑉𝑜𝑢𝑡 = 𝛾 ∙ 𝑅𝑔 ∙ 𝑡𝑜𝑢 𝑡
25
26 656 Finally, the determination of the nozzle exit mass flow is based on the component characteristic.
27
28
29 𝑊𝑜𝑢 𝑡 = 𝜌𝑜𝑢 𝑡 ∙ 𝑉𝑜𝑢 𝑡 ∙ 𝐴𝑜𝑢 𝑡 (B.26)
30
31 657 where 𝐴𝑜𝑢𝑡 is the cross-section area of the exhaust nozzle.
32
33 658
34
35
36 659 B.2 Map Scaling of Rotating Component at Design Point
37
38 660 The rotating component maps include the characteristic parameters, such as corrected shaft rotational speed,
39
40 661 pressure ratio, corrected mass flow rate, and component efficiency. However, the generic maps from the open literature
41
42 662 may not align with actual component characteristics in concern at the design point. Hence, the design point map
43
44 663 scaling is launched and referred to [46] to represent the actual component characteristic based on the design point
45
46 664 specification.
47
48 665 The scaling factor of pressure ratio (𝑆𝐹𝑃𝑅 ) is:
49
50
51 𝑃𝑅𝐷𝑃 − 1 (B.27)
52 𝑆𝐹𝑃𝑅 =
53 𝑃𝑅𝑚𝑎𝑝 − 1
54
55 666 where the subscript “DP” refers to the design point value, while the subscript “map” indicates the value obtained
56
57 667 through generic maps.
58
59 668 The scaling factor of corrected mass flow (𝑆𝐹𝐶𝑀 ) is given by:
60
61
62
63 35
64
65
1
2
3
4
5 𝑆𝐹𝐶𝑀 = 𝐶𝑀𝐷𝑃 𝐶𝑀𝑚𝑎𝑝 (B.28)
6
7 669 Similarly, the scaling factor of component efficiency (𝑆𝐹𝐸𝑓𝑓 ) is:
8
9
10 𝑆𝐹𝐸𝑓𝑓 = 𝐸𝑓𝑓𝐷𝑃 𝐸𝑓𝑓𝑚𝑎𝑝 (B.29)
11
12 670
13
14
15 671 B.3 Cross-Section Area of Exhaust Nozzle
16
17 672 For the design point, the nozzle exit entropy could be obtained by Eq. (B.6). By assuming that the exhaust gas is
18
19 673 expanding isentropically [63] to the ambient, the static entropy (𝑠𝑜𝑢𝑡 ) is equal to the total entropy (𝑆𝑜𝑢𝑡 ). Hence, 𝑡𝑜𝑢𝑡 ,
20
21 674 𝜌𝑜𝑢𝑡 , 𝛾 and 𝑅𝑔 are calculated by Eq. (B.24) and the 𝑉𝑜𝑢𝑡 is given by Eq. (B.25).
22
23
675 Finally, the exhaust nozzle's cross-section area is determined by Eq. (B.30), while exhaust mass flow rate is known
24
25
676 at the design point.
26
27
28 𝐴𝑜𝑢 𝑡 = 𝑊𝑜𝑢 𝑡 (𝜌𝑜𝑢 𝑡 ∙ 𝑉𝑜𝑢 𝑡 ) (B.30)
29
30 677
31
32
33 678 B.4 Gas Turbine Performance Model for RB211-24G
34
35 679 The gas turbine's performance simulation is based on the above algorithms for each engine component representing
36
37 680 any gas turbine. The balancing process of a triple-shaft industrial gas turbine engine is shown in Fig. B. 1 which is
38
39 681 adapted from [40]. Eight iteration variables are required for the iteration that are 𝑁𝐿𝑃 , 𝑃𝑅𝐿𝑃𝐶 , 𝑁𝐻𝑃 , 𝑃𝑅𝐻𝑃𝐶 , 𝑊𝐹𝑢𝑒𝑙 ,
40
41
682 𝑃𝑅𝐻𝑃𝑇 , 𝑃𝑅𝐿𝑃𝑇 , 𝑃𝑅𝐹𝑃𝑇 . Meanwhile, convergence criteria are shown in the left part of Fig. B. 1, where the subscript
42
43
44 683 “pre” refers to the mass flow from the previous block. In contrast, the subscript “loc” indicates the mass flow obtained
45
46 684 through component characteristics.
47
48 685
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63 36
64
65
1
2
3
4
5 Handle: Power setting (PS)
6
7
8 Guess: NLP , PRLPC, NHP, PRHPC , WFule, PRHPT , PRLPT, PRFPT
9
10
Data: LP shaft power
11 Intake Model
12 T1, P1, RH1 compatibility?
13
14 Data: CWLPC
15
Compressor
SFPR, SFCM,
16 Model (LPC)
SFEff, Map W3
17 W3, p re
18 compatibility?
19 Data:
Compressor
20 SFPR, SFCM, W3, loc
Model (HPC)
21 SFEff, Map
22
HP shaft power
23 Data: CWHPC
compatibility?
24 Burner Model
25 KHDP, PDP
26
27 Mixture W6
28 W6, p re
Model compatibility?
29
30
31 Data: W6, loc
32 Turbine Model
SFPR, SFCM, TWHPT
33 (HPT)
SFEff, Map
34
35
36 Mixture W8
W8, p re
37 Model compatibility?
38
39 Data: W8, loc
40 Turbine Model
SFPR, SFCM, TWLPT
41 (LPT)
SFEff, Map W9
42 W9, pre
43 compatibility?
44 Data:
45 Turbine Model W
SFPR, SFCM, 9,loc
46 (FPT)
SFEff, Map Power output
47 TWFPT
48 compatibility?
49 Data:
Duct Model
50 KHDP, PDP
51 W12
52 W12, pre
compatibility?
53
54 Data:
Nozzle Model W12, loc
55 Aout
56 PS
57 686
58
59 687 Fig. B. 1 Balancing process of the off-design performance simulation.
60
61
62
63 37
64
65
1
2
3
4 688 References
5
6
7 689 [1] Tahan M, Tsoutsanis E, Muhammad M, Abdul Karim ZA. Performance-based health monitoring, diagnostics and prognostics for
8 690 condition-based maintenance of gas turbines: A review. Appl Energy 2017;198:122–44.
9 691 [2] Aminyavari M, Mamaghani AH, Shirazi A, Najafi B, Rinaldi F. Exergetic, economic, and environmental evaluations and multi-objective
10 692 optimization of an internal-reforming SOFC-gas turbine cycle coupled with a Rankine cycle. Appl Therm Eng 2016;108:833–46.
11
693 [3] Marinai L, Probert D, Singh R. Prospects for aero gas-turbine diagnostics: A review. Appl Energy 2004;79:109–26.
12
13 694 [4] Hanachi H, Liu J, Kim IY, Mechefske CK. Hybrid sequential fault estimation for multi-mode diagnosis of gas turbine engines. Mech
14 695 Syst Signal Process 2019;115:255–68.
15 696 [5] Talebi SS, Tousi AM. The effects of compressor blade roughness on the steady state performance of micro-turbines. Appl Therm Eng
16 697 2017;115:517–27.
17
698 [6] Urban LA. Gas Turbine Engine Parameter Interrelationships. 2nd ed. Hamilton Standard Division of United Aircraft Corporation; 1969.
18
19 699 [7] Safiyullah F, Sulaiman SA, Naz MY, Jasmani MS, Ghazali SMA. Prediction on performance degradation and maintenance of centrifugal
20 700 gas compressors using genetic programming. Energy 2018;158:485–94.
21 701 [8] Kang DW, Kim TS. Model-based performance diagnostics of heavy-duty gas turbines using compressor map adaptation. Appl Energy
22 702 2018;212:1345–59.
23
24 703 [9] Sogut MZ, Yalcin E, Karakoc TH. Assessment of degradation effects for an aircraft engine considering exergy analysis. Energy
25 704 2017;140:1417–26.
26 705 [10] Hanachi H, Mechefske C, Liu J, Banerjee A, Chen Y. Performance-Based Gas Turbine Health Monitoring, Diagnostics, and Prognostics:
27 706 A Survey. IEEE Trans Reliab 2018;67:1340–63.
28
707 [11] Tsoutsanis E, Meskin N, Benammar M, Khorasani K. A dynamic prognosis scheme for flexible operation of gas turbines. Appl Energy
29
30 708 2016;164:686–701.
31 709 [12] Tsoutsanis E, Hamadache M, Dixon R. Real-Time Diagnostic Method of Gas Turbines Operating Under Transient Conditions in Hybrid
32 710 Power Plants. J Eng Gas Turbines Power 2020;142:1–10.
33 711 [13] Orozco DJR, Venturini OJ, Escobar Palacio JC, del Olmo OA. A new methodology of thermodynamic diagnosis, using the
34
35 712 thermoeconomic method together with an artificial neural network (ANN): A case study of an externally fired gas turbine (EFGT). Energy
36 713 2017;123:20–35.
37 714 [14] Tang S, Tang H, Chen M. Transfer-learning based gas path analysis method for gas turbines. Appl Therm Eng 2019;155:1–13.
38 715 [15] Korbicz J, Kościelny JM, Kowalczuk Z, Cholewa W. Fault Diagnosis: Models, Artificial Intelligence, Applications. 2th ed. New York,
39
716 US: Springer; 2012.
40
41 717 [16] Ying Y, Cao Y, Li S, Li J, Guo J. Study on gas turbine engine fault diagnostic approach with a hybrid of gray relation theory and gas-
42 718 path analysis. Adv Mech Eng 2016;8:1–14.
43 719 [17] Fentaye AD, Baheta AT, Gilani SI, Kyprianidis KG. A review on gas turbine gas-path diagnostics: State-of-the-art methods, challenges
44 720 and opportunities. Aerospace 2019;6.
45
46 721 [18] Hanachi H, Liu J, Mechefske C. Multi-mode diagnosis of a gas turbine engine using an adaptive neuro-fuzzy system. Chinese J Aeronaut
47 722 2018;31:1–9.
48 723 [19] Simon DL, Rinehart AW. Sensor Selection for Aircraft Engine Performance Estimation and Gas Path Fault Diagnostics. J Eng Gas
49 724 Turbines Power 2016;138:1–11.
50
725 [20] Jasmani MS, Li YG, Ariffin Z. Measurement selections for multi-component gas path diagnostics using analytical approach and
51
52 726 measurement subset concept. Proc ASME Turbo Expo 2010;3:569–79.
53 727 [21] Pinelli M, Spina PR, Venturini M. Gas turbine health state determination: Methodology approach and field application. Int J Rotating
54 728 Mach 2012;2012.
55 729 [22] Hanachi H, Liu J, Banerjee A, Chen Y, Koul A. A physics-based modeling approach for performance monitoring in gas turbine engines.
56
730 IEEE Trans Reliab 2015;64:197–205.
57
58 731 [23] Lu F, Ju H, Huang J. An improved extended Kalman filter with inequality constraints for gas turbine engine health monitoring. Aerosp
59 732 Sci Technol 2016;58:36–47.
60
61
62
63 38
64
65
1
2
3
4 733 [24] Mohammadi E, Montazeri-Gh M. Performance enhancement of global optimization-based gas turbine fault diagnosis systems. J Propuls
5 734 Power 2016;32:214–24.
6
735 [25] Yang Q, Li S, Cao Y, Zhao N. Full and part-load performance deterioration analysis of industrial three-shaft gas turbine based on genetic
7
8 736 algorithm. Proc. ASME Turbo Expo, vol. 6, Seoul, South Korea: ASME Turbo Expo 2016: Turbomachinery Technical Conference and
9 737 Exposition; 2016, p. 1–12.
10 738 [26] Sun J, Zuo H, Liang K, Chen Z. Bayesian Network-Based Multiple Sources Information Fusion Mechanism for Gas Path Analysis. J
11 739 Propuls Power 2016;32:611–9.
12
13 740 [27] Yang Q, Li S, Cao Y. An IMM-GLR Approach for Marine Gas Turbine Gas Path Fault Diagnosis. Math Probl Eng 2018;2018.
14 741 [28] Yang Q, Li S, Cao Y. Multiple model-based detection and estimation scheme for gas turbine sensor and gas path fault simultaneous
15 742 diagnosis. J Mech Sci Technol 2019;33:1959–72.
16 743 [29] Cherchi E, Guevara CA. A Monte Carlo experiment to analyze the curse of dimensionality in estimating random coefficients models with
17
744 a full variance-covariance matrix. Transp Res Part B Methodol 2012;46:321–32.
18
19 745 [30] Lei L, Xu H, Xiong X, Zheng K, Xiang W, Wang X. Multiuser Resource Control With Deep Reinforcement Learning in IoT Edge
20 746 Computing. IEEE Internet Things J 2019;6:10119–33.
21 747 [31] Daroogheh N, Meskin N, Khorasani K. A Dual Particle Filter-Based Fault Diagnosis Scheme for Nonlinear Systems. IEEE Trans Control
22 748 Syst Technol 2017;26:1317–34.
23
24 749 [32] Tsoutsanis E, Meskin N, Benammar M, Khorasani K. A component map tuning method for performance prediction and diagnostics of
25 750 gas turbine compressors. Appl Energy 2014;135:572–85.
26 751 [33] Yang Q, Li S, Cao Y. A strong tracking filter based multiple model approach for gas turbine fault diagnosis. J Mech Sci Technol
27 752 2018;32:465–79.
28
753 [34] Lu F, Wang Y, Huang J, Huang Y, Qiu X. Fusing unscented Kalman filter for performance monitoring and fault accommodation in gas
29
30 754 turbine. Proc Inst Mech Eng Part G J Aerosp Eng 2018;232:556–70.
31 755 [35] John S. Microsoft Visual C# Step by Step. 9 edition. Hoboken, USA: Microsoft Press; 2018.
32 756 [36] Sun W. Research on Performance Calculation Model of RB211-24G Gas Turbine. MEng Thesis, China University of Petroleum Beijing,
33 757 China, 2017.
34
758 [37] Tsoutsanis E, Meskin N. Derivative-driven window-based regression method for gas turbine performance prognostics. Energy
35
36 759 2017;128:302–11.
37 760 [38] Zhou D, Yao Q, Wu H, Ma S, Zhang H. Fault diagnosis of gas turbine based on partly interpretable convolutional neural networks.
38 761 Energy 2020;200:117467.
39 762 [39] Plis M, Rusinowski H. A mathematical model of an existing gas-steam combined heat and power plant for thermal diagnostic systems.
40
41 763 Energy 2018;156:606–19.
42 764 [40] Song Y, Gu C, Ji X. Development and validation of a full-range performance analysis model for a three-spool gas turbine with turbine
43 765 cooling. Energy 2015;89:545–57.
44 766 [41] NASA. NASA computer program for calculating of the Chemical Equilibrium with Application, Cleveland, OH: NASA Glenn Research
45
767 Center 2015.
46
47 768 [42] Li Y-G. Gas Turbine Diagnostics. Cranfield, Bedford, UK: Cranfield University Msc Thermal Power Course Notes; 2019.
48 769 [43] SIEMENS Products & Services. SGT-A35 (Industrial RB211) aeroderivative gas turbine 2020.
49 770 https://new.siemens.com/global/en/products/energy/power-generation/gas-turbines/sgt-a30-a35-rb.html (accessed July 2, 2020).
50 771 [44] Kurzke J. GasTurb 13: A Program to Calculate Design and Off-Design Performance of Gas Turbines 2017.
51
52 772 [45] Li J, Ying Y. A Method to Improve the Robustness of Gas Turbine Gas-Path Fault Diagnosis Against Sensor Faults. IEEE Trans Reliab
53 773 2018;67:3–12.
54 774 [46] Ogaji SOT, Sampath S, Singh R, Probert SD. Parameter selection for diagnosing a gas-turbine’s performance-deterioration. Appl Energy
55 775 2002;73:25–46.
56
776 [47] Li J, Ying Y. Gas turbine gas path diagnosis under transient operating conditions: A steady state performance model based local
57
58 777 optimization approach. Appl Therm Eng 2020;170.
59 778 [48] Chen YZ, Li YG, Newby MA. Performance simulation of a parallel dual-pressure once-through steam generator. Energy 2019;173:16–
60 779 27.
61
62
63 39
64
65
1
2
3
4 780 [49] Chen YZ, Li YG, Newby MA. Gas path diagnostics for a once-through steam generator. Proc. ASME Turbo Expo, vol. 3, Phoenix,
5 781 Arizona, USA: 2019, p. 1–11.
6
782 [50] Amirkhani S, Chaibakhsh A, Ghaffari A. Nonlinear robust fault diagnosis of power plant gas turbine using Monte Carlo-based adaptive
7
8 783 threshold approach. ISA Trans 2019.
9 784 [51] Kim P. Kalman filter for beginners: with MATLAB examples. North Charleston, S.C., United States: CreateSpace Independent
10 785 Publishing Platform; 2011.
11 786 [52] Hu RL, Granderson J, Auslander DM, Agogino A. Design of machine learning models with domain experts for automated sensor selection
12
13 787 for energy fault detection. Appl Energy 2019;235:117–28.
14 788 [53] Palmé T, Fast M, Thern M. Gas turbine sensor validation through classification with artificial neural networks. Appl Energy
15 789 2011;88:3898–904.
16 790 [54] Kim S, Kim K, Son C. A new transient performance adaptation method for an aero gas turbine engine. Energy 2020;193:116752.
17
791 [55] Zagorowska M, Schulze Spüntrup F, Ditlefsen AM, Imsland L, Lunde E, Thornhill NF. Adaptive detection and prediction of performance
18
19 792 degradation in off-shore turbomachinery. Appl Energy 2020;268.
20 793 [56] Aretakis N, Roumeliotis I, Doumouras G, Mathioudakis K. Compressor washing economic analysis and optimization for power
21 794 generation. Appl Energy 2012;95:77–86.
22 795 [57] Kotowicz J, Brzęczek M, Job M. The thermodynamic and economic characteristics of the modern combined cycle power plant with gas
23
24 796 turbine steam cooling. Energy 2018;164:359–76.
25 797 [58] Owebor K, Oko COC, Diemuodeke EO, Ogorure OJ. Thermo-environmental and economic analysis of an integrated municipal waste-
26 798 to-energy solid oxide fuel cell, gas-, steam-, organic fluid- and absorption refrigeration cycle thermal power plants. Appl Energy
27 799 2019;239:1385–401.
28
800 [59] Linares JI, Montes MJ, Cantizano A, Sánchez C. A novel supercritical CO2 recompression Brayton power cycle for power tower
29
30 801 concentrating solar plants. Appl Energy 2020;263:114644.
31 802 [60] Bracco S, Delfino F. A mathematical model for the dynamic simulation of low size cogeneration gas turbines within smart microgrids.
32 803 Energy 2017;119:710–23.
33 804 [61] Tsoutsanis E, Meskin N. Dynamic performance simulation and control of gas turbines used for hybrid gas/wind energy applications. Appl
34
805 Therm Eng 2019;147:122–42.
35
36 806 [62] Kim MJ, Kim JH, Kim TS. The effects of internal leakage on the performance of a micro gas turbine. Appl Energy 2018;212:175–84.
37 807 [63] Kurzke J, Halliwell I. Propulsion and Power: An Exploration of Gas Turbine Performance Modeling. Cham, Switzerland: Springer
38 808 International Publishing AG, part; 2018.
39 809 [64] Buck AL. New equations for computing vapour pressure and enhancement factor. J Appl Meteorol 1981;20:1527–32.
40
41 810
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63 40
64
65

You might also like