Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Aerospace: Probabilistic Model For Aero-Engines Fleet Condition Monitoring

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

aerospace

Article
Probabilistic Model for Aero-Engines Fleet
Condition Monitoring
Valentina Zaccaria 1, *, Amare D. Fentaye 1 , Mikael Stenfelt 1,2 and
Konstantinos G. Kyprianidis 1
1 Simulation and Optimization for Future Industrial Applications, Mälardalen University,
72123 Västerås, Sweden; amare.desalegn.fentaye@mdh.se (A.D.F.); mikael.stenfelt@mdh.se (M.S.);
konstantinos.kyprianidis@mdh.se (K.G.K.)
2 SAAB Aeronautic, 58254 Linköping, Sweden
* Correspondence: valentina.zaccaria@mdh.se

Received: 1 May 2020; Accepted: 25 May 2020; Published: 26 May 2020 

Abstract: Since aeronautic transportation is responsible for a rising share of polluting emissions, it is
of primary importance to minimize the fuel consumption any time during operations. From this
perspective, continuous monitoring of engine performance is essential to implement proper corrective
actions and avoid excessive fuel consumption due to engine deterioration. This requires, however,
automated systems for diagnostics and decision support, which should be able to handle large
amounts of data and ensure reliability in all the multiple conditions the engines of a fleet can be found
in. In particular, the proposed solution should be robust to engine-to-engine deviations and different
sensors availability scenarios. In this paper, a probabilistic Bayesian network for fault detection
and identification is applied to a fleet of engines, simulated by an adaptive performance model.
The combination of the performance model and the Bayesian network is also studied and compared
to the probabilistic model only. The benefit in the suggested hybrid approach is identified as up to
50% higher accuracy. Sensors unavailability due to manufacturing constraints or sensor faults reduce
the accuracy of the physics-based method, whereas the Bayesian model is less affected.

Keywords: diagnostics; performance model; Bayesian network; turbofan; fleet

1. Introduction
The stringent goals of reduced CO2 emissions set by the International Civil Aviation Organization
(ICAO) represent a challenge for aero-engine manufacturers, who are pushing the design to extreme
conditions and incrementing the use of alternative fuels [1]. With increasing design complexity
and more challenging operating conditions, airlines must rely on effective systems for monitoring,
diagnostics, and prognostics. Accurate diagnosis and prognosis of engine health condition is surely
necessary to reduce maintenance cost, since engines are currently accounting for about 25% of the
operating cost of an aircraft [2]. Furthermore, it is also essential to optimize flight operations depending
on the health status of the engine and avoid unnecessary efficiency decrement and consequent excessive
fuel consumption.
Airlines have to deal with large fleets of engines. Advances in acquisition systems and
computational tools have rendered available a huge amount of data for diagnostics: high-resolution,
on-wing data are now often available instead of a single flight data point. However, this demands
fully automated diagnostic systems that can effectively replace human reasoning. The challenges that
such a diagnostic system must face are multiple. Besides the capability of processing large amounts of
data in a reasonable time, robustness toward engine-to-engine variations due to serial deviations and
different degradation history is a pressing demand.

Aerospace 2020, 7, 66; doi:10.3390/aerospace7060066 www.mdpi.com/journal/aerospace


Aerospace 2020, 7, 66 2 of 19

Adaptive and regression models to mitigate the effects of engine-to-engine variations have been
proposed in the literature [3–6]. Chu et al. developed scalable regression models for a fleet of aircraft
engines for anomaly detection to solve two major issues: performance variability within the fleet
due to production scatter and potential computational burden due to the large amount of collected
data [3,4]. The algorithm was tested on three scenarios: performance anomaly detection, performance
shift detection over different flights, and a malfunctioning aircraft, implanting six cases over 200
aircraft. A scalable approach for anomaly detection of a large population of aircraft was also proposed
by Ohlsson et al., with the important advantage of no training needed for the algorithm [6].
Diagnostics of a gas turbine comprises of three fundamental steps: first, anomaly detection,
in which a malfunction or an anomalous condition is detected based on measured data; second, fault
isolation, which determines the component(s) where the possible malfunction is located; and third, fault
identification, which determines the fault severity and hence provides precious information for possible
corrective actions [7]. The numerous techniques for fault detection, isolation, and identification can be
generally grouped into physics-based methods [8,9] (mostly based on the gas path analysis proposed
by Urban [10]), data-driven methods [11], and hybrid methods [12]. Extensive reviews of gas turbine
diagnostic methods for aircraft engines and land-based turbomachines exist in the literature [7,13].
The most common faults that affect aero-engine performance over the lifetime are compressor fouling,
which is usually characterized by a decrease in both compressor efficiency and flow capacity, and
turbine erosion, which often leads to decreased efficiency and increased flow capacity.
Gas path analysis (GPA) based on performance models of the engine is still one of the most
common techniques. Various non-linear and adaptive models [14,15], also in combination with machine
learning methods [16,17], have been proposed to overcome the typical shortcomings of GPA, such as
smearing effect and high sensitivity to measurement noise [18,19]. Many efforts over the years have
been made on identifying optimal sets of measurements and optimal combinations of measured values
and performance factors to achieve desired diagnostics capability [9,20]. A limitation in the use of
performance models is the frequent lack of proprietary information about commercial units, in case the
service provider is not the manufacturer. A way around this limitation and retaining high diagnostics
accuracy has been achieved by employing machine learning techniques such as neural network [21],
support vector machine [22], Bayesian network [23], etc. A common challenge when applying any
data-driven method to a fleet of engines is the large amount of data needed for training, and the limited
robustness toward serial deviations, as previously discussed. Rarely the proposed solutions in the
open literature are tested on a (real or simulated) fleet.
Due to its intrinsic capability to deal with uncertain information, a Bayesian network (BN) has
been used very often for constructing a probabilistic model for diagnostics purpose. The advantage
of BN over other data-driven methods is that expert knowledge can be combined with information
from data, reducing the need for historical data and relying more on the engineering knowledge.
The applications on fault classification for gas turbines have demonstrated the success of this
method [23–26]. Sensor validation through hierarchical BN models was also proposed by different
authors [27,28]. One additional application explored in recent decades is the use of BN for information
fusion; for instance, Kyriazis et al. fed the outcomes of various diagnostics methods to a BN to
establish the most likely fault scenario [29,30]. All these examples were proven successful when
applied to a single engine, but no study on the effect of engines serial deviation or operating conditions
was performed.
Among the work done on fleet monitoring, Martinez et al. proposed a fuzzy logic-based technique
for engine health monitoring (EHM) that classified the engines of the fleet in four classes: good, normal,
highly deteriorated, and bad [31]. Their approach was a significant step toward a method to minimize
maintenance cost in a fleet. A GPA approach was presented and tested on flight data from a KLM
aircraft assessing the effect of sensor noise, but complete fleet monitoring and diagnostics were not
presented [32]. Kraft et al. proposed a method to optimize the maintenance schedule in a fleet based
Aerospace 2020, 7, 66 3 of 19

on a novel high-fidelity modeling approach [33]. The authors claimed that a reduction of 10% in
maintenance cost was achieved with their solution.
In this work, we are presenting the development of a tool for gas turbine diagnostics that
combines an adaptive performance model and a probabilistic BN for fault detection and identification.
The advantages of both strategies are exploited by a combination of the two:

• Model errors and uncertainties are reduced in the final outcome because of model adaptation to
baseline values;
• Smearing effects are mitigated in the BN layer;
• Data correction is provided via model adaptation, improving the BN performance;
• Soft sensors from the model can be included in the BN to increase the accuracy.

The robustness and flexibility of the integrated hybrid system are demonstrated by testing it
on a fleet of engines characterized by serial deviations, different deterioration levels, and various
flight conditions. The novelty of the work is to compare benefits and shortcomings of combined
physics-based and data-driven methods when applied to the whole fleet rather than a single engine.

2. Diagnostics Framework Description


The diagnostic framework used in this work consists of a multi-layer approach. First, the unreliable
measurements such as values out of physical boundaries (e.g., a negative pressure) or frozen
measurements are discarded, and subsequently, the remaining measurements are fed to a physics-based
adaptive model for a first fault detection and isolation procedure. Finally, a probabilistic BN is used in
combination with the performance model for fault isolation and identification. In this work, the benefits
of the proposed solution for fleet monitoring and diagnostics will be demonstrated.
The purpose of the adaptive model is twofold. During the first performance test, the measured data
from the healthy engine are compared to the model outputs and the model performance parameters
vary to minimize the error between the two; the obtained adapted model at baseline conditions
(denoted as AD,ref in the following sections) includes serial deviations from the nominal engine.
This adaptive baseline can be used to calculate measurement deviations (or residuals) during the
engine lifetime. The second purpose is to correct the measurements with respect to ambient conditions
and flight conditions, as further explained in the following sub-sections.
In this work, three methods are compared: the use of the BN method only, the use of the adaptive
model for gas path analysis, and the hybrid method that combines both.

2.1. The Adaptive Model


The model under consideration simulates the performance of a three-shaft turbofan engine.
The model was developed with the in-house software EVA and previously presented [34]. The various
components of the engine, depicted in Figure 1, were modeled according to mass and energy balance
equations, Gibbs free energy equations, and representative components maps, under the assumption
of ideal gases. The adaptation scheme was based on the minimization of residuals between target
values and model outputs, which form a square Jacobian matrix.
Aerospace 2020, 7, 66 4 of 19

Figure 1. Schematic diagram of the modeled engine.

The rows of the matrix represent the residuals between model outputs and measurements while
the columns represent state variables that are varied at each iteration until all the residuals are below a
predefined threshold. The Jacobian is a square matrix, i.e., the number of state variables and target
values has to be the same. Each state variable is automatically associated with a residual depending on
the impact that the state variable has on the corresponding output. For diagnostic purposes, the state
variables that the model has to adjust include the performance deviation factors, i.e., the deviation
factor in isentropic efficiency and flow capacity for each component:
• ∆ηIPC , ∆W IPC
• ∆ηHPC , ∆W HPC
• ∆ηHPT , ∆W HPT
• ∆ηIPT , ∆W IPT
• ∆ηLPT , ∆W LPT
Please note that in this work, only the intermediate and high pressure compressors and the three
turbines were considered to illustrate the proposed method. Deviation factors are defined as:

∆η = η f ault − ηdesign (1)

W f ault
∆W = (2)
W design
The target values are instead coming from the sensor measurements; it is hence notable in this
approach that the number of available measurements needs to match the number of considered
performance deviation factors. If this is not the case, relationships between components performance
deviations have to be assumed to reduce the number of state variables (for example the ratio between
efficiency deterioration in the three turbines could be fixed, or the ratio between efficiency and flow
capacity deterioration in each component). The effect of varying the number of sensors was evaluated
in this work.
The selection of sensors in an ideal case can be based on the following criteria:
• Eliminating measurements with high mutual correlation (>90%), since no additional useful
information is provided by a measurement that is highly correlated with an already existing one;
• Selection of the measured values that are mostly affected by a change in the considered
state variables.
The model can run ten thousands of data points in a few seconds on a normal computer, making
it a suitable tool for monitoring and diagnostics of an entire fleet.

2.2. The Matching Scheme


Typically, a diagnostics problem consists of finding a state vector x that contains all the performance
deviations (e.g., ∆η and ∆W for each component) from the measurement deviations vector z, where
Aerospace 2020, 7, 66 5 of 19

all the measurements are normalized with the value at reference condition, as illustrated in Figure 2.
As shown in [35], we can define an influence matrix H as in Equation (3).

z = Hx (3)

There are two aspects in measurement selection, one purely mathematical concerning which
variables would be theoretically more suitable for identification of specific faults [36,37], and the
other practical concerning which sensors are actually available on an aircraft. Measurements selection
was first performed on a theoretical basis to evaluate faults observability; in this step, all the output
variables were considered even if normally not measured in a real engine. Each deviation parameter
was varied by 1% one at a time and the resulting influence matrix was analyzed, which contains
deviations in measured values caused by a unitary variation in state variable, as illustrated in Figure 2.

Figure 2. Illustrative example of influence matrix H.

Correlation between measurement vectors was calculated (i.e., between the matrix columns) as
in [38]. Correlation between columns i and j of the influence matrix is expressed in Equation (4), where
M is the length of the measurement vectors zi and zj .

zi ·z j
cij = qP (4)
M 2
PM 2
m=1 zi,m m=1 z j,m

A high correlation level was considered if larger than 90%; the measurements with a higher
number of high correlation with other sensor measurements were discarded, until no correlated
measurements remained. Successively, the influence of each performance deviation factor on the
remaining measurements was analyzed, i.e., on each element of the influence matrix in the sensor’s
column. Measurements were discarded if a deviation smaller than three times the sensor standard
deviations was observed for all simulated faulty cases, because the induced deviations were comparable
to measurement noise. In other words, for a simulated fault case to be included, its induced deviations
must be larger than sensors uncertainty.
Finally, the correlation between matrix rows was calculated (i.e., between measurements deviations
induced by different faults) also with Equation (4), which showed a quite large correlation between
∆ηHPC and ∆ηHPT (89%). This suggests that correct identification of these two deviation parameters
may be challenging in this engine configuration. The sensors selection results for this engine are shown
in Section 4.
The adaptation scheme was also successfully applied in previous work to a 50 MW industrial gas
turbine [39] and a micro gas turbine for heat and power generation [40], demonstrating the flexibility
of the approach. This represents the first use of the adaptive model.
Aerospace 2020, 7, 66 6 of 19

The second use of the adaptive model, as presented in [39], is for data correction with respect to
ambient and flight conditions, which allows combining this physics-based approach with a data-driven
diagnostic method for fault classification and identification. As such, the adaptation scheme can run in
each flight condition, since the values of ambient temperature and pressure are set as inputs in the
model, from which altitude and Mach number can be calculated. Once the performance deviation
factors (∆η,∆W)det are estimated by the adapted model at flight condition, AD,flight*, those factors
are used in the model to simulate the same engine at reference conditions (nominal flight condition
at ISA temperature), which is denoted as AD,ref*. With this approach, the effect of varying ambient
temperature and flight parameters such as altitude or Mach number is isolated and removed from
the detected performance deviations. The flowchart of the algorithm is shown in Figure 3, where
solid arrows represent data from measurements and dashed arrows represent the estimated deviation
factors in efficiency and flow capacity.

Figure 3. Flowchart of fault detection and data correction procedure with the adaptive model.

2.3. Bayesian Network Fault Classification


Once the measured data are corrected with respect to ambient and flight conditions through the
model at reference conditions, their residuals with respect to healthy reference conditions can be used
in a data-driven algorithm to identify the cause of the engine performance deviation. The advantage
of coupling the adaptive model with a data-driven algorithm is that other malfunctions such as
mechanical failures (e.g., a valve leakage) or sensor failures could also be detected. In fact, numerous
mechanical faults show up as deviations in thermodynamic parameters, and a further analysis is
needed to isolate the correct cause. In addition, more measurements could be included in this second
step e.g., vibration measurements but also soft sensors (i.e., model output variables that are not
measured), which enhances the capability of the diagnostic system.
A Bayesian network is a graphical probabilistic model that represents the probabilistic relationships
between causes and effects according to the Bayes theorem, reported in Equation (5). The equation
represents a casual statement of the kind X →Y, where X causes Y and Y takes the role of an observable
effect of X. P(Y) is called the prior probability, while P(Y|X) is called the posterior probability. The factor
that relates the two, P(X|Y)/P(X), is called the likelihood ratio.

P(X|Y)
P(Y|X) = P(Y ) (5)
P(X )
Aerospace 2020, 7, 66 7 of 19

A BN comprises of a direct acyclic graph (DAG) that includes nodes and edges, and a conditional
probability table (CPT) for each node. The DAG is the graphical representation of the model, where
the causes are represented with parent nodes and the effects with child nodes; the edges represent
conditional dependencies and connect causes and effects in an acyclic way, i.e., it is impossible to start
from a node and traverse the entire network following edges and subsequent nodes.
Both DAG structure and CPTs can be either constructed by experience or learned from data.
Hybrid methods are also possible, where the human experience intervenes for incomplete or incorrect
data. In this work, the BN structure was built manually, while the CPTs were fitted from data generated
by the performance model. The selected BN model included five parent nodes, one for each component
(IPC, HPC, HPT, IPT, and LPT), and one child node for each sensor measurement. Parent nodes were
defined with four possible states, corresponding to different deterioration levels: Normal (N) for no
performance deviation, Very Low (VL) for performance deviations within manufacturing tolerance,
Low (L) for fault severity between 1.1% and 2.2%, and Medium (M) for fault severity between 2.2% and
3.3%. The fault severity S was a function of efficiency and flow capacity deviation calculated according
to Simon [41] as in Equation (6). This work focused on small fault severity magnitude because they are
usually more difficult to detect. Fault severities higher than 3% could also be included by increasing
the number of states in the parent nodes.
s
!2
∆w − 1
S = −∆η· 1+ (6)
∆η

The ratio between the flow capacity deviation percentage ∆W − 1 and the efficiency deviation
was assumed constant and equal to 2 [41]. Each of the child nodes was instead defined with 20 states.
These states represent non-overlapping intervals of measurements deviations from the reference
value, and were selected by dividing the span between minimum and maximum deviations into
20 intervals. The network was built in the HUGIN Expert software environment [42] and the Expectation
Maximization algorithm was applied to estimate the CPTs from training data.
An additional parent node was added for subsequent tests representing a bleed valve (BV) leakage,
with three states: N (0% leak), L (up to 2% leak), and M (between 2% and 4% leak). These tests will
illustrate how additional faults such as valves leakage or mechanical failures can be included in the BN
classification and in the hybrid approach.
The BN inference can of course be applied directly to measured data instead of adapted data from
the performance model; this scenario was compared to the proposed strategy to assess the benefit
of a hybrid method over a purely data-driven one. When measured data were used directly for the
data-driven fault classification, data correction with ambient conditions was performed as suggested by
Volponi [43]. Measured parameters were corrected with temperature and pressure ratios (θ = Tamb /Tref ,
∆ = pamb /pref ) according to Equation (7), where the exponents a and b were set as suggested in [43].

z
zcorr = (7)
θa δb

3. Methodology
The objective of the work was to test the diagnostic framework on a fleet of engines to
demonstrate the robustness of the approach. Different units of the same engine differ because
of manufacturing tolerances, which make them perform differently even when brand new. During
operations, the differences are enlarged because of different flight conditions experienced. Generally,
data-driven techniques have to be trained on a space that considers all possible serial deviations,
measurements noise, and flight conditions, and their performance reduces if the engine conditions
change significantly. These reasons make the use of such techniques for diagnostics challenging.
Aerospace 2020, 7, 66 8 of 19

Fleet data were generated by applying a Gaussian distribution to each performance deviation
factor as shown in Table 1 and simulating in this way 200 engines that differ in design efficiency and
flow capacity for each rotating component. Standard deviation values were chosen consistently with
common engine performance as per the Author’s experience. It is important to note that a deviation
factor on the isentropic efficiency was added to the reference value (ηref + ∆η), while a deviation factor
in flow capacity was a coefficient multiplying the desired mass flow (W ref *∆W), hence, the nominal
value is 1.

Table 1. Gaussian distribution characteristics for engine serial deviations simulations (Case 1).

Parameter Mean 3σ
∆η 0.00 0.5%
∆W 1.00 1%

To simulate variations in flight conditions, the deviations in efficiency and flow capacity were kept
constant and some flight conditions parameters were varied with a uniform distribution as illustrated
in Table 2. It is important to note that altitude and Mach number were set as model inputs to simulate
the faulty engines, while for model adaptation, these parameters were calculated from static and total
pressure measurements and hence subject to uncertainty.

Table 2. Uniform distributions characteristics for flight parameters (Case 2).

Parameter Max Value Min Value


Altitude 11,000 m 11,000 m
Ambient temperature TISA + 15 K TISA − 10 K
Mach number 0.85 0.82

The nominal values at Top of Climb (ToC) for the considered engine are listed in Table 3. The model
is representative of a 311 kN engine with technology levels consistent with entry into service in 1995.

Table 3. Nominal conditions at ToC.

Variable Value
Flight altitude 10,668 m
Flight Mach no. 0.82
Bypass ratio 4.7
Overall pressure ratio 34
HPT inlet temperature 1625 K
Specific thrust 167 N·s/kg

For Case 1, half of the engines were used for training the BN (i.e., for estimating the CPTs) and half
for test. For each of the 100 training engines, a deterioration pattern was simulated for each of the five
components including 21 levels of fault severity ranging from 0% to 3.3%. As a result, 10,500 points
formed the dataset for the BN. For Case 2, 100 random combinations of flight parameters were picked
from the selected range and the same fault pattern (21 severity levels * 5 components) was simulated
for each flight, for a total of other 10,500 data points.
For the tests, the 100 engines and the 100 flight parameters combinations that were not used for
CPTs estimation were employed. For each engine and each flight condition, one healthy and three fault
scenarios were simulated for each component, as presented in Table 4, for a total of 3200 tested points.
Aerospace 2020, 7, 66 9 of 19

Table 4. Simulated fault scenarios.

Fault Severity ∆η ∆W
0% (N) 0% 1.00
0.994 IPC/HPC
0.7% (VL) −0.3%
1.006 HPT/IPT/LPT
0.984 IPC/HPC
1.8% (L) −0.8%
1.016 HPT/IPT/LPT
0.973 IPC/HPC
2.9% (M) −1.35%
1.027 HPT/IPT/LPT

Gaussian noise was added to all sensor data. As mentioned in Section 2, the results of the hybrid
adaptive model and BN approach were compared with the results obtained by the BN alone and the
outcomes of the adaptive model alone, following the scheme of Figure 4.

Figure 4. Schematic diagram of the three compared approaches: BN only (a), adaptive model only (b),
and hybrid method with adaptive model and BN (c).

To test the capability to detect a mechanical fault, the training space was enlarged to include bleed
valve leakage in each engine of the fleet and at each flight conditions ranging from 0 to 4%. Then, a BV
leakage was simulated for the tested engines at 1.5% and 3% of the nominal flow.

4. Results
First, the most suitable sensor set for the model adaptation procedure was identified based on the
method described in Section 2. This step was applied to determine which measurements would be
required in an ideal case, and more realistic sensor sets were successively considered for comparison.
The influence of each performance deviation factor on the uncorrelated measurements is shown in
Figure 5 (for the variables number, refer to the diagram in Figure 1). The bars in the Figure represent
the deviation in the normalized measurements as in the influence matrix of Figure 2 induced by 1%
performance change. Where the induced deviation was smaller than 3σ (i.e., sensor standard deviation)
for all simulated faulty cases, the sensors were discarded. For example, T43, T46, W2, and W13 were
discarded at this stage. Measurements of fan and IPC inlet flow are often available during pass-off
tests and it was hence interesting to evaluate their potential use in the adaptation scheme for baseline
generation. However, their impact in the influence matrix led to discard them.
Aerospace 2020, 7, 66 10 of 19

From the plots of Figure 5, it was noted that ∆W HPC had a very small influence on the available
measurements; therefore, detection of HPC fault was expected to be more difficult. The sensors selected
for the matching scheme are listed in Table 5 with considered noise level together with the sensors for
the fuel flow and for ambient conditions.

Figure 5. Influence of performance parameters deviation on measurements of compressors (a) and


turbines (b).

The fuel flow was not used for model adaptation purpose but rather as input to run the model in
the same condition as the real engine. It can be observed from Table 1 that all the sensors necessary
to build the required adaptation scheme could be normally installed on a civil aircraft. Please note
that in this work, diagnostics was performed with simulated flight data; analyses performed during
pass-off tests on the ground would benefit from a larger sensor suite and possibly higher measurements
accuracy, hence a different matching scheme could be selected.
The sensor set in Table 5 is denoted as Set1. In this work, the influence of sensors selection was
studied by implementing the proposed method on other two possible sets of available measurements.
Set2 includes the measurements in Set1 except N46 and p43, while Set3 lacks N44 and p46 measurements.
Hence, the adaptation scheme has to discard two performance deviation factors for each of these two
sensors sets. According to the methodology illustrated in [38] and the influence diagrams in Figure 5,
Aerospace 2020, 7, 66 11 of 19

∆ηHPT and ∆ηLPT were removed from the Set2 scheme, while ∆W IPC and ∆W LPT were removed from
the Set3 scheme.

Table 5. Sensors list with considered noise level.

Sensor Name Measured Variable Noise (σ)


T1 Ambient temperature ±0.3%
p1 Ambient pressure (static) ±0.2%
p1t Ambient pressure (total) ±0.2%
T25 IPC outlet temperature ±0.4%
p25 IPC outlet pressure ±0.25%
T3 HPC outlet temperature ±0.4%
p3 HPC outlet pressure ±0.25%
p43 HPT outlet pressure ±0.25%
p46 IPT outlet pressure ±0.25%
T5 LPT outlet temperature ±0.4%
N46 LP shaft speed ±0.1%
N44 IP shaft speed ±0.1%
N4 HP shaft speed ±0.1%
Wf Fuel flow rate ±2%

4.1. Case 1—Production Scatter Effect


The results for the faulty engines with production scatter only are presented in Table 6.
The percentage of correctly identified data points for each fault is shown for the 1600 simulated
engines in the fleet (100 different engines with 16 levels of degradation severity). Although Normal and
Very Low were considered two separate states in the BN nodes to increase the CPT estimation accuracy,
the results are here combined for practical reasons, as any deviation within the production scatter can
be considered negligible. The first thing observed was that the identification of HPC deterioration was
less accurate, mainly due to the low influence that ∆W HPC had on the available measurements.
Both BN methods presented a very low percentage of false alarms, which is desirable. The BN
alone was quite effective in identifying L level of IPC deterioration (M level was mostly misclassified
as L), M level for HPT and IPT, and any level of deterioration for LPT.

Table 6. Correctly identified faults with production scatter.

BN Only Adaptive Model Only Hybrid Method


IPC N/VL 99% 86% 99%
L 72% 84% 93%
M 52% 88% 98%
HPC N/VL 99% 70% 98%
L 27% 73% 71%
M 30% 74% 87%
HPT N/VL 99% 75% 99%
L 63% 76% 97%
M 80% 72% 92%
IPT N/VL 99% 73% 99%
L 66% 69% 91%
M 80% 72% 98%
LPT N/VL 99% 94% 100%
L 75% 95% 99%
M 86% 94% 99%
BV N 99% - 92%
L 93% - 94%
M 94% - 92%
Aerospace 2020, 7, 66 12 of 19

Because of the correlation between some of the deviation parameters, up to 31% of the faults were
misclassified by the adaptive model only. It is also interesting to note that by only using the model
adaptation procedure, some extent of smearing effect was present, with up to 27% of the engines for
different fault cases showing additional faults together with the correct one, in a component that was
actually healthy.
The hybrid method considerably improved the diagnostics accuracy compared to the single BN
and reduced the number of false alarms compared to the adaptive model alone. Further improvements
in the BN classification could be achieved by e.g., increasing the number of states or the training space,
but the purpose of this work was to assess the improvements given by the adaptive model layer with
the same BN parameters and CPTs. An advantage compared to the adaptive model alone is that a fault
in the bleed valve could not be isolated by the model adaptation approach, which simply detected a
variation in performance characterized by deviations in several efficiency parameters. The combination
of adaptive model and BN could however detect and identify a BV leakage although the accuracy was
slightly lower than the single BN in case of engine production scatter.
Since the identification accuracy for some components was low, it is important to show also the
detection and isolation accuracy, reported in Table 7. This refers to correct detection and isolation of
the different faults irrespectively of the severity, i.e., it includes misclassifications between L and M.
It is possible to see that the majority of misclassifications in Table 6 were due to a wrong fault severity
attribution, which is not a major concern even though it would affect a downstream maintenance
schedule analysis. Only for the HPC, a fairly high percentage of misdetections (i.e., faulty engines
considered healthy) was observed with both BN methods.

Table 7. Detection and isolation accuracy with production scatter.

BN Only Adaptive Model Only Hybrid Method


IPC 91% 89% 92%
HPC 72% 85% 77%
HPT 92% 85% 100%
IPT 92% 83% 99%
LPT 91% 92% 99%
BV 97% - 95%

4.2. Case 2—Different Flight Conditions


Table 8 shows the diagnostics results from different flights of the same engine, i.e., with no
production scatter. Data from different flight conditions were corrected back to reference conditions
following two methods. For the “BN only”, measurements were corrected with respect to ambient
temperature and pressure as suggested by Volponi [43]. The results for the “Adaptive model only”
were obtained by running the adaptive model in flight conditions (AD,flight*); while in the “Hybrid
method”, the full procedure illustrated previously in Figure 3 was applied to correct the data with
respect to reference condition (AD,ref*).
In this case, the BN alone did not perform well because, even though the effect of different
ambient temperature and altitude was in part compensated by the parameters’ correction, the scatter
in measured data was still significant. It is expected that a larger and denser training space would be
needed to achieve higher accuracy. The adaptive model showed a higher accuracy, always above 70%,
but a notable smearing effect was still present, induced by sensor noise and mainly the uncertainty in
calculated altitude and Mach number. The hybrid method outperformed both the single BN and the
adaptive model, combining the benefits of both. This was true also for BV leakage identification.
As before, Table 9 presents the detection and isolation rate for the various faults. It is still evident
from Table 9 how the BN structure used for the hybrid method, if employed alone, did not achieve
satisfactory results. The hybrid method improved the accuracy even compared to the adaptive model.
Aerospace 2020, 7, 66 13 of 19

Table 8. Correctly identified faults in different flight conditions.

BN Only Adaptive Model Only Hybrid Method


IPC N/VL 95% 76% 98%
L 50% 78% 85%
M 53% 80% 93%
HPC N/VL 96% 77% 99%
L 40% 81% 70%
M 41% 83% 97%
HPT N/VL 96% 74% 99%
L 45% 77% 99%
M 47% 80% 97%
IPT N/VL 96% 71% 99%
L 45% 75% 99%
M 50% 74% 98%
LPT N/VL 96% 90% 99%
L 49% 91% 100%
M 51% 90% 100%
BV N 100% - 99%
L 23% - 98%
M 45% - 100%

Table 9. Detection and isolation accuracy in different flight conditions.

BN Only Adaptive Model Only Hybrid Method


IPC 55% 86% 92%
HPC 48% 86% 90%
HPT 47% 86% 99%
IPT 49% 84% 99%
LPT 50% 94% 100%
BV 37% - 99%

4.3. Effect of Sensors Set


The Case 1 and Case 2 tests were repeated with reduced sensors sets. Table 10 presents the
results for Case 1, i.e., in case of engine serial deviations, when Set2 and Set3 are employed for the BN
only, the adaptive model (AM), and the combined method (AM+BN). When Set2 is used, the ∆ηHPC
and ∆ηHPT rows in the influence matrix show 94% correlation, and the ∆ηIPC and ∆ηIPT rows have a
correlation of 88%. Hence, in the BN alone, IPC, HPC, and IPT faults were mostly misclassified with
Set2. A similar observation can be done for Set3, for which ∆ηHPC and ∆ηHPT present 94% correlation,
which affects the classification in this case of HPC faults. Overall however, the accuracy for the BN
alone was not much lower than with the full sensor Set1, with a significant decrement only in IPT and
LPT. When the adaptive model alone was used with Set2 and Set3, the overall diagnostics accuracy
was reduced, especially in the components whose deviation factor was removed from the matching
scheme. A considerable degradation in accuracy compared to Set1 was also observed for the hybrid
method with both sensor sets, since the errors in model prediction are propagated to the BN. However,
the results were in most cases better than the BN alone, in particular for Set3, because although the
matching scheme run with a reduced number of variables, the model outputs could still be used as soft
sensors in the BN to provide the missing information.
Aerospace 2020, 7, 66 14 of 19

Table 10. Identified faults with different sensors sets and production scatter.

BN Set2 BN Set3 AM Set2 AM Set3 AM+BN Set2 AM+BN Set3


IPC N/VL 99% 99% 84% 76% 95% 90%
L 50% 44% 82% 11% 61% 57%
M 54% 56% 86% 1.5% 81% 63%
HPC N/VL 98% 98% 77% 76% 95% 90%
L 44% 33% 79% 77% 36% 29%
M 52% 46% 82% 81% 65% 53%
HPT N/VL 99% 99% 98% 75% 95% 90%
L 70% 69% 55% 76% 63% 72%
M 84% 88% 48% 72% 81% 75%
IPT N/VL 98% 98% 41% 61% 95% 90%
L 50% 41% 45% 63% 68% 55%
M 53% 68% 53% 65% 84% 71%
LPT N/VL 99% 99% 98% 92% 95% 90%
L 52% 49% 54% 14% 67% 63%
M 62% 58% 48% 1% 70% 59%

The results for Case 2 with sensors Set2 and 3 are summarized in Table 11. The same considerations
on sensors influence on detectable performance deviations can be made, and the same trends in accuracy
loss were observed also for the engine in different flight conditions. It was evident how IPC and LPT
deterioration were hard to detect with Set3. The hybrid method provided also in this case an overall
higher accuracy than the BN alone, although the loss in accuracy from the results obtained with Set1
was higher.

Table 11. Identified faults with different sensors sets and flight conditions.

BN Set2 BN Set3 AM Set2 AM Set3 AM+BN Set2 AM+BN Set3


IPC N/VL 80% 79% 74% 75% 96% 80%
L 32% 15% 76% 11% 83% 16%
M 43% 38% 78% 2% 96% 31%
HPC N/VL 80% 79% 78% 77% 95% 75%
L 37% 23% 83% 81% 74% 29%
M 36% 25% 85% 83% 91% 57%
HPT N/VL 78% 78% 99% 74% 97% 84%
L 37% 36% 49% 77% 89% 64%
M 44% 43% 49% 80% 61% 64%
IPT N/VL 78% 79% 42% 63% 94% 78%
L 48% 46% 47% 71% 100% 66%
M 41% 40% 57% 67% 99% 82%
LPT N/VL 80% 80% 100% 93% 95% 90%
L 37% 33% 50% 9% 96% 21%
M 41% 39% 50% 0% 73% 33%

5. Discussion
The proposed hybrid method for fault diagnostics was proven to achieve satisfactory accuracy in
the isolation and identification of various components deterioration in a fleet of engines, although
severity identification for HPC deterioration presented up to 30% error. The combined adaptive model
and BN approach showed an overall accuracy of 95% for detection and isolation, with 94% correct
severity classifications for different flight conditions, and 92% for engines serial deviations, with
particularly poor performance for the HPC. The results were higher in both cases than for the same BN
applied directly on measured data without going through the first layer of model adaptation. Since the
model provided a correction with flight conditions and a first step of fault isolation and identification,
the BN in the hybrid method ended up working on better data, from which the effect of measurement
Aerospace 2020, 7, 66 15 of 19

noise, production scatter, and ambient conditions was already removed. The advantage of the hybrid
approach is also that a reduced dataset is necessary for estimating the CPT, since variations in flight
conditions do not need to be accounted for, which reduces the initial computational burden. Several
actions can of course be taken to improve the accuracy of the standalone BN, for example increasing
the training space, pre-treating the data with more advanced correction methods, or increasing the
number of nodes states. This work demonstrates however that if the BN layer is combined with
a model adaptation layer, the final accuracy is higher even with limited training data and reduced
network complexity.
The use of an adaptive performance model for estimating performance parameters deviations
resulted in successful fault identification for most of the engines. However, a smearing effect was
always observed due to the strong correlation between various performance parameters, and up to 39%
of false alarms occurred. In addition, mechanical faults could not be isolated. Overall, the outcome of
the model adaptation approach strongly depends on the available sensors; since the sensors-states
matrix needs to be square in the identification problem, if a sensor is malfunctioning one state needs
to be dropped, with a consequent reduction in diagnostics accuracy. For this reason, also the hybrid
approach suffered with a limited sensor set whereas the BN alone was observed to be more robust to
sensor unavailability. However, the overall accuracy of the hybrid method was still superior to the BN
alone even with limited sensors.
Although the overall accuracy was higher, the hybrid method did not give better results in all
cases, which leads to think that multiple solutions should run in parallel and a downstream information
fusion system should use conflicting suggestions for making a final decision.
The proposed strategy can be a solution for both on-line and off-line monitoring and diagnostics,
since the computational time for analyzing thousands of points is few seconds on a normal computer.
However, building the BN requires a significant effort and experience, or a considerable amount of
data for the network to learn by itself. A reliable and validated model may be essential to generate the
required data since large amounts of real deterioration data are seldom available.
A limitation of this work is that only single component faults were considered, while in reality,
several components are expected to degrade at the same time at different rates. Nonetheless,
the combination of adaptive model and BN could be proven beneficial also in this case. As such, once
a certain level of deterioration is estimated in one component, this can be incorporated into the model
to generate a new baseline that filters out that component degradation, and the BN could identify any
deviation from the new baseline caused for example by another part starting degrading or an abrupt
fault in a valve or others. This has to be demonstrated however with a dedicated study.

6. Conclusions and Future Work


In this work, a combination of physics-based gas path analysis and a probabilistic approach for
gas turbine diagnostics was presented and tested on a simulated fleet of engines. The fleet data were
generated by means of an adaptive performance model and comprised of engine-to-engine serial
deviations, different deterioration levels, and flight-to-flight variations. Performance model adaptation
and a Bayesian network classifier were tested separately and in combination. The use of the adaptive
model alone showed good results but a significant smearing effect was observed and the accuracy
dropped if less sensors than performance deviation factors were used. The Bayesian network alone
was more robust toward sensors availability but the results were not satisfactory, the detection and
isolation rate being 89% for production scatter but only 47% for different flight conditions. The hybrid
method was proven to be superior since it included the benefits of both and successful identification of
components degradation was demonstrated, with 94% and 96% correct isolation rate in presence of
engine-to-engine variations and flight-to-flight variations respectively.
For the future, a more advanced reasoning method will be developed to fuse information from
the single approaches and the hybrid one, as well as information coming from cross-correlations
of measurements in the fleet. Baseline adaptation over time to distinguish multiple components
Aerospace 2020, 7, 66 16 of 19

degradation and other mechanical failures will also be included in future work. It is expected that an
improved identification of simultaneous multiple faults and in presence of gradual deterioration will
be achieved.

Author Contributions: Conceptualization, V.Z., A.D.F. and K.G.K.; methodology, V.Z., A.D.F. and M.S.;
software, M.S. and K.G.K.; data processing and analysis, V.Z. and A.F.; writing—original draft preparation, V.Z.;
writing—review and editing, A.D.F., M.S. and K.G.K.; supervision and project administration, K.G.K. All authors
have read and agreed to the published version of the manuscript.
Funding: This research was funded by the Swedish Research Foundation under the national project DIAGNOSIS.
Acknowledgments: The authors would like to acknowledge the other project partners for their advices and
support, and Xin Zhao for the support with the software development.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of the
study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to
publish the results.

Nomenclature
Acronyms
AD Adapted
BN Bayesian network
BV Bleed valve
CPT Conditional probability table
DAG Direct acyclic graph
GPA Gas path analysis
HPC, HPT High pressure compressor/turbine
IPC, IPT Intermediate pressure compressor/turbine
ISA International standard atmosphere
LPC, LPT Low pressure compressor/turbine
Abbreviations
corr Corrected
det Deteriorated condition
flight Flight condition
flight* Flight faulty/deteriorated condition
ref Reference condition
ref* Reference faulty/deteriorated condition
Symbols
c Correlation factor
H Influence matrix
L Low
M Medium
N Normal
N Shaft speed
P Probability
p Pressure
S Fault severity
T Temperature
VL Very low
W Flow capacity
x Performance deviation vector
z Measurements deviation vector
Greek letters
δ Pressure coefficient
∆ Deviation
η Efficiency
θ Temperature coefficient
Aerospace 2020, 7, 66 17 of 19

References
1. “Carbon Offsetting and Reduction Scheme for International Aviation (CORSIA)”, in ICAO Annual Report 2017.
Available online: https://www.icao.int/annual-report-2017/Pages/default.aspx (accessed on 30 May 2019).
2. Marinai, L.; Probert, D.; Singh, R. Prospects for aero gas-turbine diagnostics: A review. Appl. Energy 2004, 79,
109–126. [CrossRef]
3. Chu, E.; Gorinevsky, D.; Boyd, S. Scalable Statistical Monitoring of Fleet Data. In Proceedings of the 18th World
Congress, The International Federation of Automatic Control, Milano, Italy, 28 August–2 September 2011.
4. Chu, E.; Gorinevsky, D.; Boyd, S. Detecting aircraft performance anomalies from cruise flight data.
In Proceedings of the AIAA Infotech@Aerospace, Atlanta, GA, USA, 20–22 April 2010.
5. Borguet, S.; Léonard, O.; Dewallef, P. Regression-Based Modeling of a Fleet of Gas Turbine Engines for
Performance Trending. J. Eng. Gas Turb. Power 2016, 138. [CrossRef]
6. Ohlsson, H.; Chen, T.; Khoshfetratpakazad, S.; Ljung, L.; Sastry, S.S. Scalable anomaly detection in large
homogeneous populations. Automatica 2014, 50, 1459–1465. [CrossRef]
7. Fentaye, A.D.; Baheta, A.T.; Gilani, S.I.; Kyprianidis, K.G. A Review of Gas Turbine Gas-Path Diagnostics:
State-of-the-Art Methods, Challenges and Opportunities. Aerospace 2019, 6, 83. [CrossRef]
8. Hosseini, S.H.R.; Khaledi, H.; Solani, M.R. New Model Based Gas Turbine Fault Diagnostics Using 1D Engine
Model and Nonlinear Identification Algorithms. In Proceedings of the ASME Turbo Expo 2009: Power for
Land, Sea and Air, Orlando, FL, USA, 8–12 June 2009; GT2009-59439.
9. Mathioudakis, K.; Kamboukos, P.; Stamatis, A. Gas turbine component fault detection from a limited number
of measurements. J. Power Energy 2014, 218, 609–618. [CrossRef]
10. Urban, L.A. Gas path analysis applied to turbine engine condition monitoring. J. Aircr. 1973, 10. [CrossRef]
11. Vatani, A.; Khorasani, K.; Meskin, N. Health monitoring and degradation prognostics in gas turbine engines
using dynamic neural networks. In Proceedings of the ASME Turbo Expo 2015: Turbine Technical Conference
and Exposition, Montréal, QC, Canada, 15–19 June 2015. GT2015-44101.
12. Romesis, C.; Aretakis, N.; Roumeliotis, I.; Alexiou, A.; Tsalavoutas, A.; Stamatis, A.; Mathioudakis, K.
Experience with condition-based maintenance related methods and tools for gas turbines. In Proceedings
of the Future of Gas Turbine Technology, 7th International Gas Turbine Conference, Brussels, Belgium,
14–15 October 2014.
13. Zhao, N.; Wen, X.; Li, S. A review on gas turbine anomaly detection for implementing health management.
Proceedings of ASME Turbo Expo 2016: Turbomachinery Technical Conference and Exposition, Seoul, Korea,
13–17 June 2016. Paper number GT2016-58135.
14. Mathioudakis, K. Comparison of Linear and Nonlinear Gas Turbine Performance Diagnostics. J. Eng. Gas
Turb. Power 2005, 127. [CrossRef]
15. Tsoutsanis, E.; Meskin, N.; Benammar, M.; Khorasani, K. A component map tuning method for performance
prediction and diagnostics of gas turbine compressors. Appl. Energy 2014, 135, 572–585. [CrossRef]
16. Kyriazis, A.; Mathioudakis, K. Enhanced fault localization using probabilistic fusion with gas path analysis
algorithms. J. Eng. Gas Turb. Power 2009, 131, 051601. [CrossRef]
17. Roumeliotis, I.; Aretakis, N.; Alexiou, A. Industrial gas turbine health and performance assessment with
field data. J. Eng. Gas Turb. Power 2017, 139, 051202. [CrossRef]
18. Ntantis, E. The Impact of Measurement Noise in GPA Diagnostic Analysis of a Gas Turbine Engine. J. Aeronaut
Aerosp. Eng. 2013, 2. [CrossRef]
19. Verbist, M.L.; Visser, W.P.J.; Van Buijtenen, J.P. Experience with gas path analysis for on-wing turbofan
condition monitoring. In Proceedings of the ASME Turbo Expo 2013: Turbine Technical Conference and
Exposition, San Antonio, TX, USA, 3–7 June 2013.
20. Chen, M.; Quan, H.L.; Tang, H. An Approach for Optimal Measurements Selection on Gas Turbine Engine
Fault Diagnosis. J. Eng. Gas Turb. Power 2015, 137, 071203. [CrossRef]
21. Tayarani-Bathaie, S.S.; Vanini, Z.S.; Khorasani, K. Dynamic neural network-based fault diagnosis of gas
turbine engines. Neurocomputing 2014, 125, 153–165. [CrossRef]
22. Xu, J. PHM-Oriented Integrated Fusion Prognostics for Aircraft Engines Based on Sensor Data. IEEE Sens. J.
2014, 14. [CrossRef]
Aerospace 2020, 7, 66 18 of 19

23. Romessis, C.; Mathioudakis, K. Bayesian network approach for gas path fault diagnosis. J. Eng. Gas Turb.
Power 2004, 128, 64–72. [CrossRef]
24. Romessis, C.; Stamatis, A.; Mathioudakis, K. Setting up a belief network for turbofan diagnosis with the aid
of an engine performance model. Int. Symp. Air Breath. Engines ISABE 2001, 1032, 19–26.
25. Lee, Y.K.; Mavris, D.N.; Volovoi, V.V.; Yuan, M.; Fisher, T. A fault diagnosis method for industrial gas turbines
using Bayesian data analysis. J. Eng. Gas Turbines Power 2010, 132, 041602. [CrossRef]
26. Kestner, B.K.; Lee, Y.K.; Voleti, G.; Mavris, D.N.; Kumar, V.; Lin, T. Diagnostics of highly degraded industrial
gas turbines using Bayesian networks. In Proceedings of the ASME 2011 Turbo Expo: Turbine Technical
Conference and Exposition, Vancouver, BC, Canada, 6–10 June 2011.
27. Ibarguengoytia Gonzales, P.H. Any Time Probabilistic Reasoning for Sensor Validation. Ph.D. Thesis,
University of Salford, Salford, UK, 1997.
28. Losi, E.; Venturini, M.; Manservigi, L.; Ceschini, G.F.; Bechini, G. Anomaly Detection in Gas Turbine Time
Series by Means of Bayesian Hierarchical Models. J. Eng. Gas Turb. Power 2019, 141. [CrossRef]
29. Kyriazis, A.; Arethkis, N.; Mathioudakis, K. Gas turbine fault diagnosis from fast response data using
probabilistic methods and information fusion. In Proceedings of the ASME Turbo Expo 2006: Power for
Land, Sea and Air, Barcelona, Spain, 8–11 May 2006. GT2006-90362.
30. Kyriazis, A.; Mathioudakis, K. Gas turbines diagnostics using weighted parallel decision fusion framework.
In Proceedings of the 8th European Turbomachinery Conference, Graz, Austria, 23–27 March 2009.
31. Martinez, A.; Sanchez, L.; Couso, I. Improved life cycle cost–reduced engine maintenance through engine
health monitoring genetic fuzzy system–Method validation and case study. In Proceedings of the ASME
Turbo Expo 2014: Turbine Technical Conference and Exposition, Düsseldorf, Germany, 16–20 June 2014;
Paper number GT2014-25639.
32. Verbist, M.L.; Visser, W.P.J.; Van Buijtenen, J.P. Gas path analysis on KLM in-flight engine data. In Proceedings
of the ASME Turbo Expo 2011: Turbine Technical Conference and Exposition, Vancouver, BC, Canada,
6–10 June 2011; Paper number GT2011-45625.
33. Kraft, J.; Sethi, V.; Singh, R. Optimization of Aero Gas Turbine Maintenance Using Advanced Simulation and
Diagnostic Methods. J. Eng. Gas Turb Power 2014, 136. [CrossRef]
34. Najafi Saatlou, E.; Kyprianidis, K.G.; Sethi, V.; Abu, A.O.; Pilidis, P. On the trade-off between minimum fuel
burn and maximum time between overhaul for an intercooled aeroengine. Proc. Inst. Mech. Eng. Part G J.
Aerosp. Eng. 2014, 228, 2424. [CrossRef]
35. Ganguli, R. Introduction. In Gas Turbine Diagnostics: Signal Processing and Fault Isolation; CRC Press: Boca
Raton, FL, USA, 2017; ISBN 9781138074422.
36. Bechini, G.; Ameyugo, G.; Marinai, L.; Singh, R. Gas path diagnostics: The importance of measurement
selection in the monitoring process. In Proceedings of the International Symposium of Air Breathing Engine,
Munich, Germany, 4–9 September 2005; ISABE-2005-1281.
37. Sowers, T.S.; Fittje, J.E.; Kopasakis, G.; Simon, D.L. Expanded application of the systematic sensor selection
strategy for turbofan engine diagnostics. In Proceedings of the ASME Turbo Expo 2009: Turbine Technical
Conference and Exposition, Orlando, FL, USA, 8–12 June 2009; GT2009-59251.
38. Stenfelt, M.; Zaccaria, V.; Kyprianidis, K. Automatic gas turbine matching scheme adaptation for robust GPA
diagnostics. In Proceedings of the ASME Turbo Expo 2019: Power for Land, Sea and Air, Phoenix, AZ, USA,
11–15 June 2019; GT2019-91018.
39. Zaccaria, V.; Stenfelt, M.; Sjunnesson, A.; Hansson, A.; Kyprianidis, K. A model-based solution for gas
turbine diagnostics: Simulations and experimental verification. In Proceedings of the ASME TURBO EXPO
2019: Power for Land, Sea and Air, Phoenix, AZ, USA, 11–15 June 2019; GT2019-90858.
40. Rahman, M.; Zaccaria, V.; Zhao, X.; Kyprianidis, K. Diagnostics-oriented modelling of micro gas turbines for
fleet monitoring and maintenance optimization. Processes 2018, 6, 1–22. [CrossRef]
41. Simon, D.L. Propulsion Diagnostic Method Evaluation Strategy (ProDiMES) User’s Guide. NASA/TM—2010-215840.
June 2010. Available online: https://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/20100005639.pdf
(accessed on 30 March 2020).
Aerospace 2020, 7, 66 19 of 19

42. Hugin. Hugin Expert A/S, Hugin Version 8.8. Available online: http://www.hugin.com (accessed on
30 March 2020).
43. Volponi, A.J. Gas Turbine Parameter Corrections. ASME J. Eng. Gas Turb. Power 1999, 121, 613–621. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like