Next Article in Journal
Energy-Efficient Trajectory Optimization for UAV-Enabled Cellular Communications Based on Physical-Layer Security
Next Article in Special Issue
Semi-Physical Simulation of Fan Rotor Assembly Process Optimization for Unbalance Based on Reinforcement Learning
Previous Article in Journal
An End-to-End UAV Simulation Platform for Visual SLAM and Navigation
Previous Article in Special Issue
Machine Learning Applications in Modelling and Analysis of Base Pressure in Suddenly Expanded Flows
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Neural Networks and Transfer Learning to Turbomachinery Heat Transfer

Institute of Aircraft Propulsion Systems, University of Stuttgart, Pfaffenwaldring 6, 70569 Stuttgart, Germany
*
Author to whom correspondence should be addressed.
Aerospace 2022, 9(2), 49; https://doi.org/10.3390/aerospace9020049
Submission received: 23 November 2021 / Revised: 11 January 2022 / Accepted: 13 January 2022 / Published: 20 January 2022

Abstract

:
Model-based predictive maintenance using high-frequency in-flight data requires digital twins that can model the dynamics of their physical twin with high precision. The models of the twins need to be fast and dynamically updatable. Machine learning offers the possibility to address these challenges in modeling the transient performance of aero engines. During transient operation, heat transferred between the engine’s structure and the annulus flow plays an important role. Diabatic performance modeling is demonstrated using non-dimensional transient heat transfer maps and transfer learning to extend turbomachinery transient modeling. The general form of such a map for a simple system similar to a pipe is reproduced by a Multilayer Perceptron neural network. It is trained using data from a finite element simulation. In a next step, the network is transferred using measurements to model the thermal transients of an aero engine. Only a limited number of parameters measured during selected transient maneuvers is needed to generate suitable non-dimensional transient heat transfer maps. With these additional steps, the extended performance model matches the engine thermal transients well.

1. Introduction

Today, more data sampled at higher frequency is available from engine in-flight operation. To use this data for predictive maintenance, online fault diagnostics, and fleet management, digital twins are of paramount importance. By their definition, digital twins fully describe their physical twin using so-called virtual information constructs [1]. Therefore, digital twins need to model their physical twin and dynamically update the model as new data becomes available [2,3]. To perform model-based analysis using high-frequency in-flight data, the model needs to represent the engine’s dynamics with high precision [4]. Additionally, the transient calculations need to be fast, as potentially hundreds of different operating scenarios of hundreds of engines have to be simulated.
During aero gas turbine engine transient operation, significant amounts of heat are transferred between the engine structure and the annulus flow [5,6]. These heat flows change the heat balance of the components and the engine. They affect tip clearances, as well as seal clearances, and result in changes of the component characteristics, the secondary air system, and the power delivered by the engine [7,8]. Today’s aero gas turbine engines compensate these detrimental effects by over-fueling, which leads to the well-known temperature overshoots [9], reduced temperature margins, and a reduction in on-wing life.
Early approaches to model the transient heat transfer reduce the involved engine structures to equivalent flat plates of uniform temperature [10]. Further discretization of the substitute structures improved the approximation of the engine’s time response [11,12]. Using more complex substitute structures increases the computational efforts and requires measured or computed data of high spatial and time resolution to derive the local heat transfer parameters [13,14]. State space modeling of heat transfer and tip clearances allows the direct identification of the model parameters either by test or by using calibrated finite element models [15]. Yet, the identified parameters are difficult to interpret and data from a comprehensive set of maneuvers is required for the identification. Riegler [16] proposed modeling the heat transfer from the flow to the engine’s structure using non-dimensional heat transfer maps. Those maps can be implemented as look-up tables into the underlying performance program to facilitate fast calculations while keeping the required accuracy. Furthermore, the maps can be well interpreted as they use a reduced number of non-dimensional variables describing the problem while still reflecting the fundamental physics [17]. However, the maps must be generated using an extensive amount of experimental data or higher-order computations, such as simulations using the finite element method (FEM).
In our previous work [18], the steady state heat transfer map of a pipe was adapted to accurately model the heat transfer of a small turboshaft engine. Hence, less data is needed to accurately predict the thermal response of the engine. Consequently, the same approach is proposed to model the transient heat transfer between the engine’s fluid flow and its structure. To derive the necessary non-dimensional variables, the substitute system is introduced, and a dimensional analysis is performed. Then, the map representing the transient heat transfer of a pipe is generated. This data is used to train a neural network. Using transfer learning, this map can be transferred to match the thermal transients of a small turboshaft engine using measurements from a limited amount of transient maneuvers.

2. Substitute System and Dimensional Analysis

As described in Reference [18], a generalized flow-through system with input or output of shaft power P is used as a base substitute system for a turbomachine. It is characterized by the dimensional parameters shown in Figure 1 and Table 1. The inner heat flow Q ˙ i t and the outer heat flow Q ˙ o t differ in a time-dependent way in the present case of transient heat transfer. As shown in Equation (1), they are related by the rate of change in internal heat energy of the pipe structure,
Q ˙ i Q ˙ o = ρ c p T mat t .
The thermal properties of the substitute system are characterized by the inner and outer heat transfer coefficients h i and h o , the thermal conductivity of the material k mat , the volumetric heat capacity ρ c p , and the volume of the structure V. The input or output of shaft power is reflected by differing values of T min and T max . As in Reference [16], the inner heat flow Q ˙ i ( t ) is divided into a steady state and a transient component,
Q ˙ i t = Q ˙ i , total t Q ˙ i t .
Consequently, the transient component of the inner heat flow can be expressed as a function of the dimensional parameters as:
Q ˙ i = f t , T min , T max , T 0 , δ , A i , A o , h i , h o , k mat , ρ c p , V .
In Equation (3), the surface areas A i and A o are directly related to r i and r o , respectively, as well as the length l . The characteristic thickness δ of an arbitrary structure is the ratio of its volume to its surface area relevant to heat transfer [19].
Using the dimensional matrix shown in Table 1, the non-dimensional inner heat flow becomes a function of the non-dimensional parameters defined in Equations (4)–(8).
Q ˙ i T max T 0 l ref k mat = f Fo , Bi i , Bi o , ϑ char , A o A i , A i δ 2 , V δ 3 ,
Fo = k mat t ρ c p δ 2 ,
Bi i = h i δ k mat ,
Bi o = h o δ k mat ,
ϑ char = T min T 0 T max T 0 .
The non-dimensional temperature difference ratio ϑ char represents the operating point of the machine [18]. Since the Biot numbers can be expressed using the engine’s mass flow [16,18] and surface temperatures [18], they depend on the engine’s operating point, too. Hence, ϑ char , Bi i , Bi o are time-dependent parameters. The parameters A o A i , A i δ 2 , V δ 3 describe the geometry. The reference length l ref is defined as
l ref = A i δ .
The basic form of the resulting non-dimensional transient heat transfer maps is derived assuming a straight pipe as a simple structural element.

3. Simulation

A two-dimensional finite element model is set up to derive the non-dimensional transient heat transfer maps based on the parameters shown in Figure 2. Assuming rotational symmetry, the simulation is carried out using MATLAB’s partial differential equation toolbox. For the simulation, a triangular mesh was generated. It consists of 25 470 elements with six nodes per element. The nodes are located at the corners and edge midpoints. To generate the mesh, the growth rate was set to 1.5, the target minimum element size to 0.05, and the target maximum element size to 0.1.
The non-dimensional flow temperature Θ describes the temperature of the fluid flow at a given position within the pipe. It is formed using the ambient temperature T 0 and the initial maximum temperature T max , ini as references:
Θ = T T 0 T max , ini T 0 .
To verify the simulation setup, the results were compared to the heat transfer of a lumped capacitance, for which an analytical solution exists [19]. To do so, uniform temperature around the system was assumed, and a unit-step was simulated by setting
t = 0 : Θ min = Θ max = Θ 0 = 0 ,
t > 0 : Θ min = Θ max = Θ 0 = 1 .
Furthermore, the Biot numbers Bi i and Bi o are set equally. The material properties and the non-dimensionals describing the geometry are kept constant. The Root Mean Squared Percentage Error of the verification was smaller than 0.1% This indicates a good match between the predicted heat transfer of the FEM simulation and the lumped capacitance.
After verifying the model, it is now used to derive the non-dimensional transient heat transfer maps. The material properties and the non-dimensionals describing the geometry are kept constant for the following steps. The non-dimensional flow temperature is assumed to vary inside the system linearly from Θ min to Θ max . The non-dimensional temperature difference ϑ char is the ratio of the non-dimensional flow temperatures Θ min to Θ max ,
ϑ char = Θ min Θ max = T min T 0 T max , ini T 0 T max T 0 T max , ini T 0 .
The transient response of the system is identified by simulating step changes in Θ max of different sizes Θ step :
t = 0 : Θ max = 1 ,
t > 0 : Θ max = 1 + Θ step = 1 + T max T max , ini T max , ini T 0 .
This approach maps step changes in thermal load of different sizes. The temperature Θ min follows according to Equation (13). With constant ambient temperature T 0 , the outside temperature ratio stays constant at Θ 0 = 0 .
Bi i , Bi o , and ϑ char are kept constant during the step changes. They are varied for the individual step changes such that the operating ranges of the compressor and the turbine of the investigated engine are well covered by the simulation. Consequently, the Biot numbers are chosen to be distributed evenly in the logarithm to base 10 (lg) space,
3.8 lg Bi i 0.8 ,
5.4 lg Bi o 2.4 .
ϑ char is selected from an interval from −0.05 to 0.15 and from 0.65 to 1 to reflect possible values for the investigated compressor and turbine, respectively.
The data is sampled for each simulation at different Fourier numbers. The rationale behind the distribution of the sample points is based on the transient response of a lumped capacitance, which can be expressed as [19]:
Q ˙ = f exp BiFo .
At constant Biot number Bi , the heat flow Q ˙ reduces with increasing Fourier number Fo . As indicated in Figure 3, sampling the transient response at constant intervals of Fourier number Fo would cluster the sampling points at low heat flows. Consequently, the part of the step response featuring significant change in heat flow rates would be underrepresented in the data. Hence, the data points were sampled linearly with respect to 1 exp BiFo .
The non-dimensional transient heat transfer map derived from FEM simulation for a given step change of Θ step = 1.0 is shown in Figure 4 for sets of parameters Bi i , Bi o , and ϑ char . It becomes obvious that the variation of the non-dimensional heat transfer map as a function of these parameters is relevant.
The shapes of the non-dimensional heat transfer maps resulting from different step responses are shown in Figure 5a. A change of sign of the step change results in a different shape of the maps. As shown in Figure 5b, these lines collapse with variation due to the parameters Bi i , Bi o , and ϑ char remaining, if the non-dimensional heat flow is referred to the size of the non-dimensional temperature step. This normalizing step leads to
Q ˙ i Θ step T max T 0 l ref k mat = Q ˙ i T max T max , ini l ref k mat ,
which is in accordance with the approach shown in References [16,20]. Furthermore, the heat transfer map shown in Figure 5b is insensitive to variation in geometry.
The maps obtained from the FEM simulation are used to train a neural network. This is the basis for transient heat transfer modeling of turbomachines using transfer learning.

4. Encoding the Non-Dimensional Maps Using Neural Networks

In a first step a suitable type of neural network is selected. With the thermal transients representing a time series possible choices for the networks are:
  • Feedforward Networks, e.g., Multilayer Perceptron (MLP);
  • Recurrent Neural Networks:
    -
    Shallow Networks, e.g., Non-linear Autoregressive Network with Exogenous Inputs (NARX);
    -
    Deep Networks, e.g., Long Short-Term Memory (LSTM).
Both shallow and deep recurrent neural networks directly model the time dependency in their architecture. However, transfer learning using a recurrent neural network is challenging. Using a feedforward neural network, such as a Multilayer Perceptron, follows the principle of directly modeling the non-dimensional transient heat transfer maps. They cover the step responses of the non-dimensional heat flow
Q ˙ par = Q ˙ i Θ step T max T 0 l ref k mat
at constant Bi i , Bi o , and ϑ char and varying Fourier numbers Fo . Since a Multilayer Perceptron can approximate any function [21], the transient heat transfer map of the pipe can be modeled by an MLP. Therefore, an MLP was chosen with Fo , lg Bi i , lg Bi o , and ϑ char as features, and Q ˙ par as target, for the following steps.
A total of 4096 different step responses were simulated. Sixty percent of the data was used for training, and 20% for test and validation, respectively.
Random sampling of training and test data sets generally guarantees that the trained network can interpolate points on the maps. However, if the trained network is able to predict a whole step response, which it has previously not seen, it can be assumed that the network captures the underlying physics.
Without an additional transformation, the data is clustered towards small values of Q ˙ par , as can be seen in Figure 6a. To approximate a normal distribution of the data, the Box-Cox transform, as implemented in SciPy [22], is used on the training data. It is defined as
Q ˙ par , trans = Q ˙ par 1 λ if   λ 0 , log Q ˙ par if   λ = 0 .
The value of the parameter λ was estimated, using maximum likelihood estimation, to be λ = 0.053 . Additionally, the targets are standardized to zero mean and unit variance to facilitate better training. The histogram of the transformed Q ˙ par is shown in Figure 6b. The features Fo , lg Bi i , lg Bi o , and ϑ char are all normalized to fall in the interval 0 , 1 .
A grid search was carried out to obtain a good architecture. For this purpose the number of hidden layers was varied from one to three. The number of neurons per layer was varied, too. For all architectures, the number of neurons decreases the deeper the layer, resulting in a pyramid-shaped architecture. This is done to compress the information from hidden layer to hidden layer to predict Q ˙ par . All hidden layers feature rectified linear units as activation functions.
Training was performed using PyTorch [23] and fastai v1 [24]. The 1-cycle policy [25,26] was applied to training, where the learning rate is first increased and then decreased during training, while the opposite is done for momentum. The maximum value for the learning rate was set using a learning rate finder as described in Reference [24]. Weight decay was chosen to achieve the highest learning rates [26]. A total number of 120 epochs was sufficient to achieve good results.
The performance of the different architectures was evaluated using the Root Mean Squared Error (RMSE) on the validation data. Table 2 shows results of investigated architectures.
Since all layers in the network are fully connected, the number of trainable parameters can be calculated using the number of inputs n in and the number of outputs n out of each layer by
n param = j = 1 k n in , j · n out , j + j = 1 k n out , j .
In Equation (22), k is the number of layers in the architecture, which includes the output layer. The first sum reflects the trainable weights, and the second sum reflects the trainable bias. The same total number of neurons in the hidden layers can result in a different number of trainable parameters. This becomes obvious by comparing, e.g., architecture No. 2 with architecture No. 14, shown in Table 2. For three-layer architectures the RMSE of the validation data shown in Table 2 falls to a minimum and then increases with increasing number of parameters. This indicates that there is an optimum for the number of neurons per layer. Furthermore, three hidden layers perform better than two hidden layers with a similar amount of trainable parameters. Overall, good performance was achieved using a three-layer MLP with 250, 150, and 50 neurons in the first, second, and third hidden layer, respectively. The quality of the achieved prediction is documented in Figure 7, where four step responses that are all part of the test set are shown. As can be seen, the network approximates the results from the FEM simulation well.

5. Measured Data Used for Transfer Learning

The test rig as described in Reference [18] is used to provide measured data for transfer learning, as well as testing and validation, of the model. It comprises the engine, whose transient thermal response will be modeled using transfer learning. Temperatures are measured using Typ K thermocouples, while Netscanner pressure sensor were used to measure the gas path pressures. The engine’s mass flow m ˙ is measured using an inlet nozzle. Flow around the engines is blocked by the bulkhead. This ensures that only natural convection and thermal radiation have to be considered when calculating Bi o . The transmission system features a dedicated torque sensor. The experimental setup, the instrumentation, and the systematic parts of the sensor uncertainties are shown in Figure 8.
To use the measurements from this rig for transfer learning, a total of 20 different maneuvers were carried out. The maneuvers were designed to cover the operating range of the engine. As mentioned earlier, the Biot numbers are a function of the engine’s mass flow and the temperatures. Additionally, ϑ char is directly related to the temperatures present in the engine (see Equation (8)). Hence, varying acceleration and deceleration maneuvers were carried out along the operating line with low shaft power output P Brk . The minimum rotational speed was at 60 000 min 1 , while the maximum rotational speed was at 94 000 min 1 . This varies the engine’s mass flow and, also, moderately, the temperatures. For this particular engine, the temperatures vary significantly when the engine is loaded. Therefore, the engine was loaded or unloaded transiently at maximum rotational speed during some maneuvers by varying the shaft power output P Brk using the test rig’s eddy current brake. The power level at the end of the loading maneuvers was varied, as well as the the time to load or unload the engine. All of the measured maneuvers resemble step responses. Since the maps obtained from the FEM simulation represent the unit step response of a pipe, data obtained from these maneuvers is well suited for transfer learning.
The time series of parameters measured during a typical acceleration maneuver is shown in Figure 9. The acceleration is started at t = 9 s. At t = 26.5 s, the compressor has reached steady state non-dimensional speed n / R T t 2 and mass flow parameter m ˙ R T t 2 / p t 2 . Until t = 140 s, the compressor exit temperature T t 3 keeps creeping towards its steady state level T t 3 , . This indicates continuing heat transfer, since tip clearance effects are assumed to be negligible for this type of engines [27].
At a given point in time t, the heat flow from the fluid to the compressor structure is computed using
Q ˙ ( t ) = m ˙ t h t 3 t h t 3 t .
The total enthalpies in Equation (23) are calculated using the respective total temperatures and a fluid model typical for today’s gas turbine engine performance programs. The temperatures used to calculate h t were measured directly. The temperature readings are corrected to account for the thermal inertia of the thermocouples [28]. The actual temperature T actual is related to the reading T reading by
T actual = t 63.2 % d T reading d t + T reading .
The time constant t 63.2 % is calculated according to Reference [28] using the values of the time constants given in the data sheet of the thermocouples. Only the part of the maneuvers during thermal stabilization, as indicated in Figure 9, is used to calculate the heat flow.
The characteristic non-dimensional temperature difference ϑ char directly results from the measurements. According to Equation (13), it is zero for the investigated compressor, since the entry temperature into the compressor T t , min T 0 .
Calculation of the Biot numbers follows the same principles as described in Reference [18]. For Bi o of the compressor, only free convection is considered, due to the special setting at the test bed [18].
The two parts of the centrifugal compressor’s casing feature different radii. The casing is, therefore, modeled as an equivalent pipe (see Figure 10). This approach ensures consistency with the assumptions underlying the non-dimensional transient heat transfer maps. This equivalent radius r equ is calculated by requiring that the heat flow due to free convection stays the same for the equivalent pipe,
Q ˙ equ , free = Q ˙ in , free + Q ˙ out , free .
Assuming a constant temperature difference and constant k mat , Equation (25) can be written as
Nu equ A equ l equ = Nu in A 1 l in + Nu out A out l out .
Finally, Equation (27) is derived using the Nusselt correlation for an isothermal horizontal cylinder [29],
r equ 0.25 = r in 0.75 l in + r out 0.75 l out r in l in + r out l out .
The inner and outer Biot numbers both vary in a limited range for the example radial compressor (see Figure 11).
The Biot numbers of the turbine are identified as described in Reference [18]. The heat flow Q ˙ is derived indirectly since the combustor exit temperature T t 4 cannot be measured with adequate accuracy. T t 4 ( t ) is synthesized as a function of time t from the combustor heat balance reflecting the compressor heat transfer,
m ˙ t h t 3 t + m ˙ f u e l t L H V η C C = m ˙ 4 t h t 4 t ,
m ˙ 4 ( t ) = m ˙ ( t ) + m ˙ f u e l ( t ) .
It becomes obvious that T t 4 only depends on the transient performance of the compressor and, hence, is not influenced by the thermal transient of the turbine. Consequently, the heat flow of the turbine can be calculated using
m ˙ 4 h t 4 h t 5 dia Comp = m ˙ 4 h t 4 h t 5 dia + Q ˙ ,
Q ˙ = m ˙ 4 h t 5 , dia . Comp . h t 5 , measured .
The temperature measurements used to calculate h t 5 , measured are again corrected to account for the thermal inertia (see Equation (24)). Both T t 5 , dia . Comp . and T t 5 , measured were filtered to reduce the high amount of noise. Additionally, T t 5 , measured was shifted, so that
T t 5 , dia . Comp . t = T t 5 , measured t .
This is done to account for deviations in the stationary performance model, thereby ensuring that Q ˙ t = 0 . The experimentally-identified data is now used to transfer the non-dimensional heat transfer maps encoded by the MLP from the pipe to the engine.

6. Transfer Learning

First, a constant scaling factor representing the circumference of the equivalent pipe is introduced at the end of the network to consider three-dimensional effects not covered by the simulation. This enabled easier and faster training.
In general, transfer learning is done in three steps. In the first step, additional hidden layers are added to the end of the network. For both the compressor and the turbine, one additional layer consisting of 25 neurons and rectified linear unit activations was sufficient. In the second step, only the additional hidden layers are trained using the experimentally-identified data. In the third step, the rest of the network is unfrozen, and all the layers are trained with a low learning rate. This third step is often referred to as fine tuning.
Since there are only 20 maneuvers, 16 were used for training, while 4 were used as test set. It was made sure that maneuvers from similar operating conditions were grouped together. This prevents data leaking from the training set into the test set and ensures that the network properly generalizes.
Following Equations (28)–(31), the thermal transient of the compressor is required to derive the thermal transient of the turbine. Hence, results achieved for the compressor are discussed first. The measurements did not show large variation in the non-dimensional parameters, as shown for the Biot numbers in Figure 11.
One experimentally-identified step response of the compressor’s non-dimensional transient heat flow is shown in Figure 12a. It features significant noise. This is attributed to the correction of the temperature readings, as described in Equation (24). In Figure 12a, the prediction of the original network, that represents the thermal transient of a pipe, is compared to the prediction of the transferred neural network. The prediction of the original network needs to be scaled to meet the range of the measured data. The scaling factor is displayed in the legend of Figure 12. However, the thermal transient of the compressor is slower due to the large thermal mass of the compressor’s impeller and its bladed diffuser. Hence, it is the step of transfer learning which improves the prediction by matching not only the measured range but also the time constant. This is visible by comparing the rise times of the two different predictions shown in Figure 12a.
The thermal transient of the turbine is identified based on these results. Despite filtering the temperatures, the scatter of the experimentally-identified step response of the turbine shown in Figure 12b is still clearly visible. Again, the introduction of a constant scaling factor ensures the prediction using the thermal transient of a pipe meet the measured range. The time constant is better matched if transfer learning is applied.

7. Application to Gas Turbine Prediction

The neural networks representing the non-dimensional transient heat transfer maps of compressor and turbine were implemented into a state of the art gas turbine performance program to model the investigated shaft power engine. As shown in Equation (19), the transferred maps refer the heat flow to a jump in temperature. Therefore, the total heat flow of a component is modeled by superposing the heat flows resulting from their respective temperature jumps at the different time steps. This follows the same approach used in Reference [16]. The performance model is run to the measured fuel flow, which eliminates potential uncertainties in the modeling of the gas turbine control system. The changes in shaft power are measured on the testbed. They are an input to the performance model. The achieved quality of prediction is demonstrated for two exemplary maneuvers:
-
Deceleration from 94 000 min 1 to 70 000 min 1 ;
-
Acceleration from 80 000 min 1 to 94 000 min 1 , followed by a change in shaft power of 1 s duration to a load of 1000 W .
The comparison of the predicted and measured total temperature at compressor exit T t 3 indicates the quality of the diabatic compressor modeling achieved by the transferred maps. The comparison of the predicted and measured total temperatures at turbine exit T t 5 is used to indicate the quality of diabatic turbine modeling. It further embraces the quality of diabatic compressor modeling and the quality of the assumptions needed for the combustor energy balance documented in Equation (28).
The results obtained for both maneuvers are documented in Figure 13, where the measured total temperatures are compared to the diabatic and adiabatic prediction. Here, adiabatic prediction refers to the performance model without using the neural network to account for the engine’s thermal response. In Figure 13c,d, the end of the acceleration is marked by a vertical line. To analyze the quality of the transient model, the predictions are corrected for deviations in the stationary performance model, so that
T t , prediction t = T t , measured t .
It becomes obvious that adiabatic modeling of the compressor does not lead to a time series of T t 3 , which matches the time constant and the temperature level of the measurements. Diabatic compressor modeling using the non-dimensional transient heat transfer maps significantly improves the prediction of time constant and temperature level.
Consequently, the discussion regarding the turbine exit temperature T t 5 is based only on diabatic compressor modeling. The measured time series of turbine exit temperature T t 5 is matched well by the enhanced gas turbine model. The deviations visible can be explained with higher measurement uncertainty in T t 5 and more noise present in the signal. This makes transfer learning more difficult, which is also evident in Figure 12b.
Figure 13 shows that the thermal transient of the turbine is faster than the thermal transient of the compressor. This is in accordance with the larger thermal mass of the centrifugal compressor.

8. Generalization of the Method

Generally, the proposed method is directly usable for large gas turbine engines. It relies on typical measurement positions in such turbines. Different methods to derive the combustor exit temperature T t 4 , such as combustor heat balance or turbine capacity method, are directly applicable.
However, the ranges for Biot number and ϑ char are quite different for larger engines [18]. This can be accounted for by expanding the ranges of the parameters in the two-dimensional FEM simulation. Additionally, tip clearance changes may play an important role in larger aero engines [9]. Hence, so-called cold stabilization maneuvers are not suited to support the required transfer learning. Hot reslam maneuvers featuring high heat flows are suggested for transfer learning. Different start and end conditions of the hot reslams should be chosen to increase the variability in the data. Furthermore, slam accelerations might be used, if the effects of heat transfer is significantly bigger than the effect of tip clearance changes.
The most time-consuming part of the proposed approach is the FEM simulation and the training of the original network encoding the results of the FEM. Transfer learning takes only a small amount of time. If the parameter range of the FEM simulation is chosen to be wide enough, the first two steps have to be carried out only once. The results of these steps could then be used to transfer the thermal behavior of an entire fleet.
Due to the dimensional analysis, this method is applicable to any system that can be be based on a system according to Figure 1 and Table 1. Additionally, the system and the dimensional analysis can be expanded to explicitly model further effects, such as cooling flows, by including the relevant parameters. The resulting non-dimensionals from the expanded dimensional analysis need to be included into the FEM simulation and used as features of the neural network used to encode the simulation results.

9. Conclusions and Outlook

Modeling transient turbomachinery heat transfer using non-dimensional heat transfer maps and neural networks is successfully demonstrated using a very small gas turbine as an example. Such a small gas turbine represents a worst case scenario, since the heat transfer within the components is superimposed by significant heat conduction between the components. Scaling of the non-dimensional transient heat transfer maps and transfer learning are required to match the engine thermal transients. The approach to use transfer learning has the advantage that only a limited amount of data is needed to obtain sufficient accuracy.
In a next step, the transferred networks can be used to regenerate the transient heat transfer maps. Those maps can then be directly implemented into the performance program as lookup tables. This should lead to very fast calculation times of the thermal transients. In addition, the results of the transfer learning and, therefore, the thermal transients can be interpreted by looking at the regenerated maps.

Author Contributions

Conceptualization, M.B. and St.S.; methodology, M.B., St.S. and C.K.; software, M.B.; validation, M.B., St.S. and C.K.; formal analysis, M.B.; investigation, M.B.; data curation, M.B.; writing—original draft preparation, M.B. and St.S.; writing—review and editing, M.B. and St.S.; visualization, M.B. and St.S.; supervision, St.S. and C.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting the findings of this paper are available upon request.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

The following nomenclature are used in this manuscript:
Symbols
ASurface area
Bi Biot number
c p Specific heat at constant pressure
dDiameter
Fo Fourier number
hHeat transfer coefficient
h t Specific total enthalpy
klTotal number of layers in neural network (hidden layers + output layer)
k fl Thermal conductivity of fluid
k mat Thermal conductivity of solid material
lgLogarithm to base 10
logNatural logarithm
lLength
l ref Reference length of non-dimensional heat flow l ref = δ / A i
L H V Lower heating value
m ˙ Mass flow rate
M ax Axial Mach number
nRotational speed
n in , j Number of inputs of j-th layer
n out , j Number of outputs of j-th layer
n param Number of trainable parameters
N u Nusselt number
PPower
pStatic pressure
p t Total pressure
Q ˙ Heat flow rate
Q ˙ par Nondimensional heat flow rate
RSpecific gas constant
rRadius
TTemperature
tTime
t 63.2 % Time constant
T t Total Temperature
U 95 Expanded uncertainty
VVolume
V ˙ fuel Volumetric fuel flow rate
Δ p Differential pressure of inlet nozzle
δ Characteristic thickness
η Efficiency
λ Fitting parameter of Box-Cox transform
ρ Density
τ Torque
Θ Nondimensional temperature
ϑ char Characteristic nondimensional temperature difference
Subscripts
Comp Compressor
dia Diabatic
equ Equivalent
free Free convection
ini Initial
i Inner
mat Material
max Maximum
min Minimum
o Outer
surf Surface
trans Transformed
0Ambient
2Compressor inlet
3Compressor exit
4Turbine inlet
5Turbine exit
Acronyms
AccelAcceleration
BrkBrake
FEMFinite Element Method
LSTMLong Short-Term Memory
MLPMultilayer Perceptron
NARXNonlinear Autoregressive Network with Exogenous Inputs
RMSERoot Mean Squared Error
RNNRecurrent Neural Network

References

  1. Grieves, M.; Vickers, J. Digital Twin: Mitigating Unpredictable, Undesirable Emergent Behavior in Complex Systems. In Transdisciplinary Perspectives on Complex Systems; Kahlen, F.J., Flumerfelt, S., Alves, A., Eds.; Springer: Cham, Switzerland, 2017; pp. 85–113. [Google Scholar]
  2. AIAA Digital Engineering Integration Committee. Digital Twin: Definition & Value—An AIAA and AIA Position Paper; American Institute of Aeronautics and Astronautics (AIAA): Reston, VA, USA, 2020. [Google Scholar]
  3. Wright, L.; Davidson, S. How to tell the difference between a model and a digital twin. Adv. Model. Simul. Eng. Sci. 2020, 7, 13. [Google Scholar] [CrossRef]
  4. Volponi, A.J. Gas Turbine Engine Health Management: Past, Present, and Future Trends. J. Eng. Gas Turbines Power 2014, 136, 051201. [Google Scholar] [CrossRef]
  5. Bauerfeind, K. A New Method For The Determination of Transient Jet Engine Performance Based on The Nonstationary Characteristics of The Components. In AGARD Conference Proceedings; Number 34 Part 2. AGARD-CP-34, Paper 34; NATO: Neuilly-sur-Seine, France, 1968. [Google Scholar]
  6. Thomson, B. Basic Transient Effects of Aero Gas Turbines. In AGARD Conference Proceedings; AGARD-CP-151, Paper 2; NATO: Neuilly-sur-Seine, France, 1974; Number 151; pp. 2.1–2.16. [Google Scholar]
  7. Maccallum, N.R.L. Thermal Influences in Gas Turbine Transients: Effects of Changes in Compressor Characteristics. In Proceedings of the ASME 1979 International Gas Turbine Conference and Exhibit and Solar Energy Conference, San Diego, CA, USA, 12–15 March 1979; Volume 1B. [Google Scholar] [CrossRef] [Green Version]
  8. MacCallum, N.R.L. Further Studies of the Influence of Thermal Effects on the Predicted Acceleration of Gas Turbines. In Proceedings of the ASME 1981 International Gas Turbine Conference and Products Show, Houston, TX, USA, 9–12 March 1981. Volume 1: Aircraft Engine; Marine; Turbomachinery; Microturbines and Small Turbomachinery. [Google Scholar] [CrossRef] [Green Version]
  9. Nielsen, A.E.; Moll, C.W.; Staudacher, S. Modeling and Validation of the Thermal Effects on Gas Turbine Transients. ASME J. Eng. Gas Turbines Power 2005, 127, 564–572. [Google Scholar] [CrossRef]
  10. Fawke, A.J.; Saravanamutto, H.I.H. Digital Computer Methods for Prediction of Gas Turbine Dynamic Response. SAE Trans. 1971, 80, 1805–1813. [Google Scholar]
  11. Khalid, S.J.; Hearne, R.E. Enhancing Dynamic Model Fidelity for Improved Prediction of Turbofan Engine Transient Performance. In Proceedings of the 16th Joint Propulsion Conference, Hartford, CT, USA, 30 June–2 July 1980. [Google Scholar] [CrossRef]
  12. Pilidis, P.; Maccallum, N.R.L. A Study of the Prediction of Tip and Seal Clearances and Their Effects in Gas Turbine Transients. In Proceedings of the ASME 1984 International Gas Turbine Conference and Exhibit, Amsterdam, The Netherlands, 4–7 June 1984; Volume 1. [Google Scholar] [CrossRef] [Green Version]
  13. Schmidt, K.J. Experimentelle und Theoretische Untersuchungen zum Instationären Betriebsverhalten von Gasturbinentriebwerken. Ph.D. Thesis, Universität der Bundeswehr, Hamburg, Germany, 1992. [Google Scholar]
  14. Fiola, R. Berechnung des Instationären Betriebsverhaltens von Gasturbinen unter Besonderer Berücksichtigung von Sekundäreffekten. Ph.D. Thesis, Technische Universität München, München, Germany, 1993. [Google Scholar]
  15. Moll, C.W.; Nielsen, A.E.; Staudacher, S. Derivation and Validation of State Space Models to Predict Heat Transfer and Clearance Changes on Gas Turbine Transients. In Proceedings of the 9th CEAS European Propulsion Forum, Rome, Italy, 15–17 October 2003. [Google Scholar]
  16. Riegler, C. Modulares Leistungsberechnungsverfahren für Turboflugtriebwerke mit Kennfelddarstellung für Wärmeübertragungsvorgänge. Ph.D. Thesis, Universität Stuttgart, Düsseldorf, Germany, 1997. [Google Scholar]
  17. Simon, V.; Weigand, B.; Gomaa, H. Dimensional Analysis for Engineers; Mathematical Engineering; Springer: Cham, Switzerland, 2017. [Google Scholar] [CrossRef]
  18. Baumann, M.; Koch, C.; Staudacher, S. Experimental Identification of Steady-State Turbomachinery Heat Transfer Using Nondimensional Groups. J. Heat Transf. 2020, 142, 061806. [Google Scholar] [CrossRef]
  19. Bergman, T.L.; Lavine, A.S.; Incropera, F.P.; DeWitt, D.P. Fundamentals of Heat and Mass Transfer, 7th ed.; International Student Version; John Wiley & Sons: Hoboken, NJ, USA, 2011. [Google Scholar]
  20. Riegler, C. Correlations to include heat transfer in gas turbine performance calculations. Aerosp. Sci. Technol. 1999, 3, 281–292. [Google Scholar] [CrossRef]
  21. Hornik, K.; Stinchcombe, M.; White, H. Multilayer feedforward networks are universal approximators. Neural Netw. 1989, 2, 359–366. [Google Scholar] [CrossRef]
  22. Virtanen, P.; Gommers, R.; Oliphant, T.E.; Haberland, M.; Reddy, T.; Cournapeau, D.; Burovski, E.; Peterson, P.; Weckesser, W.; Bright, J.; et al. SciPy 1.0: Fundamental algorithms for scientific computing in Python. Nat. Methods 2020, 17, 261–272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Paszke, A.; Gross, S.; Massa, F.; Lerer, A.; Bradbury, J.; Chanan, G.; Killeen, T.; Lin, Z.; Gimelshein, N.; Antiga, L.; et al. PyTorch: An Imperative Style, High-Performance Deep Learning Library. In Advances in Neural Information Processing Systems; Wallach, H., Larochelle, H., Beygelzimer, A., d’Alché-Buc, F., Fox, E., Garnett, R., Eds.; Curran Associates, Inc.: Red Hook, NY, USA, 2019; Volume 32. [Google Scholar]
  24. Howard, J.; Gugger, S. Fastai: A Layered API for Deep Learning. Information 2020, 11, 108. [Google Scholar] [CrossRef] [Green Version]
  25. Smith, L.N. Cyclical Learning Rates for Training Neural Networks. In Proceedings of the 2017 IEEE Winter Conference on Applications of Computer Vision (WACV), Santa Rosa, CA, USA, 24–31 March 2017; pp. 464–472. [Google Scholar] [CrossRef] [Green Version]
  26. Smith, L.N. A disciplined approach to neural network hyper-parameters: Part 1—Learning rate, batch size, momentum, and weight decay. arXiv 2018, arXiv:1803.09820. [Google Scholar]
  27. Schilling, F.U. Untersuchungen zum Betriebsverhalten von Mikrogasturbinen. Ph.D. Thesis, Universität Stuttgart, Stuttgart, Germany, 2009. [Google Scholar]
  28. Bernhard, F. Handbuch der Technischen Temperaturmessung, 2nd ed.; VDI-Buch; Springer: Berlin/Heidelberg, Germany, 2014. [Google Scholar] [CrossRef]
  29. Morgan, V.T. The Overall Convective Heat Transfer from Smooth Circular Cylinders. In Advances in Heat Transfer; Irvine, T.F., Hartnett, J.P., Eds.; Elsevier: Amsterdam, The Netherlands, 1975; Volume 11, pp. 199–264. [Google Scholar] [CrossRef]
Figure 1. Model of generalized turbomachine.
Figure 1. Model of generalized turbomachine.
Aerospace 09 00049 g001
Figure 2. Non-dimensional parameters used for FEM simulation of the diabatic pipe flow.
Figure 2. Non-dimensional parameters used for FEM simulation of the diabatic pipe flow.
Aerospace 09 00049 g002
Figure 3. Effect of linearly spacing Fo .
Figure 3. Effect of linearly spacing Fo .
Aerospace 09 00049 g003
Figure 4. Non-dimensional transient heat transfer map for a given step change of Θ step = 1 .
Figure 4. Non-dimensional transient heat transfer map for a given step change of Θ step = 1 .
Aerospace 09 00049 g004
Figure 5. (a) Non-dimensional heat transfer map for varying step changes; (b) non-dimensional heat transfer map after referring heat flow to step size.
Figure 5. (a) Non-dimensional heat transfer map for varying step changes; (b) non-dimensional heat transfer map after referring heat flow to step size.
Aerospace 09 00049 g005
Figure 6. Histogram of targets in training set (a) before scaling and (b) after scaling.
Figure 6. Histogram of targets in training set (a) before scaling and (b) after scaling.
Aerospace 09 00049 g006
Figure 7. Quality of transient heat transfer map prediction using Multilayer Perceptron.
Figure 7. Quality of transient heat transfer map prediction using Multilayer Perceptron.
Aerospace 09 00049 g007
Figure 8. Experimental setup, instrumentation, and systematic part of sensor uncertainties.
Figure 8. Experimental setup, instrumentation, and systematic part of sensor uncertainties.
Aerospace 09 00049 g008
Figure 9. Compressor data for a typical acceleration maneuver.
Figure 9. Compressor data for a typical acceleration maneuver.
Aerospace 09 00049 g009
Figure 10. Modeling the compressor casing as an equivalent pipe.
Figure 10. Modeling the compressor casing as an equivalent pipe.
Aerospace 09 00049 g010
Figure 11. Scatter plot of experimentally-identified Biot numbers in the compressor.
Figure 11. Scatter plot of experimentally-identified Biot numbers in the compressor.
Aerospace 09 00049 g011
Figure 12. Comparison of experimentally-identified non-dimensional transient heat flow to scaled non-transferred network, that encodes the thermal transient of a pipe and to transfer-learned network for (a) compressor and (b) turbine, experimentally-identified step response filtered.
Figure 12. Comparison of experimentally-identified non-dimensional transient heat flow to scaled non-transferred network, that encodes the thermal transient of a pipe and to transfer-learned network for (a) compressor and (b) turbine, experimentally-identified step response filtered.
Aerospace 09 00049 g012
Figure 13. Comparison of measurement, adiabatic prediction (performance model without the neural network to model heat transfer), and diabatic prediction (same performance model with the neural network). Predictions are corrected for deviations in the stationary performance model; see Equation (33). (a) Compressor exit temperature, deceleration, (b) turbine exit temperature, deceleration, (c) compressor exit temperature, acceleration (accel), and loading, and (d) turbine exit temperature, acceleration (accel), and loading.
Figure 13. Comparison of measurement, adiabatic prediction (performance model without the neural network to model heat transfer), and diabatic prediction (same performance model with the neural network). Predictions are corrected for deviations in the stationary performance model; see Equation (33). (a) Compressor exit temperature, deceleration, (b) turbine exit temperature, deceleration, (c) compressor exit temperature, acceleration (accel), and loading, and (d) turbine exit temperature, acceleration (accel), and loading.
Aerospace 09 00049 g013
Table 1. Dimensional matrix used to derive the non-dimensional heat flow.
Table 1. Dimensional matrix used to derive the non-dimensional heat flow.
Q ˙ i T min T 0 T max T 0 δ A i A o h i h o k mat ρ c p Vt
L200122001−130
T−300000−3−3−3−201
M100000111100
Θ011000−1−1−1−100
Table 2. Root Mean Squared Error (RMSE) of different architectures. The architecture with the lowest RMSE is marked in bold font.
Table 2. Root Mean Squared Error (RMSE) of different architectures. The architecture with the lowest RMSE is marked in bold font.
Architecture No.Number of Hidden LayersNumber of Neurons in the Hidden LayersNumber of ParametersRMSE
1125015010.0527
2145027010.0627
32300, 10031 7010.0133
42250, 20051 6510.0112
52400, 20082 4010.0143
62500, 250128 0010.0169
7350, 25, 1017960.0212
83100, 50, 2568510.0524
93150, 100, 5020 9510.0136
103200, 100, 5026 2010.0104
113250, 100, 1027 3710.0117
123200, 150, 5038 7510.0126
133250, 150, 1040 4210.0120
143250, 150, 5046 5010.0095
153200, 200, 5051 3010.0106
163250, 200, 5061 5510.0115
173300, 200, 10081 9010.0124
183500, 200, 100122 9010.0120
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Baumann, M.; Koch, C.; Staudacher, S. Application of Neural Networks and Transfer Learning to Turbomachinery Heat Transfer. Aerospace 2022, 9, 49. https://doi.org/10.3390/aerospace9020049

AMA Style

Baumann M, Koch C, Staudacher S. Application of Neural Networks and Transfer Learning to Turbomachinery Heat Transfer. Aerospace. 2022; 9(2):49. https://doi.org/10.3390/aerospace9020049

Chicago/Turabian Style

Baumann, Markus, Christian Koch, and Stephan Staudacher. 2022. "Application of Neural Networks and Transfer Learning to Turbomachinery Heat Transfer" Aerospace 9, no. 2: 49. https://doi.org/10.3390/aerospace9020049

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop