You are on page 1of 16

Computers and Structures 289 (2023) 107163

Contents lists available at ScienceDirect

Computers and Structures


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

Data-informed statistical finite element analysis of rail buckling


Fuzheng Sun a,∗ , Eky Febrianto b,e , Heshan Fernando c , Liam J. Butler d , Fehmi Cirak b ,
Neil A. Hoult a
a
Department of Civil Engineering, Queen’s University, 58 University Ave., Kingston, K7L 3N6, ON, Canada
b
Department of Engineering, University of Cambridge, Trumpington Street, Cambridge, CB2 1PZ, UK
c
Ingenuity Labs Research Institute, Queen’s University, 69 Union St W, Kingston, K7L 3N6, Canada
d
Department of Civil Engineering, York University, 4700 Keele St, Toronto, M3J 1P3, ON, Canada
e
James Watt School of Engineering, University of Glasgow, Glasgow, G12 8QQ, UK

A R T I C L E I N F O A B S T R A C T

Keywords: In this paper, the statistical finite element method is developed further to synthesize distributed rail response
Distributed fiber optic sensing data with nonlinear finite element model predictions within and outside the measured loading range. In the
Statistical finite element method data-generating model of the statistical finite element method, the distributed sensing data is decomposed into a
Polynomial chaos expansion
finite element model component, a model-reality mismatch component, and a noise component. Each component
Bayesian learning
Progressive rail buckling
is considered uncertain and is represented as a Gaussian random vector with a corresponding prior density.
The finite element prior density is updated using the Bayesian statistical framework in light of the distributed
fiber optic sensing data. The calculated posterior density enables one to infer the true structural response. The
required finite element prior density is determined by solving a conventional stochastic forward problem using
a polynomial chaos expansion of random fields and a non-intrusive pseudo-spectral projection approach. In
this paper, a lab test involving a section of a rail (i.e., a beam-column structural member) instrumented with
distributed fiber optic sensors and subjected to axial load causing progressive lateral buckling is considered to
demonstrate the use of the statistical finite element method to improve rail response prediction. The results
show improved prediction of true rail strain in linear and nonlinear regimes compared with a pure finite element
model-based prediction.

1. Introduction under future extreme loading events. This paper investigates the blend-
ing of distributed fibre optic sensors (DFOS) measurement data with
Railways play an essential role in the economies of many countries a conventional finite element (FE) model to obtain improved structural
by facilitating the transport of goods and people. Continuous welded response predictions with sharp confidence bounds. To this end, the sta-
rail (CWR) has gradually replaced the traditional jointed rail as the pre- tistical finite element method (statFEM) recently proposed by Girolami
ferred track type since the early 1950s due to the advantages of higher et al. [1] is used and developed further, and the theoretical derivations
operating speeds, lower maintenance costs, increased passenger com- are validated with a lab-based rail progressive buckling test.
fort, and noise reduction. However, due to the elimination of expansion Traditionally, railway engineers have relied on FE modelling to help
joints in CWR, longitudinal compression forces develop when the track better understand the rail response under longitudinal loads and predict
temperature exceeds the stress-free temperature (also known as the rail the critical buckling loads [2–5]. FE modelling allows the rail buck-
neutral temperature). Once the compression force in the rail exceeds a ling response to be studied without needing expensive and sometimes
critical value, the rail will buckle, and large lateral displacements will impractical experiments. Lim et al. [2] developed a three-dimensional
develop that can alter the gauge of the tracks and potentially lead to FE model for extensive buckling analysis of a CWR track system sub-
a train derailment. To avoid train derailments, rail network operators jected to longitudinal load and considered the geometrical and material
do not only need to monitor the railway infrastructure continuously, nonlinearity in the FE modelling. Pucillo et al. [3] studied the thermal
but they often need to understand how the infrastructure will perform buckling and post-buckling responses of CWR through FE modelling

* Corresponding author.
E-mail addresses: 18fs26@queensu.ca (F. Sun), eky.febrianto@glasgow.ac.uk (E. Febrianto), heshan.fernando@queensu.ca (H. Fernando),
liam.butler@lassonde.yorku.ca (L. Butler), fc286@cam.ac.uk (F. Cirak), neil.hoult@queensu.ca (N. Hoult).

https://doi.org/10.1016/j.compstruc.2023.107163
Received 25 January 2023; Accepted 6 September 2023
Available online 20 September 2023
0045-7949/© 2023 Elsevier Ltd. All rights reserved.
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

and conducted sensitivity analyses to investigate the influence of crit- provides a coherent formalism to determine their respective posterior
ical structural parameters on the track buckling response. Kang et al. densities while using the likelihood of the observations. An assumed
[4] investigated the rail response due to compressive axial force on bal- statistical data-generating model determines the probability or the like-
lastless track systems through both experimental investigation and FE lihood that the model produced the observed data. See relevant books,
modelling. They concluded that the rail behaviour in the lateral direc- e.g. [31,32] for an introduction to Bayesian statistics and data analy-
tion is less stiff than the longitudinal direction under axial load and sis. The statFEM generating model is inspired by the seminal Kennedy
that the rail deforms non-linearly and rapidly once it reaches a particu- and O’Hagan [33] paper on Bayesian calibration, which introduces a
lar load. Miri et al. [5] used FE models to study the effect of the shape data-driven Gaussian process-based framework for blending data from
of concrete sleepers on track buckling load and concluded that winged a black-box simulator and observations. Since its inception, numerous
and frictional sleepers could provide more significant lateral resistance extensions and applications of the Kennedy and O’Hagan framework
to the track system and hence increase the thermal buckling load ca- have been proposed, e.g. [34–38]. A crucial component of statFEM is
pacity. These FE models involve assumptions and simplifications about a random discrepancy or misspecification, a term that takes into ac-
the track system (e.g., the use of beam elements for modelling the rail) count the mismatch between the finite element model and the actual
and estimates of the geometry (e.g., using sine curves for modelling system. In practice, such a mismatch is inevitable because of the many
the initial rail misalignment), material properties (e.g., elastic or elas- engineering assumptions and simplifications necessary in creating a FE
tic and perfectly plastic springs for modelling the lateral resistance), model. The conditioning of the statistical statFEM model on the ob-
and boundary conditions, which inevitably introduce uncertainties and servation data yields a posterior probability density that consistently
inaccuracy into the FE prediction of the rail response due to the fact blends the finite element prior with the observation data. In statFEM,
that in-service operation of the railway system takes place in an ever- the likelihood and the priors are all considered to be Gaussians so that
changing environment. In order to provide a predictive rail manage- the posterior probability has a closed form and is a Gaussian, hence
ment tool, these uncertainties and inaccuracies must be accounted for simplifying its computation.
when analyzing and interpreting the FE modelling results so that the In the original statFEM paper [1], the prior probability density of
potential variation in results can be quantified. the FE solution was obtained by solving a probabilistic FE problem us-
Recent advances in structural health monitoring (SHM) enable the ing a perturbation technique, which is insufficient for structures with
deployment of advanced distributed sensing systems in railways, which highly nonlinear responses. Alternatively, the prior probability density
has introduced new opportunities for more comprehensive monitoring can be obtained using commercially available software (e.g. Abaqus)
of the structural performance [6–12]. The application of DFOS in civil through Monte Carlo (MC) sampling. However, in the case of rail buck-
engineering has overcome some disadvantages associated with tradi- ling, due to the nonlinearity of the FE model, evaluating the uncer-
tional discrete sensors by enabling long-distance, high-resolution, high- tainties in the FE prediction through commercial software is extremely
accuracy, durable, and robust strain measurements. The data collected time-consuming and impractical. In recent years, surrogate models de-
from the sensors provide the actual rail response information under op- veloped using machine learning or deep learning, such as ANN [39,40],
erational loads, which were so far interpreted or used to validate FE Gaussian processes (GP) [41,42], polynomial chaos expansion (PCE)
models for structural assessment in an ad-hoc manner to support op- [43,44], have been increasingly used to approximate complex prob-
eration and maintenance decisions [13–15]. However, this traditional lems in engineering. Those surrogate models can provide sufficiently
approach of updating FE models using SHM data is inefficient when accurate results and have been used to bypass the traditional MC ap-
the amount of structural response data from advanced sensor networks proach for uncertainty quantification. In the present work, a PCE-based
is large, and the modelled structural response is complex. This has re- surrogate model is constructed to approximately model the rail progres-
sulted in the development of data-driven approaches to directly model sive buckling response under axial load. This approach enables faster
structural performance by learning the nonlinear mapping between the uncertainty quantification in the modelled rail response, which signif-
sensor data and the structural response, using machine learning and icantly reduced the computational cost of modelling the geometrical
statistical learning approaches, see the recent reviews [16,17]. Fur- nonlinearity in the rail response. A simplified rail FE model is first es-
thermore, Shen et al. [18], for instance, employed a Gaussian process tablished with beam elements in the commercial software package (i.e.,
regression (GPR) model to directly infer rail pad and ballast stiffness us- Abaqus) to generate samples to train the surrogate model, which will be
ing frequency response data from field hammer tests. Do et al. [19] used further used to obtain the uncertainties in the modelled rail response.
an artificial neural network (ANN) to directly evaluate the track modu- The uncertain rail response is then combined with the DFOS data from
lus using the continuous deflection data from a train-mounted vertical previously performed lab-based rail progressive buckling experiments
track deflection measurement system. The fidelity of purely data-driven under controlled conditions. This process enables improved predictions
approaches usually depends closely on the quality of the training data, of the rail response and buckling assessment under axial load using
and usually, it requires large amounts of information-rich data to train a the data-informed statistical finite element analysis. By introducing the
sophisticated data-driven model. The challenges associated with purely classical Euler-Bernoulli beam theory, statFEM also enables the accu-
data-driven approaches have instigated research to assimilate numer- rate prediction of lateral rail deflection, which is valuable rail response
ical modelling based on physical laws with data-driven approaches to information that railway owners rely on to make decisions on railway
provide improved predictions of the structural response. GPR with phys- operation and maintenance.
ical constraints [20–24] and physics-informed neural networks [25–29]
are two typical approaches to effectively learn from limited data by 2. Strain data acquisition
explicitly taking into account the underlying physical laws governing
the behaviour of structures. The mentioned methods are framed as a Compared to traditional jointed rail, longitudinal compression forces
replacement for prevalent finite element analysis techniques and are can be developed in CWR due to the elimination of joints, which results
currently only suitable for domains with elementary geometries. in rail buckling once this load is higher than the load-carrying capacity
In contrast, the recently proposed statistical construction of the FE of the CWR system. In the current proof of concept work, the force in
method, dubbed statFEM, allows predictions to be made about the ac- the rail was applied using an actuator rather than due to thermal effects.
tual system behaviour in light of measurement data and a conventional However, the buckling mechanisms are similar and DFOS data can be
FE model [1,30]. Adopting a Bayesian viewpoint in statFEM, uncertain- used to capture the mechanism in both cases. A 3.048 m ASCE 12 lb/yd
ties in the data and model are treated as random variables with suitably rail was used in this work in order to limit the buckling load of the
chosen prior probability densities, which consolidate any knowledge rail to an achievable level based on the available space and actuator
available. Starting from the prior probability densities, the Bayes rule capacity. Two different distributed fiber optic measurement systems—a

2
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 1. The setup for the pinned-pinned rail loading experiment. (1) Hydraulic Jack; (2) Load cell; (3) Pin supports; (4) Rail; (5) Linear potentiometers.

Table 1 Table 2
Geometric and material properties of ASCE 12 lb/yd rail. Properties of the two FOS measurement systems used in the rail loading exper-
iments.
Property Value
Property BOTDA Rayleigh
Elastic modulus 207 GPa
Cross-sectional area 761 mm2 Sensor resolution 50 mm 2.6 mm
Minimum second moment of area 53000 mm4 Measurement accuracy (1.96*𝜎𝑒 ) ± 15 𝜇𝜀 ± 3 𝜇𝜀
Maximum sensing distance 100 km 50 m

long-range Brillouin-based system and a short-range, but more accurate,


Rayleigh-based system—were used to measure the strains during the data during the loading experiments. One system was the Neubrex NBX-
loading test. The following subsections describe the experimental setup, 6050A BOTDA analyzer, which is referred to as the “BOTDA system” in
properties of the two analyzers, and the strain data set. Further details this paper; and the second system was a LUNA ODiSI 6104 Rayleigh
of the experiment can be found in Sun et al. [12]. backscatter analyzer, which is referred to as the “Rayleigh system” in
this paper. Strain measurement properties for both analyzers are pro-
2.1. Experimental test setup vided in Table 2. The BOTDA measurements have a larger gauge length
and lower measurement precision compared to the Rayleigh system.
The laboratory setup for the rail loading experiments is shown in However, the BOTDA system is capable of measuring up to 100 km,
Fig. 1. The geometric and material properties of the ASCE 12 lb/yd which is more practical for real rail implementations.
rail are provided in Table 1. The rail was held in a self-reacting frame,
which consisted of two steel plates connected by four threaded rods. 2.3. Experiments and data sets
The axial load was applied using an Enerpac hydraulic jack that was
aligned with the instrumented rail along its neutral axis. A Strainsert Two loading experiments were conducted with the instrumented rail
universal flat load cell with a 222 kN loading capacity was connected setup: one test was conducted with the BOTDA measurement system
to the hydraulic jack to measure the load applied to the rail. and the second test was conducted with the Rayleigh system. In both
Two rollers were placed at either end of the rail specimen, and the experiments, the compressive load was applied using load steps of 1 kN
axial load was applied to the rail through the rollers enabling rotation until the rail buckled or the maximum strain reached 1000 microstrain
about the weak axis of bending but restraining rotation about the strong (𝜇𝜀) to avoid plastically deforming the rail. In this range, the non-linear
axis. In all tests, the rail was also supported against strong axis buck- response of the rail could be measured while keeping the rail within its
ling by two rollers at 1 m and 2 m along the length. A small amount elastic stress range to avoid causing irreversible damage. It should be
of axial compressive load (approximately 0.4 kN) was applied to the noted that both tests were completed in a closed indoor environment
rail before the start of the test to hold the rail in place. A linear po- within one hour, such that environmental changes during the test were
tentiometer (LP) was installed at the rail’s mid-length to measure the neglectable. The rail is considered to be buckled when there is a sudden
lateral deflection. The axial load and LP measurements were recorded increase in the lateral deflection accompanied by a drop in the axial
using a System 7000 StrainSmart data acquisition system at a rate of load (explosive buckling). In a less stiff railway system, the rail is also
1 Hz. considered to be buckled when the axial loads reach a plateau while
A 35 m long nylon-coated fiber optic sensing cable was installed the lateral deflections increase significantly (progressive buckling) [45].
along eight different paths on the surface of the rail as shown in Fig. 2b. Measurements from the load cell and LP corresponding to every DFOS
To ensure enough data was obtained for use with the statFEM, the fiber sampling interval were also recorded in each test.
was installed at 8 separate locations on the rail cross-section (F1 to Fig. 3 presents the axial load versus the lateral displacement at the
F8, as shown in Fig. 2) and monitored with one of either the Brillouin mid-length of the rail as measured with the linear potentiometers for
Optical Time Domain Analysis (BOTDA) or Rayleigh system (it is not the two experiments. For the Rayleigh experiment, the nonlinear be-
possible to attach both systems to the same fiber simultaneously), which haviour begins at an axial load, 𝜆, of approximately 6 kN, whereas
enables the strain plane at any point along the length of the rail to be for the BOTDA experiment, the nonlinear behaviour begins at an axial
fitted (see Sun et al. [12] for details). As a result of this and the fact that load, 𝜆, of approximately 10 kN. This difference in buckling response is
no two tests had the same geometric properties, as will be discussed in considered to be the result of different load eccentricities and initial ge-
greater detail later, it was not possible to compare the results of the two ometrical imperfections in each loading experiment, as the rail needed
tests directly. to be reinstalled and realigned before each experiment [12]. Since the
geometry changed for every test, it was not possible to conduct repli-
2.2. Measurement systems cate experiments, and thus only one test for each DFOS system (i.e., the
BOTDA and Rayleigh system) was used in the development of the cur-
In this study, two DFOS measurement systems with different mea- rent model. As seen in Fig. 3, the rail buckled at around 12 kN in the
surement precision and sensor gauge lengths were used to acquire strain BOTDA and 11.2 kN in the Rayleigh tests, respectively. In a real railway

3
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 2. Schematic fiber location a) on cross-section, b) along the rail length.

𝒚(𝜆) = 𝒛(𝜆) + 𝒆 . (1)


The Gaussian measurement noise 𝒆 ∼ 𝑝(𝒆) =  (𝟎, 𝜎𝑒2 𝑰) is independent
of the loading and the sensor locations. Furthermore, the parameter 𝜎𝑒
is the standard deviation suggested by the manufacturer of the DFOS
sensing system and 𝑰 is the identity matrix.
The unknown true strains 𝒛(𝜆) at the 𝑛𝑦 sensor positions are assumed
to be given by

𝒛(𝜆) = 𝑷 𝒖(𝜆) + 𝒅(𝜆) , (2)


where the matrix 𝑷 projects the random FE solution 𝒖(𝜆) at the
mesh nodes to the FE strains at locations corresponding to the mea-
sured strains 𝒚(𝜆). The random model-reality mismatch vector 𝒅(𝜆)
Fig. 3. Axial load vs mid-length lateral displacement in compressive buckling takes into account the inadequacy of the FE model in reproducing
tests. the true structural response and is assumed to have the zero-mean
Gaussian probability distribution 𝑝(𝒅(𝜆)) =  (𝟎, 𝑪 𝑑 (𝜆, 𝜆)). As will be
specified in Section 3.3, the covariance matrix 𝑪 𝑑 (𝜆, 𝜆) is obtained
system, the buckling load is expected to be higher due to the size of the
from a kernel function. The probability distribution of the FE solu-
rail and lateral resistance provided by the sleepers and ballast. Addi-
tion 𝒖(𝜆) is approximated with the Gaussian probability distribution
tionally, in the field, the buckling load is created by restrained thermal
𝑝(𝒖(𝜆)) =  (𝒖(𝜆), 𝑪 𝑢 (𝜆, 𝜆)) as discussed next.
expansion. However, the fundamental basis of the sensor system (DFOS
strain measurements along the rail) and the structural behaviour (lat-
3.2. Probability distribution of the finite element solution
eral weak-axis buckling) are the same in both this study and in the field.
Thus, the experimental results enable the goal of this research (i.e. to
The FE probability distribution 𝑝(𝒖(𝜆)) is one of the key components
perform a proof-of-concept evaluation of the application of statFEM to a
in statFEM and represents the uncertainty in the FE solution due to
non-linear structural system monitored with DFOS) to be achieved. The
the known uncertainties in model parameters 𝜽 ∼ 𝑝(𝜽), like load eccen-
restraint provided by the ballast and the thermal loading effects will be
tricities or material properties. The probability distribution 𝑝(𝒖(𝜆)) is
considered in a future study.
obtained by propagating 𝑝(𝜽) through the FE discretized equilibrium
Although the data at different fiber locations were collected, the
equations
data used for the statFEM analysis is the strain measured at the neutral
axis for weak axis bending (F3 in Fig. 2a). The strain at this location
𝑮 (𝒖, 𝜽) + 𝜆𝑭 = 𝟎 , (3)
is purely due to axial compression and weak axis bending. This strain
location was chosen to avoid the need to remove the effects of strong where 𝑮 (𝒖, 𝜽) is the vector of the internal forces, 𝑭 is the reference
axis bending from the measurements. external force vector and 𝜆 is the scalar axial load parameter. In this
paper, the FE solution 𝒖(𝜆∗ ) at a fixed 𝜆∗ is obtained by modelling the
3. Review of the statistical finite element method rail as a geometrically nonlinear beam and solving the respective equa-
tions with the commercial FE package Abaqus [46,47]. To obtain the FE
In this section, the statFEM is reviewed and its application to rail probability distribution 𝑝(𝒖(𝜆∗ )) often the MC method is employed with
progressive buckling is discussed. To begin with, a statistical model de- samples drawn from the input probability distribution 𝑝(𝜽). The MC
method usually requires a large number of samples to obtain a good es-
scribing the relationship between the experimental DFOS strain data
timate for 𝑝(𝒖(𝜆∗ )). Consequently, the MC approach is computationally
and FE data is proposed. Next, the computation of the prior probabil-
intractable for most engineering applications requiring large FE models
ity distribution of the FE data using a PCE is considered. Finally, the
or nonlinear problems requiring iterative solution techniques.
synthesis of the DFOS strain data and the FE model using Bayes rule is
Therefore, a PCE is used to approximate the uncertainties in the FE
described. The possibility to predict the structural response for loadings
solution 𝒖(𝜆∗ ) at a fixed axial load parameter 𝜆∗ resulting from the un-
where only FE prediction but no measurements are available is high-
certainties in the input parameters 𝜽. In PCE a set of orthogonal Hermite
lighted. { }𝑛𝑃
polynomials 𝚿𝑗 𝑗=0 is considered for approximating

3.1. Statistical model 𝑛𝑃



𝒖(𝜆∗ , 𝜽) ≈ 𝜶 𝑗 (𝜆∗ ) 𝚿𝑗 (𝜽) , (4)
𝑗=0
As introduced in Section 2, the strain distribution along the rail spec-
imen is determined by measuring the strains at a fixed set of sensor where 𝑛𝑃 is equal to the degree of the orthogonal polynomials. The
{ }𝑛𝑃
locations and axial loads using the DFOS sensing system. The observed unknown expansion coefficients 𝜶 𝑗 𝑗=0 are determined following a
strain vector 𝒚(𝜆) collects the strains at the 𝑛𝑦 sensor locations and de- pseudo-spectral projection approach [47,48]. Specifically, at first, a set
{ }𝑛PCE
pends on the (scalar) axial load parameter 𝜆. The observed strains are of a few quadrature points 𝜽𝑘 𝑘=0 are sampled from the distribution
assumed to be additively composed of the true strains 𝒛(𝜆) and the mea- 𝑝(𝜽), and subsequently for each 𝜽𝑘 the respective finite element solution
surement noise 𝒆, i.e., 𝒖(𝜆∗ , 𝜽𝑘 ) is computed. It is worth emphasizing that the total number of

4
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

samples is in the order of 𝑛PCE ≈ 10, compared with tens of thousands of Moreover, the inferred true strain probability distribution is given by
samples that are required in standard MC. As mentioned, each FE solu- the conditional probability distribution 𝑝(𝒛𝜆∗ |) =  (𝒛| (𝜆∗ ), 𝑪 𝑧| (𝜆∗ ,
tion 𝒖(𝜆∗ , 𝜽𝑘 ) was determined with the commercial FE package Abaqus. 𝜆∗ ). The conditioned mean 𝒛| (𝜆∗ ) and covariance 𝑪 𝑧| (𝜆∗ , 𝜆∗ ) are ob-
Subsequently, the expansion coefficients {𝜶𝑗 (𝜆∗ )} are determined with tained from Equation (6) with the aid of standard relations for Gaussian
the open-source library Chaospy [49]. probability densities, see e.g. [51].
Once the polynomial chaos expansion (4) is established, a large set In summary, statFEM can be used to synthesize the FE prediction
of samples 𝜽𝑘 ∼ 𝑝(𝜽) is sampled to determine the approximate Gaus- and the available monitoring data to infer the true structural response
sian probability distribution 𝑝(𝒖(𝜆∗ )) =  (𝒖(𝜆∗ ), 𝑪 𝑢 (𝜆∗ , 𝜆∗ )) by comput- of an instrumented structure. The overall procedure for implementing
ing the empirical mean 𝒖(𝜆∗ ) and covariance 𝑪 𝑢 (𝜆∗ , 𝜆∗ ) of the samples. statFEM is outlined in Table 3. For non-Gaussian prior distributions, the
In the present study, the probability distribution 𝑝(𝜽) is a Gaussian and posterior can be, for instance, obtained by Laplace approximation [52,
can be easily sampled. 53], variational Bayes [54–56] or Markov chain Monte Carlo (MCMC)
sampling [57,58].
3.3. Bayesian inference
4. Results and discussion

By virtue of the postulated statistical model in Equations (1) and (2), In this study, the prediction of true strains for a section of rail sub-
the Gaussian approximate FE probability distribution 𝑝(𝒖(𝜆∗ ) and the jected to increasing axial loads using uncertain FE predictions and the
Gaussian model-mismatch distribution 𝑝(𝒅(𝜆∗ )), the true strain proba- DFOS strain measurements from either the BOTDA system (longer mea-
bility distribution at a fixed 𝜆∗ reads surement length) or the Rayleigh system (higher accuracy and sensor
( ) resolution) is presented. In future field applications, only the measure-
𝑝(𝒛(𝜆∗ )) =  𝒛(𝜆∗ ), 𝑪 𝑧 (𝜆∗ , 𝜆∗ ) ments from one DFOS system will be considered for the analysis. The
( ) analysis involves the distributed rail surface strain measurement from
=  𝑷 𝒖(𝜆∗ ), 𝑷 𝑪 𝑢 (𝜆∗ , 𝜆∗ )𝑷 𝖳 + 𝑪 𝑑 (𝜆∗ , 𝜆∗ ) . (5)
fiber F3 (see Fig. 2) across multiple discrete axial loads and FE mod-
Note that this represents a prior probability distribution and is indepen- elling of the physical test, see Section 2. All the analyses conducted in
dent of measurement data. this paper are carried out using a computer with 4 cores, 8 GB Random-
{ }𝑛𝜆
The measured strain and axial load parameter pairs  = 𝒚 𝑖 , 𝜆𝑖 𝑖=1 access memory (RAM), and an i5-1035G1 CPU.
are only known for 𝑛𝜆 distinct axial load values. The probability dis-
tribution of the true strain 𝒛(𝜆∗ ) at an axial load value 𝜆∗ can be 4.1. Forward finite element prediction
determined using the Bayes formula. However, because all the involved
densities are Gaussians, it is more expedient to consider instead the fol- In this section, the prior FE strain distribution 𝑝(𝑷 𝒖) at the location
lowing joint probability distribution of F3 (see Fig. 2) due to randomized input parameters is presented. As
( ) ( ) mentioned in Section 3.2, Abaqus is used as the forward FE solver. Fig. 4
⎛ 𝒛 𝜆∗ ⎞ ⎛ ⎛ 𝑷 𝒖(𝜆∗ ) ⎞ presents the geometrical model used in the simulations. The geometry
⎜ 𝒚 ⎟ ⎜ ⎜ 𝑷 𝒖 𝜆1 ⎟
⎜ 1 ⎟∼ ⎜ ⎜ ⎟, is discretized into 10 three-node quadratic beam elements (Abaqus B22
⎜ ⋮ ⎟ ⎜ ⎜ (⋮ ) ⎟ elements) with 21 nodes and 63 degrees of freedom. The large displace-
⎜ 𝒚 ⎟ ⎜ ⎜𝑷𝒖 𝜆 ⎟ ment effect is considered in the analysis in order to include geometrical
⎝ 𝑛𝜆 ⎠ ⎝⎝ 𝑛𝜆 ⎠
( ) ( ) ( ) nonlinearity under increasing axial loads. Similar to the experimental
⎛ 𝑪 𝑧 (𝜆∗ , 𝜆∗ ) (𝑪 𝑧 𝜆∗ ,)𝜆1 … 𝑪 𝑧 (𝜆∗ , 𝜆𝑛𝜆 ) ⎞⎞
⎜ 𝑪 𝑧 𝜆1 , 𝜆∗ ⎟⎟ specimen, the left end of the beam (Point A) is pinned (displacement
𝑪 𝑧 𝜆1 , 𝜆1 + 𝜎𝑒2 𝑰 … 𝑪 𝑧 𝜆1 , 𝜆𝑛𝜆
⎜ ⋮ ⎟⎟ , constraint in x and y directions with free rotation), while the right end
⎜ ( ) ( ⋮ ) ⋱ ( ⋮ )
⎟⎟
⎝ 𝑪 𝑧 𝜆𝑛𝜆 , 𝜆∗ 𝑪 𝑧 𝜆𝑛𝜆 , 𝜆1 … 𝑪 𝑧 𝜆𝑛𝜆 , 𝜆𝑛𝜆 + 𝜎𝑒2 𝑰 ⎠⎠ (Point B) is supported with a roller constraint (only the y displacement
(6) is constrained).
In this work, uncertainty inputs with three statistically independent
and to obtain the inferred true strain probability distribution by com-
components 𝜽 = (𝛿𝐿 , 𝛿𝑅 , 𝜅) are considered, where 𝛿𝐿 and 𝛿𝑅 denote the
puting the conditional distribution 𝑝(𝒛𝜆∗ |) [50].
( ) load eccentricities on the left and right end of the beam, respectively,
For the covariance matrices 𝑪 𝑧 (𝜆𝑖 , 𝜆𝑗 ) = 𝑷 𝑪 𝑢 (𝜆𝑖 , 𝜆𝑗 )𝑷 𝖳 + 𝑪 𝑑 𝜆𝑖 , 𝜆𝑗
( ) and 𝜅 encapsulates the initial geometric imperfections along the rail.
the FE covariance matrix 𝑪 𝑢 𝜆𝑖 , 𝜆𝑗 is required. As described above, the
Specifically, the load eccentricity 𝛿 is defined as the distance between
FE covariance matrix is obtained by first computing the PCE approxima-
the location where the axial load is applied and the centroid of the
tions 𝒖(𝜆𝑖 , 𝜽) and 𝒖(𝜆𝑗 , 𝜽), and subsequently calculating their empirical
beam’s cross-section, while the initial geometrical imperfection 𝜅 repre-
covariances 𝑪 𝑢 (𝜆𝑖 , 𝜆𝑗 ) using the PCE solution. The covariance matrix
( ) sents the initial misalignment defect of the track due to manufacturing,
𝑪 𝑑 𝜆𝑖 , 𝜆𝑗 is determined from the squared exponential kernel
see Fig. 4. The initial geometry imperfection 𝜅 is applied by assign-
( ) ing the scaled first buckling mode to the geometry of the FE model.
( ) ‖𝒙𝑚 − 𝒙𝑛 ‖2 ‖𝜆𝑖 − 𝜆𝑗 ‖2
𝑐𝑑 (𝒙𝑚 , 𝜆𝑖 ) , (𝒙𝑛 , 𝜆𝑗 ) = 𝜎𝑑 exp −
2
− , (7) The probability distribution of the three random inputs is defined as a
2𝓁𝑑2 2𝓁𝜆2 multivariate normal distribution where the means (in millimeters) and
where 𝒙𝑚 and 𝒙𝑛 denote the coordinates of the measurement locations. variances are determined empirically as
The scaling parameter 𝜎𝑑 and the length scale parameters 𝓁𝑑 and 𝓁𝜆 are ⎛⎛ 1.78 ⎞ ⎛ 0.52 0 0 ⎞⎞
the three underlying hyperparameters of the proposed statistical model. 𝑝(𝜽) =  ⎜⎜ 0.27 ⎟ , ⎜ 0 0.52 0 ⎟⎟ . (9)
The hyperparameters are learned from the data by minimizing the nega- ⎜⎜ ⎟ ⎜ ⎟⎟
⎝⎝ −3.0 ⎠ ⎝ 0 0 0.52 ⎠⎠
tive of the log marginal likelihood − log 𝑝(). According to Equation (6)
and the marginalization property of Gaussians, the marginal likelihood A first-order Hermite polynomial was used to obtain a surrogate
is given by model for computing the FE solution 𝒖(𝜆∗ , 𝜽) (see Equations (4)) [59].
This surrogate can approximate the probability distribution of the FE
( ) ( ) ( )
⎛⎛ 𝑷 𝒖 𝜆1 ⎞ ⎛ 𝑪 𝑧 𝜆1 , 𝜆1 + 𝜎𝑒2 𝑰 … 𝑪 𝑧 𝜆1 , 𝜆𝑛𝜆 ⎞⎞ strain that one would get from a full standard MC analysis up to an error
⎜⎜ ⎟ ⎜ ⎟⎟ of less than 1%. The explicit form of the first order PCE is Ψ0 (𝜽) = 1 and
𝑝() =  ⎜⎜ (⋮ ) ⎟ , ⎜ ( ⋮ ) ⋱ ( ⋮ ) ⎟⎟ .
⎜⎜ 𝑷 𝒖 𝜆𝑛 ⎟ ⎜ 𝑪 𝜆 , 𝜆 Ψ1 (𝜽) = 𝜽 = (𝛿𝐿 , 𝛿𝑅 , 𝜅). The respective unknown coefficients associated
𝑪 𝜆 , 𝜆 𝜎 2 𝑰 ⎟⎟
⎝⎝ 𝜆 ⎠ ⎝ 𝑧 𝑛𝜆 1 … 𝑧 𝑛𝜆 𝑛𝜆 + 𝑒 ⎠⎠ with the first order PCE, 𝜶 ∶= (𝛼 0 , 𝛼𝜀1 , 𝛼𝜀1 , 𝛼𝜅1 ), are obtained through
𝐿 𝑅
(8) numerical integration based on Gauss quadrature. Since a first-order

5
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Table 3
Summary of statFEM analysis.
{ }𝑛𝜆 𝑛𝑦
Input: Data: DFOS strain measurements  = 𝒚 𝑖 , 𝜆𝑖 𝑖=1 at 𝑛𝜆 discrete load parameters; standard deviation of the measurement noise 𝜎𝑒 ; coordinates  = {𝒙𝑘 }𝑘=1 of
the 𝑛𝑦 measurement locations.
FE model: A geometrically nonlinear beam FE model of the rail; axial loading 𝑭 ; probability distribution of the input parameters 𝑝(𝜽).

Step 1: Establish the strain projection matrix 𝑷 from the sensor coordinates 𝒙𝑖 , nodal FE coordinates, and FE basis functions.
Step 2: Determine the PCE approximations 𝒖(𝜆𝑖 , 𝜽) and 𝒖(𝜆∗ , 𝜽) in Equation (4) at the 𝑛𝜆 load parameter values and the load parameter of interest 𝜆∗ .
( ) ( )
Step 3: Calculate the empirical mean vectors 𝒖(𝜆𝑖 ) and 𝒖(𝜆∗ ), and the covariance matrices 𝑪 𝑢 𝜆𝑖 , 𝜆𝑗 and 𝑪 𝑢 𝜆∗ , 𝜆𝑗 by sampling from the PCE approximations. Note that
the covariance matrices are symmetric.
Step 4: Construct the model-reality mismatch covariance matrices 𝑪 𝑑 (𝜆𝑖 , 𝜆𝑗 )) by introducing the sensor coordinates 𝒙𝑚 and load values 𝜆𝑖 into the covariance
function (7).
Step 5: Learn the values of the hyperparameters 𝜽 by minimizing the negative log marginal likelihood − log 𝑝().
Step 6: Infer true probability distribution 𝑝(𝒛𝜆∗ |) =  (𝒛| (𝜆∗ ), 𝑪 𝑧| (𝜆∗ , 𝜆∗ ) of the strains by determining the conditional distribution of the joint distribution in
Equation (6).

Fig. 4. Mathematical model (a) of the experiments and the corresponding FE mesh (b) with the applied loading and boundary conditions.

PCE was used, a total of 8 points needs to be evaluated for 3 uncertain distribution shows good agreement with the data in the regime with
inputs (2 Gauss quadrature points for each input). less nonlinearity (below 6 kN) and deviates from the experimental data
Once the PCE surrogate is obtained, the evaluation of 𝒖(𝝀∗ , 𝜽) be- in the highly nonlinear stage (above 6 kN). Moreover, the data in the
comes computationally inexpensive, allowing for evaluation with more highly nonlinear regime is not captured within the 95% confidence in-
samples. Here 𝑛𝑠 = 1000 pseudo-random samples are used drawn from terval of the prior FE strain distribution. This deviation clearly indicates
the joint distribution 𝑝(𝜽). The Gaussian probability distribution 𝑝(𝑷 𝒖) the presence of a mismatch between the reality (experiment) and the FE
of the FE strain can now be obtained by computing the empirical mean model, which will be captured in the random variable 𝒅 in the statis-
and covariance of the samples. Fig. 5 presents the comparison of strain tical model, see Equations (1) and (2). The use of PCE as a surrogate
distribution 𝑝(𝑷 𝒖(11 kN)) at 𝑥 = {0 m, 1.524 m, 3.048 m} along the length model to obtain the FE strain distribution can significantly decrease the
of the rail evaluated with first order PCE and using the MC approach. computational cost while still maintaining high accuracy for the cur-
As evident, the first-order PCE produces a similar strain probability rent problem. This can decrease the computational costs associated with
distribution at different locations in comparison with the MC results, evaluating the uncertain FE predictions to an acceptable level in order
which evaluate the strain distribution through exhaustively sampling to apply the statFEM approach to practical large engineering problems.
the individual FE solution obtained with Abaqus. The average percent-
4.2. Prediction of rail strain in the linear regime
age difference between the two methods is 0.06% in the mean 𝑷 𝒖 and
0.5% in the standard deviation. It is worth noting that at 𝜆∗ = 11 kN
In a railway system, the lateral bending of the rail track under axial
the rail load-deflection response is highly nonlinear and it requires ap-
force is associated with rail lateral deflection, which alters the gauge
proximately 54 seconds to run an individual FE analysis using Abaqus.
distance between two adjacent rails or alignments and can potentially
Evaluating the probabilistic distribution of the FE strain at 11 kN using
result in the derailment of a train.
PCE takes 9 minutes and 27 seconds, which is about 100 times faster The true rail strain 𝒛(𝝀∗ ) underlying the sensor measurement at the
than computing 𝑛𝑠 = 1000 times through MC (16 h 40 min). Therefore, axial load of interest 𝜆∗ is unavailable and only the strain measurements
in the subsequent analysis, the first-order PCE will be used to evaluate at 𝑛𝑦 locations and the corresponding 𝑛𝜆 lower axial loads pairs  are
the prior FE strain distribution 𝑝(𝑷 𝒖). available from both the Rayleigh and BOTDA measurement systems. In
Fig. 6 shows the comparison between the prior FE strain distribution this section, the statFEM approach is used to infer the true rail strain
and the experimental strain measurements obtained with the Rayleigh distribution 𝑝(𝒛|) at the axial load of interest 𝜆∗ based on the uncer-
and BOTDA systems at axial loads 𝜆∗ between 0 kN and 11 kN. The tain FE strain predictions 𝑝(𝑷 𝒖(𝝀∗ )), the DFOS strain measurements 
rail enters the nonlinear regime at different loads in the two progres- by determining the conditional distribution of the joint distribution in
sive buckling tests, which is thought to be the result of different initial Equation (6).
imperfections in the two tests. According to the strain data acquired
from the experiments, the rail response showed evident nonlinearity 4.2.1. Hyper-parameter learning
above 6 kN during the Rayleigh system test, and above 10 kN during Prior to computing the inferred true strain distribution, the hyper-
the BOTDA system test. It is evident from Fig. 6 that the prior FE strain parameters associated with the model-reality mismatch term 𝒅 in the

6
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 5. FE strain probability distribution at 11 kN evaluated using MC (red) and 1𝑠𝑡 order PCE (blue), a) at left end, b) at mid-span, c) at right end.

Fig. 6. FE strain at the mid-length of the rail a) for axial loading from 0 kN to 11 kN, b) over the rail length at an axial load of 6 kN. The blue line represents the
mean of the FE solution. The shaded area represents the 95% confidence interval. The points represent strain measurements at the mid-length of the rail (‘+’ for the
Rayleigh system and ‘⋅’ for the BOTDA system).

statistical model (Equation (2)) need to be derived. For the current im- Before proceeding, it should be emphasized that the DFOS strain
plementation, a bivariate squared exponential kernel (Equation (7)) was data obtained from the experiments and the corresponding 𝑛𝜆 axial
used for the mismatch term 𝒅 [60]. In the nonlinear regime, displace- load parameter pair  were used in the hyperparameter inference.
ments and strains at lower axial loads 𝜆 affect the behaviour at higher Specifically, the strain measurements obtained at particular axial loads
axial loads. Hence, we consider a bivariate squared exponential kernel 𝜆 = {1, 2, … , 6} kN were used. The choice of data constrained the rail
(𝒙 and 𝜆, see Equation (7)) instead of its univariate version (only 𝒙) response under chosen axial loads within the linear regime, where the
to account for this effect. In total three hyper-parameters (𝜎𝑑 , 𝓁𝑑 , 𝓁𝜆 ) strain increased linearly with the axial load. The justification for using
are associated with the bivariate squared exponential kernel, which are only linear data is that in most structural monitoring applications, in-
collected in vector 𝒘. cluding rail monitoring, nonlinear response data is usually unavailable.

7
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 7. Inferred true rail strain with 400 sensors for Rayleigh system under an axial load of a) 2 kN, b) 4 kN, c) 6 kN; and 48 sensors for BOTDA system under an
axial load of d) 2 kN, e) 4 kN, f) 6 kN. The blue ‘+’ represents the Rayleigh measurements and the black ‘.’ represent the BOTDA measurements. The blue lines
represent the mean 𝑷 𝒖(𝜆∗ ) of the FE strain along the rail length, and the black lines the conditioned mean 𝒛| (𝜆∗ ). The shaded areas denote the corresponding 95%
confidence bounds.

Table 4 Table 5
Optimized hyper-parameters for the Rayleigh system and BOTDA system using Optimized hyper-parameters for the BOTDA system and Rayleigh system using
400 sensors and 48 sensors, respectively. a reduced number of sensors 𝑛𝑦 = {3, 10, 40}.

Rayleigh BOTDA 𝜎𝑑 (𝜇𝜀) 𝓁𝑑 (𝑚) 𝓁𝜆 (𝑘𝑁) 𝑅𝑢𝑛𝑛𝑖𝑛𝑔 𝑡𝑖𝑚𝑒(𝑠𝑒𝑐)

𝜎𝑑 (𝜇𝜀) 1.56 23.70 Rayleigh 𝑛𝑦 = 3 2.07 2.10 4.40 1


𝓁𝑑 (𝑚) 0.25 0.88 𝑛𝑦 = 10 2.60 0.98 8.70 6
𝓁𝜆 (𝑘𝑁) 0.91 3.72 𝑛𝑦 = 40 1.73 0.70 0.86 39
Computation time 54 min 04 sec 1 min 36 sec
BOTDA 𝑛𝑦 = 3 9.12 0.36 431.20 4
𝑛𝑦 = 10 16.90 1.35 8.98 10
𝑛𝑦 = 40 18.47 1.38 10.23 98

The hyperparameter associated with the sensor measurement noise 𝜎𝑒


was estimated empirically using the reference strain measurement un- 4.2.2. Influence of the number of sensors and strain measurement accuracy
der zero load. For the Rayleigh system, the value is 𝜎𝑒 = 1.7𝜇𝜀 and for In this section, the use of a reduced number of sensor data points to
BOTDA system 𝜎𝑒 = 7.7𝜇𝜀 (See Table 2). infer the true system strain 𝑝(𝒛|) is considered, for both the Rayleigh
Due to the lack of information about the hyperparameters, the prior and BOTDA systems. It should be emphasized that the two DFOS mea-
probability distribution of the hyperparameters was assumed to be surement systems have characteristic measurement uncertainty 𝜎𝑒 =
𝑝(𝒘) ∝ 1 when inferring the 𝑝(𝒘|), resulting in 𝑝(𝒘|) ∝ 𝑝(|𝒘). The 1.7 𝜇𝜀 and 𝜎𝑒 = 7.7 𝜇𝜀 for Rayleigh and BOTDA, respectively. In prac-
point estimation of the hyperparameters 𝒘 can be obtained through ei- tice, using reduced sensor data is desired to improve the efficiency of
ther MCMC or optimization [1,30]. In the current implementation, the the overall monitoring system by reducing the cost of investment, oper-
limited-memory BFGS with boundary constraints (L-BFGS-B) method ation, and data analysis. For this purpose, three sets of sensor clusters
was used to optimize the hyperparameters, due to its efficiency and ro- 𝑛𝑦 = {3, 10, 40} are considered to infer the true system response. More-
bustness in solving large-scale nonlinear problems [61]. Table 4 gives over, the inference for different loads 𝜆∗ = {2, 4, 6} kN is also compared.
the optimized values of hyperparameters 𝒘 using Rayleigh strain mea- As usual, prior to the inference of the true strain 𝑝(𝒛|), hyperparam-
surements with 400 sensors and BOTDA measurements with 48 sensors. eter estimation is performed. In this set of simulations, the same set of
Once the optimal values of the hyperparameters are obtained, the true hyperparameters (𝜎𝑑 , 𝓁𝑑 , 𝓁𝜆 ) are inferred using L-BFGS-B. The point
strain distribution can be inferred by conditioning on the joint distribu- estimates of the hyperparameters for both the Rayleigh and BOTDA
tion in Equation (6). Fig. 7 shows the FE strain prediction p(𝒖(𝝀∗ )) and systems using a reduced number of sensors are shown in Table 5. Ac-
the inferred true strain prediction 𝑝(𝒛𝜆∗ |) with the 95% confidence in- cording to Table 5, the value of the 𝜎𝑑 from the BOTDA system is much
terval using statFEM at the axial load of interest 𝜆∗ = {2, 4, 6} kN for higher than that from the Rayleigh system. This is expected due to the
both the BOTDA system and Rayleigh system. It is evident from Fig. 6 sensor uncertainty of the BOTDA system being substantially larger than
that the confidence interval of 𝑝(𝒛𝜆∗ |) (grey region) obtained using the Rayleigh system. The optimal values of 𝓁𝑑 are of similar order be-
the Rayleigh system is narrower than the one using the BOTDA system. tween the two systems. Another key observation is that as the sensor
One possible explanation is that the BOTDA data have less precision, resolution increases from 10 to 40, the optimized value of 𝓁𝑑 does not
and the number of data points is approximately 10 times smaller than change significantly. The value of 𝓁𝑑 for the Rayleigh system decreases
the Rayleigh system. by 50% as 𝑛𝑦 increases from 3 to 10, while it decreases by only 20%

8
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

when 𝑛𝑦 increases from 10 to 40. In comparison, the optimized value of to calibrate the statistical model. However, a key observation is that
𝓁𝑑 using sensor measurements from the BOTDA system varies around the predictive nonlinear strain distribution inferred using statFEM devi-
1.35 as 𝑛𝑦 increased from 10 to 40. The computational cost of optimiz- ates from the pure nonlinear FE prediction 𝑝(𝑷 𝒖) towards the measured
ing the hyperparameters decreases significantly compared to the time nonlinear response strain, due to the incorporation of the linear strain
spent on optimization using more sensors (see Table 4), which is ex- data. And with an increasing number of sensors, the prediction gets
pected considering the reduced effort required for the matrix inversion closer to the measured nonlinear response strain and further from the
when evaluating 𝑝(𝒛|). The difference in runtime for the Rayleigh and FE prior prediction. When using the number of sensors 𝑛𝑦 = 40 for
BOTDA systems is considered to be the result of the different number the Rayleigh system, the measured strain recorded during the non-
of iteration steps required during the optimization as a result of differ- linear loading stage falls within the 95% confidence interval of the
ent levels of precision. These findings indicate that it is possible to use predicted strain. It should be noted here the real rail bending strain
a reduced number of sensors (i.e., data points) for both systems without measurements in the nonlinear regime are not used to calibrate the
affecting the learning of the hyperparameters. statistical model and are only used to compare with the predictive non-
Fig. 8 and Fig. 9 present the inferred true strain distribution linear strain from statFEM. The prediction using the number of sensors
𝑝(𝒛|) under the axial load of interest 𝜆∗ = {2, 4, 6} kN using 𝑛𝑦 = 𝑛𝑦 = {3, 10, 40} from the BOTDA system demonstrates similar perfor-
{3, 10, 40} sensors from the Rayleigh and BOTDA measurements, re- mance to the Rayleigh system. The nonlinear predictive strain response
spectively. As seen in Fig. 8, strain prediction 𝑝(𝑷 𝒖) is less significant at is improved by incorporating the linear strain data into the FE pre-
lower axial load levels. This deviation gradually increases as the axial diction. The 95% confidence interval of the predictive strain for the
load increases. The inferred true strain distribution using 3 sensors does BOTDA system is larger than that for the Rayleigh system due to the
not capture this deviation effectively due to the limited strain informa- larger value of the hyperparameter 𝜎𝑑 for the BOTDA system. In the lin-
tion from the 3 sensors. In comparison, the increase in the number of ear region, the strain distribution prediction using statFEM lies through
sensors enables a more accurate inference of the true strain response. By the measurement data. The predictive nonlinear strain distribution is
comparing the results for 3, 10, and 40 sensors, one can notice that the significantly improved over the FE prior prediction, although this im-
rail response using 3 sensors demonstrates a parabolic response along provement is less significant when only using 3 sensors. For 𝑛𝑦 = 10
the rail length as the axial load increases, while the inference using and 𝑛𝑦 = 40 sensors, the predictive mean of the strain response before
10 and 40 sensors indicates a sinusoidal rail response along the rail 9 kN agrees well with the strain measurements during the experiment.
length. As the axial load increases, the rail response transforms from Considering strain predictions in the non-linear loading regime (i.e.,
a sinusoidal to a parabolic response. The mean of the standard devia- between 9 and 11 kN), even though the predictive mean deviates from
tion of the inferred true strain decreases from 0.7 to 0.4. Fig. 9 shows the strain measurements due to the lack of data at those axial loads, the
the inferred rail response using the strain data from the BOTDA system. strain prediction using the statFEM and the linear strain measurements
Compared to Fig. 8, the inferred strain distribution using 3 sensors has were still better compared to the FE prediction. The available data in
relatively large uncertainties at non-sensor locations while the uncer- practice is usually limited to the linear regime as most railway tracks
tainty decreases with an increasing number of sensors. The increase in are operated under normal (unbuckled) conditions, so using the linear
the number of sensors from 3 to 10 to 40 results in a decrease in the data to achieve improved predictions of the nonlinear response is de-
standard deviation from 6.8 to 1.2 to 1.1, respectively. It is worth not- sired. Understanding the nonlinear rail response under potential loads
ing that with a smaller number of sensors, the inferred rail response is critical for helping to inform rail asset managers when to implement
may be more influenced by individual sensor measurements. These re- future maintenance and repair work. The results presented in this sec-
sults indicate that for both systems it is possible to use fewer sensors to tion indicate that the statFEM overall can provide improved prediction
obtain reliable true strain predictions, but for the BOTDA system, fewer of nonlinear strain by synthesizing the linear measurement data and the
sensors might result in less reliable inference due to the lower precision nonlinear FE predictions, although the improvement is less significant
of the measurements. when the number of sensors is small.

4.4. Prediction of the lateral displacement


4.3. Prediction of rail strain in the nonlinear regime
In this section, the inferred true strain distribution from the statFEM
When performing structural monitoring of actual rail systems, the
is used to derive the rail lateral displacement. Ten thousand groups of
captured strain data is usually limited to the linear regime, while data
strain profiles are sampled based on the inferred true strain distribution
captured during the nonlinear rail response that is indicative of buckling
𝑝(𝒛|) at a specific axial load 𝜆∗ obtained from the statFEM in the previ-
is rarely captured. To overcome this challenge, the statistical model
ous sections, and the curvature profile 𝒌 corresponding to each sampled
in Equations (1) and (2) was calibrated in the linear regime and used
strain profile 𝝐 is further evaluated, based on elastic beam theory, i.e.,
to extrapolate the prediction of the strain response into the nonlinear ( )
stage. This extrapolation is assisted with the prior FE strain distribution 1 𝜆
𝒌= 𝝐− ∗ (10)
prediction 𝑝(𝑷 𝒖) in the nonlinear regime (i.e., 𝜆∗ > 6 kN) obtained using 𝑦fiber − 𝑦NA 𝐸𝐴
the PCE as described in Section 3.2. where 𝑦fiber − 𝑦NA represents the distance of fiber location away from
The predictions of the true rail strain 𝑝(𝒛|) at the rail mid-length the neutral axis (NA); 𝜆∗ represents axial load and 𝐸, 𝐴 represents the
under axial load 𝜆∗ = {0, 1, … , 11} kN with the 95% confidence in- elastic modulus of steel and the cross-sectional area of the rail. It should
tervals using the reduced number of sensors 𝑛𝑦 = {3, 10, 40} from the be noted here that the 𝐸, 𝐴, 𝑃 , 𝑦fiber − 𝑦NA are assumed to be determin-
Rayleigh and BOTDA systems are presented in Fig. 10. The predictive istic.
strain distribution 𝑝(𝒛|) is evaluated based on the joint distribution in With the curvature profile 𝒌, the corresponding displacement pro-
Equation (6), where the point estimates of hyperparameters for 𝑪 𝒅 ob- file 𝒇 is further evaluated based on double integration of the curvature
tained in Section 4.2.2 and the same strain measurements obtained at profile 𝒌 using numerical trapezoid integration, assuming no boundary
particular axial loads 𝜆 = {1, 2, … , 6} kN were used. movement. The displacement distribution 𝑝(𝒇 ) is obtained by evalu-
As shown in Fig. 10, the predictive strain with the 95% confidence ating displacement profiles corresponding to the sampled curvature
interval between 0 and 11 kN covers most of the strain measurements profiles. The predicted displacements with the 95% confidence inter-
below 6 kN, and the mean strain prediction lies mostly through the val were compared with the actual LP-measured displacements during
mean of the data in the linear stage and close to the nonlinear FE the experiments.
prediction in the nonlinear regime. This is expected as only the DFOS Fig. 11 shows the displacement means and the 95% confidence in-
strain measurements within the linear regime (below 6 kN) were used terval. Here, the displacements at an axial load of 6 kN are taken as an

9
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 8. Inferred true rail strain with the reduced number of sensors 𝑛𝑦 for the Rayleigh system under axial load 𝜆∗ kN: a) 𝑛𝑦 = 3, 𝜆∗ = 2, b) 𝑛𝑦 = 3, 𝜆∗ = 4, c)
𝑛𝑦 = 3, 𝜆∗ = 6, d) 𝑛𝑦 = 10, 𝜆∗ = 2, e) 𝑛𝑦 = 10, 𝜆∗ = 4, f) 𝑛𝑦 = 10, 𝜆∗ = 6, g) 𝑛𝑦 = 40, 𝜆∗ = 2, h) 𝑛𝑦 = 40, 𝜆∗ = 4, i) 𝑛𝑦 = 40, 𝜆∗ = 6. The black ‘+’ represents the
Rayleigh measurements. The blue lines represent the mean 𝑷 𝒖(𝜆∗ ) of the FE strain, and the black lines the conditioned mean 𝒛| (𝜆∗ ). The shaded areas denote the
corresponding 95% confidence bounds.

example for both the Rayleigh and BOTDA systems. The actual displace- ferred displacement with the 95% confidence interval enables buckling
ments measured at the mid-length of the rail recorded by the LP are 2.2 assessment and track maintenance to be undertaken with greater confi-
mm and 0.66 mm, respectively. According to Fig. 11 and Table 6, us- dence.
ing a larger number of sensors (i.e., 10 and 40 sensors) overall results
in the improved estimation of lateral displacement, compared with the 4.5. StatFEM using pre-processed data
estimation with only 3 sensors. The lateral displacements inferred us-
The above analysis presents the inferred true strain distribution from
ing only 3 sensors fail to capture the LP-measured displacements for the
statFEM using the distributed strain measurement from a single fiber
Rayleigh system and demonstrate a significantly higher standard devi-
F3 (see Fig. 2) and the corresponding FE strain distribution. The lat-
ation (i.e., lower confidence) for the BOTDA system. In comparison,
eral displacement distribution is evaluated from the inferred true strain
the LP-measured displacements fall within the 95% confidence bounds
distribution, see Section 4.4. However, due to the limited strain in-
of the inferred lateral displacements using 10 and 40 sensors for both
formation provided by a single fiber, one has to incorporate other
the Rayleigh and BOTDA systems. The predicted mean is also closer to
information (i.e., the value of axial loads and the coordinates of the
the LP-measured displacements with a smaller standard deviation than fiber) to obtain the curvature and displacement, which inevitably in-
that of the predicted displacement using only 3 sensors. In practical duces uncertainties when evaluating the curvature and displacement.
operation, the CWR is considered to be at high risk of buckling once Failing to consider those uncertainties would result in a deviation be-
the lateral displacement reaches the critical value (i.e., 12 mm), above tween the predicted displacement means and the actual displacement
which substantial track maintenance work and the appropriate decision (see Fig. 11 d, e, f). One way to avoid considering the uncertainties in
regarding train operation would be required to prevent track buckling. axial loads and the fiber coordinates is to use the curvature data in stat-
The inferred true strain distribution using the statFEM enabled accu- FEM obtained from pre-processing of the strain data from multiple fiber
rate displacement evaluations with the inherent uncertainties from the measurements. With more strain measurements from multiple fibers (F1
sensor measurements and FE predictions, although the confidence is to F8, see Fig. 2), one can pre-process the strain measurements of the
lower when a smaller number of sensors is used in statFEM. The in- fibers (at least three fibers) to derive the curvature data including noise

10
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 9. Inferred true rail strain with the reduced number of sensors 𝑛𝑦 for the BOTDA system under axial load 𝜆∗ kN: a) 𝑛𝑦 = 3, 𝜆∗ = 2, b) 𝑛𝑦 = 3, 𝜆∗ = 4, c)
𝑛𝑦 = 3, 𝜆∗ = 6, d) 𝑛𝑦 = 10, 𝜆∗ = 2, e) 𝑛𝑦 = 10, 𝜆∗ = 4, f) 𝑛𝑦 = 10, 𝜆∗ = 6, g) 𝑛𝑦 = 40, 𝜆∗ = 2, h) 𝑛𝑦 = 40, 𝜆∗ = 4, i) 𝑛𝑦 = 40, 𝜆∗ = 6. The black ‘+’ represents the
BOTDA measurements. The blue lines represent the mean 𝑷 𝒖(𝜆∗ ) of the FE strain, and the black lines the conditioned mean 𝒛| (𝜆∗ ). The shaded areas denote the
corresponding 95% confidence bounds.

Table 6
Mid-length lateral displacement prediction using inferred true strain distribution from the statFEM analysis (in millimeters)).

Predicted displacement Standard deviation Difference from LP measurements

Rayleigh 𝑛𝑦 = 3 2.80 0.46 0.60


𝑛𝑦 = 10 2.21 0.21 0.01
𝑛𝑦 = 40 2.25 0.15 0.05

BOTDA 𝑛𝑦 = 3 3.09 1.39 2.43


𝑛𝑦 = 10 1.08 0.87 0.42
𝑛𝑦 = 40 -0.27 0.49 0.93

along the rail length [12]. This section presents the use of curvature is used to obtain the inferred true bending curvature using the curva-
data with noise and the prior distribution of FE bending curvature in ture data with noise from the experiments conducted with the Rayleigh
statFEM to infer true rail bending curvature and lateral displacement system and BOTDA system, respectively.
distribution at 6 kN. As the curvature is derived from pre-processing Fig. 12 presents the inferred true bending curvature distribution
strain data from multiple fibers, the standard deviation of the noise 𝜎𝑒 obtained through statFEM using curvature data from 𝑛𝑦 = {3, 10, 40}
in the curvature data is not directly available from the manufacturer, sensors in the experiment conducted with the Rayleigh system and
hence it needs to be treated as another hyper-parameter that has to be BOTDA system, respectively. As seen in Fig. 12, the inferred mean
derived from the data. The PCE is used again to approximate the prior of the true bending curvature passes through the mean of the data
FE curvature distribution, and the data used in statFEM is the DFOS cur- and the 95% confidence interval covers the variation of the data
vature data with noise from pre-processing of strain measurements of due to noise from its mean for both systems. As the number of sen-
F1-F8. The same process of statFEM analysis introduced in Section 4.2 sors 𝑛𝑦 increases from 3 to 10 to 40, the standard deviation of in-

11
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 10. Predicted true strain distribution at the rail mid-length between 0-11 kN (black line + grey shaded area) based on linear strain measurements (black
crossing) and nonlinear FE strain prediction (blue line + blue shaded area) using the number of sensors 𝑎) 𝑛𝑦 = 3, 𝑏) 𝑛𝑦 = 10, 𝑐) 𝑛𝑦 = 40 from the Rayleigh test; and
𝑑) 𝑛𝑦 = 3, 𝑒) 𝑛𝑦 = 10, 𝑓 ) 𝑛𝑦 = 40 from the BOTDA test. The blue ‘+’ and ‘.’ represent the nonlinear strain measurements in the Rayleigh and BOTDA test, and are
presented for comparison with the predicted true nonlinear strain distribution.

Fig. 11. Lateral displacement probability distribution evaluated from inferred posterior true strain probability distribution for the Rayleigh system with the number
of sensors 𝑎) 𝑛𝑦 = 3, 𝑏) 𝑛𝑦 = 10, 𝑐) 𝑛𝑦 = 40; and for the BOTDA system with the number of sensors 𝑑) 𝑛𝑦 = 3, 𝑒) 𝑛𝑦 = 10, 𝑓 ) 𝑛𝑦 = 40. The black line represents the inferred
mean of displacement and the grey area represents the 95% confidence interval, compared with the actual displacement ‘+’ measured during the experiments with
LP.

ferred true curvature distribution of the Rayleigh system gradually does not demonstrate significant change as the number of sensors 𝑛𝑦
decreases from 0.39 km−1 to 0.26 km−1 to 0.09 km−1 , indicating the increases, with the standard deviation changing from 0.47 km−1 to
decreased uncertainties in the inferred true curvature distribution as 0.62 km−1 to 0.56 km−1 . Overall, the standard deviation of the in-
the number of sensors increases. In comparison, the standard devia- ferred true curvature distribution decreases as the number of sensors
tion of the inferred true curvature distribution for the BOTDA system increases.

12
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 12. The inferred true curvature distribution at axial load 𝜆∗ = 6 kN for the Rayleigh system with the number of sensors a) 𝑛𝑦 = 3, b) 𝑛𝑦 = 10, c) 𝑛𝑦 = 40; and for
the BOTDA system with the number of sensors d) 𝑛𝑦 = 3, e) 𝑛𝑦 = 10, f) 𝑛𝑦 = 40; the blue line and blue shaded area represent the FE curvature prior 𝑝(𝒌) with 95%
confidence interval; the black line and gray shadow area represents the inferred FE curvature prior; the black crossing ‘+’ represents the curvature data used for
updating.

Table 7
Mid-length lateral displacement prediction using inferred true strain distribution from the statFEM analysis (in millimeters)).

Predicted displacement Standard deviation Difference from LP measurements

Rayleigh 𝑛𝑦 = 3 2.02 0.45 0.18


𝑛𝑦 = 10 2.15 0.30 0.05
𝑛𝑦 = 40 2.23 0.10 0.03

BOTDA 𝑛𝑦 = 3 0.35 0.54 0.31


𝑛𝑦 = 10 0.87 0.73 0.21
𝑛𝑦 = 40 0.63 0.66 0.03

Fig. 13 presents the inferred lateral displacement with the 95% con- lateral displacement but generally does decrease their standard devia-
fidence interval at an axial load of 6 kN obtained using the inferred true tion, although the decrease is less significant for the BOTDA system due
curvature distribution from statFEM, and Table 7 summarizes the in- to its lower precision.
ferred mean and standard deviation of the lateral displacement. As seen
in Fig. 13, the probabilistic distribution of lateral displacement with 5. Conclusions
the 95% confidence interval can successfully capture the LP-measured
displacement. The difference between the mean of the inferred lateral In this paper, the use of the statFEM with DFOS data measured
within the linear range to infer and predict structural response in both
displacement at the mid-length and the LP-measured displacement are
the linear and non-linear range is demonstrated for the first time.
within 0.13 mm for the Rayleigh system and 0.30 mm for the BOTDA
The use of DFOS measurements from a rail buckling experiment is
system. As the number of sensors 𝑛𝑦 increases from 3 to 10 to 40, the
used to demonstrate the approach. By synthesizing available DFOS
standard deviation for the lateral displacement for the Rayleigh test
data and uncalibrated FE model predictions, statFEM provides an ap-
decreases from 0.45 mm to 0.3 mm to 0.1 mm. In comparison, the stan-
proach that leverages both monitoring data and a physical model to
dard deviation for the BOTDA test varies from 0.54 to 0.73 to 0.66 as
achieve better rail strain and lateral displacement predictions. Based
the number of sensors increased from 3 to 10 to 40. Generally, the lat- on the results presented in this study, the following conclusions can be
eral displacement with the 95% confidence bound integrated using the drawn:
inferred true curvature distribution in the test with the Rayleigh system
shows a smaller level of uncertainty than that with the BOTDA sys- 1. Using PCE in statFEM provides an accurate and computationally
tem, which is expected considering the higher precision of the Rayleigh efficient way to obtain the prior FE probability distribution (while
system versus the BOTDA system. Fig. 12 and Fig. 13 show that with accounting for the relative levels of uncertainties from the govern-
pre-processing curvature data from multiple fiber measurements, using ing FE model parameters), especially for large nonlinear FE models.
3 curvature data points only can provide a good inference of the true For the lab model discussed in this paper, using PCE as a surrogate
curvature and lateral displacement. Using more than 10 sensors does model significantly decreased the computational time versus using
not significantly influence the mean of the inferred true curvature and MC simulation from 16 h 40 min to 9 min. The accuracy of the

13
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

Fig. 13. Inferred lateral displacement at axial load 𝜆 = 6 kN evaluated from inferred posterior curvature probability distribution for the Rayleigh test with the number
of sensors, a) 𝑛𝑦 = 3, b) 𝑛𝑦 = 10, c) 𝑛𝑦 = 40; and for the BOTDA test with the number of sensors, 𝑑) 𝑛𝑦 = 3, 𝑒) 𝑛𝑦 = 10, 𝑓 ) 𝑛𝑦 = 40. The black line represents the inferred
mean of lateral displacement and the grey area represents the 95% confidence interval, compared with the actual displacement (‘+’) measured during the experiment
with LP.

mean and standard deviation was maintained to be within 99.94% buckling evaluation. Building on the Bayesian framework, the stat-
and 99.50% compared with the mean and standard deviation eval- FEM enables the fusion of data from different sources and accuracy
uated from the MC simulation. (i.e., sensor data and modelling data) for the improved prediction of
2. The effect of both the number of sensing points included and the ac- true structural response. As the first step toward constructing a digital
curacy of the sensor measurements on the statFEM results was eval- twin of an existing railway system, this paper presented the challenges
uated by considering reduced numbers of sensors 𝑛𝑦 = {3, 10, 40}. associated with the application of statFEM to evaluating rail systems
The use of higher accuracy measurements and more sensors en- monitored using DFOS and offered several solutions. In future work,
ables a more detailed response to be inferred and also decreases the influence of temperature on longitudinally restrained rail systems
the associated uncertainty bounds. However, the confidence inter- will be incorporated into both lab tests and the accompanying stat-
val decreased less significantly when considering a larger num- FEM models. After this further evaluation work is conducted, a digital
ber of sensing points, indicating that even for a reduced number twin of an existing railway system will be constructed to help signif-
of sensing points, adequate rail response inference can still be icantly improve current railway monitoring and buckling evaluation
achieved. processes.
3. By employing the statFEM, it was possible to improve the pre-
dictions of rail strains in the nonlinear loading regime. StatFEM
Declaration of competing interest
provided improved predictions by synthesizing the strain measure-
ment data captured in the linear loading regime with the non-linear
FE model predictions, compared to the predictions from the non- The authors have no conflict of interest/competing interests.
linear FE model alone. Although the prediction deviates from the
experimental measurements as the extrapolation gradually devi- Data availability
ates from the training data, the use of a larger number of sensors
tends to result in improved predictions that are closer to the mea- Data will be made available on request.
sured nonlinear strain for both of the DFOS sensing systems used
in this study.
Acknowledgements
4. The results from statFEM enabled the evaluation of the lateral
displacement along the rail and the corresponding levels of un-
certainty. The evaluation of the lateral displacement was affected This project was supported in part by collaborative research funding
by the number of sensors included and the accuracy of the sens- from the National Research Council of Canada’s Artificial Intelligence
ing systems. When using the same number of strain sensors to for Logistics Program. This research was also supported by the Natural
make predictions, the measurements captured using the Rayleigh Sciences and Engineering Research Council of Canada and Transport
system resulted in inferred lateral displacements with improved Canada. E.F. and F. C. were supported by Wave 1 of The UKRI Strate-
confidence and a closer match to the LP measurements versus the gic Priorities Fund under the EPSRC Grant EP/T001569/1, particularly
BOTDA system. the “Digital twins for complex engineering systems” theme within that
grant, and The Alan Turing Institute. The authors also appreciate the
The results of this research have highlighted the application of stat- support of Merrina Zhang, Alireza Roghani, and Robert Caldwell from
FEM for rail strain and displacement predictions for the purpose of the National Research Council.

14
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

References [26] Han J, Jentzen A, E W. Solving high-dimensional partial differential equations us-
ing deep learning. Proc Natl Acad Sci 2018;115:8505–10. https://doi.org/10.1073/
pnas.1718942115.
[1] Girolami M, Febrianto E, Yin G, Cirak F. The statistical finite element method (stat-
FEM) for coherent synthesis of observation data and model predictions. Comput [27] Raissi M, Perdikaris P, Karniadakis GE. Physics-informed neural networks: a deep
Methods Appl Mech Eng 2021;375:113533. https://doi.org/10.1016/j.cma.2020. learning framework for solving forward and inverse problems involving nonlinear
113533. partial differential equations. J Comput Phys 2019;378:686–707. https://doi.org/
10.1016/j.jcp.2018.10.045.
[2] Lim N-H, Park N-H, Kang Y-J. Stability of continuous welded rail track. Comput
[28] Lu L, Meng X, Mao Z, Karniadakis GE. Deepxde: a deep learning library for solv-
Struct 2003;81:2219–36. https://doi.org/10.1016/S0045-7949(03)00287-6.
ing differential equations. SIAM Rev 2021;63:208–28. https://doi.org/10.1137/
[3] Pucillo GP. Thermal buckling and post-buckling behaviour of continuous welded rail
19M1274067.
track. Veh Syst Dyn 2016;54:1785–807. https://doi.org/10.1080/00423114.2016.
[29] Yang L, Meng X, Karniadakis GE. B-PINNs: Bayesian physics-informed neural net-
1237665.
works for forward and inverse pde problems with noisy data. J Comput Phys
[4] Kang C, Bode M, Wenner M, Marx S. Experimental and numerical investigations
2021;425:109913. https://doi.org/10.1016/j.jcp.2020.109913.
of rail behaviour under compressive force on ballastless track systems. Eng Struct
[30] Febrianto E, Butler L, Girolami M, Cirak F. Digital twinning of self-sensing structures
2019;197:109413. https://doi.org/10.1016/j.engstruct.2019.109413.
using the statistical finite element method. Data-Cent Eng 2022;3. https://doi.org/
[5] Miri A, Zakeri JA, Thambiratnam D, Chan T. Effect of shape of concrete sleepers for
10.1017/dce.2022.28.
mitigating of track buckling. Constr Build Mater 2021;294:123568. https://doi.org/
[31] Barber D. Bayesian reasoning and machine learning. Cambridge University Press;
10.1016/j.conbuildmat.2021.123568.
2012.
[6] Wang P, Xie K, Shao L, Yan L, Xu J, Chen R. Longitudinal force measurement in
[32] Gelman A, Carlin JB, Stern HS, Dunson DB, Vehtari A, Rubin DB. Bayesian data
continuous welded rail with bi-directional FBG strain sensors. Smart Mater Struct
analysis. third ed. CRC Press; 2014.
2015;25:015019. https://doi.org/10.1088/0964-1726/25/1/015019.
[33] Kennedy MC, O’Hagan A. Bayesian calibration of computer models. J R Stat Soc, Ser
[7] Butler LJ, Lin W, Xu J, Gibbons N, Elshafie MZ, Middleton CR. Monitoring, model-
B, Stat Methodol 2001;63:425–64. https://doi.org/10.1111/1467-9868.00294.
ing, and assessment of a self-sensing railway bridge during construction. J Bridge
[34] Higdon D, Kennedy M, Cavendish JC, Cafeo JA, Ryne RD. Combining field data
Eng 2018;23:04018076. https://doi.org/10.1061/(ASCE)BE.1943-5592.0001288.
and computer simulations for calibration and prediction. SIAM J Sci Comput
[8] Qiushi M, Xiaorong G, Hongna Z, Zeyong W, Quanke Z. Composite railway health
2004;26:448–66. https://doi.org/10.1137/S1064827503426693.
monitoring system based on fiber optic Bragg grating sensing array. In: 2014 IEEE
[35] Bayarri MJ, Berger JO, Paulo R, Sacks J, Cafeo JA, Cavendish J, et al. A framework
far East forum on nondestructive evaluation/testing. IEEE; 2014. p. 259–64.
for validation of computer models. Technometrics 2007;49:138–54. https://doi.org/
[9] Wheeler LN, Take WA, Hoult NA, Le H. Use of fiber optic sensing to measure dis-
10.1198/004017007000000092.
tributed rail strains and determine rail seat forces under a moving train. Can Geotech
[36] Xiong Y, Chen W, Tsui K-L, Apley DW. A better understanding of model updat-
J 2019;56:1–13. https://doi.org/10.1139/cgj-2017-0163.
ing strategies in validating engineering models. Comput Methods Appl Mech Eng
[10] Wheeler LN, Pannese E, Hoult NA, Take WA, Le H. Measurement of distributed
2009;198:1327–37. https://doi.org/10.1198/004017007000000092.
dynamic rail strains using a Rayleigh backscatter based fiber optic sensor: lab and
[37] Yang X, Barajas-Solano D, Tartakovsky G, Tartakovsky AM. Physics-informed cokrig-
field evaluation. Transp Geotech 2018;14:70–80. https://doi.org/10.1016/j.trgeo.
ing: a Gaussian-process-regression-based multifidelity method for data-model con-
2017.10.002.
vergence. J Comput Phys 2019;395:410–31. https://doi.org/10.1016/j.jcp.2019.06.
[11] Barker C, Hoult NA, Zhang M. Development of an axial strain measurement sys- 041.
tem for rails. J Perform Constr Facil 2021;35:04020145. https://doi.org/10.1061/ [38] Jiang C, Hu Z, Liu Y, Mourelatos ZP, Gorsich D, Jayakumar P. A sequential calibra-
(ASCE)CF.1943-5509.0001559. tion and validation framework for model uncertainty quantification and reduction.
[12] Sun F, Hoult NA, Butler LJ, Zhang M. Distributed monitoring of rail lateral buckling Comput Methods Appl Mech Eng 2020;368:113172. https://doi.org/10.1016/j.cma.
under axial loading. J Civ Struct Health Monit 2021:1–18. https://doi.org/10.1007/ 2020.113172.
s13349-021-00504-w. [39] Cao B-T, Obel M, Freitag S, Mark P, Meschke G. Artificial neural network surro-
[13] Pimentel R, Ribeiro D, Matos L, Mosleh A, Calçada R. Bridge weigh-in-motion system gate modelling for real-time predictions and control of building damage during
for the identification of train loads using fiber-optic technology. Structures, vol. 30. mechanised tunnelling. Adv Eng Softw 2020;149:102869. https://doi.org/10.1016/
Elsevier; 2021. p. 1056–70. j.advengsoft.2020.102869.
[14] Ye C, Butler L, Huseynov F, Middleton C. Evaluating in-service structural behaviour [40] Han X, Xiang H, Li Y, Wang Y. Predictions of vertical train-bridge response using
of an operational railway bridge using fibre optic sensing and structural model artificial neural network-based surrogate model. Adv Struct Eng 2019;22:2712–23.
updating. Eng Struct 2021;247:113116. https://doi.org/10.1016/j.engstruct.2021. https://doi.org/10.1177/1369433219849809.
113116. [41] Santner TJ, Williams BJ, Notz WI, Williams BJ. The design and analysis of computer
[15] Xu J, Butler LJ, Elshafie MZ. Experimental and numerical investigation of the per- experiments. Springer; 2003.
formance of self-sensing concrete sleepers. Struct Health Monit 2020;19:66–85. [42] Bessa MA, Glowacki P, Houlder M. Bayesian machine learning in metamaterial
https://doi.org/10.1177/147592171983450. design: fragile becomes supercompressible. Adv Mater 2019;31:1904845. https://
[16] Azimi M, Eslamlou AD, Pekcan G. Data-driven structural health monitoring doi.org/10.1002/adma.201904845.
and damage detection through deep learning: state-of-the-art review. Sensors [43] Hariri-Ardebili MA, Sudret B. Polynomial chaos expansion for uncertainty quan-
2020;20:2778. https://doi.org/10.3390/s20102778. tification of dam engineering problems. Eng Struct 2020;203:109631. https://
[17] Sun L, Shang Z, Xia Y, Bhowmick S, Nagarajaiah S. Review of bridge struc- doi.org/10.1016/j.engstruct.2019.109631.
tural health monitoring aided by big data and artificial intelligence: from condi- [44] Ni P, Xia Y, Li J, Hao H. Using polynomial chaos expansion for uncertainty and sen-
tion assessment to damage detection. J Struct Eng 2020;146:04020073. https:// sitivity analysis of bridge structures. Mech Syst Signal Process 2019;119:293–311.
doi.org/10.1061/(ASCE)ST.1943-541X.0002535. https://doi.org/10.1016/j.ymssp.2018.09.029.
[18] Shen C, Dollevoet R, Li Z. Fast and robust identification of railway track stiff- [45] Kish A, Samavedam G, et al. Track buckling prevention: theory, safety concepts,
ness from simple field measurement. Mech Syst Signal Process 2021;152:107431. and applications, Technical Report, Washington D.C.: John A. Volpe National Trans-
https://doi.org/10.1016/j.ymssp.2020.107431. portation Systems Center (US); 2013.
[19] Do NT, Gül M, Nafari SF. Continuous evaluation of track modulus from a moving [46] Rubinstein RY, Kroese DP. Simulation and the Monte Carlo method, vol. 10. John
railcar using ANN-based techniques. Vibration 2020;3:149–61. https://doi.org/10. Wiley & Sons; 2016.
3390/vibration3020012. [47] Gottlieb D, Orszag SA. Numerical analysis of spectral methods: theory and applica-
[20] Graepel T. Solving noisy linear operator equations by Gaussian processes: applica- tions. SIAM; 1977.
tion to ordinary and partial differential equations. In: ICML, vol. 3. 2003. p. 234–41. [48] Xiu D. Numerical methods for stochastic computations: a spectral method approach.
[21] Särkkä S. Linear operators and stochastic partial differential equations in Gaussian Princeton University Press; 2010.
process regression. In: International conference on artificial neural networks; 2011. [49] Feinberg J, Langtangen HP. Chaospy: an open source tool for designing methods of
p. 151–8. uncertainty quantification. J Comput Sci 2015;11:46–57. https://doi.org/10.1016/
[22] Raissi M, Perdikaris P, Karniadakis GE. Machine learning of linear differential j.jocs.2015.08.008.
equations using Gaussian processes. J Comput Phys 2017;348:683–93. https:// [50] Rasmussen CE. Gaussian processes in machine learning. In: Summer school on ma-
doi.org/10.1007/978-3-642-21738-8. chine learning. Springer; 2003. p. 63–71.
[23] Raissi M, Karniadakis GE. Hidden physics models: machine learning of nonlinear [51] Williams CKI, Rasmussen CE. Gaussian processes for machine learning. MIT Press;
partial differential equations. J Comput Phys 2018;357:125–41. https://doi.org/10. 2006.
1016/j.jcp.2017.11.039. [52] Bishop CM. Pattern recognition and machine learning. Information science and
[24] Gregory A, Lau FD-H, Girolami M, Butler LJ, Elshafie MZ. The synthesis of data from statistics. New York: Springer; 2006.
instrumented structures and physics-based models via Gaussian processes. J Comput [53] Bui-Thanh T, Ghattas O, Martin J, Stadler G. A computational framework for
Phys 2019;392:248–65. https://doi.org/10.1016/j.jcp.2019.04.065. infinite-dimensional Bayesian inverse problems part I: the linearized case, with ap-
[25] Lagaris IE, Likas A, Fotiadis DI. Artificial neural networks for solving ordinary and plication to global seismic inversion. SIAM J Sci Comput 2013;35:A2494–523.
partial differential equations. IEEE Trans Neural Netw 1998;9:987–1000. https:// [54] Jordan MI, Ghahramani Z, Jaakkola TS, Saul LK. An introduction to variational
doi.org/10.1109/72.712178. methods for graphical models. Mach Learn 1999;37:183–233.

15
F. Sun, E. Febrianto, H. Fernando et al. Computers and Structures 289 (2023) 107163

[55] Blei DM, Kucukelbir A, McAuliffe JD. Variational inference: a review for statisti- [59] Wiener N. The homogeneous chaos. Am J Math 1938;60:897–936. https://doi.org/
cians. J Am Stat Assoc 2017;112:859–77. 10.2307/2371268.
[56] Povala J, Kazlauskaite I, Febrianto E, Cirak F, Girolami M. Variational Bayesian ap- [60] Schulz E, Speekenbrink M, Krause A. A tutorial on Gaussian process regression: mod-
proximation of inverse problems using sparse precision matrices. Comput Methods elling, exploring, and exploiting functions. J Math Psychol 2018;85:1–16. https://
Appl Mech Eng 2022;393:114712. doi.org/10.1016/j.jmp.2018.03.001.
[57] Gelman A, Carlin JB, Stern HS, Rubin DBR. Bayesian data analysis, texts in statistical [61] Zhu C, Byrd RH, Lu P, Nocedal J. Algorithm 778: L-BFGS-B: fortran subrou-
science. 2nd ed. Chapman and Hall/CRC; 2004. tines for large-scale bound-constrained optimization. ACM Trans Math Softw
[58] Robert C, Casella G. Monte Carlo statistical methods. second ed. Springer; 2004. 1997;23:550–60. https://doi.org/10.1145/279232.279236.

16

You might also like