Bayesian inference of physics-based models of acoustically-forced laminar premixed conical flames

Alessandro Giannotta Matthew Yoko Stefania Cherubini Pietro De Palma Matthew P. Juniper mpj1001@cam.ac.uk
Abstract

We perform twenty experiments on an acoustically-forced laminar premixed Bunsen flame and assimilate high-speed footage of the natural emission into a physics-based model containing seven parameters. The experimental rig is a ducted Bunsen flame supplied by a mixture of methane and ethylene. A high-speed camera captures the natural emission of the flame, from which we extract the position of the flame front. We use Bayesian inference to combine this experimental data with our prior knowledge of this flame’s behaviour. This prior knowledge is expressed through (i) a model of the kinematics of a flame front moving through a model of the perturbed velocity field, and (ii) a priori estimates of the parameters of the above model with quantified uncertainties. We find the most probable a posteriori model parameters using Bayesian parameter inference, and quantify their uncertainties using Laplace’s method combined with first-order adjoint methods. This is substantially cheaper than other common Bayesian inference frameworks, such as Markov Chain Monte Carlo. This process results in a quantitatively-accurate physics-based reduced-order model of the acoustically forced Bunsen flame for injection velocities ranging from 1.75m/s1.75m/s1.75\,\mbox{m/s}1.75 m/s to 2.99m/s2.99m/s2.99\,\mbox{m/s}2.99 m/s and equivalence ratio values ranging from 1.261.261.261.26 to 1.471.471.471.47, using seven parameters. We use this model to evaluate the heat release rate between experimental snapshots, to extrapolate to different experimental conditions, and to calculate the flame transfer function and its uncertainty for all the flames. Since the proposed model relies on only seven parameters, it can be trained with little data and successfully extrapolates beyond the training dataset. Matlab code is provided so that the reader can apply it to assimilate further flame images into the model.

keywords:
Thermoacoustics , Data Assimilation , Laplace’s Method , Bayesian inference
journal: arXiv
\affiliation

[a]organization=Department of Mechanics, Mathematics and Management, Politecnico di Bari, addressline=via Re David, 200, city=Bari, postcode=70125, country=Italy

\affiliation

[b]organization=Department of Engineering, University of Cambridge, addressline=Trumpington Street, city=Cambridge, postcode=CB2 1PZ, country=UK

Novelty and Significance Statement

Thermoacoustic systems tend to be extremely sensitive to small parameter changes, which makes them difficult to model a priori from existing models in the literature. This means, however, that thermoacoustic models tend to be easy to train using data-driven methods because, with well-chosen experiments, their parameters can be easily observed from experimental data. This paper presents a novel use of Bayesian inference to combine experimental measurements, numerical simulations, and prior knowledge about flame behaviour. We outline our methodology and demonstrate its effectiveness using a laminar premixed Bunsen flame. Our approach yields a quantitatively-accurate physics-based model that predicts the expected value and uncertainty bounds of the flame transfer function between velocity and heat release rate perturbations. The proposed model contains only seven physical parameters, which is fewer parameters than non-physics-based models, and can therefore be trained on relatively little data. We also illustrate how the trained model effectively extrapolates beyond the training dataset. Our numerical code and experimental data are open access.

1 Introduction

The efficiency of the mechanism driving thermoacoustic oscillations depends strongly on the phase difference between the heat release rate (h.r.r) and pressure oscillations [1]. In turn, this phase difference usually depends strongly on the flame’s dynamics [2] and on the acoustics of a combustion chamber, particularly if time delays in the flame are similar to or greater than the period of acoustic chamber modes [3]. Many factors affect the flame dynamics and combustion chamber acoustics, meaning that a combustor’s thermoacoustic behaviour tends to be extremely sensitive to small changes [4]. For the same reason, the outputs of faithful thermoacoustic models are sensitive to small changes in models and their parameters. On the negative side, this makes thermoacoustic systems difficult to model a priori from existing models in the literature. On the positive side, this makes thermoacoustic models easy to train from experimental data because their parameters tend to be easily observable from experimental data. With well-chosen experiments, we can therefore (i) tune the parameters of candidate models and (ii) compare candidate models against each other and select the one with most evidence, given the experimental data [5, 6]. Also on the positive side, this extreme sensitivity means that thermoacoustic systems can often be stabilized by making small changes, which is attractive in industrial settings. The challenge is to model systems accurately and to determine stabilizing modifications as quickly and accurately as possible.

The most influential component of a thermoacoustic system is the mechanism linking acoustic velocity and pressure fluctuations at the base of the flame to subsequent h.r.r. fluctuations in the body of the flame. (In most settings, only the acoustic velocity fluctuations are influential.) This mechanism is difficult to model or simulate [7], so researchers often rely on experimental measurements of the h.r.r. as a function of the acoustic velocity. The fluctuating natural emission from the flame is not, however, a reliable measurement of the fluctuating h.r.r.[8]. Alternatives, such as PLIF to identify reaction zones [9] are more accurate but are technically difficult and, in large systems, impractical. This paper seeks to circumvent the above problems by combining experimental measurements with numerical simulations of a flame model. The flame model is physics-based and qualitatively-accurate. It gives, amongst other things, the flame front dynamics under harmonic forcing. The physics-based model has several parameters, whose expected values and uncertainties are inferred from the flame position, as found from experimental natural emission measurements. We then use the model to calculate the h.r.r., and its uncertainty, as a function of the acoustic velocity perturbations.

In previous work, Juniper & Yoko [5] applied this process to a hot wire Rijke tube. From several candidate models in the literature, they selected those with the most evidence, given the data, and created a quantitatively accurate physics-based model of this thermoacoustic system. In subsequent work on the same system, they used Bayesian optimal experimental design [10] to evaluate the most informative experiments. Following this, they extended the method to infer the flame response of laminar flames from pressure measurements [11]. This study revealed that, in some cases, the uncertainty in the inferred flame responses could be large when the source of information was pressure measurements alone. In this paper we revisit the flames studied in [11] in order to assess an alternative source of information about the flame response: the dynamics of the flame front position. We use a high-speed camera to record the natural emission of the flames in steady conditions and then during a forced limit cycle. We assimilate this data into a physics-based model of a conical Bunsen flame from a previous study [12]. This model contains sufficient parameters to describe the forced experimental conical flame and is differentiable with respect to its parameters. We can therefore obtain the first derivatives of the model outputs with respect to model parameters and use these gradients to (i) find the most probable parameter values, given some data, and (ii) quantify our uncertainty in those parameters and the resulting uncertainty in the model predictions. This process creates a quantitatively-accurate physics-based model of the h.r.r. fluctuations as a function of the velocity perturbation. Unlike the output of many other data-driven approaches, the resulting model is interpretable, trustworthy and extrapolatable.

The paper is structured as follows: in section 2 we explain the experiments and the image processing; in section 3 we report the key feature of the model and how we generate the gradients of the model output with respect to its parameters; in section 4 we explain the Bayesian inference methodology and Laplace’s approximation; in section 5.1 and 5.2 we show how we assimilate the experimental data to create the quantitatively-accurate physics-based model of the flame; in section 5.3 we show how we employ the model to predict the flame transfer function and its uncertainty; in section 5.4 we show how the model is capable of extrapolating outside the training dataset; in section 6 we discuss the conclusions and outline some future work.

2 Experiments

The experimental observations consist of high-speed footage of a conical flame under steady and acoustically-forced conditions. The footage is processed to isolate the position of the flame front.

2.1 Data Collection

The experimental configuration is a laminar premixed conical flame inserted into a vertical duct, as illustrated in figure 1. The lower end of the duct is fixed to a plenum chamber, through which co-flow air is supplied. The upper end is open to the atmosphere. The duct is a 0.8 mmeter\mathrm{m}roman_m long section of quartz tube with an internal diameter of 75 mmmillimeter\mathrm{mm}roman_mm. Quartz is used to allow for optical access to the flame.

Refer to caption
Figure 1: Diagram of the experimental rig.

The burner is a 0.85 mmeter\mathrm{m}roman_m long section of brass tubing with an internal diameter of 14 mmmillimeter\mathrm{mm}roman_mm. The outlet of the burner is fitted with a nozzle that is chosen such that the system could become thermoacoustically unstable. At the injection plane, the nozzle diameter is 9.35 mmmillimeter\mathrm{mm}roman_mm. The burner is fuelled by a mixture of methane and ethylene over a wide range of equivalence ratios and fuel mass flow rates, which allow us to study flames with a wide range of thermoacoustic responses. The reactants are metered using a set of mass flow controllers, from which they flow through a choke plate, decoupling the supply lines from the acoustic fluctuations in the rig.

We use a high-speed camera to record the flame under both steady and perturbed conditions. If the system is linearly stable, we achieve the perturbed condition by using a loudspeaker to harmonically force the flame near the first resonant frequency of the system, f=230Hz𝑓230Hzf=230\mbox{Hz}italic_f = 230 Hz. If the system is self-excited, we achieve the steady condition using active control with a phase-shift amplifier.

We study 20 flames, which we select to produce a wide range of flame responses. We parameterize the flames based on (i) the convective time delay, τc=Lf/u¯subscript𝜏𝑐subscript𝐿𝑓¯𝑢\tau_{c}=L_{f}/\bar{u}italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT / over¯ start_ARG italic_u end_ARG, which is the time taken for a perturbation travelling at the bulk velocity in the burner tube, u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG, to traverse the length of the flame, Lfsubscript𝐿𝑓L_{f}italic_L start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, and (ii) the mean heat release rate of the inner cone, q¯¯𝑞\bar{q}over¯ start_ARG italic_q end_ARG. We group the 20 flames into five sets of four flames each, and select the composition of each flame such that the time delay is constant within a group and the mean heat release rate varies. The convective time delays range from 9.5 ms to 16 ms, and the mean heat release rates range from 375 W to 600 W. These flames produce a thermoacoustic behaviour ranging from strongly damped, to neutral, to strongly driven. The flame properties are summarized in table 1 in A, and illustrated in figure 2.

2.2 Data Processing

The raw footage is processed in order to (i) isolate the flame front position and (ii) produce a Euclidian distance field with the flame front as the zeroth level set. The images are first averaged to improve the signal-to-noise ratio. For the steady flames, this involves averaging over 200 frames. For the unsteady flames, it involves phase-averaging over 20 frames captured at 10 phase angles. The averaged images are straightened and centred so that the symmetry axis of the flame is vertical and in the centre of the frame. This is necessary to undo any misalignment in the camera setup. To do this, we exploit the symmetry of the flame and find the rigid transformation that maps the image to its mirror. From this, we can determine the transformation that moves the symmetry axis of the flame to be vertical and centred in the frame. Once the flame is centred, we perform an Abel deconvolution to undo the line-of-sight integration of the natural luminosity of the flame. We use an implementation of the ‘onion-peeling’ deconvolution [13]. This process produces a clearly defined flame front but accumulates noise towards the centre line of the image. We perform image segmentation on the Abel deconvolved image using Chan-Vese segmentation [14]. This separates the image signal from the background and allows us to interrogate the connectivity map of each cluster of pixels. We discard weakly connected clusters to denoise the image, leaving just the flame front as the dominant image structure. Finally, we perform a Euclidian distance transform on the binary image of the isolated flame front. This produces a matrix of the same dimensions as the original image, where each element contains the normal distance between the corresponding pixel and the flame front. We use this in our data assimilation process to quantify the discrepancy between the predicted and observed flame front positions.

Refer to caption
Figure 2: Processed steady flame images from the 20 flames. Images are artificially coloured according to the approximate convective time delay. Flames with equal mean heat release rates are overlaid.

3 Reduced-Order Model of a laminar premixed Bunsen flame

3.1 Model description

The flame model is detailed in Ref. [12] and can be expressed compactly through:

𝐱˙(t)=𝐟(𝐱(t),t;𝐚),𝐱(0)=𝐱0,formulae-sequence˙𝐱𝑡𝐟𝐱𝑡𝑡𝐚𝐱0subscript𝐱0\dot{\mathbf{x}}(t)=\mathbf{f}(\mathbf{x}(t),t;\mathbf{{\color[rgb]{0,0,0}% \definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}{a}}}),\quad{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{\mathbf{x}(0)=\mathbf{x}_{0}}},over˙ start_ARG bold_x end_ARG ( italic_t ) = bold_f ( bold_x ( italic_t ) , italic_t ; bold_a ) , bold_x ( 0 ) = bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , (1)

where 𝐱(t)=(𝐫(t),𝐳(t))𝐱𝑡𝐫𝑡𝐳𝑡\mathbf{x}(t)=\left(\mathbf{r}(t),\mathbf{z}(t)\right)bold_x ( italic_t ) = ( bold_r ( italic_t ) , bold_z ( italic_t ) ) is the state vector defining the flame front position in terms of radial coordinates, 𝐫𝐫\mathbf{r}bold_r, and longitudinal coordinates, 𝐳𝐳\mathbf{z}bold_z, at a given time, t𝑡titalic_t, and 𝐱𝟎𝐱𝟎\mathbf{x_{0}}bold_x start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT is the initial state. The vector 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a contains the physics-based parameters, while 𝐟𝐟\mathbf{f}bold_f is the physics-based nonlinear operator encapsulating the flame front dynamics. The main feature of the model is that it has been designed to be differentiable with respect to the state and the parameters. The heat release rate q(t)𝑞𝑡q(t)italic_q ( italic_t ) can also be expressed in compact form through:

q(t)=g(𝐱(t);𝐚),𝑞𝑡𝑔𝐱𝑡𝐚q(t)=g(\mathbf{x}(t);\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{a}}}),italic_q ( italic_t ) = italic_g ( bold_x ( italic_t ) ; bold_a ) , (2)

where g𝑔gitalic_g is a nonlinear operator that links the heat release rate q(t)𝑞𝑡q(t)italic_q ( italic_t ) to the flame front position and shape defined by 𝐱(t)𝐱𝑡\mathbf{x}(t)bold_x ( italic_t ). The model parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a are: (i) the flame aspect ratio β=β(u¯,SL)𝛽𝛽¯𝑢subscript𝑆𝐿\beta=\beta(\bar{u},S_{L})italic_β = italic_β ( over¯ start_ARG italic_u end_ARG , italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ), which depends on the bulk velocity in the burner tube u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG and the unstretched laminar flame speed SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT; (ii) the nondimensional Markstein length \mathcal{M}caligraphic_M, which depends on the thermal expansion parameter, the effective Lewis number of the mixture, the Zel’dovich number and the thermal conductivity of the mixture [15]; (iii) the shape parameter μ𝜇\muitalic_μ, which linearly combines a uniform and a parabolic mean velocity profile [16]; (iv) the wavelength of the harmonic perturbation velocity field K𝐾Kitalic_K; (v) the amplitude of the acoustic forcing εVsubscript𝜀𝑉\varepsilon_{V}italic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT; (vi) the amplitude of the flame base oscillations λ���\lambdaitalic_λ; (vii) the initial phase of the flame base oscillations ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. These seven parameters are sufficient to describe the flame front dynamics qualitatively. All lengths, including the Markstein length, are normalised by the nozzle radius R=4.675𝑅4.675R=4.675italic_R = 4.675 mmmillimeter\mathrm{mm}roman_mm.

3.2 Obtaining periodic solutions and their sensitivities using adjoint methods

All flames are perturbed by a harmonic acoustic forcing with period T𝑇Titalic_T. We consider the following vector-valued function:

𝐅(𝐱0;𝐚)=𝐱(T)𝐱0,where 𝐱(T)=0T𝐟(t,𝐱;𝐚)dt+𝐱0.formulae-sequence𝐅subscript𝐱0𝐚𝐱𝑇subscript𝐱0where 𝐱𝑇superscriptsubscript0𝑇𝐟𝑡𝐱𝐚d𝑡subscript𝐱0\mathbf{F}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}};% \mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}})=\mathbf{x}(T)-{% \color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}},\quad% \text{where }\mathbf{x}(T)=\int_{0}^{T}\mathbf{f}(t,\mathbf{x};\mathbf{{\color% [rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}})\,\mbox{d}t+{% \color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}}.bold_F ( bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ; bold_a ) = bold_x ( italic_T ) - bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , where bold_x ( italic_T ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_f ( italic_t , bold_x ; bold_a ) d italic_t + bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT . (3)

For a given parameter set 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a, a solution 𝐱(t)𝐱𝑡\mathbf{x}(t)bold_x ( italic_t ) of equation (1) with an initial state 𝐱0subscript𝐱0{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}}bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is periodic with period T𝑇Titalic_T if:

𝐅(𝐱0;𝐚)=𝟎.𝐅subscript𝐱0𝐚𝟎\mathbf{F}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}};% \mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}})=\mathbf{0}.bold_F ( bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ; bold_a ) = bold_0 . (4)

We solve equation (4) using the trust-region dogleg algorithm and by providing the partial derivative of 𝐅𝐅\mathbf{F}bold_F with respect to the initial state 𝐱0subscript𝐱0{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}}bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT

𝐅𝐱0=𝐱(T)𝐱0𝐈,partial-derivativesubscript𝐱0𝐅partial-derivativesubscript𝐱0𝐱𝑇𝐈\partialderivative{\mathbf{F}}{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{\mathbf{x}_{0}}}}=\partialderivative{\mathbf{x}(T)}{{\color[rgb]{0,0,0}% \definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}}}-\mathbf{I},divide start_ARG ∂ start_ARG bold_F end_ARG end_ARG start_ARG ∂ start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG = divide start_ARG ∂ start_ARG bold_x ( italic_T ) end_ARG end_ARG start_ARG ∂ start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG - bold_I , (5)

where 𝐈𝐈\mathbf{I}bold_I is the identity matrix, and 𝐱(T)/𝐱0𝐱𝑇subscript𝐱0\partial{\mathbf{x}(T)}/\partial{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{\mathbf{x}_{0}}}}∂ bold_x ( italic_T ) / ∂ bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the partial derivative of the state at time T𝑇Titalic_T with respect to changes in the initial state 𝐱0subscript𝐱0{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}}bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, which we compute using adjoint methods. We can also obtain the gradient of the periodic solution 𝐱(t)𝐱𝑡{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}(t)}}bold_x ( italic_t ) with respect to the parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a, by differentiating equation (3) with respect to 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a:

d𝐅d𝐚=𝐱(T)𝐚+𝐱(T)𝐱0d𝐱0d𝐚d𝐱0d𝐚=𝟎d𝐱0d𝐚=(𝐱(T)𝐱0𝐈)1𝐱(T)𝐚,derivative𝐚𝐅partial-derivative𝐚𝐱𝑇partial-derivativesubscript𝐱0𝐱𝑇derivative𝐚subscript𝐱0derivative𝐚subscript𝐱0𝟎derivative𝐚subscript𝐱0superscriptpartial-derivativesubscript𝐱0𝐱𝑇𝐈1partial-derivative𝐚𝐱𝑇\begin{split}\derivative{\mathbf{F}}{\mathbf{{\color[rgb]{0,0,0}\definecolor[% named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}{a}}}}=&\partialderivative{\mathbf{x}(T)}{\mathbf{{% \color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}}+% \partialderivative{\mathbf{x}(T)}{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{\mathbf{x}_{0}}}}\derivative{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{\mathbf{x}_{0}}}}{\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{a}}}}-\derivative{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{% rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{% 0}}}}{\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{% 0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}}=\mathbf{0}% \\ \implies&\derivative{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{% rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{% 0}}}}{\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{% 0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}}=-\left(% \partialderivative{\mathbf{x}(T)}{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{\mathbf{x}_{0}}}}-\mathbf{I}\right)^{-1}\partialderivative{\mathbf{x}(T)}{% \mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}},\end{split}start_ROW start_CELL divide start_ARG roman_d start_ARG bold_F end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG = end_CELL start_CELL divide start_ARG ∂ start_ARG bold_x ( italic_T ) end_ARG end_ARG start_ARG ∂ start_ARG bold_a end_ARG end_ARG + divide start_ARG ∂ start_ARG bold_x ( italic_T ) end_ARG end_ARG start_ARG ∂ start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG divide start_ARG roman_d start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG - divide start_ARG roman_d start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG = bold_0 end_CELL end_ROW start_ROW start_CELL ⟹ end_CELL start_CELL divide start_ARG roman_d start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG = - ( divide start_ARG ∂ start_ARG bold_x ( italic_T ) end_ARG end_ARG start_ARG ∂ start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG - bold_I ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG ∂ start_ARG bold_x ( italic_T ) end_ARG end_ARG start_ARG ∂ start_ARG bold_a end_ARG end_ARG , end_CELL end_ROW (6)

and then by differentiating:

d𝐱(t)d𝐚=𝐱(t)𝐚+𝐱(t)𝐱0d𝐱0d𝐚.derivative𝐚𝐱𝑡partial-derivative𝐚𝐱𝑡partial-derivativesubscript𝐱0𝐱𝑡derivative𝐚subscript𝐱0\derivative{\mathbf{x}(t)}{\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{a}}}}=\partialderivative{\mathbf{x}(t)}{\mathbf{{\color[rgb]{0,0,0}% \definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}{a}}}}+\partialderivative{\mathbf{x}(t)}{{\color[rgb% ]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}}}% \derivative{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\mathbf{x}_{0}}}}{% \mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}}.divide start_ARG roman_d start_ARG bold_x ( italic_t ) end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG = divide start_ARG ∂ start_ARG bold_x ( italic_t ) end_ARG end_ARG start_ARG ∂ start_ARG bold_a end_ARG end_ARG + divide start_ARG ∂ start_ARG bold_x ( italic_t ) end_ARG end_ARG start_ARG ∂ start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG divide start_ARG roman_d start_ARG bold_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG . (7)

Therefore we compute the gradients of the h.r.r. with respect to the parameters by differentiating (2):

dq(t)d𝐚=g(𝐱(t);𝐚)𝐚+g(𝐱(t);𝐚)𝐱(t)d𝐱(t)d𝐚.derivative𝐚𝑞𝑡partial-derivative𝐚𝑔𝐱𝑡𝐚partial-derivative𝐱𝑡𝑔𝐱𝑡𝐚derivative𝐚𝐱𝑡\derivative{q(t)}{\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{a}}}}=\partialderivative{g(\mathbf{x}(t);\mathbf{{\color[rgb]{0,0,0}% \definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}{a}}})}{\mathbf{{\color[rgb]{0,0,0}\definecolor[% named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}{a}}}}+\partialderivative{g(\mathbf{x}(t);\mathbf{{% \color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}})}{\mathbf{x}(t)}% \derivative{\mathbf{x}(t)}{\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{% pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill% {0}{a}}}}.divide start_ARG roman_d start_ARG italic_q ( italic_t ) end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG = divide start_ARG ∂ start_ARG italic_g ( bold_x ( italic_t ) ; bold_a ) end_ARG end_ARG start_ARG ∂ start_ARG bold_a end_ARG end_ARG + divide start_ARG ∂ start_ARG italic_g ( bold_x ( italic_t ) ; bold_a ) end_ARG end_ARG start_ARG ∂ start_ARG bold_x ( italic_t ) end_ARG end_ARG divide start_ARG roman_d start_ARG bold_x ( italic_t ) end_ARG end_ARG start_ARG roman_d start_ARG bold_a end_ARG end_ARG . (8)

This quantity is essential for quick calculation of the uncertainties of the heat release rate and therefore the flame transfer function.

4 Bayesian data assimilation

Bayesian data assimilation provides us with a set of tools that allow us to (i) infer the most probable parameters of a physics-based model, given some data, (ii) quantify the uncertainty in the model parameters, and estimate the systematic uncertainty in the model and data, (iii) rank several candidate models to select the best model, given the data, and (iv) identify optimal experiment designs and sensor placements. In this paper, we use points (i) and (ii), so these are described in detail. A demonstration of model ranking can be found in [11], and a demonstration of optimal experiment design can be found in [10].

In this section, we introduce the following notation to outline the Bayesian inference methodology: the data 𝐃𝐃\mathbf{D}bold_D contains experimental observations of the flame position in steady and/or unsteady configurations; the model \mathcal{H}caligraphic_H encodes a physics-based reduced-order model, such that for a given set of parameters, 𝐚𝐚\mathbf{a}bold_a, the model (𝐚)𝐚\mathcal{H}(\mathbf{a})caligraphic_H ( bold_a ) gives a prediction of the data 𝐃𝐃\mathbf{D}bold_D.

4.1 Parameter inference

When performing parameter inference, we assume that the model is structurally correct and we infer its most probable parameters, 𝐚MPsubscript𝐚MP\mathbf{a}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT. The model is rarely free of structural error, however, and we will revisit this assumption later. We encode our level of uncertainty in the parameter values through a probability distribution, which we denote P()𝑃P(\bullet)italic_P ( ∙ ). Using any prior knowledge we have about the unknown parameters (which may be none at all), we propose a prior probability distribution over the parameter values, P(𝐚|)𝑃conditional𝐚P(\mathbf{a}|\mathcal{H})italic_P ( bold_a | caligraphic_H ). We then assimilate the data, 𝐃𝐃\mathbf{D}bold_D, by performing a Bayesian update on the parameter values:

P(𝐚|𝐃,)=P(𝐃|𝐚,)P(𝐚|)P(𝐃|)𝑃conditional𝐚𝐃𝑃conditional𝐃𝐚𝑃conditional𝐚𝑃conditional𝐃P(\mathbf{a}|\mathbf{D},\mathcal{H})=\frac{P(\mathbf{D}|\mathbf{a},\mathcal{H}% )P(\mathbf{a}|\mathcal{H})}{P(\mathbf{D}|\mathcal{H})}italic_P ( bold_a | bold_D , caligraphic_H ) = divide start_ARG italic_P ( bold_D | bold_a , caligraphic_H ) italic_P ( bold_a | caligraphic_H ) end_ARG start_ARG italic_P ( bold_D | caligraphic_H ) end_ARG (9)

The quantity on the left-hand side of equation (9) is the posterior probability of the parameters, given the data. It is generally computationally intractable to calculate the full posterior, because this requires integration over a large parameter space. The integral typically cannot be evaluated analytically, and requires thousands of model evaluations to compute numerically. At the parameter inference stage, however, we are only interested in finding the most probable parameters, which are those that maximize the posterior. We therefore use an optimization algorithm to find the peak of the posterior without evaluating the full distribution. This process is made computationally efficient by (i) assuming that the experimental uncertainty is Gaussian distributed, and (ii) choosing the prior parameter distribution to be Gaussian. Assumption (i) is reasonable for well-designed experiments in which the uncertainty is dominated by random error, which is typically Gaussian distributed. For assumption (ii) we note that the choice of prior is often the prerogative of the researcher, and we are free to exploit the mathematical convenience offered by the Gaussian distribution. The correlations between model parameters are rarely known a-priori, so an independent Gaussian distribution is often used for the prior, as in this paper. When finding the most probable parameters, we neglect the denominator of the right-hand side of equation (9), because it does not depend on the parameters. It is then convenient to define a cost function, 𝒥𝒥\mathcal{J}caligraphic_J, to be the negative log of the numerator of equation (9), which we minimize. In this study, the data, 𝐃𝐃\mathbf{D}bold_D, consists of a set of experimental observations of the flame position at various time frames and experimental configurations. In its most general form, the cost function for a single image of a flame is:

𝒥(𝐚)=log({P(𝐃|𝐚,)P(𝐚|))}=12(𝐬(𝐚)𝐳)T𝐂ee1(𝐬(𝐚)𝐳)+12(𝐚𝐚𝐩)T𝐂aa1(𝐚𝐚𝐩)+const.\begin{split}\mathcal{J}(\mathbf{a})=&-\log{\{P(\mathbf{D}|\mathbf{a},\mathcal% {H})P(\mathbf{a}|\mathcal{H})}\}\\ =&\frac{1}{2}(\mathbf{s}(\mathbf{a})-\mathbf{z})^{T}\mathbf{C}_{ee}^{-1}(% \mathbf{s}(\mathbf{a})-\mathbf{z})\\ &+\frac{1}{2}(\mathbf{\mathbf{a}-\mathbf{a_{p}}})^{T}\mathbf{C}_{aa}^{-1}(% \mathbf{\mathbf{a}-\mathbf{a_{p}}})+\text{const.}\end{split}start_ROW start_CELL caligraphic_J ( bold_a ) = end_CELL start_CELL - roman_log ( start_ARG { italic_P ( bold_D | bold_a , caligraphic_H ) italic_P ( bold_a | caligraphic_H ) end_ARG ) } end_CELL end_ROW start_ROW start_CELL = end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( bold_s ( bold_a ) - bold_z ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_C start_POSTSUBSCRIPT italic_e italic_e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_s ( bold_a ) - bold_z ) end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( start_ID bold_a - start_ID bold_a start_POSTSUBSCRIPT bold_p end_POSTSUBSCRIPT end_ID end_ID ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_C start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( start_ID bold_a - start_ID bold_a start_POSTSUBSCRIPT bold_p end_POSTSUBSCRIPT end_ID end_ID ) + const. end_CELL end_ROW (10)

where: 𝐬(𝐚)𝐬𝐚\mathbf{s}(\mathbf{a})bold_s ( bold_a ) and 𝐳𝐳\mathbf{z}bold_z are column vectors of the predicted and measured flame front positions respectively; 𝐂eesubscript𝐂𝑒𝑒\mathbf{C}_{ee}bold_C start_POSTSUBSCRIPT italic_e italic_e end_POSTSUBSCRIPT is the covariance matrix describing the experimental uncertainty; 𝐚𝐚\mathbf{a}bold_a and 𝐚psubscript𝐚𝑝\mathbf{a}_{p}bold_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT are column vectors of the current and prior parameter values respectively; 𝐂aasubscript𝐂𝑎𝑎\mathbf{C}_{aa}bold_C start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT is the covariance matrix describing the uncertainty in the prior, and const. is a constant that arises from the Gaussian pre-exponential factors, which does not impact the maximum posterior parameter estimation 𝐚MPsubscript𝐚MP\mathbf{a}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT. In this study, we evaluate the discrepancy between the model predictions and the experimental observations by projecting the predicted flame front position onto the Euclidian distance field produced from the corresponding frame of the high speed footage. This produces a column vector in which each entry is the normal distance between the predicted and observed flame front position at each point in the model.

We see from equation (10) that assuming Gaussian distributions for the data and prior reduces the task of parameter inference to a quadratic optimization problem. We solve this optimization problem using the BFGS algorithm, where the gradient information is provided using adjoint methods.

The parameter inference process is illustrated in figure 3 for a simple system with a single parameter, a𝑎aitalic_a, and a single observable variable, z𝑧zitalic_z. In (a) we show the marginal probability distributions of the prior, p(a)𝑝𝑎p(a)italic_p ( italic_a ), and the data, p(z)𝑝𝑧p(z)italic_p ( italic_z ). The prior and data are independent, so we construct the joint distribution, p(a,z)𝑝𝑎𝑧p(a,z)italic_p ( italic_a , italic_z ) by multiplying the two marginals. In (b), we overlay the model predictions, s𝑠sitalic_s, for various values of a𝑎aitalic_a. Marginalizing along the model predictions yields the true posterior, p(a|z)𝑝conditional𝑎𝑧p(a|z)italic_p ( italic_a | italic_z ). This is possible for a cheap model with a single parameter, but exact marginalization quickly becomes intractable as the number of parameters increases. In (c) we plot the cost function, 𝒥𝒥\mathcal{J}caligraphic_J, which is the negative log of the unnormalized posterior. We show the three steps of gradient-based optimization that are required to find the local minimum, which corresponds to the most probable parameters, aMPsubscript𝑎MPa_{\mathrm{MP}}italic_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT.

Refer to caption
Figure 3: Illustration of parameter inference on a simple univariate system. (a) the marginal probability distributions of the prior and data, p(a)𝑝𝑎p(a)italic_p ( italic_a ) and p(z)𝑝𝑧p(z)italic_p ( italic_z ), as well as their joint distribution, p(a,z)𝑝𝑎𝑧p(a,z)italic_p ( italic_a , italic_z ) are plotted on axes of parameter value, a𝑎aitalic_a, vs observation outcome, z𝑧zitalic_z. (b) the model, \mathcal{H}caligraphic_H, imposes a functional relationship between the parameters, a𝑎aitalic_a, and the predictions, s𝑠sitalic_s. Marginalizing along the model predictions yields the true posterior, p(a|z)𝑝conditional𝑎𝑧p(a|z)italic_p ( italic_a | italic_z ). This is computationally intractable for even moderately large parameter spaces. (c) instead of evaluating the full posterior, we use gradient-based optimization to find its peak. This yields the most probable parameters, aMPsubscript𝑎MPa_{\mathrm{MP}}italic_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT

4.2 Uncertainty quantification

Uncertainty quantification can be split into two steps: (i) quantifying the parametric uncertainty and propagating it to the model prediction uncertainty, and (ii) estimating the systematic and structural uncertainty in the experiments and model predictions. We will deal with these separately.

4.2.1 Parameteric uncertainty

Once we have found the most probable parameters, 𝐚MPsubscript𝐚MP\mathbf{a}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT, we estimate the uncertainty in these parameter values using Laplace’s method [17, 18, 5]. This method approximates the posterior as a Gaussian distribution with a mean of 𝐚MPsubscript𝐚MP\mathbf{a_{\mathrm{MP}}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT, and a covariance given by the Hessian of the cost function:

𝐂aaMP12𝒥aiajsuperscriptsubscript𝐂𝑎𝑎MP1superscript2𝒥subscript𝑎𝑖subscript𝑎𝑗\begin{split}\mathbf{C}_{aa}^{\mathrm{MP}-1}&\approx\frac{\partial^{2}\mathcal% {J}}{\partial a_{i}\partial a_{j}}\end{split}start_ROW start_CELL bold_C start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_MP - 1 end_POSTSUPERSCRIPT end_CELL start_CELL ≈ divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_J end_ARG start_ARG ∂ italic_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∂ italic_a start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG end_CELL end_ROW (11)

The Hessian of the cost function is estimated by the BFGS algorithm during the minimization process. This estimate is based on the change in the value and gradient of the cost function over the previous iteration.

The accuracy of the Laplace approximation depends on the functional dependence between the model predictions and the parameters. This is shown graphically in figure 4 for three univariate systems. In (a), the model is linear in the parameters. Marginalizing a Gaussian joint distribution along any intersecting line produces a Gaussian posterior, so Laplace’s method is exact. In (b), the model is weakly nonlinear in the parameters. The true posterior is skewed, but the Gaussian approximation is still reasonable. This panel also shows a geometric interpretation of Laplace’s method: the approximate posterior is given by linearizing the model around 𝐚MPsubscript𝐚MP\mathbf{a}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT, and marginalizing the joint distribution along the linearized model. In (c), the model is strongly nonlinear in the parameters, so the true posterior is multi-modal and the main peak is highly skewed. In this case, the gradient-based optimization algorithm will only find a single local minimum, which will depend on the choice of initial condition for the optimization. Furthermore, we see that the posterior estimated using Laplace’s method is a poor approximation of the true posterior. This problem can be avoided by reducing the extent of the nonlinearity captured by the joint distribution by (i) shrinking the joint distribution by providing more precise prior information or more precise experimental data, or (ii) re-parameterizing the model to reduce the strength of the nonlinearity [18, Chapter 27].

Refer to caption
Figure 4: Illustration of uncertainty quantification for three univariate systems, comparing the true posterior, p(a|z)𝑝conditional𝑎𝑧p(a|z)italic_p ( italic_a | italic_z ) to the approximate posterior from Laplace’s method, p(a|z)L𝑝subscriptconditional𝑎𝑧𝐿p(a|z)_{L}italic_p ( italic_a | italic_z ) start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. (a) the model is linear in the parameters, so the true posterior is Gaussian and Laplace’s method is exact. (b) the model is weakly nonlinear in the parameters, the true posterior is slightly skewed, but Laplace’s method yields a reasonable approximation. (c) the model is strongly nonlinear in the parameters, the posterior is multi-modal and Laplace’s method underestimates the uncertainty

In many cases, the use of this approximate inference framework approximation will be justifiable, given the substantial reduction in computational cost compared to sampling methods, which are the only viable alternative for constructing the posterior. In previous work, we compared the computational cost of our framework to two sampling approaches [11]. The comparison was done on a computationally cheap thermoacoustic network model. Applying our framework to this model, we can compute the posterior probability of five unknown parameters in under 5 seconds on a single core on a laptop. The same inference problem takes 35 CPU hours running on a workstation when solved with Markov Chain Monte Carlo, and 22 CPU hours when solved with Hamiltonian Monte Carlo.

4.2.2 Uncertainty propagation

To quantify the uncertainty in the model predictions due to the parametric uncertainty, we linearise the model around 𝐚MPsubscript𝐚MP\mathbf{a}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT and propagate the parameter uncertainties through the linearised model [19]. The uncertainty in the model prediction is given by:

𝐂xx=𝐉T𝐂aa𝐉subscript𝐂𝑥𝑥superscript𝐉𝑇subscript𝐂𝑎𝑎𝐉\mathbf{C}_{xx}=\mathbf{J}^{T}\mathbf{C}_{aa}\mathbf{J}bold_C start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT = bold_J start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_C start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT bold_J (12)

where 𝐂xxsubscript𝐂𝑥𝑥\mathbf{C}_{xx}bold_C start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT represents the covariance in the model predictions, and 𝐉𝐉\mathbf{J}bold_J denotes the Jacobian matrix of the model predictions with respect to the parameters 𝐚𝐚\mathbf{a}bold_a. The marginal uncertainty in each model prediction, (σxi)2superscriptsubscript𝜎subscript𝑥𝑖2(\sigma_{x_{i}})^{2}( italic_σ start_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, is given by the diagonal entries of 𝐂xxsubscript𝐂𝑥𝑥\mathbf{C}_{xx}bold_C start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT.

4.2.3 Systematic uncertainty

In most cases, experimental data will contain some systematic uncertainty, and models will contain some structural uncertainty. These uncertainty sources cannot be quantified a-priori, and are often referred to as “unknown unknowns”. We can, however, construct a total covariance matrix, 𝐂ttsubscript𝐂𝑡𝑡\mathbf{C}_{tt}bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT, which encodes the total uncertainty due to (i) the known experimental uncertainty, (ii) the unknown systematic experimental uncertainty, and (iii) the unknown structural model uncertainty. We can then estimate this total covariance from the posterior discrepancy between the model and the data. This must be done simultaneously with parameter inference, because the posterior parameter distribution depends on the total uncertainty in the model and data. We therefore replace 𝐂eesubscript𝐂𝑒𝑒\mathbf{C}_{ee}bold_C start_POSTSUBSCRIPT italic_e italic_e end_POSTSUBSCRIPT with 𝐂ttsubscript𝐂𝑡𝑡\mathbf{C}_{tt}bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT in equation (10), and estimate the total uncertainty by simultaneously minimizing 𝒥𝒥\mathcal{J}caligraphic_J with respect to 𝐚𝐚\mathbf{a}bold_a and 𝐂tt1superscriptsubscript𝐂𝑡𝑡1\mathbf{C}_{tt}^{-1}bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

We begin by calculating the derivative of 𝒥𝒥\mathcal{J}caligraphic_J with respect to 𝐂tt1superscriptsubscript𝐂𝑡𝑡1\mathbf{C}_{tt}^{-1}bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, assuming that the observed variables are uncorrelated, and keeping in mind that the normalizing constant, K𝐾Kitalic_K, depends on 𝐂ttsubscript𝐂𝑡𝑡\mathbf{C}_{tt}bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT:

𝒥=12(𝐬(𝐚)𝐳)T𝐂tt1(𝐬(𝐚)𝐳)𝒥12superscript𝐬𝐚𝐳𝑇superscriptsubscript𝐂𝑡𝑡1𝐬𝐚𝐳\displaystyle\mathcal{J}=\frac{1}{2}(\mathbf{s}(\mathbf{a})-\mathbf{z})^{T}% \mathbf{C}_{tt}^{-1}(\mathbf{s}(\mathbf{a})-\mathbf{z})caligraphic_J = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( bold_s ( bold_a ) - bold_z ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_s ( bold_a ) - bold_z ) +log((2π)k|𝐂tt|)superscript2𝜋𝑘subscript𝐂𝑡𝑡\displaystyle+\log\left(\sqrt{(2\pi)^{k}|\mathbf{C}_{tt}|}\right)+ roman_log ( square-root start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT | bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT | end_ARG ) (13)
+12(𝐚𝐚𝐩)T𝐂aa1(𝐚𝐚𝐩)12superscript𝐚subscript𝐚𝐩𝑇superscriptsubscript𝐂𝑎𝑎1𝐚subscript𝐚𝐩\displaystyle+\frac{1}{2}(\mathbf{a}-\mathbf{a_{p}})^{T}\mathbf{C}_{aa}^{-1}(% \mathbf{a}-\mathbf{a_{p}})+ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( bold_a - bold_a start_POSTSUBSCRIPT bold_p end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_C start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_a - bold_a start_POSTSUBSCRIPT bold_p end_POSTSUBSCRIPT ) +log((2π)k|𝐂aa|)superscript2𝜋𝑘subscript𝐂𝑎𝑎\displaystyle+\log\left(\sqrt{(2\pi)^{k}|\mathbf{C}_{aa}|}\right)+ roman_log ( square-root start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT | bold_C start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT | end_ARG )
𝒥𝐂tt1=12(𝐬(𝐚)𝐳)(𝐬(𝐚)𝐳)T𝐈12𝐂tt𝒥superscriptsubscript𝐂𝑡𝑡112𝐬𝐚𝐳superscript𝐬𝐚𝐳𝑇𝐈12subscript𝐂𝑡𝑡{\frac{\partial\mathcal{J}}{\partial\mathbf{C}_{tt}^{-1}}=\frac{1}{2}(\mathbf{% s}(\mathbf{a})-\mathbf{z})(\mathbf{s}(\mathbf{a})-\mathbf{z})^{T}\circ\mathbf{% I}-\frac{1}{2}\mathbf{C}_{tt}}divide start_ARG ∂ caligraphic_J end_ARG start_ARG ∂ bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( bold_s ( bold_a ) - bold_z ) ( bold_s ( bold_a ) - bold_z ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ∘ bold_I - divide start_ARG 1 end_ARG start_ARG 2 end_ARG bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT (14)

where 𝐈𝐈\mathbf{I}bold_I is the identity matrix, and \circ denotes the Hadamard product. For a given set of parameters, the most probable 𝐂ttsubscript𝐂𝑡𝑡\mathbf{C}_{tt}bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT sets equation (14) to zero. This gives the estimate:

𝐂tt=(𝐬(𝐚)𝐳)(𝐬(𝐚)𝐳)T𝐈subscript𝐂𝑡𝑡𝐬𝐚𝐳superscript𝐬𝐚𝐳𝑇𝐈{\mathbf{C}_{tt}=(\mathbf{s}(\mathbf{a})-\mathbf{z})(\mathbf{s}(\mathbf{a})-% \mathbf{z})^{T}\circ\mathbf{I}}bold_C start_POSTSUBSCRIPT italic_t italic_t end_POSTSUBSCRIPT = ( bold_s ( bold_a ) - bold_z ) ( bold_s ( bold_a ) - bold_z ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ∘ bold_I (15)

which is the expected result that the total variance in the model and data is the square of the discrepancy between the model predictions and the data. Although we cannot directly identify the source of the unknown uncertainty because the experimental and model uncertainties cannot be disentangled, the inferred total uncertainty can assist the researcher with identifying potential error sources. For example, if the unknown error in a single sensor is unexpectedly large, this could indicate a faulty sensor or bad installation. If the unknown error at a certain experimental operating condition is large, this could prompt the researcher to repeat that experiment. If the unknown error grows with one of the input variables, the researcher might investigate the model to see if any important physical phenomena has been neglected.

5 Creating the quantitatively-accurate physics-based model

In this study, we perform Bayesian data assimilation in two steps. In the first step, we assimilate the flame front position data of each flame individually. Therefore, for each experimental configuration (i.e. for each combination of mean flow velocity u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG and equivalence ratio ϕitalic-ϕ\phiitalic_ϕ), we determine the most probable set of model parameters 𝐚=(,μ,β,K,εV,λ,ϕ0)𝐚𝜇𝛽𝐾subscript𝜀𝑉𝜆subscriptitalic-ϕ0\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}=(\mathcal{M},\mu,% \beta,K,\varepsilon_{V},\lambda,\phi_{0})bold_a = ( caligraphic_M , italic_μ , italic_β , italic_K , italic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT , italic_λ , italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), which have been introduced in section 3. This process does not create a general model that can predict the behaviour of an arbitrary flame, but is useful for assessing the influence of the model parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a on the model predictions. After completion of this step, we need to introduce a set of hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α, such that 𝐚=𝐚(ϕ,u¯;𝜶)𝐚𝐚italic-ϕ¯𝑢𝜶\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}=\mathbf{{\color[% rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}(\phi,\bar{u};% \boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}})bold_a = bold_a ( italic_ϕ , over¯ start_ARG italic_u end_ARG ; bold_italic_α ). We then simultaneously assimilate the flame front position data of all flames. This allows us to determine the most probable set of hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α so that, for each combination of mean flow velocity u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG and equivalence ratio ϕitalic-ϕ\phiitalic_ϕ, the model 𝐚=𝐚(ϕ,u¯;𝜶)𝐚𝐚italic-ϕ¯𝑢𝜶\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}=\mathbf{{\color[% rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}(\phi,\bar{u};% \boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}})bold_a = bold_a ( italic_ϕ , over¯ start_ARG italic_u end_ARG ; bold_italic_α ) provides the best prediction of the physics-based model parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a and, consequently, of the flame front dynamics. In figure 5, we show a diagram describing the methodology followed in this study. The end goal is to create a general model that provides the flame transfer function, along with quantified uncertainty bounds, for any given pair of equivalence ratio and bulk velocity in the burner tube.

Refer to caption
Figure 5: Illustration showing the methodology for constructing a quantitatively-accurate reduced-order model. The general model depends on the hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α and comprises two components: the physics-based reduced order model, which depends on the physical parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a, and a linking model 𝐚=𝐚(ϕ,u¯;𝜶)𝐚𝐚italic-ϕ¯𝑢𝜶\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}=\mathbf{{\color[% rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}(\phi,\bar{u};% \boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}})bold_a = bold_a ( italic_ϕ , over¯ start_ARG italic_u end_ARG ; bold_italic_α ) that establishes the relationship between these physical parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a and the hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α. The model is fully differentiable and provides the flame front dynamics and the flame transfer function. By comparing the model’s predictions of the flame front with experimental images, we refine our hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α through Bayesian updates. The resulting model predicts the flame transfer function, with confidence bounds, for any combination of equivalence ratio ϕitalic-ϕ\phiitalic_ϕ and the mean flow velocity u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG. This model is both interpretable and capable of extrapolation.

5.1 Step 1: assimilation of each flame individually

First, we assimilate the steady-state snapshots in order to infer \mathcal{M}caligraphic_M, β𝛽\betaitalic_β and μ𝜇\muitalic_μ. Second, we use these inferred values as priors for the second step, in which we assimilate the unsteady flame images and infer all parameters simultaneously: \mathcal{M}caligraphic_M, β𝛽\betaitalic_β, μ𝜇\muitalic_μ, K𝐾Kitalic_K, εVsubscript𝜀𝑉\varepsilon_{V}italic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT, λ𝜆\lambdaitalic_λ, ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Here we show the process for flame 17 in Table 1, which has an equivalence ratio ϕ=1.34italic-ϕ1.34\phi=1.34italic_ϕ = 1.34, a flow rate of u¯=2.86m/s¯𝑢2.86m/s\bar{u}=2.86\,\mbox{m/s}over¯ start_ARG italic_u end_ARG = 2.86 m/s and an equal volume fraction of methane (CH4)subscriptCH4(\mathrm{CH_{4}})( roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ) and ethylene (C2H4)subscriptC2subscriptH4(\mathrm{C_{2}H_{4}})( roman_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ).

5.1.1 Steady-state data assimilation

The steady flame front position depends on β𝛽\betaitalic_β, \mathcal{M}caligraphic_M and μ𝜇\muitalic_μ. We can include prior knowledge about the parameter β=β(u¯,SL)𝛽𝛽¯𝑢subscript𝑆𝐿\beta=\beta(\bar{u},S_{L})italic_β = italic_β ( over¯ start_ARG italic_u end_ARG , italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) because it depends on the unstretched flame speed SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, which can be estimated using the open source chemical kinetics code Cantera [20], and the mean flow velocity u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG, which can be measured during the experiments. Furthermore, we can estimate the value of the Markstein length by using Matalon’s formulation [15]. Following this process for flame 17171717, we assign a prior estimate to the unstretched laminar flame speed of SL=0.4036ms1subscript𝑆𝐿0.4036superscriptms1S_{L}=0.4036\,\mbox{ms}^{-1}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = 0.4036 ms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and a dimensional Markstein length of ~=0.2579mm~0.2579mm\tilde{\mathcal{M}}=0.2579\,\mbox{mm}over~ start_ARG caligraphic_M end_ARG = 0.2579 mm, corresponding to =~/R=0.0552~𝑅0.0552\mathcal{M}=\tilde{\mathcal{M}}/R=0.0552caligraphic_M = over~ start_ARG caligraphic_M end_ARG / italic_R = 0.0552. To construct the parameter covariance matrix, we must assign confidence bounds to these values, which is naturally a subjective process. In this case, we assign an uncertainty of 6σSL=10%SL6subscript𝜎subscript𝑆𝐿percent10subscript𝑆𝐿6\sigma_{S_{L}}=10\%S_{L}6 italic_σ start_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 10 % italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and 6σ=100%6subscript𝜎percent1006\sigma_{\mathcal{M}}=100\%\mathcal{M}6 italic_σ start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT = 100 % caligraphic_M to indicate that we are relatively more confident in our ability to predict laminar flame speed than Markstein length. The only prior knowledge available for the velocity shape parameter μ𝜇\muitalic_μ is that it can vary between 00 and 1111, so we set a prior with large uncertainty for this parameter. From the steady-state flame snapshot, we can infer the maximum a posteriori probability estimate of the aspect ratio, β𝛽\betaitalic_β, the nondimensional Markstein length, \mathcal{M}caligraphic_M, and the velocity shape parameter, μ𝜇\muitalic_μ. Figure 6(a-b) shows the prior and posterior probability distributions of the model parameters. The contours show 1,2121,21 , 2 and 3333 standard deviations from the prior 𝐚psubscript𝐚𝑝\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}_{p}bold_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (blue) and posterior 𝐚MPsubscript𝐚MP\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT (red). We see that the experimental observations have reduced the uncertainties in the model parameters. Figure 6(c) shows the model prediction ±3plus-or-minus3\pm 3± 3 standard deviations overlaid on the steady flame image. The largest uncertainties are found at the flame tip, which is the region that is most sensitive to the parameters. The model prediction uncertainty is found by solving equation (12).

Refer to caption
Figure 6: (a-b) Probability distributions of parameters in the steady case: prior (blue) and posterior (red). Contours denote 1, 2, and 3 standard deviations from the prior 𝐚psubscript𝐚𝑝\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}_{p}bold_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and posterior 𝐚MPsubscript𝐚MP\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT estimations. Parameters include the flame aspect ratio β𝛽\betaitalic_β, which is a function of the unstretched laminar flame speed SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and the bulk velocity in the burner tube u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG, the shape parameter for the velocity field μ𝜇\muitalic_μ, and the nondimensional Markstein length \mathcal{M}caligraphic_M. Both panels represent the same quantities but panel (a) is scaled to the prior distributions, while panel (b) is scaled to the posterior distributions. (c) Comparison between model predictions before (red line) and after (blue line) assimilating the experimental steady-state flame image. Confidence intervals, spanning 3 standard deviations, are illustrated by the red and blue shadings overlaying the steady flame image.

5.1.2 Unsteady-state (acoustically forced)

We repeat the process with the images of the acoustically forced flame. In this case, we set the values of the three steady parameters using the information gained from the previous step, and we set large uncertainties in the prior values of K𝐾Kitalic_K, εVsubscript𝜀𝑉\varepsilon_{V}italic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT, λ𝜆\lambdaitalic_λ and ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Figure 7(a-b) shows the prior and posterior distribution of the parameters before and after assimilating the unsteady flame images. The results show that the uncertainty in the values of K𝐾Kitalic_K, εVsubscript𝜀𝑉\varepsilon_{V}italic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT, λ𝜆\lambdaitalic_λ and ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT has been reduced significantly through the assimilation of this data, and the values of \mathcal{M}caligraphic_M and μ𝜇\muitalic_μ have shifted slightly from the values found during the assimilation of the steady data, without exceeding the 3-standard deviation bounds. Figure 7(c) shows the model prediction ±3plus-or-minus3\pm 3± 3 standard deviations, plotted against the unsteady flame images before (blue) and after (red) assimilating the experimental video footage. As in the steady-state case, the largest uncertainties are found at the flame tip. We see that the model matches the data well at each frame. With this, we have produced a digital twin of the flame, which we can use to estimate additional quantities of interest such as the unstretched laminar flame speed SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, the Markstein length \mathcal{M}caligraphic_M, the convective speed of velocity perturbations K𝐾Kitalic_K and the heat release rate, which were not directly measured in the experiments. Furthermore, because the model is probabilistic, we can also quantify our uncertainty in these estimates.

Refer to caption
Figure 7: (a-b) Probability distributions of parameters for the acoustically forced case: prior (blue) and posterior (red). Contours depict 1, 2, and 3 standard deviations from the prior 𝐚psubscript𝐚𝑝\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}_{p}bold_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT and posterior 𝐚MPsubscript𝐚MP\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT estimations. Both panels represent the same quantities; however, panel (a) is scaled to the prior distributions, while panel (b) is scaled to the posterior distributions. Parameters include the flame aspect ratio β𝛽\betaitalic_β, which is a function of the unstretched laminar flame speed SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and the bulk velocity in the burner tube u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG, the shape parameter for the velocity field μ𝜇\muitalic_μ, the nondimensional Markstein length \mathcal{M}caligraphic_M, the nondimensional velocity perturbation convective speed K𝐾Kitalic_K, and the amplitude of velocity perturbations εVsubscript𝜀𝑉\varepsilon_{V}italic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT. (c) Comparison of model predictions before (red line) and after (blue line) assimilating the experimental video footage. The confidence interval, spanning 3 standard deviations, is indicated by red and blue shadings overlaid on the experimental flame images.

For the example of flame 17, the unstretched laminar flame speed and the Markstein length are estimated to be SL=0.4191m/ssubscript𝑆𝐿0.4191m/sS_{L}=0.4191\,\mbox{m/s}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = 0.4191 m/s and =0.2430mm0.2430mm\mathcal{M}=0.2430\,\mbox{mm}caligraphic_M = 0.2430 mm compared to the value of SLc=0.4036m/ssubscriptsuperscript𝑆𝑐𝐿0.4036m/sS^{c}_{L}=0.4036\,\mbox{m/s}italic_S start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = 0.4036 m/s and c=0.2579mmsuperscript𝑐0.2579mm\mathcal{M}^{c}=0.2579\,\mbox{mm}caligraphic_M start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT = 0.2579 mm obtained from Cantera and the value of the convective speed is estimated to be K=0.8874𝐾0.8874K=0.8874italic_K = 0.8874 compared to K=1𝐾1K=1italic_K = 1 in Schuller et al. [21], K=0.9𝐾0.9K=0.9italic_K = 0.9 in Kashinath et al. [22] and K=0.83𝐾0.83K=0.83italic_K = 0.83 in Orchini & Juniper [23]. In this case, we were able to provide reasonable prior information for \mathcal{M}caligraphic_M and SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. Similar values are, however, inferred even if the prior is inaccurate and uncertain. This methodology could therefore be used to cheaply and easily estimate combustion properties of fuels for which data are not available.

The proposed model contains few parameters, allowing it to be trained on a relatively small amount of data. Furthermore, the data required to train the model is easy to collect from a simple experiment because only snapshots of the natural emission of the flame are required.

Once the most probable parameter values are inferred and the flame front position has been predicted, we can estimate the heat release rate using equation (2). The resulting heat release rate estimation is shown in figure 8. Panel (a) shows the model prediction (red lines) with a confidence of 3 standard deviations (red shading) overlaid on the experimental observations (grayscale). We also plot the predictions between two observations (blue lines and shading). Panel (b) shows the h.r.r. perturbation qsuperscript𝑞q^{\prime}italic_q start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT normalised by its mean value q¯¯𝑞\bar{q}over¯ start_ARG italic_q end_ARG, computed using the model. The black dots correspond to the time frames observed in the experiments. Figure 8 highlights the advantages of the Bayesian inference approach: using experimental flame images we know the flame front position at ten different time points during a period of oscillation; we use this data to infer the seven parameters of a physics-based reduced order model, which gives the flame front position and the heat release rate at any point in time with their uncertainties.

Refer to caption
Figure 8: Diagram showing the concept of using Bayesian inference to create a quantitatively accurate model capable of providing physical information not directly available from experiments. The experimental data consists of ten frames during one period of oscillation. Panel (a) shows the model prediction (red line) with a confidence interval of 3 standard deviations (red shading) plotted on top of the experimental flame images, as in 7. We also use the model to infer the flame position (blue lines) and its uncertainty (blue shading) at moments between observations. Panel (b) shows the h.r.r. perturbation normalised by its mean value, computed using the model. The black dots correspond to the timeframes observed in the experiments

5.2 Step 2: assimilation of all flames into a general model

At this stage, we introduce a linking model 𝐚=𝐚(ϕ,u¯;𝜶)𝐚𝐚italic-ϕ¯𝑢𝜶\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}=\mathbf{{\color[% rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}(\phi,\bar{u};% \boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}})bold_a = bold_a ( italic_ϕ , over¯ start_ARG italic_u end_ARG ; bold_italic_α ), designed to determine the model parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a based on the equivalence ratio ϕitalic-ϕ\phiitalic_ϕ and bulk flow rate in the burner tube u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG, given any specified set of hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α. Figure 9 shows the maximum a posteriori probability distribution of the model parameters 𝐚MPsubscript𝐚MP\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}_{\mathrm{MP}}bold_a start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT assimilated in the first step, denoted by blue markers with error bars representing 3 standard deviations. Panel (a) shows the assimilated unstretched laminar flame speed (SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT) against the equivalence ratio ϕitalic-ϕ\phiitalic_ϕ. Panel (b) shows the assimilated Markstein length (\mathcal{M}caligraphic_M) against the equivalence ratio ϕitalic-ϕ\phiitalic_ϕ. Panel (c) shows the assimilated shape parameter (μ𝜇\muitalic_μ) against the bulk velocity (u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG) in the burner tube. The parameter μ𝜇\muitalic_μ decreases linearly with increasing u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG. Panel (d) shows the assimilated convective speed ratio (K𝐾Kitalic_K) against the acoustic-forcing Strouhal number St. The plot shows that K𝐾Kitalic_K falls within the range of 0.80.80.80.8 to 1111, with no discernible pattern with respect to St. The remaining parameters εVsubscript𝜀𝑉\varepsilon_{V}italic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT, λ𝜆\lambdaitalic_λ, and ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (not shown) depend on experimental conditions and vary according to external acoustic forcing, both of which are independent of the flame properties.

Based on these observations, the model 𝐚=𝐚(ϕ,u¯;𝜶)𝐚𝐚italic-ϕ¯𝑢𝜶\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}=\mathbf{{\color[% rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}(\phi,\bar{u};% \boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}})bold_a = bold_a ( italic_ϕ , over¯ start_ARG italic_u end_ARG ; bold_italic_α ) is formulated as follows: (i) the laminar flame speed SL=SLc+α1ϕ+α2subscript𝑆𝐿superscriptsubscript𝑆𝐿𝑐subscript𝛼1italic-ϕsubscript𝛼2S_{L}=S_{L}^{c}+{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{% 0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}_{1}\phi+% {\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}_{2}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT + italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ϕ + italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is equal to the laminar flame speed computed using Cantera SLcsuperscriptsubscript𝑆𝐿𝑐S_{L}^{c}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT plus a linear correction that depends on the equivalence ratio; (ii) the Markstein length =α3csubscript𝛼3superscript𝑐\mathcal{M}={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}_{3}\mathcal{M}% ^{c}caligraphic_M = italic_α start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT caligraphic_M start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT is equal to the Markstein length computed using Cantera multiplied by the parameter α3subscript𝛼3{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}_{3}italic_α start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT; (iii) the shape parameter μ=α4u¯+α5𝜇subscript𝛼4¯𝑢subscript𝛼5\mu={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}_{4}\bar{u}+{% \color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}_{5}italic_μ = italic_α start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT over¯ start_ARG italic_u end_ARG + italic_α start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT is assumed to vary linearly with the bulk flow velocity and therefore with the Reynolds number of the bulk flow in the burner tube; (iv) the value K=α6𝐾subscript𝛼6K={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}_{6}italic_K = italic_α start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT is assumed to be constant across all flame cases. We keep εV,λsubscript𝜀𝑉𝜆\varepsilon_{V},\lambdaitalic_ε start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT , italic_λ and ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT equal to the values obtained in the previous assimilation step because the amplitudes of the acoustic forcing and of the flame base oscillations depend on the specific experimental conditions and are not general properties of the flame.

We then assimilate the most probable hyperparameter set, denoted as 𝜶MPsubscript𝜶MP\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}_{\mathrm{MP}}bold_italic_α start_POSTSUBSCRIPT roman_MP end_POSTSUBSCRIPT, by minimizing the cost function expressed by equation (10) for all flames at once. The results of this step are shown in red in figure 9. The red lines display the estimates of the physical parameters 𝐚𝐚\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}bold_a as functions of the equivalence ratio ϕitalic-ϕ\phiitalic_ϕ, the mean flow velocity u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG, and the Strouhal number St, after the assimilation of the hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α. The three red contours show 1, 2, and 3 standard deviations from the posterior values.

Figure 10 displays the predicted flame front dynamics for all 20 flames in the library at a single timestep. The experimental observations are shown as grayscale images, and the model predictions are overlaid as red lines with confidence intervals of 1, 2, and 3 standard deviations (depicted as red shading). The numbers at the base of the flames indicate the corresponding flame numbers referenced in Table 1. In the supplementary material, we provide the animated version of this figure, showing the predictions for all timesteps of the experimental video footage.

Refer to caption
Figure 9: Results of the second step of the data assimilation process. From the first step of the process, we obtain the trends of the physical parameters with respect to the equivalence ratio ϕitalic-ϕ\phiitalic_ϕ and the bulk velocity u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG, which is proportional to the Reynolds number, Re, inside the burner tube. Then we propose a general model for the parameters and assimilate all the flame conditions together. The blue markers represent the posterior values from the first step of the data assimilation process and the error bars correspond to 3 standard deviations. The red dotted lines correspond to the posterior values from the second step and the shaded areas represent 1, 2 and 3 standard deviations. Panel (a) shows the assimilated values of the unstretched laminar flame speed SLsubscript𝑆𝐿S_{L}italic_S start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT plotted against the equivalence ratio ϕitalic-ϕ\phiitalic_ϕ. Panel (b) shows the assimilated values of the dimensional Markstein length ~~\tilde{\mathcal{M}}over~ start_ARG caligraphic_M end_ARG plotted against the equivalence ratio ϕitalic-ϕ\phiitalic_ϕ. Panel (c) shows the assimilated values of the velocity shape parameter μ𝜇\muitalic_μ plotted against the bulk velocity in the burner tube u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG. Panel (d) shows the assimilated values of the convective speed ratio K𝐾Kitalic_K plotted against the acoustic perturbation Strouhal number St.
Refer to caption
Figure 10: The general model predictions for all flames in the library are represented by the red line, with confidence intervals of 3 standard deviations (depicted as red shading), overlaid on the experimental flame images. The numbers at the base of the flames indicate the corresponding flame numbers referenced in Table 1. Additionally, in the supplementary material, we provide an animated version of this figure, showing predictions at all time steps of the experimental video footage.

5.3 Flame transfer function predictions

We introduce the flame transfer function \mathcal{F}caligraphic_F as:

(St;ϕ,u¯,𝜶)=q^/q¯u^ref/u¯,Stitalic-ϕ¯𝑢𝜶^𝑞¯𝑞subscript^𝑢𝑟𝑒𝑓¯𝑢\mathcal{F}(\mbox{St};\phi,\bar{u},\boldsymbol{{\color[rgb]{0,0,0}\definecolor% [named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}% \pgfsys@color@gray@fill{0}{\alpha}}})=\frac{\hat{q}/\bar{q}}{\hat{u}_{ref}/% \bar{u}},caligraphic_F ( St ; italic_ϕ , over¯ start_ARG italic_u end_ARG , bold_italic_α ) = divide start_ARG over^ start_ARG italic_q end_ARG / over¯ start_ARG italic_q end_ARG end_ARG start_ARG over^ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT / over¯ start_ARG italic_u end_ARG end_ARG , (16)

where ^^\hat{\star}over^ start_ARG ⋆ end_ARG denotes the Fourier component of \star at the Strouhal number St; q𝑞qitalic_q denotes the fluctuating heat release rate and urefsubscript𝑢𝑟𝑒𝑓u_{ref}italic_u start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT denotes the velocity fluctuations at a reference position. In this study, we define the reference velocity urefsubscript𝑢𝑟𝑒𝑓u_{ref}italic_u start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT as the component of the velocity field perpendicular to the flame front at its base, adhering to the definition provided by Orchini & Juniper [24]. The heat release rate q𝑞qitalic_q is normalised by the mean heat release rate, while the velocity u𝑢uitalic_u is normalised by the mean flow velocity in the burner tube u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG. We use the calibrated general model obtained in the second step to predict the gain and phase of the flame transfer function at the excitation Strouhal number St for all flames in Table 1.

In figure 11(a), we present the model predictions along with their uncertainties in a polar plot. Each flame transfer function is coloured according to its corresponding flame group as in figure 2. It is worth noting that the flames within the same group, which are grouped based on convective time delay τ𝜏\tauitalic_τ, exhibit similar phase delays on the polar plot, as one would expect. We extend these findings by assuming that the hyperparameters 𝜶𝜶\boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}}bold_italic_α remain constant across different forcing frequencies. Figure 11(b) illustrates the predictions of flame transfer functions along with their uncertainties over a wide range of Strouhal numbers (St). In future research, we will evaluate the validity of this assumption by assimilating data collected over a range of Strouhal numbers. Should it prove untenable, we will enhance the complexity of the linking model 𝐚=𝐚(ϕ,u¯;𝜶)𝐚𝐚italic-ϕ¯𝑢𝜶\mathbf{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}=\mathbf{{\color[% rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{a}}}(\phi,\bar{u};% \boldsymbol{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}% \pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\alpha}}})bold_a = bold_a ( italic_ϕ , over¯ start_ARG italic_u end_ARG ; bold_italic_α ) to incorporate the impact of the forcing frequency, by following the same methodology presented in this paper.

Refer to caption
Figure 11: Inferred flame transfer functions presented as (a) polar plots, which show gain on the radial axis and phase on the angular axis, and (b) Bode plots, which show (i) gain and (ii) phase against Strouhal number St. In (a) we plot flame transfer functions as shaded patches, which denote 1-3 standard deviations around the expected value. In (b) we plot the flame transfer functions as solid lines with a shaded patch representing the confidence bounds of 3 standard deviations, whereas the coloured dots represent the values corresponding to the experimental forcing frequency.

5.4 Extrapolation

The model in this paper contains a small number of parameters because it is physics-based. Consequently, the training process requires only a small amount of data. In this section, we show the model’s ability to predict the flame front dynamics of the 20 flames in our library (see Table 1) using data from just four flames. We train the model’s hyperparameters on flames 1, 6, 11, and 16 of Table 1 and predict the remaining flames with this model. These flames have the same convective time delay, but different powers. We therefore test the ability of the model to extrapolate outside the range of trained convective time delays. Figure 12 shows the predicted flame front dynamics for all flames at a single timestep. Red lines show the flames used for training. Blue lines show the flames used for testing. Red and blue shading denotes 1, 2, and 3 standard deviations from the predicted flame front position. All model predictions are overlaid onto the experimental flame images for comparison. The number at the base of each flames indicates the flame numbers in Table 1. In the supplementary material, we provide the animated version of this figure, showing the predictions for all timesteps of the experimental video footage. We now compare figure 12 with figure 10. When only one-fifth of the available data has been used, the predicted flame front positions are less accurate and have larger uncertainty. This is particularly clear for flame 7, which has a large acoustic amplitude and is close to pinch off in these figures. Indeed, the model shows significant sensitivity to the model parameters as the acoustic amplitude increases. Similarly, the prediction for flame 20 is less accurate because the flame length is much larger than the length of the four flames used for model training. Even with such a small amount of training data, however, the data still falls within the uncertainty bounds of the model predictions, even when the data is well outside the training range. We further note that figure 12 shows the timestep with the largest model misfit and uncertainties. In the supplementary material, we provide an animated version of this figure showing the predictions for all timesteps of the video footage.

Refer to caption
Figure 12: Predicted flame front dynamics for all flames in the library at a single timestep. Red lines show flames used for training, while blue lines show flames used for testing. The shaded regions in red and blue indicate 1, 2, and 3 standard deviations from the optimal flame front estimation. Additionally, all model predictions are overlaid onto the experimental flame images for comparison. In the supplementary material, we provide an animated version of this figure, presenting predictions for all time steps of the experimental video footage.

6 Conclusions

In this study, we perform experiments on an acoustically forced, laminar premixed conical flame in a duct. We use a high-speed camera to record snapshots of the natural emission of the flame while steady and forced. We propose a physics-based reduced-order model that provides, amongst other things, the flame front dynamics under harmonic forcing. We combine the model output with the experimental images of the flame to infer the most probable model parameters. This process (i) turns a qualitatively-accurate model into a quantitatively-accurate model, and (ii) quantifies the uncertainty in the inferred model parameters and the model predictions.

The inference process produces a digital twin of the flame, which provides access to quantities that were not directly measured in the experiments. From observations of the steady flame, we can estimate combustion characteristics such as the laminar flame speed and the Markstein length. This can be used to characterise the combustion properties of a fuel for which data are not available. From observations of the perturbed flame, we can infer the fluctuating heat release rate as a response to the velocity perturbation, which is then used to calculate the flame’s thermoacoustic response through, for example, the flame transfer function. The proposed model contains few parameters and can therefore be trained on relatively little data. To demonstrate this we infer the model’s parameters using one-fifth of the available data and successfully predict the remaining flames in the dataset. The model is also differentiable with respect to the parameters meaning that it can be used as a design optimization tool.

In future work, the flame model will be included within a larger model of the thermoacoustic behaviour of the system, whose other parameters are tuned in the same way. The final result will be a quantitatively-accurate physics-based model of the flame and the thermoacoustic system that is interpretable, trustworthy, and extrapolatable.

The Matlab code developed in this study is made available to the reader, who is invited to modify and apply it to assimilate other flame image data into the flame model.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the research reported in this paper.

Data availability

The experimental data is openly available at http://doi.org/10.17863/CAM.106960. The Matlab code is available at https://github.com/aleGiannotta/Bayesian_Inference_Lam_Flame.git

Acknowledgements

This research was partly supported under the National Recovery and Resilience Plan (NRRP), Mission 4 Component 2 Investment 1.4 - Call for “National Centres” from research to business, funded by the European Union – NextGenerationEU. Project code CN00000023, Concession Decree No. 1033 adopted by Ministero dell’Università e della Ricerca (MUR), CUP - D93C22000410001, Project title “MOST - National Center for Sustainable Mobility”. This research was partly carried out within the NEST - Network 4 Energy Sustainable Transition (D.D. 1243 02/08/2022, PE00000021) and received funding under the National Recovery and Resilience Plan (NRRP), Mission 4 Component 2 Investment 1.3, funded from the European Union - NextGenerationEU. This manuscript reflects only the authors’ views and opinions, neither the European Union nor the European Commission can be considered responsible for them. Matthew Yoko acknowledges funding for his PhD from The Cambridge Trust, The Skye Foundation and the Oppenheimer Memorial Trust.

Appendix A Properties of the flames studied in this paper

Table 1: Summary of the properties of the 20 flames studied. We show the average measured flow rates of air, methane (CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT) and ethylene (C2H4subscriptC2subscriptH4\mathrm{C_{2}H_{4}}roman_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT), the equivalence ratio (ϕitalic-ϕ\phiitalic_ϕ), the bulk velocity in the burner tube (u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG), the measured flame length (Lfsubscript𝐿𝑓L_{f}italic_L start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT), and the mean heat release rate (Q¯¯𝑄\bar{Q}over¯ start_ARG italic_Q end_ARG).
Flame No. Air CH4subscriptCH4\mathrm{CH_{4}}roman_CH start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT C2H4subscriptC2subscriptH4\mathrm{C_{2}H_{4}}roman_C start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ϕitalic-ϕ\phiitalic_ϕ u¯¯𝑢\bar{u}over¯ start_ARG italic_u end_ARG Lfsubscript𝐿𝑓L_{f}italic_L start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT meas. Q¯¯𝑄\bar{Q}over¯ start_ARG italic_Q end_ARG
[-] [ln/min] [ln/min] [ln/min] [-] [m/s] [mm] [W]
1 6.049 0.325 0.325 1.28 1.75 16.6 375
2 6.147 0.348 0.348 1.35 1.79 20.0 375
3 6.219 0.364 0.364 1.40 1.82 23.1 375
4 6.283 0.379 0.379 1.44 1.84 26.4 375
5 6.338 0.391 0.391 1.47 1.86 28.9 375
6 7.246 0.387 0.387 1.27 2.10 19.8 450
7 7.369 0.416 0.416 1.34 2.15 23.8 450
8 7.459 0.436 0.436 1.39 2.18 27.4 450
9 7.537 0.454 0.454 1.43 2.21 30.9 450
10 7.603 0.468 0.468 1.47 2.24 34.2 450
11 8.444 0.449 0.449 1.27 2.45 23.1 525
12 8.594 0.484 0.484 1.34 2.51 27.6 525
13 8.699 0.508 0.508 1.39 2.55 31.7 525
14 8.790 0.529 0.529 1.43 2.58 36.4 525
15 8.868 0.546 0.546 1.47 2.61 39.2 525
16 9.644 0.512 0.512 1.26 2.80 25.9 600
17 9.818 0.553 0.553 1.34 2.86 30.1 600
18 9.939 0.580 0.580 1.39 2.91 34.8 600
19 10.045 0.604 0.604 1.43 2.95 39.9 600
20 10.134 0.624 0.624 1.47 2.99 43.6 600

References

  • [1] J. W. S. B. Rayleigh, The explanation of certain acoustical phenomena, Nature 18 (1878) 319–321.
  • [2] T. C. Lieuwen, V. Yang, Combustion Instabilities In Gas Turbine Engines: Operational Experience, Fundamental Mechanisms, and modelling, Vol. 210, American Institute of Aeronautics and Astronautics, 2005.
  • [3] M. P. Juniper, R. I. Sujith, Sensitivity and Nonlinearity of Thermoacoustic Oscillations, Annual Review of Fluid Mechanics 50 (2018) 661–89.
  • [4] H. C. Mongia, T. Held, G. Hsiao, R. Pandalai, Challenges and progress in controlling dynamics in gas turbine combustors, Journal of Propulsion and Power 19 (2003) 822–829.
  • [5] M. P. Juniper, M. Yoko, Generating a physics-based quantitatively-accurate model of an electrically-heated Rijke tube with bayesian inference, Journal of Sound and Vibration 535 (2022) 117096. doi:10.1016/j.jsv.2022.117096.
  • [6] M. Yoko, M. P. Juniper, Minimizing the data required to train a physics-based thermoacoustic model, in: 29th International Congress on Sound and Vibration, 2023.
  • [7] T. Poinsot, Prediction and control of combustion instabilities in real engines, Proceedings of the Combustion Institute 36 (2017) 1–28.
  • [8] Z. Han, S. Balusamy, S. Hochgreb, Spatial analysis on forced heat release response of turbulent stratified flames, Journal of Engineering for Gas Turbines and Power 137 (2015).
  • [9] R. Yuan, J. Kariuki, A. Dowlut, R. Balachandran, E. Mastorakos, Reaction zone visualisation in swirling spray n-heptane flames, Proceedings of the Combustion Institute 35 (2015) 1649–1656.
  • [10] M. Yoko, M. P. Juniper, Optimal experiment design with adjoint-accelerated Bayesian inference, Data-Centric Engineering 5 (2024) 1–28.
  • [11] M. Yoko, M. P. Juniper, Adjoint-accelerated Bayesian inference applied to the thermoacoustic behaviour of a ducted conical flame, Journal of Fluid Mechanics 985 (2024) A38. doi:10.1017/jfm.2024.276.
  • [12] A. Giannotta, S. Cherubini, P. De Palma, M. P. Juniper, The effect of flame curvature and flame base movement on the frequency response of a conical bunsen flame, Combustion and Flame 259 (2024) 113179.
  • [13] C. J. Dasch, One-dimensional tomography: a comparison of Abel, onion-peeling, and filtered backprojection methods, Applied Optics 31 (1992) 1146. doi:10.1364/ao.31.001146.
  • [14] T. Chan, L. Vese, An active contour model without edges, Lecture Notes in Computer Science (including subseries Lecture Notes in Artificial Intelligence and Lecture Notes in Bioinformatics) 1682 (1999) 141–151. doi:10.1007/3-540-48236.
  • [15] M. Matalon, Intrinsic flame instabilities in premixed and nonpremixed combustion, Annu. Rev. Fluid Mech. 39 (2007) 163–191.
  • [16] H. Yu, M. P. Juniper, L. Magri, A data-driven kinematic model of a ducted premixed flame, Proceedings of the Combustion Institute 38 (4) (2021) 6231–6239.
  • [17] H. Jeffreys, Scientific Inference, 3rd Edition, Cambridge University Press, 1973.
  • [18] D. J. MacKay, Information-based objective functions for active data selection, Neural computation 4 (1992) 590–604.
  • [19] M. M. Putko, A. C. Taylor, P. A. Newman, L. L. Green, Approach for input uncertainty propagation and robust design in CFD using sensitivity derivatives, Journal of Fluids Engineering, Transactions of the ASME 124 (2002) 60–69. doi:10.1115/1.1446068.
  • [20] D. G. Goodwin, H. K. Moffat, R. L. Speth, Cantera: An object-oriented software toolkit for chemical kinetics, thermodynamics, and transport processes (2018).
  • [21] T. Schuller, S. Ducruix, D. Durox, S. Candel, Modelling tools for the prediction of premixed flame transfer functions, Proceedings of the Combustion Institute 29 (2002) 107–113.
  • [22] K. Kashinath, I. C. Waugh, M. P. Juniper, Nonlinear self-excited thermoacoustic oscillations of a ducted premixed flame: bifurcations and routes to chaos, Journal of Fluid Mechanics 761 (2014) 399–430.
  • [23] A. Orchini, M. P. Juniper, Linear stability and adjoint sensitivity analysis of thermoacoustic networks with premixed flames, Combustion and Flame 165 (2015) 97–108. doi:10.1016/j.combustflame.2015.10.011.
  • [24] A. Orchini, M. P. Juniper, Flame double input describing function analysis, Combustion and Flame 171 (2016) 87–102.