State estimation in minimal turbulent channel ow A comparative study of 4DVar and PINN Yifan Du Mengze Wang Tamer A. Zaki

2025-05-03 0 0 2.68MB 28 页 10玖币
侵权投诉
State estimation in minimal turbulent channel flow:
A comparative study of 4DVar and PINN
Yifan Du, Mengze Wang, Tamer A. Zaki
Department of Mechanical Engineering, Johns Hopkins Univeristy, Baltimore, MD 21218, USA
Abstract
The state of turbulent, minimal-channel flow is estimated from spatio-temporal sparse
observations of the velocity, using both a physics-informed neural network (PINN) and
adjoint-variational data assimilation (4DVar). The performance of PINN is assessed against
the benchmark results from 4DVar. The PINN is efficient to implement, takes advantage
of automatic differentiation to evaluate the governing equations, and does not require the
development of an adjoint model. In addition, the flow evolution is expressed in terms of
the network parameters which have a far smaller dimension than the predicted trajectory
in state space or even just the initial condition of the flow. Provided adequate observations,
network architecture and training, the PINN can yield satisfactory estimates of the the flow
field, both for the missing velocity data and the entirely unobserved pressure field. However,
accuracy depends on the network architecture, and the dependence is not known a priori.
In comparison to 4DVar estimation which becomes progressively more accurate over the
observation horizon, the PINN predictions are generally less accurate and maintain the
same level of errors throughout the assimilation time window. Another notable distinction
is the capacity to accurately forecast the flow evolution: while the 4DVar prediction depart
from the true flow state gradually and according to the Lyapunov exponent, the PINN is
entirely inaccurate immediately beyond the training time horizon unless re-trained. Most
importantly, while 4DVar satisfies the discrete form of the governing equations point-wise
to machine precision, in PINN the equations are only satisfied in an L2sense.
1 Introduction
Simulations and experiments are complementary approaches in the study of wall turbulence.
The respective benefits of both methods can be combined synergistically, when simulations
are infused with experimental measurements. The assimilation of the observations endows the
simulations with a unique level of fidelity, because their predictions are transformed from being
a generic trajectory in state space to the exact one realized in the experiment. In addition, the
simulations compute the flow state at full resolution, far beyond the original measurements, and
can also directly evaluate quantities that are not measured such as pressure. Data assimilation,
or flow estimation from limited observations, can be performed using a variety of techniques
including filtering [1, 2], smoothing using adjoint [3] and ensemble variational algorithms [4, 5]
and recently using physics-informed machine learning [6]. In this study, we consider sparse
observations from an independent simulation of minimal turbulent channel flow, and perform
state estimation using both adjoint-variational data assimilation (4DVar) and physics informed
neural network (PINN). Using the benchmark results from 4DVar, we provide a detailed analysis
on the accuracy of the estimation using PINN.
Email address for correspondence: t.zaki@jhu.edu
1
arXiv:2210.09424v1 [physics.flu-dyn] 17 Oct 2022
Data assimilation methods for flow reconstruction can be classified based on a number of
criteria, perhaps most important among them in fundamental studies of turbulence is whether
the governing equations are enforced. If the governing equations are not enforced, a statistical
or dynamical model is adopted instead. For example, linear stochastic estimation [7, 8] uses
prior knowledge of two-point correlations to estimate the full velocity field from observations;
Extended Kalman filter provides an optimal update to the state by combining a prediction from
the linearized equations and the observations [9, 10], while ensemble Kalman filter advances
an ensemble of solutions and optimally weighs them based on the observations to estimate
the new state [11]. Some recent machine-learning algorithms also belong in this class, when
neural networks learn a mapping from limited measurements to full flow field, based on training
data from simulations and without direct enforcement of the Navier-Stokes equations. In such
instances, there is no guarantee that predictions, especially outside the training space, satisfy
the governing equations. For example, Fukami et al. [12] trained a convolutional neural network
to map from coarse-grained to full-resolution velocity fields based on training data from two-
dimensional homogeneous turbulence; In Gundersen et al. [13], a semi-conditional variational
auto-encoder was developed to perform flow reconstruction from sparse measurements in a
probabilistic framework, which predicts the full flow field as well as it uncertainty. In these
methods, the predicted fields do not satisfy the physical governing equations, and some of the
estimated flow structures could be the outcome of generalization errors of the surrogate model
instead of physical ones governed by the Navier-Stokes equations.
The second class of methods aims to predict a trajectory of the flow in state space, that
both satisfies the governing equations and optimally reproduces the measurements. In this class,
four-dimensional adjoint variational data assimilation [14, 15], or 4DVar, casts the problem as an
optimization constrained by the governing equations. Starting from an estimate of the unknown
flow state, a forward simulation produces the full flow trajectory over the time horizon where
measurements are available. The disparity between the measurements and their estimates from
the simulation defines the loss function, and also features in the adjoint equations which are
solved backward in time. A complete forward-adjoint loop yields the gradient of the loss with
respect to the unknown flow state, which can be adopted to improve the estimate of the state.
Since the governing equations are nonlinear, the procedure is repeated till convergence, whereby
the optimal flow state is identified and accurately reproduces the entire history of observations.
The method has been adopted in a wide range of applications, including prediction of scalar
sources from remote measurements [16], estimation of transitional and turbulent Taylor-Couette
flows from limited observations [17], and estimation of turbulence in channel flow [18, 19, 20, 21].
An important property of 4DVar is that the computational cost of evaluating the gradient of
the loss function is one forward-adjoint loop, independent of the size of the control vector being
optimized. This efficiency presents an advantage compared to other optimization algorithms
that only adopt the forward model. For example, in ensemble-variational (EnVar) approaches
[22, 23] the gradient of the loss function is evaluated from an ensemble of forward solutions
whose size, and hence the associated computational cost, are proportional to the dimension of
the control vector. Mons et al. [24] performed a comparison of 4DVar, ensemble-variational
(EnVar) method and ensemble Kalman filter, and concluded that 4DVar is the most accurate.
In this study, the adjoint reconstruction of flow field will be adopted as the benchmark for
evaluating the performance of the physics-informed, machine-learning approach.
Recent innovation in machine learning has presented new opportunities for data assimilation
and flow estimation [25, 26, 27, 28, 29]. Our primary focus will be on physics-informed neural
networks (PINNs). Similar to the adjoint-variational approach, flow estimation using PINNs
is formulated as a minimization problem. The network inputs are the spatial and temporal
coordinates and its outputs are the flow variables at the corresponding coordinates. The loss
function for training the PINN is comprised of different parts: (a) The first part is due to
the mismatch between the network predictions and the available flow measurements; (b) The
2
Figure 1: Schematic of the reference simulation of turbulent channel flow and sample observa-
tions. Right two panels show a zoomed-in region of the flow, and contrast the fine and coarse
observations (color rectangle) to the background line contours of the full, unknown flow field.
second part is in terms of the residuals of the governing equations and other constraints, e.g. the
boundary conditions. The minimization of such loss function leads to a neural network which
predicts correct observation values and satisfies the governing equations at the training spatio-
temporal locations, in a weak sense. Raissi et al. [6] introduced the notion of PINNs for
solving partial differential equations (PDEs). The ease of implementation led to a number
of applications with different types of PDEs and flow conditions, for example Mao et al. [25]
considered forward and inverse problems in high-speed flows; Lou et al. [26] examined rarified
flow with Bolzmann-BGK formulation; Mowlavi and Nabi [30] adopted the PINN framework for
optimal control and validated the approach against adjoint-based nonlinear control; and Yang
et al. [31] developed the Bayesian PINN methodology for inverse PDE problems with noisy
data.
In contrast to adjoint methods which enforce the governing equations exactly and preserve
causality in the predicted state-space trajectory, PINNs methods use the residual of the govern-
ing equation as a penalty in the optimization problem. As such, the PINN-estimated flow only
satisfy the governing equations in L2sense. The advantage of PINN, however, is the ease of
implementation for any set of forward evolution equations, and without the need for an adjoint
model. We will provide a practical guide for application of PINN in estimation of turbulent
flows from limited observations, and discuss various properties of these networks.
In this study, we perform estimation of turbulent flow in a minimal channel configuration,
from sparse velocity measurements using 4DVar and PINNs. The results from the adjoint-
variational approach provide the benchmark for evaluating the accuracy of the PINN predic-
tions. In §2, we formulate the data assimilation problem and present the details of the 4DVar
and PINNs algorithms. The results are presented and discussed in §3, followed by a summary
of the main conclusions in §4.
2 Methodology
2.1 The state-estimation problem
In this section, we formulate the data-assimilation problem, where we attempt to estimate the
initial state of channel-flow turbulence from sparse measurements. The channel half height h
and bulk flow velocity Uare adopted as the reference scales, where the star denotes dimensional
quantities. The flow domain, shown in figure 1, is a minimal rectangular flow unit, =
{(x, y, z)[0, Lx]×[0,2] ×[0, Lz]}. The fluid motion is bounded by two fixed walls at y= 0
and y= 2, is periodic on both the horizontal xand zdirections, and is governed by the
3
Domain Size Grid points Grid resolution Reynolds number
L+
xL+
yL+
z
314 360 157
NxNyNz
65 385 65
x+y+
min y+
max z+
4.90 0.40 1.40 2.45
Re Reτ
2800 180
Table 1: Domain sizes and grid resolutions of reference DNS.
non-dimensional incompressible Navier-Stokes equations,
∇ · u= 0 (1)
u
t +u· ∇u+p1
Re2u= 0 (2)
where u= (u, v, w) is the velocity, and pis the pressure. The bulk Reynolds number is
Re Uh, where νis the fluid kinematic viscosity. The velocity satisfies no slip at the
walls, and is periodic on xand zboundaries. The flow is statistically stationary, and is sustained
by a known constant mean pressure gradient in the xstreamwise direction. The total pressure
is decomposed into p=P+p0, where Pis the mean pressure and p0is the fluctuating pressure
that is periodic in xand zdirections. The velocity and pressure boundary conditions can be
expressed in operator form as,
B[u, p0] = 0.(3)
The friction velocity is defined as u
τpτ
w, where τ
wis the mean shear stress on the
walls, and the friction Reynolds number is given by Reτu
τh. The discussion in the main
text will be focused on Reτ= 180, and the effect of a higher Reynolds number (Reτ= 392)
will be discussed in Appendix B. When viscous scaling is adopted for non-dimensionalization,
quantities will be marked by superscript ‘+’. For example, the domain sizes in viscous units,
L+=Lu
τ, are reported in table 1. The present values are motivated by previous studies of
the minimal flow unit, which determined the domain sizes required to sustain wall turbulence
[32, 33, 34]. Specifically, the statistics of minimal channel in the near-wall region agree with
those in large channels when the spanwise and streamwise sizes of the former are larger then
100 wall units [32].
The flow estimation problem relies on the availability of some measurements, or observations.
In the present study, the measurements are spatio-temporal sparse velocity data, without any
knowledge of the pressure field. Our goal is to reconstruct the full spatio-temporal velocity and
pressure fields. The objective is therefore to identify the initial flow state that, when evolved
using the governing equations (1), reproduces all the available measurements. The estimation
of the initial condition can therefore be cast as the constrained optimization problem:
min
u,p Jm=1
2kM (u)mk2,
subject to N[u, p]=0,B[u, p]=0,
(4)
where Jmis the measurement loss, mrepresents sparse measurement velocity data from exper-
iments or reference numerical simulations, and Mis the measurement operator that maps the
continuous spatio-temporal solution uto its values at the same discrete measurement locations.
The operator Nmaps functions uand pto the residuals of equations (1). When the equations
are strongly enforced, the residuals are exactly zero, and the instantaneous forward velocity
fields are fully determined by the initial condition, u=u(u0). As such, the loss function is
fully determined by the initial condition, Jm=Jm(u(u0)), and its derivative is denoted as
DJm/Du0.
In order to ensure accuracy and have full control of the resolution of measurements, we
acquired them from an independent reference direct numerical simulation (DNS). Once the
4
Case ∆MxMyMzMtx+
my+
mz+
mt+
mT+
F 6 24 4 32 32 [9.6, 33.2] 10 2.27 50.0
C 12 48 8 64 64 [19.2, 66.4] 20 4.54 50.0
Table 2: Parameters of observations extracted from the reference simulation. The sampling rate
relative to the DNS is ∆M, with a subscript that denotes the spatial or temporal dimension.
The spatio-temporal resolution of the measurements (subscript ‘m’) and the time horizon Tare
also provided, in viscous units (superscript ‘+’).
measurements are collected, the true flow is hidden throughout the data assimilation procedure,
and is only revisited to quantify the accuracy of the estimated initial conditions. The algorithm
for the reference DNS was used in a number of earlier studies of wall turbulence, and has
been extensively validated [35, 36]. It adopts a fractional-step approach with a local volume
flux formulation on a staggered grid [37]. The advection terms are discretized by the Adams-
Bashforth scheme, and the Crank-Nicolson scheme is adopted for the diffusion terms. The
pressure Poisson equation is solved using Fourier transforms in the periodic directions and
tri-diagonal inversion in the wall-normal direction. The reference solution uris then sub-
sampled in both space and time. The sampling gaps are expressed in terms of the number of
discretization points (∆Mx,My,Mz,Mt) in table (2). Previous studies have demonstrated
that the maximum observation gap that allows accurate reconstruction of the full velocity field
corresponds to the correlation lengthscale of the turbulence [20, 38], which is measured by the
Taylor microscale:
Λξ
r=1
2
d2Rξ(r)
dr21/2
(5)
where Rξ(r) is the two-point correlation function of ξvelocity component in the rcoordinate
direction.
Two resolutions of the measurements will be considered, the finer of which is denoted case
‘F’ and the coarse as case ‘C’ (see table 2 and also figure 1). The gap between observations in the
former case is one Taylor microscale, and twice that scale in case ‘C’. In all cases, observations
are available within a time horizon 0 tT, where in viscous units T+= 50. This duration
is approximately one Lyapunov timescale (τ+
σ48 for Reτ= 180), which has weak Reynolds-
number dependence in the range considered [21, 39]. In all cases, the spatio-temporal resolution
of the observations is not sufficient to achieve synchronization of the estimated state to the true
flow [38]. As a result, the 4DVar and PINN are not expected to perfectly reproduce the true
flow, and we can compare their prediction accuracy.
2.2 Adjoint-variational state estimation
The optimization problem (4) can be reformulated in terms of the Lagrangian,
L=Jm+Dq,N[u, p]E+Dβ,B[u, p]E(6)
where the forward state is q(u, v, w, p) and its adjoint is q= (p, u, v, w). For convenience,
we also introduce forward and adjoint variables, βand β, for enforcing the boundary conditions.
The first order optimality condition yields the following equalities:
DL
Dq= 0,DL
Dβ= 0,
DL
Dq= 0,DL
Dβ= 0,
(7)
where Ddenotes the functional derivative. The first two conditions are gradients of the La-
grangian with respect to the adjoint variables, which lead to the Navier-Stokes equations (1) and
5
摘要:

Stateestimationinminimalturbulentchannelow:Acomparativestudyof4DVarandPINNYifanDu,MengzeWang,TamerA.Zaki*DepartmentofMechanicalEngineering,JohnsHopkinsUniveristy,Baltimore,MD21218,USAAbstractThestateofturbulent,minimal-channelowisestimatedfromspatio-temporalsparseobservationsofthevelocity,usingbotha...

展开>> 收起<<
State estimation in minimal turbulent channel ow A comparative study of 4DVar and PINN Yifan Du Mengze Wang Tamer A. Zaki.pdf

共28页,预览5页

还剩页未读, 继续阅读

声明:本站为文档C2C交易模式,即用户上传的文档直接被用户下载,本站只是中间服务平台,本站所有文档下载所得的收益归上传人(含作者)所有。玖贝云文库仅提供信息存储空间,仅对用户上传内容的表现方式做保护处理,对上载内容本身不做任何修改或编辑。若文档所含内容侵犯了您的版权或隐私,请立即通知玖贝云文库,我们立即给予删除!
分类:图书资源 价格:10玖币 属性:28 页 大小:2.68MB 格式:PDF 时间:2025-05-03

开通VIP享超值会员特权

  • 多端同步记录
  • 高速下载文档
  • 免费文档工具
  • 分享文档赚钱
  • 每日登录抽奖
  • 优质衍生服务
/ 28
客服
关注