PreprintPDF Available

Addressing challenges in uncertainty quantification. The case of geohazard assessments

Authors:

Abstract and Figures

By describing critical tasks in quantifying uncertainty using geohazard models, we analyse some of the challenges involved. Under the often-seen condition of very limited data and despite the availability of recently developed sophistications to parameterise models, a major challenge that remains is the constraining of the many model parameters involved. However, challenges also lie in the credibility of predictions required in the assessments, the uncertainty of input quantities, and the conditional nature of the quantification on the choices and assumptions made by analysts. Addressing these challenges calls for more insightful approaches that are yet to be developed; however, clarifications and reinterpretations of some fundamental concepts together with practical simplifications may be required first, and these are discussed in this paper. The research aims at strengthening both the foundation of geohazard risk assessments and its practice.
Content may be subject to copyright.
1
Addressing challenges in uncertainty quantification. The case
of geohazard assessments
Ibsen Chivata Cardenas1, Terje Aven1, Roger Flage1
1Department of Safety, Economics and Planning, University of Stavanger, Stavanger, 4021, Norway
Correspondence to: Ibsen Chivata Cardenas (ibsen.chivatacardenas@uis.no)
5
Abstract. By describing critical tasks in quantifying uncertainty using geohazard models, we analyse some of the
challenges involved. Under the often-seen condition of very limited data and despite the availability of recently
developed sophistications to parameterise models, a major challenge that remains is the constraining of the many
model parameters involved. However, challenges also lie in the credibility of predictions required in the
assessments, the uncertainty of input quantities, and the conditional nature of the quantification on the choices and
10
assumptions made by analysts. Addressing these challenges calls for more insightful approaches that are yet to be
developed; however, clarifications and reinterpretations of some fundamental concepts together with practical
simplifications may be required first, and these are discussed in this paper. The research aims at strengthening both
the foundation of geohazard risk assessments and its practice.
1 Introduction
15
Uncertainty quantification, UQ, helps determine the uncertainty of a system’s responses when some quantities and
events in such system are unknown. Using models, systems responses can be calculated analytically, numerically,
or by random sampling (Monte Carlo method, rejection sampling, Monte Carlo sampling using Markov chains,
importance sampling, subset simulation) (Metropolis and Ulam, 1949; Brown, 1953; Ulam, 1961; Hastings, 1970).
Given the high-dimensional and spatial nature of hazard events and associated quantities, sampling methods are
20
frequently used because they result in a less expensive and more tractable uncertainty quantification in comparison
with analytical and numerical methods. In the sampling procedure, specified distributions of the input quantities
and parameters are sampled and respective outputs of the model are recorded, then the process is repeated as many
times as may be required to achieve the desired accuracy (Vanmarcke, 1984). Eventually, the distribution of the
outputs can be used to calculate probability-based metrics, such as expectations or probabilities of critical events.
25
Model-based uncertainty quantification using sampling is now more often used in geohazard assessments, e.g.,
Uzielli and Lacasse (2007), Wellmann and Regenauer-Lieb (2012), Rodríguez-Ochoa et al. (2015), Pakyuz-
Charrier et al. (2018), Huang et al. (2021), Luo et al. (2021), Sun et al. (2021a).
In this paper, we consider recent advances in UQ and analyse some remaining challenges. For instance, we note
that, despite the availability of sophistications to parameterise models used for UQ, a major problem persists,
30
namely the constraining of the many parameters involved. In practice, based solely on historical data, only some
parameters can be constrained (e.g., Albert, Callies, and von Toussaint, 2022). Another challenge is that model
outputs are not only conditional on the choice of model parameters, but also on input quantities, including initial
and boundary conditions. For example, a geological system model could be specified to include some structures
in the ground and geological boundary conditions (Juang et al., 2019). Such systems are usually time dependent
35
and spatial in nature involving, e.g., possibly changing conditions (e.g., Chow, Li, and Koh, 2019). Incorporating
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
2
uncertainties related to such conditions complicates the modelling and demands further acquisition of data. Next,
models could be accurate at reproducing data from past events but may be inadequate for unobserved outputs or
predictions, as might be the case when predicting, e.g., extreme velocities in marine turbidity currents, which are
driven by emerging and little understood soil and fluid interactions (Vanneste et al., 2019). Overlooking these
40
challenges in a geohazard assessment implies that the quantification will only reflect some aspects of the
uncertainty involved. These challenges are, unfortunately, neither exhaustively nor clearly discussed in the
geohazard literature. Options and clarifications addressing these challenges are underreported in the field , yet
analysing these challenges can be useful in treating uncertainties consistently and providing meaningful results in
an assessment. This papers aim is thus to bridge the gap in the literature by providing an analysis and clarifications
45
enabling a useful quantification of uncertainty.
It should be emphasised that, in this paper, we consider uncertainty quantification in terms of probabilities. Other
approaches to measure or represent uncertainty that have been studied by, for example, Zadeh (1968), Shafer
(1976), Ferson and Ginzburg (1996), Helton and Oberkampf (2004), Dubois (2006), Aven (2010), Flage et al.,
(2013), Shortridge, Aven, and Guikema (2017), Flage, Aven, and Berner (2018), and Gray et al. (2022a,b), will
50
not be discussed here. The discussion about the complications in UQ related to computational issues generated by
numerical approximations including, for instance, sampling procedures are also beyond the scope of the current
work.
The remainder of the paper is as follows. In Section 2, based on recent advances, we describe how uncertainty
quantification using geohazard models can be conducted. Next, some remaining challenges in UQ are identified
55
and illustrated. Options to address the challenges in UQ are discussed in Section 3. A simplified example, further
illustrating the discussion, is found in Section 4, while the final section provides some conclusions from this study.
2 Quantifying uncertainty using geohazard models
In this section, we make explicit critical steps in uncertainty quantification, UQ. We describe a general approach
to UQ that considers uncertainty as the analysts incomplete knowledge about quantities or events. The UQ
60
approach described is restricted to probabilistic analysis. Emphasis is made on the choices and assumptions usually
made by analysts.
A geohazard model can be described as follows. We consider a system (e.g., debris flow) with a set of input
quantities X (e.g., sediment concentration, entrainment rate) whose relationships to the output quantities Y (e.g.,
runout volume, velocity, or height of flow) can be expressed by a set of models . Analysts identify or specify X,
65
Y, and . A vector Θɱ (including, e.g., friction, viscosity, turbulence coefficients) parameterises a model ɱ in .
The parameters Θɱ determine specific functions among a family of potential functions modelling the system.
Accordingly, a model ɱ can be described as a multi-output function with, e.g., Y = {runout volume, velocity,
height of flow}, and we can write Eq. (1) based on Lu and Lermusiaux (2021):
ɱ: Xs, ΘɱYs,t (1)
70
ɱ (Eɱ, SGɱ, BCɱ, ICɱ) (2)
where y as realisations of Y are the model responses when elements in X take the values x at a spatial location s
S and a specific time t T, and parameters θɱ Θɱ are used. In Eq. (1), Xis the set of input quantities,
T is the time domain, S is the spatial domain, Θɱ󰡁 corresponds to a parameter vector, and
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
3
Y is the set of output quantities, with d = 0, 1, 2, or 3. The system is fully described if ɱ is specified in terms
75
of a set of equations Eɱ (e.g., conservation equations), the spatial domain geometry SGɱ (e.g., extension, soil
structure), the boundary conditions BCɱ (e.g., downstream flow), and the initial conditions ICɱ (e.g., flow at t =
t0), see Eq. (2).
In the sampling approach to uncertainty quantification, specified probability distributions reflecting analysts
uncertainty about input quantities are sampled many times, and the distribution of the produced outputs can be
80
calculated. The output probability distribution for a model ɱ can be denoted as f(y|x,θɱ,ɱ) for realisations y, x, θɱ,
ɱ of Y, X, Θɱ, and , respectively.
Betz (2017) has suggested that the parameter set is fully described by a parameter vector Θ in Eq. (3):
Θ={Θɱ,ΘX,Θε,Θo} (3)
in which, Θɱ relates to parameters of the model ɱ, ΘX are parameters linked to the input X, Θε is the vector of the
85
output-prediction error ε, and Θo is the vector associated with observation/measurement errors, when historical
records are used. More explicitly, to compute an overall joint probability distribution, we may have the following
distributions:
f(y|x,θɱ,ɱ) is the distribution of Y when X takes the values x, and parameters θɱ Θɱ and a model ɱ
are used to compute y;
90
f(x|θX,ɱ) is the conditional distribution of X given the parameters θX ΘX and the model ɱ. Note that
each ɱ defines which elements in X are to be considered in the analysis;
f(x|x
,θo) is a distribution of X given the observed quantities X
= x
and the observations/measurement error
parameters θo Θo;
additionally, one can consider f(y*|yε,ɱ), which is a distribution of Y*, the future system’s response,
95
conditioned on the model output y and the output-prediction error vector θε Θε. The output-prediction
error ε is the mismatch between the model predictions and non-observed systems responses y*. ε is used
to correct the imperfect model output y (Betz, 2017; Juang et al., 2019).
If, for example, the parameters Θɱ are poorly known, a prior distribution π(θɱ|ɱ) weighing each parameter value
θɱ for a model ɱ is usually specified. A prior is a subjective probability distribution quantified by expert judgement
100
that represents uncertainty about the parameters prior to considering information in data (Raices-Cruz, Troffaes,
and Sahlin, 2022). When some data is available, e.g., in the form of measurements ɗ = {Ŷ = ŷ, X
= x
}, as a part of
different sources of data, i.e., ɗ Ɗ, such parameter values θɱ or their distributions π(θɱ|ɱ) can be constrained by
back-analysis methods. Back analysis methods include matching experimental measurements ŷ and calculated
model outputs y using different assumed values 󰡁
󰆒, i.e., more formally, values for θɱ can be calculated as follows
105
(based on Liu et al., 2022):
θɱ = argmin[ŷ-y(x
,󰡁
󰆒)] (4)
The revision or updating of the prior π(θɱ|ɱ) with data ɗ to obtain a posterior distribution described as π(θɱ|ɗ,ɱ)
is also an option in back analysis. The updating can be calculated as follows (based on Juang et al., 2019; Liu et
al., 2022):
110
󰇛󰔿󰡁󰇜 󰇛󰔿󰇜󰇛󰡁󰇜
󰕏󰇛󰔿󰇜󰇛󰡁󰇜
(5)
where (θm|ɗ) = f(ɗ|θm) is a likelihood function, a distribution which weighs ɗ given θɱ.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
4
Similarly, we can constrain any of the distributions above, e.g., f(y|x,θɱ,ɱ), or f(x|θX,ɱ) to obtain f(y|x,θɱ,ɗ,ɱ) and
f(x|θX,ɗ,ɱ), respectively.
For a geohazard problem, it is possible to specify several competing models, e.g., distinct geological models with
115
diverse boundary conditions, see Eq. 2. If available knowledge is insufficient to determine the best model, different
models ɱ can be considered, and the respective overall output probability distribution for, e.g., uncorrelated
models is computed as (Betz, 2017; Juang et al., 2019):
f(y|x,Θ,Ɗ,) = f(y|x,θ,ɗ,ɱ)ω(ɱ|Ɗ,) (6)
f(y|x,θ,ɗ,ɱ) = ʃ f(y|x,θ,ɱ)π(θ|ɗ,ɱ)dθ (7)
120
In Eq. (6), ω(ɱ|Ɗ,) is another distribution weighing each model ɱ in .
The various models , their inputs X, parameters Θ, and outputs Y, experimental data ɗ can be coupled all together
through a Bayesian network, as has been suggested by Sankararaman and Mahadevan (2015) or Betz (2017). One
possible configuration of a network coupling elements in , X, Θ, Y is illustrated in Figure 1.
Figure 1: A configuration of a network coupling some elements in , X, Θ, Y, Y*
The previous description of a general approach to UQ considers uncertainty as that reflected in the analysts
125
incomplete knowledge about quantities or events. In UQ, to measure or describe uncertainty, subjective
probabilities can be used and constrained using historical observations ɗ. It is also explicitly shown that model
outputs are conditional on historical observations ɗ made available and models chosen by analysts including
the selection of several parameters Θ and initial and boundary conditions, BCɱ and ICɱ. Based on the above
description, in the following, we analyse some of the challenges that arise when conducting UQ.
130
As mentioned, back-analysis methods help constrain some elements in Θ; however, given the considerable number
of parameters (see Eq. 1-3) and data scarcity, constraining Θ is often only achieved in a limited fashion. Back-
analysis is further challenged by the potential dependency among Θ or and between Θ and SGɱ, BCɱ, ICɱ. We
also note that, back analysis, or more specifically, inverse analysis, faces problems regarding non-identifiability,
non-uniqueness, and instability. Non-identifiability occurs when some parameters do not drive changes on the
135
inferred quantities. Non-uniqueness arises because there may be more than one set of fitted or updated parameters
that adequately reproduce observations. Instability in the solution arises from errors in observations and the non-
linearity of models (Carrera and Neuman, 1986). Alternatively, in specifying, e.g., a joint distribution f(x,θ) to be
sampled, analysts may consider the use of, e.g., Bayesian networks (Albert, Callies, and von Toussaint 2022).
However, under the usual circumstance of a lack of information, establishing such a joint distribution is
140
challenging and requires, in many instances, that analysts encode many additional assumptions (e.g., prior
distributions, likelihood functions, independence, linear relationships, normality, stationarity of the quantities and
X
θo
X
θX
Y
θε
Y*
ɱ:x˟θɱ
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
5
parameters considered), see e.g., Tang, Wang, and Li (2020); Sun et al. (2021b); Albert, Callies, and von Toussaint
(2022); Pheulpin, Bertrand, and Bacchi (2022). A more conventional choice is that x or θ are specified using the
maximum entropy principle in an attempt to specify the least biased distributions possible on the given information
145
(Jaynes, 1957). Such distributions are subject to the system's physical constraints based on some available data.
The information entropy of a probability distribution measures the amount of information contained in the
distribution. The larger the entropy, the less information is provided by the distribution. Thus, by maximising the
entropy over a suitable set of probability distributions, one finds the distribution that is least informative in the
sense that it contains the least amount of information consistent with the system's constraints. Note that a
150
distribution is sought over all the candidate distributions subject to a set of constraints. This principle has been
questioned, since its validity and usefulness lie in the proper choice of physical constraints (Jaynes, 1957; Yano
2019). Doubts are also raised regarding the potential information loss when using the principle. Analysts usually
strive in making use of all knowledge available and avoiding unjustified information loss (Christakos, 1990; Flage,
Aven, and Berner, 2018).
155
Options to address the parametrisation challenge also include surrogate models, parameters reduction, and model
learning (e.g., Lu and Lermusiaux, 2021; Sun et al., 2021b; Albert, Callies, and von Toussaint, 2022; Degen et al.
2022; Liu et al., 2022). Surrogate models are learnt in order to replace a complicated model with an inexpensive
and fast approximation. Parameters reduction is achieved based on either principal component analysis or global
sensitivity analysis to determine which parameters significantly impact model outputs and are essential to the
160
analysis (Degen et al., 2022; Wagener, Reinecke, and Pianosi, 2022). Remarkably, versions of the model learning
option do not need any prior information about model equations Eɱ but require local verification of conservation
laws in the data ɗ (Lu and Lermusiaux, 2021). These approaches still require large past data sets, sourced
systematically, which is a frequent limitation in geohazard assessments. More importantly, however is that, like
many models, even those exhaustively validated, the credibility of unobserved surrogate model outputs can always
165
be questioned, since, for instance, records may miss crucial events or models may fail to reproduce outputs caused
by recorded abrupt changes (e.g., extreme velocities of turbidity currents) (Alley, 2004; Woo, 2019). An additional
point is the issue of incomplete model response, which refers to a model not having a solution for some
combinations of the input variables (Cardenas, 2019; van den Eijnden, Schweckendiek, and Hicks, 2021).
In bypassing the described challenges when quantifying uncertainty, simplifications are usually enforced,
170
sometimes unjustifiably, in the form of assumptions, denoted here by the set Ą, and this set can include one or
more of the assumptions listed in Table 1. Note that the set of assumptions can be increased with those assumptions
imposed by the use of specific models (e.g., conservation of energy, momentum or mass, Mohr-Coulombs
failure criterion).
Table 1. Some enforced assumptions in UQ for geohazard assessments
Predictions (non-observed outputs) of Y are credible, despite models only reproducing responses based on
historical data ɗ={Ŷ=ŷ,X=x
}, ɗ Ɗ.
A model has a solution for any combination of the input quantities X.
Elements in X are fully specified.
Elements in X are mutually independent.
The joint distribution f(x,θ) distributes according to the maximum entropy principle.
If measurements are available, some input quantities X are set to specific values x= x
.
Input quantities X are set to constant values x0, that is X=x0.
Some θ are set to specific point values and are mutually independent.
Some θ are independent of SGɱ, BCɱ, ICɱ.
SGɱ, BCɱ, ICɱ are set to be constant.
When some data ɗ is available in the form of measurements , x
}, likelihood functions [ɱ(θ|ɗ)] are mutually
independent.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
6
3 Addressing the challenges in uncertainty quantification
175
From the previous section, we saw that it is very difficult in geohazard assessments to meet data requirements for the
ideal parameterisation of models. Further, we have noted that, although fully parameterised models could potentially be
accurate at reproducing data from past events, these may turn out to be inadequate for unobserved outputs. We also made
explicit that predictions are not only conditional on Θ (including priors, likelihood functions and linked
hyperparameters) but possibly also on SGɱ, BCɱ, ICɱ, see Eq. (1-7). Ultimately, assumptions made also condition
180
model outputs. More importantly, note that when only some model quantities or parameters can be constrained by data
ɗ, the modelling will only reflect some aspects of the uncertainty involved. If the above challenges remain unaddressed,
UQ lacks credibility. To address such challenges and provide increased credibility, clarifications and
reinterpretation of some fundamental concepts together with practical simplifications may be required, and these
are discussed in the following. Table 2 shows the major challenges found and how they are addressed in related
185
literature, while in Table 3 some clarifications or considerations put forward by us are displayed. The discussion
in this section builds on previous analysis by Aven and rn (1998), Apeland, Aven, and Nilsen (2002), Aven and
Kvaløy (2002), Nilsen and Aven (2003), Aven and Zio (2013), Khorsandi and Aven (2017), and Aven (2019).
Table 2. Major challenges and options to address them in geohazard assessments
Challenges, CH
Options to address the challenges, O
Challenges related to the model outputs and system responses
CH1. Predictions of y lack credibility since these are
model outputs not recorded in the data ɗ,
CH2. A model does not have a solution for a feasible
combination of the input quantities X.
O1. Credibility of predictions is judged in terms of physical
consistency checks (Wagener, Reinecke, and Pianosi, 2022) and
by examining the ability of models to reproduce disruptive
changes recorded in the data (Alley, 2004).
O2. Predictions by Bayesian forecasting methods. Based on a prior
distribution for y, a posterior distribution of y is obtained by
including the information provided by the model prediction in the
form of model likelihood (Montanari and Koutsoyiannis, 2012).
Challenges related to input quantities
CH3. Data available ɗ may not include all the historic
crucial events or disruptive changes,
CH4. Some input quantities X remain unknown
(unidentified) to analysts during an assessment,
CH5. The distribution f(x) or the bounds of x are
unknown,
CH6. Some input quantities X may be mutually
dependent.
O3. Using the maximum entropy principle to specify the
distributions based on the choice of constraints regarding the
physics of the phenomena involved. Constraining the distributions
by data including data other than measurements to reduce
unjustified information loss (Jaynes, 1957; Christakos, 1990; Betz,
2017; Yano 2019).
O4. Counterfactual analysis in which alternative events to
observed facts including disruptive changes are ʻimaginedʼ,
assumed, and explored to obtain alternative system’s responses
using models (Pearl, 1993; Woo, 2019).
O5. Exhaustive investigation of input uncertainty using the
assumptions deviation approach to specify input distributions
(Aven, 2013).
Challenges related to the parameters and models
CH7. The distribution f(θ) or the bounds of θ are
unknown,
CH8. Some θ may be dependent on SGɱ, BCɱ, or ICɱ,
CH9. Likelihood functions [ɱ(θ|ɗ)] may be mutually
dependent,
CH10. Models ɱ in may be mutually correlated.
O3. Using the maximum entropy principle as described above.
O6. A joint distribution of Θ, SGɱ, BCɱ, ICɱ, X for each ɱ, can be
specified by encoding other assumptions (e.g., prior distributions,
likelihood functions, independence, linear relationships,
normality, stationarity) in Bayesian networks (Albert, Callies, and
von Toussaint, 2022).
O7. Using surrogate models, parameters reduction, and model
learning (Lu and Lermusiaux, 2021; Albert, Callies, and von
Toussaint, 2022).
190
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
7
Table 3. Some clarifications and considerations to address the challenges in UQ
C1. Uncertainty reflects that analysts’ knowledge about quantities or events is incomplete.
C2. Models are simplifications, mainly used for understanding the performance of the system and approximating its
responses; they are part of the knowledge of the system, and they do not introduce uncertainty.
C3. The focus is on the quantification of the uncertainty of the system responses rather than on the accuracy of a model
reproducing recorded data.
C4. Predictions are conditional on the model(s) chosen together with a number of assumptions made by analysts.
C5. The specification of the joint distribution f(x,θ) cannot solely rely on the use of maximum entropy principle, but on the
full scrutiny of background knowledge Қ, and such a distribution is better specified using knowledge-based probabilities.
C6. Some elements in the parameter set Θ are not quantities or properties of the system as such, and there could not be
uncertainty about them.
C7. Analysts may choose a model or a set of them, which are believed or judged to be the best credible models.
Among the clarifications, we consider a major conceptualisation suggested by the literature, which is the definition
of uncertainty. Uncertainty refers to incomplete information or knowledge about a hypothesis, quantity, or the
occurrence of an event (Society for Risk Analysis, 2018). In Table 3, we denote this clarification as C1. Embracing
195
this definition has some implications for uncertainty quantification using geohazard models. We make use of these
implications to address the major complications and challenges found. For instance, if uncertainty is measured in
terms of probability, one such implication is that analysts are discouraged to use the so-called frequentist
probabilities due to the fact these probabilities do not measure uncertainty or lack of knowledge. Rather such
probabilities reflect frequency ratios representing fluctuations or variation in the outcomes of quantities.
200
Frequentist probabilities are of limited use because these assume that quantities vary in large populations of
identical settings, a condition which can hardly be justified for rather few geohazard quantities due to both the
often one-off nature of many geohazard features and the impossibility to verify or validate data by, e.g., a large
number of repeated tests. Thus, considering the usual constraints in data, as well as the nature of geohazard events,
a more meaningful and practical approach can be suggested to actually measure uncertainty, namely the use of
205
knowledge-based (also referred to as judgemental or subjective) probabilities (Aven 2019). A knowledge-based
probability is understood as an expression of the degree of belief in the occurrence of an event or quantity by a
person assigning the probability and conditional on the available knowledge Қ. Such knowledge Қ includes not
only data in the form of measurements ɗ made available to the analysts at the time of the assessment, but also other
data sources in Ɗ together with the models chosen by the analysts for the prediction, as well as the modelling
210
assumptions Ą made by those analysts. Accordingly, we have in total that, to describe uncertainty about, e.g., x or
y, probabilities are assigned based on Қ and, therefore, we acknowledge that those probabilities are conditional on
Қ. In the previous section, we have made explicit the conditional nature of uncertainty on measured data ɗ Ɗ
and models and wrote, explicitly, for the overall output probability distribution, the expression f(y|x,Θ,Ɗ,),
(see Eq. 6). If assumptions Ą are also acknowledged as conditioning uncertainty, we write more explicitly
215
f(y|x,Θ,Ɗ,,Ą) or equivalently f(y|x,Θ,Қ). We can therefore write:
f(y|x,Θ,Қ) = f(y|x,Θ,Ɗ,,Ą) (8)
The meaning of this expression can be elucidated next. If, in a specific case, we would write f(y|x,θ,Ɗ), it means
that Ɗ summarises all the knowledge that analysts have to calculate y given (realised or known) x and θ. The full
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
8
expression in Eq. 8 implies, accordingly, that to calculate y, and when knowing x and θ, the background knowledge
220
includes Ɗ, , Ą. Note that Қ can also be formed by observations, justifications, rationales, and arguments, thus
Eq. 8 can be further detailed to include these aspects of Қ. Structured methods exist to assign knowledge-based
probabilities (see, e.g., Apeland, Aven, and Nilsen, 2002; Aven 2019). Here we should note, however, that since
models form part of the available background knowledge Қ, models can also inform these knowledge-based
probability assignments. It follows that, based on knowledge-based input probabilities, an overall output
225
probability distribution calculated using models is also of subjective character or knowledge based, although
uncertainty quantification using sampling could provide some frequency resemblance or interpretation (Jaynes,
1957). Some of the implications of using knowledge-based probabilities are described throughout this section.
According to the left column in Table 2, the focus of the challenges relates to the model outputs, more specifically
predictions (CH1 and CH2), input quantities (CH3-CH6), parameters (CH7-CH9), and models (CH10). We recall
230
that as defined, uncertainty quantification helps determine the systems response uncertainty based on specified
input quantities using models. Accordingly, the focus of an assessment is on the potential systems responses.
When it comes to quantifying uncertainty, the focus is often on uncertainty about future non-observed responses
Y*, which are approximated by the model output Y considering some specified input quantities X. We recall that
then Y* and X* are quantities that are unknown at the time of the analysis, but will, if the system being analysed
235
is implemented, take some value in the future, and possibly become known. Thus, during an assessment Y* and
X* are the uncertain quantities of the system since we have incomplete knowledge about Y* and X*. Based on
these grounds, the output-prediction error ε (described in the previous section) being the mismatch between the
model prediction values y and non-observed systems responses values y*, usually suggested to correct the
imperfect model output y, can only be specified on the basis on the scrutiny of Қ.
240
Another consequence of considering the definition of uncertainty put forward, which solely links uncertainty to
quantities or events, and taking into account that models are only mathematical artefacts, is that models, as such,
are not to be linked to uncertainty. Models, per se, do not introduce uncertainty, but they are likely to be inaccurate.
Accordingly, another major distinction is to be set in place. We recall that models, by definition, are
simplifications, approximations of the system being analysed; they express or are part of the knowledge of the
245
system and should therefore be solely used for understanding the performance of the system rather than for illusory
perfect predictions. In Table 3, we denote the latter clarification as C2.
Regarding the challenges CH1 and CH2, we should note that geohazard analysts are often more interested in
predictions rather than known system outputs. For instance, predictions are usually required to be calculated for
input values that are not contained in the validation data. We consider that predictions are those model outputs not
250
observed or recorded in the data, i.e., extrapolations out of the range of values covered by observations (e.g.,
mudflow extreme velocities). Thus, the focus is on the quantification of the uncertainty of extreme systems
responses rather than on the accuracy of a model reproducing recorded data. This is the clarification C3 in Table
3. Considering this, models are yet to provide accuracy in reproducing observed outputs but, more importantly,
afford credibility in predictions. Such credibility is to be assessed, mainly, in terms of judgements since
255
conventional validation cannot be conducted using non-observed outputs. Recall that model accuracy usually
relates to the comparison of model outputs with experimental measurements (Roy and Oberkampf, 2011; Aven
and Zio, 2013) and is the basis for validating models. Regarding credibility of predictions, Wagener, Reinecke,
and Pianosi (2022) have reported that such credibility can be mainly judged in terms of the physical consistency
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
9
of the predictions, by checks rejecting physically impossible representations of the system. The credibility of
260
predictions may also include the verification of the ability of models to accurately reproduce disruptive changes
recorded in the data (Alley, 2004). However, as we have made explicit in the previous section, models’ predictions
are conditional on a considerable number of critical assumptions and choices made by analysts (see Table 1 and
clarification C4 in Table 3). Therefore, predictions can only be as good as the quality of the assumptions made.
The assumptions could be wrong, and the examination of the impact of these deviations on the predictions must
265
be assessed. To provide credibility of predictions, such assumptions and choices should be justified and scrutinised,
ref. option O5 in Table 2. The option O5 addresses the challenge CH1; however, when conducting UQ using
models, O5 has a major role when investigating input uncertainty, which is discussed next.
A critical task in uncertainty quantification is the quantification of input uncertainty. As shown in Table 2, such
uncertainty is not only to be associated with the lack of knowledge of the distribution f(x) or bounds of x, CH5,
270
and some other sources are to be considered. For example, input uncertainty may originate from the situation that
historic crucial events or disruptive changes are missing in the records, CH3, or from the condition in which critical
quantities in X may remain unidentified to analysts during an assessment, CH4. We recall that analysts can
unintendedly fail to identify relevant elements in X due to insufficiencies in data or limitations of existing models.
For example, during many assessments, trigger factors that could bring a soil mass to failure could remain unknown
275
to analysts (e.g., Hunt et al. 2013; Clare et al. 2016; Leynaud et al. 2017; Casalbore et al. 2020). Uncertainty
quantification based on models requires simulating sampled values from X, and elements in X can possibly be
mutually dependent; however, the joint distribution of X, namely f(x), is often also unknown. This is the challenge
CH6. Considering the potential challenges CH3 to CH6, to specify f(x) we cannot solely rely on the use of the
maximum entropy principle, as described in the previous section, since it may fail to advance an exhausted
280
uncertainty quantification in the input, e.g., by missing relevant values not recorded in the measured data. This
would undermine the quality of predictions and therefore uncertainty quantification. Recall that the principle
suggests the use of the least informative distribution among candidate distributions constrained solely on
measurements. Using counterfactual analysis, as described in Table 2, is an option, but this will also fail in
providing quality in predictions, since this analysis focuses on the analysis of counterfactuals (a part of the data
285
Ɗ) and little on the overall knowledge available Қ. Note that systems knowledge Қ includes, e.g., among others,
the assumptions made in the quantification of uncertainty, such as those shown in Table 1. Further note that such
assumptions not only relate to data, but also to input quantities, modelling, and predictions. Thus, it appears that
the examination of these assumptions should, desirably, be at the core in quantifying uncertainty in geohazard
assessments, as suggested in Table 2, option O5. The assessment of deviations of assumptions has been suggested
290
originally by Aven (2013) and Khorsandi and Aven (2017). An assumption deviation risk assessment evaluates
different deviations, their associated probabilities of occurrence, and the effect of the deviations. The major
distinctive features of the assumption deviation risk assessment approach are the evaluation of the credibility of
the knowledge Қ supporting the assumptions made, together with the questioning of the justifications supporting
the potential for deviations. The examination of Қ can be achieved by assessing the justifications for the
295
assumptions made, the amount and relevance of data or information, the degree of agreement among experts, and
the extent to which the phenomena involved are understood and can be modelled accurately. Note that justifications
might be in the form of direct evidence becoming available, indirect evidence from other observable quantities,
supported by modelling results, or possibly inferred by assessments of deviations of assumptions. This approach
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
10
is succinctly demonstrated in the following section. Accordingly, we suggest specifying f(x) in terms of
300
knowledge-based probabilities in conjunction with the investigation of input uncertainty using the assumptions
deviation approach. This is identified as consideration C5 in Table 3.
Another point to consider is that, when uncertainty reflects analysts (lack of) knowledge about quantities or events
and is measured in terms knowledge-based probabilities, analysts should be aware that conditionality among
elements in X only implies that increased knowledge about, e.g., a quantity X1 will change the uncertainty about
305
another quantity X2, if X2 is conditional on X1. The expression that denotes this is conventionally written as X2|X1.
This interpretation may be exploited by analysts when specifying, e.g., the joint distribution f(x,θ). Analysts, for
example, may simplify the analysis when, according to the scrutiny of Қ, increased knowledge about, e.g., X1 will
not result in increased knowledge about another quantity, e.g., X2, and, for example, if a distribution f(y|x1,x2) is to
be specified, we may have then that f(y|x1,x2) reduces to f(y|x1)f(y|x2), according to probability theory. Apeland,
310
Aven, and Nilsen (2002) have illustrated how conditionality in the setting of knowledge-based probabilities can
inform the specification of a joint distribution.
The parameterisation problem, which involves the challenges CH7 to CH9 in Table 2, warrants exhaustive
consideration. Addressing these challenges also requires some reinterpretation. To start, we recall that, parameters,
by definition, are coefficients determining specific functions among a family of potential functions modelling the
315
system. Those parameters, eventually, constrain a model’s responses. Recall that, y as realisations of Y are the
model responses when X takes the values x and some parameters θ Θ and models ɱ  are used. Thus, as
shown in the previous section, any output y is conditional on θ and so is the uncertainty attached to y. Further note
that, taking into consideration the definition of uncertainty previously stated, which attributes uncertainty to events,
quantities, or hypothesis solely, we may say that, unless a parameter is associated with a quantity or property of
320
the system, non-quantity parameters, being merely artefacts in the models, are not to be linked to any uncertainty.
This is identified as clarification C6 in Table 3. Using the uncertainty definition previously put forward and looking
at Eq. 3, analysts may consider that, for instance, parameters θε, which are linked to the output-prediction error ε,
some model parameters in Θɱ, the vector associated with observation/measurement errors Θo and the overall
attached hyperparameters linked to probability distributions (including priors, likelihood functions), are not
325
quantities or properties of the system as such; they are modelling artefacts, and therefore, it is questionable to
consider uncertainty about them. Thus, focused on the uncertainty of the system responses rather than model
inaccuracies, the analysts may consider strictly assigning uncertainty to those parameters that represent physical
quantities and fix with single values those that are not quantities as such. To help identify those parameters to
which some uncertainty can be linked, we can scrutinise, e.g., the physical nature of these. Note, however, that in
330
fixing parameters to a single value, we still can make use of back analysis procedures, as mentioned previously.
Analysts may have some additional basis to specify parameters values when the background knowledge available
Қ is scrutinised to verify that Қ, including not only data measurements but other sources of data, models,
assumptions made, and arguments, strongly supports a specific parameter value. Based on this interpretation,
setting the values of these non-quantity-based parameters to a single value reduces, considerably, the complications
335
in quantifying uncertainty. It also follows that, analysts are encouraged to make explicit that model outputs are
conditional on these fixed parameters, as well as on the model or models chosen, as we have shown in the previous
section. The latter interpretation also leads us to argue that the focus of the quantification is on the uncertainty of
the system response rather than the inaccuracies of the models. This, of course, implies that, in a practical sense
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
11
and in the context of geohazard assessments, once a clear differentiation between parameters and input quantities
340
is made, and a model or a set of them believed or judged to be the best models providing the most credible
predictions are chosen, uncertainty quantification can then proceed. This parsimonious modelling approach is
identified as consideration C7 in Table 3. This latter consideration addresses, to an extent, the challenge CH10.
In the following section, we further illustrate the above discussion by analysing a documented case in which UQ
in a geohazard assessment was informed by modelling using sampling procedures.
345
4 Case analysis
To further describe the proposed considerations, we analyse a case reported in the specialised literature. The case
deals with the quantification of uncertainty of geological structures, namely uncertainty about the subsurface
stratigraphic configuration. Conditions in the subsurface are highly variable, whereas site investigations only
provide sparse measurements. Consequently, subsurface models are usually inaccurate. At a given location,
350
subsurface conditions are unknown until accurately measured. Soil investigation at all locations is usually
impractical and uneconomical, and point-to-point condition variation cannot therefore be known (Vanmarcke,
1984). Such uncertainty means significant engineering and environmental risk to, e.g., infrastructure built on the
surface. One way to quantify this uncertainty is calculating the probability of every possible configuration of the
geological structures (Tacher et al. 2006; Thiele et al. 2016; Pakyuz-Charrier et al. 2019). Sampling procedures
355
for UQ are helpful in this undertaking. We use an analysis and information by Zhao et al. (2021), which refer to a
site located in the Central Business District, Perth, Western Australia, where 6 boreholes were executed. The case
has been selected taking into account its simplicity to illustrate the points of this paper, but at the same time, it
provides details to allow some discussion. Figure 2 displays the system being analysed.
Figure 2: Borehole logs in colours and longitudinal section reported by Zhao et al. (2021) located in the Central Business
District, Perth, Western Australia. The records correspond to information of the six boreholes. Three types of materials
are revealed by the boreholes including sand (yellow), clay (magenta), and gravel (blue).
In the system under consideration, a particular material type to be found in a non-bored point, a portion of terrain
360
that is not penetrated during soil investigation, is unknown and thus uncertain. The goal is, therefore, the
computation of the probability of encountering a given type of soil in these points. In Zhao et al. (2021) the focus
is on calculating the probabilities of encountering clay in the subsurface, and the approach advocated was a
sampling procedure to generate many plausible configurations of the geological structures and evaluating their
4
5
6
Vertical direction (metres)
Horizontal direction (metres)
40
60
80
0
-10
-20
-30
0
20
1
2
3
y=?
y=?
y=?
y=?
y=?
One of many potential
configurations of the clay
SGm
BCm=?
X
θh=?
θv=?
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
12
probabilities based on their frequencies. To calculate the probability of encountering a given type of soil c, p(y=c),
365
in a non-penetrated point in the ground, Zhao et al., (2021) used a function that depends on two correlation
parameters, namely the horizontal and vertical scale of fluctuation θh and θv. Note that spatial processes and their
properties are conventionally assumed as spatially correlated. Such spatial variation may presumably be
characterised by correlation functions, which depend on a scale of fluctuation parameter. The scale of fluctuation
measures the distance within which points are significantly correlated (Vanmarcke, 1984). Eq. 9 describes the
370
basic components of the model chosen by Zhao et al., (2021) (specific details are given in the Appendix to this
paper):
ɱ: X ΘɱYs →p(y=c) (9)
where X is the collection of all quantities at borehole-points sx which can take values x from the set {sand, clay,
gravel}, according to the setting in Figure 2. Y is the collection of all quantities with values y at non-borehole
375
points sy. The values y together with the values x are sampled and probabilities computed on the basis of a chosen
model using the parameters θh=11,1 and θv=4,1 metres, θh, θv Θɱ. Using the maximum likelihood method, the
parameters were determined based on the borehole data revealed at the site. In the determination of parameters,
the sampling of uniform and mutually independent distributions of θh and θv was the procedure advocated. The
system is further described by a set of equations Eɱ (a correlation function and a probability function), the spatial
380
domain geometry SGɱ (a terrain block of 30 x 80 metres), and the boundary conditions BCɱ (the conditions at the
borders). More details are given in the Appendix to this paper. Since this system is not considered time dependent,
the initial conditions ICɱ were not specified.
The summary results reported by Zhao et al. (2021) are shown in Figure 3. In Figure 3, the most probable
stratigraphic configuration along with spatial distribution of the probability of the existence of clay is displayed.
385
The authors focused their attention on this sensitive material, which likely represents a risk to infrastructure built
on the surface.
(a)
(b)
Figure 3: Zhao et al. (2021) findings shown in their Figure 9. (a) Most probable stratigraphic configuration. (b)
Spatial distribution of the probability of the existence of clay. Authorisation of reproduction: License number
5351891245835. Distances in metres.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
13
Zhao et al., (2021) stated that characterisation results of the stratigraphic configuration and its uncertainty are
consistent with the intuition and the state of knowledge on site characterisation”. Next, throughout Zhao et al.’s
390
(2021) analysis the following assumptions were enforced (Table 4), although these were not explicitly disclosed
by the authors.
Table 4. Assumptions enforced by Zhao et al., (2021)
Predictions (non-observed outputs) are credible.
Likelihood functions [ɱ(θ|ɗ)] were set to be mutually independent.
For the determination of parameters and model, f(θhv) distributes according to the maximum entropy principle
and θh and θv are mutually independent.
Specified elements X are complete.
Input quantities X were set to the measured values, i.e., x = x
(no inaccuracies in data).
θ are independent of SGɱ and BCɱ.
SGɱ, BCɱ were set to be constant.
Unfortunately, the authors did not report enough details on how the majority of these assumptions are justified.
We, however, should note that providing these justifications was not the objective of their research. Yet, we analyse
395
here how these can be justified by both scrutinising the supportive knowledge Қ and using some elements of the
assumption deviation approach, described in the previous section. Table 5 summarises the analysis conducted and
only reflects the most relevant observations and reservations identified by us. Accordingly, the information in
Table 5 may not be exhaustive but still useful for the desired illustration. Table 5 displays some observations by
us related to the credibility of the knowledge Қ. The examination of Қ is achieved by assessing the amount and
400
relevance of data or information, the extent to which the phenomena involved are understood and can be modelled
accurately, the degree of agreement among experts, and the justifications for the assumptions made. Observations
regarding the justifications for potential deviations of assumptions also form the analysis.
Not surprisingly, the observations in our analysis are concentrated on the predictions credibility. Recall that the
focus of UQ is on the systems response, which is approximated by model predictions (considerations C2 and C3
405
in Table 3). We note, for example, that, although the use of correlations is an accepted practice and a practical
simplification, correlation functions appear to be counterintuitive to model geological structures or domains and
do not help much in understanding the system (consideration C2 in Table 3). Recall that such structures are mainly
disjoint domains linked to a finite set of possible categorical (masses of soil or rock) rather than continuous
quantities. Next, the variation of such structures can occur by abrupt changes of materials, thus the use of, for
410
example, smoothed functions, as correlation functions may be, to represent them requires additional consideration.
Further, the physical basis of the correlation functions is not clear and physical models based on deposition
processes may be suggested (e.g., Catuneanu et al., 2019). We should further note that a potential justification for
the deviation of the assumption regarding the credibility of predictions is that knowledge from additional sources
such as surface geology, sedimentology, local geomorphic setting, and structural geology was not explicitly taken
415
into account in the quantification of uncertainty. The revision of this knowledge can contribute to reduce the
probability of the deviation in predictions. Based on the observations in Table 5, we can conclude that there is
potential to improve the credibility of predictions.
420
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
14
Table 5. Examination of supporting knowledge Қ and justifications for the potential deviation of assumptions (part a)
The amount and relevance of
data or information
The extent to which the phenomena
involved are understood, and
accurate models exist
The degree of agreement
among experts
Justifications for the
assumptions made
Justifications supporting the
potential deviations
Assumption: Predictions of Y are credible
The analysis is only based
on borehole information;
however, such investigation is
exceptionally exhaustive, 6
boreholes
The physical basis for using
correlations is dubious and models based
on the deposition process can be
considered
The use of correlations is
an accepted practice in the
field
Exhaustive borehole
information
The knowledge of surface
geology, sedimentology, local
geomorphic setting, and
structural geology was not
explicitly incorporated into
UQ
Variation of geological structures can
occur by abrupt changes, thus the use of
smoothed functions to represent them
requires additional consideration
Global rather than local correlation
between spatial quantities has been used,
possibly misrepresenting geological
structures variation
Assumption: Likelihood functions [ɱ(θ|ɗ)] were set to be mutually independent
Assumption: f(θhv) distributes according to the maximum entropy principle and θh and θv are mutually independent
Data of the six boreholes has
been used to calibrate the
model chosen; however,
knowledge of surface geology,
sedimentology, local
geologic/geomorphic setting,
and structural geology was not
explicitly incorporated into the
analysis
Based on the maximum
likelihood method, a model
judged to be the best model
was chosen
Dependency between θh
and θv, cannot be supported
by general knowledge and
such dependency hardly can
be enforced
An increased revision of Қ,
could have been useful to
specify f(x) and f(θ) providing
a richer information than that
suggested by the maximum
entropy principle, regarding
how X or Θ take values
A joint distribution
f(x,θ,sgɱ,bcɱ) could have
been investigated when
determining the models and
parameters, but establishing
such a joint distribution is
challenging and requires, in
many instances, that
analysts encode many
additional assumptions,
which hardly can be justified
14
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
15
Table 5. Examination of supporting knowledge Қ and justifications for the potential deviation of assumptions (part b)
The amount and relevance of
data or information
The extent to which the phenomena
involved are understood, and
accurate models exist
The degree of agreement
among experts
Justifications for the
assumptions made
Justifications supporting the
potential deviations
Assumption: Specified elements X are complete
Input quantities were
considered fully specified
Another type of material
was revealed by other
soundings in the area, i.e., silt
(see sources in Zhao et al.
2021), and depending on the
revision of Қ
, this fourth
material could be considered.
Assumption: Input quantities X were set to measured values x=x
Errors during surveys may
have resulted in horizontal
positioning inaccuracies
Data was judged to be
accurate
There is usually not data
basis to calculate these
errors
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
16
The choices made by Zhao et al., (2021) regarding the use of parameters with fixed values together with the choice
425
for a single best model can be highlighted and illustrate the points raised in the considerations C6 and C7 (Table
3), respectively. The maximum likelihood method supported these choices, a back analysis method focused on the
matching of measurements and calculated model outputs using different assumed values for θh and θv. We highlight
that, a model judged to be the best model was chosen. This includes the specification of a particular spatial domain
geometry SGɱ. Investigating the impact of variation of SGɱ was considered unnecessary. There was no need to
430
specify several competing models, which is line with our consideration labelled as C7 in this paper.
Although Zhao et al. (2021) investigated the joint distribution f(x), which was sampled to calculate probabilities,
someone can suggest that the joint distribution f(x,θ,sgɱ,bcɱ) could have been produced when determining the
models and parameters for calculating probabilities. Nevertheless, we can argue that establishing such a joint
distribution is challenging and requires, in many instances, that analysts encode many additional assumptions (e.g.,
435
prior distributions, likelihood functions, independence, linear relationships, normality, stationarity of the quantities
and parameters considered).
A more crucial reservation derived from the analysis of potential deviations of assumptions, which might impact
considerably the credibility of predictions, comes from revisiting the knowledge sources of Zhao et al.’s (2021)
analysis, available from https://australiangeomechanics.org/downloads/. Another type of sensitive material was
440
revealed by other soundings in the area, more specifically, silt. Depending on the revision of Қ, this fourth
suspected material could be analysed in an extended uncertainty quantification of the system. Note that, originally,
the input quantities X were assumed to take values x from the set {sand, clay, gravel}. Such an assumption was
based on the records of six boreholes which were believed accurate. The latter illustrates the relevance of the
consideration C5 in Table 3.
445
Another interesting choice made by Zhao et al. (2021) is that they disregarded the possibility of incorporating into
the modelling measurement errors in the borehole data, probably because this data was judged to be accurate. We
recall in this respect that these errors reflect the inaccuracy of the ground model rather than the uncertainty about
the system. We further note that, as stated for consideration C6 (Table 3), we can hardly justify attaching
uncertainty to measurement error parameters, since measurement errors are not a property of the system. The same
450
can be said for the parameters θh and θv, which are not quantities of the system. Note that their physical basis is
questioned. We should note, however, that assuming coefficients for the parameters θh and θv is an established
practice (Vanmarcke, 1984, Lloret-Cabot et al., 2014, Juang et al., 2019). We also point out that uncertainty
quantification in this kind of systems is to an extent sensitive to the choice of scale of fluctuation values
(Vanmarcke, 1984), and that the use of a global rather than local correlation between spatial quantities can
455
potentially misrepresent geological structures variation. Accordingly, a further examination of the existing Қ can
justify undertaking an assessment of the impact of a deviation of assuming a local rather than global scale of
fluctuation.
Overall, the Zhao et al. (2021) analysis is to an extent based on the previously suggested definition of uncertainty,
ref. the consideration C1 in Table 3.
460
We should stress that the Zhao et al. (2021) uncertainty quantification refers specifically to the ground model
described at the beginning of this section. In other words, the probabilities displayed in Figure 3b are conditional
on the parameters chosen (θh=11,1 and θv=4,1 metres); the model selected (described by the Eq. 9, A-1 and A-2 in
the Appendix to this paper); the specified spatial domain geometry SGɱ (a terrain block of 30 x 80 metres); and
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
17
ultimately the assumptions made (listed in Table 4). This information is to be reported explicitly to the users of
465
the results. This reflects the clarification C4 in Table 3.
Regarding the consideration of subjective probabilities, there has been agreement, to an extent, on their use in this
kind of UQ since Vanmarcke (1984). However, the use of knowledge-based probabilities in the extension
described here is recommended given the illustrated implications to advance UQ (as discussed in the previous
section and stated in consideration C5). For example, increased examination of Қ, might have resulted in using a
470
more informative distribution f(θh,,θv) other than the uniform distribution, and in turn, different values for θh and
θv, as well as a different model. Recall that the selection of the model and determination of parameters were based
on the maximum likelihood method, which only makes use of measured data ɗ. Note, however, that when Zhao et
al. (2021) calculated the probabilities, they sampled an improved joint distribution f(x) using the parameters θh and
θv and a chosen model.
475
In our analysis of Zhaoʼs et al. (2021) assessment, the examination of supporting knowledge Қ resulted essentially
in:
(i) judging the credibility of predictions;
(ii) providing justifications for undertaking an assessment of assumption deviations considering, e.g., further
analysis involving the modelling of a fourth material;
480
(iii) considering additional data other than the borehole records, such as surface geology, sedimentology, local
geomorphic setting, and structural geology;
(iv) analysing the possibility of distinct geological models with diverse spatial domain geometry and local
correlations; and
(v) ultimately, further examining the existing Қ.
485
5 Conclusions
In this paper, we have discussed challenges in uncertainty quantification, UQ, based on models, for geohazard
assessments. Beyond the parameterisation problem, the challenges include how to assess the quality of predictions
required in the assessments, the quantification of uncertainty in the input quantities, the consideration of the impact
of choices and assumptions made by analysts. Such challenges arise from the common-place situation of limited
490
data and the one-off nature of geohazard features. If these challenges are kept unaddressed, UQ lacks credibility.
Here, we have formulated seven considerations that may contribute to providing increased credibility in the
quantifications. For example, we proposed understanding uncertainty as lack of knowledge, a condition that can
only be attributed to quantities or events in the system under consideration. Another consideration is that the focus
of the quantification should be more on the uncertainty of the system response rather than the accuracy of the
495
models used in the quantification. We drew attention to the clarification that models, in geohazard assessments,
are simplifications used for predictions approximating the systems responses. We have also considered that since
uncertainty is only to be linked to properties of the system, models, as such, do not introduce uncertainty.
Inaccurate models can, however, produce poor predictions, and under these circumstances, the uncertainty about
system response is to be judged as large. Such models should be rejected, and increased examination of background
500
knowledge will be required to credibly quantify uncertainty. We also put forward that there could not be
uncertainty about those elements in the parameter set that are not quantities or properties of the system. The latter
also has pragmatic implications concerning, e.g., how the many parameters in a geohazard system could be
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
18
constrained in a geohazard assessment.
We went into details to show that predictions, and in turn UQ, are conditional on the model(s) chosen together
505
with the assumptions made by analysts. Based on the identified limitations of measured data to support the
assessment of the quality of predictions, we have proposed that the quality of UQ is to be judged based on crucial
tasks such as the exhaustive scrutiny of the knowledge coupled to the assessment of deviations of those
assumptions made in the analysis.
Key to enacting the proposed clarifications and simplifications is the full consideration of knowledge-based
510
probabilities. Based on the proposed examination of strength of knowledge, knowledge-based probabilities can be
assigned. Considering this type of probabilities will also help overcome the identified limitations of the maximum
entropy principle or counterfactual analysis to quantify uncertainty in input quantities. We have exposed that the
latter approaches are prone to produce unexhausted uncertainty quantification due to their reliance on measured
data, which can miss crucial events or overlook relevant input quantities.
515
Appendix
In this Appendix, the necessary details of the original analysis made by Zhao et al. (2021) are given. The following
are the basic equations Eɱ used by these authors.
󰇛󰇜󰥟
󰥟
(A-1)
 󰇧
󰇨
(A-2)
where X is the collection of all quantities at borehole-points, which take values x, Y is the collection of all quantities
520
at non-borehole-points with values y,  is the value of correlation between a quantity value x at a penetrated
point sx Sx and the value y at a non-penetrated point sy Sy.
is the horizontal distance between points sx and
sy, while is the vertical one. θh and θv are the horizontal and vertical scale of fluctuation, respectively. Each
material class considered is associated exclusively with an element in the set of integers {1,2,,C}. p(y=c) is the
probability of encountering a type of material c in a point sy. Such probability is initially approximated using Eq.
525
A-1. More accurate probabilities are computed on the basis of the repeated sampling of the joint distribution f(x,y),
which was approximated using Eq. A-1. Eq. A-1, described in short, approximates probabilities as the ratio of the
sum of correlation values calculated for a penetrated point in the set Sx and the set of non-penetrated points Sy for
a given material c to the sum of correlation values for all points and all materials.
Based on the stratigraphy data collected at borehole locations, the selection of the type of correlation function and
530
the scales of fluctuation took place using the maximum likelihood method. The authors considered three types of
correlation functions, namely squared exponential, single exponential, and second-order Markov. In this case, the
likelihood function (θm|ɗ) = f(ɗ|θm) represents the likelihood of observing ɗ at borehole locations, given the
spatial correlation structure θm. The squared exponential function yielded the maximum likelihood when the
horizontal scale of fluctuation and the vertical scale of fluctuation are set to 11.1 and 4.1 metres, respectively.
535
Hence, the squared exponential function correlation whose expression is Eq. A-2 in this Appendix was selected.
Eq. A-3 and A-4 correspond to the single exponential and second-order Markov functions, respectively.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
19
 󰇧
󰇨
(A-3)

󰇧
󰇨󰇧
󰇨
(A-4)
Data availability
540
The research reported in this paper did not generated data.
Code availability
The research reported in this paper did not generate any code.
Authors' contributions
Ibsen Chivata Cardenas: Conceptualization, Methodology, Writing- Original draft preparation, Investigation,
545
Validation; Terje Aven: Investigation, Supervision, Writing- Reviewing and Editing; Roger Flage: Investigation,
Supervision, Writing- Reviewing and Editing.
Conflict of interest/Competing interests
The authors declare that they have no conflict of interest.
Funding
550
This research is funded by ARCEx partners and the Research Council of Norway (grant number 228107).
References
Albert, C. G., Callies, U., and von Toussaint, U.: A Bayesian approach to the estimation of parameters and their
interdependencies in environmental modeling, Entropy, 24(2), 231. doi:10.3390/e24020231, 2022.
Alley, R. B. Abrupt climate change, Sci. Am., 291(5), 62-69. doi:10.1126/science.1081056, 2004.
555
Apeland, S., Aven, T., and Nilsen, T.: Quantifying uncertainty under a predictive, epistemic approach to risk
analysis, Reliab. Eng. Syst. Saf., 75(1), 93-102. doi:10.1016/S0951-8320(01)00122-3, 2002.
Aven, T.: On the need for restricting the probabilistic analysis in risk assessments to variability, Risk Anal., 30(3),
354-360. doi:10.1111/j.1539-6924.2009.01314.x, 2010.
Aven, T.: Practical implications of the new risk perspectives, Reliab. Eng. Syst. Saf., 115, 136-145.
560
doi:10.1016/j.ress.2013.02.020, 2013.
Aven, T.: The science of risk analysis: Foundation and practice. Routledge, London. doi:10.4324/9780429029189,
2019.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
20
Aven, T., and Kvaløy, J. T.: Implementing the Bayesian paradigm in risk analysis, Reliab. Eng. Syst. Saf., 78(2),
195-201. doi:10.1016/S0951-8320(02)00161-8, 2002.
565
Aven, T., and rn, K.: Expressing and interpreting the results of quantitative risk analyses, Review and
discussion. Reliab. Eng. Syst. Saf., 61(1-2), 3-10. doi:10.1016/S0951-8320(97)00060-4, 1998.
Aven, T., and Zio, E.: Model output uncertainty in risk assessment, Int. J. Perform. Eng., 29(5), 475-486.
doi:10.23940/ijpe.13.5.p475.mag, 2013.
Betz, W.: Bayesian inference of engineering models (Doctoral dissertation, Technische Universität nchen),
570
2017.
Brown, G.W.: Monte Carlo methods, Modern Mathematics for the Engineers, 279-303. McGraw-Hill, New York,
1956.
Cardenas, I.: On the use of Bayesian networks as a meta-modelling approach to analyse uncertainties in slope
stability analysis, Georisk, 13(1), 5365. doi:10.1080/17499518.2018.1498524. 2019.
575
Carrera, J., and Neuman, S.: Estimation of aquifer parameters under transient and steady state conditions: 2.
Uniqueness, stability, and solution algorithms, Water Resour. Res., 22(2), 211227.
doi:10.1029/WR022i002p00211, 1986.
Casalbore, D., Passeri, F., Tommasi, P., Verrucci, L., Bosman, A., Romagnoli, C., and Chiocci, F.L.: Small-scale
slope instability on the submarine flanks of insular volcanoes: the case-study of the Sciara del Fuoco slope
580
(Stromboli), Int. J. Earth Sci., 109, (8), 2643-2658. doi:10.1007/s00531-020-01853-5, 2020.
Catuneanu, O., Abreu, V., Bhattacharya, J. P., Blum, M. D., Dalrymple, R. W., Eriksson, P. G., ... and Winker, C.:
Towards the standardization of sequence stratigraphy, Earth-Sci. Rev., 92(1-2), 1-33.
doi:10.1016/j.earscirev.2008.10.003, 2009.
Chow, Y.K., S. Li, and Koh. C.G.: A particle method for simulation of submarine landslides and mudflows, Paper
585
presented at the 29th International Ocean and Polar Engineering Conference, June 1621, in Honolulu, Hawaii,
USA, ISOPE-I-19-594, 2019.
Christakos, G.: A Bayesian/maximum-entropy view to the spatial estimation problem, Math. Geol., 22(7), 763-
777. doi:10.1007/BF00890661, 1990.
Clare, M.A., Clarke, J.H., Talling, P.J., Cartigny, M.J., and Pratomo, D.G.: Preconditioning and triggering of
590
offshore slope failures and turbidity currents revealed by most detailed monitoring yet at a fjord-head delta, Earth
Planet. Sci. Lett., 450, 208-220. doi:10.1016/j.epsl.2016.06.021, 2016.
Clare, M. A., Vardy, M. E., Cartigny, M. J., Talling, P. J., Himsworth, M. D., Dix, J. K., ... and Belal, M.: Direct
monitoring of active geohazards: Emerging geophysical tools for deepwater assessments, Near Surf.
Geophys., 15(4), 427-444. doi:10.3997/1873-0604.2017033, 2017.
595
Degen, D., Veroy, K., Scheck-Wenderoth, M., and Wellmann, F.: Crustal-scale thermal models: Revisiting the
influence of deep boundary conditions, Environ. Earth Sci., 81(3), 1-16. doi:10.1007/s12665-022-10202-5, 2022.
Dubois, D.: Possibility theory and statistical reasoning, Comput. Stat. Data Anal., 51(1), 47-69.
doi:10.1016/j.csda.2006.04.015, 2006.
Ferson, S., and Ginzburg, L. R.: Different methods are needed to propagate ignorance and variability, Reliab. Eng.
600
Syst. Saf., 54(2-3), 133-144. doi:10.1016/S0951-8320(96)00071-3, 1996.
Flage, R., Baraldi, P., Zio, E., and Aven, T.: Probability and possibilitybased representations of uncertainty in
fault tree analysis, Risk Anal., 33(1), 121-133. doi:10.1111/j.1539-6924.2012.01873.x, 2013.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
21
Flage, R., Aven, T., and Berner, C. L.: A comparison between a probability bounds analysis and a subjective
probability approach to express epistemic uncertainties in a risk assessment contextA simple illustrative example,
605
Reliab. Eng. Syst. Saf., 169, 1-10. doi:10.1016/j.ress.2017.07.016, 2018.
Gray, A., Ferson, S., Kreinovich, V., and Patelli, E.: Distribution-free risk analysis, Int. J. Approx. Reason., 146,
133-156. doi:10.1016/j.ijar.2022.04.001, 2022a.
Gray, A., Wimbush, A., de Angelis, M., Hristov, P. O., Calleja, D., Miralles-Dolz, E., and Rocchetta, R.: From
inference to design: A comprehensive framework for uncertainty quantification in engineering with limited
610
information, Mech. Syst. Signal Process., 165, 108210. doi:10.1016/j.ymssp.2021.108210, 2022b.
Hastings, W. K.: Monte Carlo sampling methods using Markov chains and their applications, Biometrika, 87, 97-
109. doi:10.2307/2334940, 1970.
Helton, J. C., and Oberkampf, W. L.: Alternative representations of epistemic uncertainty, Reliab. Eng. Syst. Saf.,
1(85), 1-10. doi:10.1016/j.ress.2011.02.013, 2004.
615
Huang, L., Cheng, Y. M., Li, L., and Yu, S. Reliability and failure mechanism of a slope with non-stationarity and
rotated transverse anisotropy in undrained soil strength, Comput. Geotech., 132, 103970.
doi:10.1016/j.compgeo.2020.103970, 2021.
Hunt, J.E., Wynn, R.B., Talling, P.J., Masson, D.G.: Frequency and timing of landslide-triggered turbidity currents
within the Agadir Basin, offshore NW Africa: Are there associations with climate change, sea level change and
620
slope sedimentation rates? Mar. Geol., 346, 274-291. doi:10.1016/j.margeo.2013.09.004, 2013.
Jaynes, E. T.: Information theory and statistical mechanics, Phys. Rev., 106(4), 620.
doi:10.1103/PhysRev.106.620, 1957.
Juang, C. H., Zhang, J., Shen, M., and Hu, J.: Probabilistic methods for unified treatment of geotechnical and
geological uncertainties in a geotechnical analysis, Eng. Geol, 249, 148-161. doi:10.1016/j.enggeo.2018.12.010,
625
2019.
Khorsandi, J., Aven, T.: Incorporating assumption deviation risk in quantitative risk assessments: A semi-
quantitative approach, Reliab. Eng. Syst. Saf., 163, 22-32. doi:10.1016/j.ress.2017.01.018, 2017.
Leynaud, D., Mulder, T., Hanquiez, V., Gonthier, E., Régert, A.: Sediment failure types, preconditions and
triggering factors in the Gulf of Cadiz, Landslides, 14(1), 233-248. doi:10.1007/s10346-015-0674-2, 2017.
630
Liu, Y., Ren, W., Liu, C., Cai, S., and Xu, W.: Displacement-based back-analysis frameworks for soil parameters
of a slope: Using frequentist inference and Bayesian inference, Int. J. Geomech., 22(4), 04022026.
doi:10.1061/(ASCE)GM.1943-5622.0002318, 2022.
Lloret-Cabot. M., Fenton, G.A., Hicks, M.A.: On the estimation of scale of fluctuation in geostatistics, Georisk
8(2):129-140. doi:10.1080/17499518.2013.871189, 2014.
635
Lu, P., and Lermusiaux, P. F.: Bayesian learning of stochastic dynamical models, Phys. D: Nonlinear
Phenom., 427, 133003. doi:10.1016/j.physd.2021.133003, 2021.
Luo, L., Liang, X., Ma, B., and Zhou, H.: A karst networks generation model based on the Anisotropic Fast
Marching Algorithm, J. Hydrol., 126507. doi:10.1016/j.jhydrol.2021.126507, 2021.
Metropolis, N., and Ulam, S.: The Monte Carlo method, J. Am. Stat. Assoc., 44(247), 335-341.
640
doi:10.1080/01621459.1949.10483310, 1949.
Montanari, A., and Koutsoyiannis, D.: A blueprint for processbased modeling of uncertain hydrological systems,
Water Resour. Res., 48(9). doi:10.1029/2011WR011412, 2012.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
22
Nilsen, T., and Aven, T.: Models and model uncertainty in the context of risk analysis, Reliab. Eng. Syst. Saf.,
79(3), 309-317. doi:10.1016/S0951-8320(02)00239-9, 200.
645
Pakyuz-Charrier, E., Lindsay, M., Ogarko, V., Giraud, J., and Jessell, M.: Monte Carlo simulation for uncertainty
estimation on structural data in implicit 3-D geological modeling, a guide for disturbance distribution selection
and parameterization, Solid Earth, 9(2), 385-402. doi:10.5194/se-9-385-2018, 2018.
Pearl, J.: Comment: graphical models, causality and intervention, Statist. Sci., 8(3), 266-269.
https://www.jstor.org/stable/2245965, 1993.
650
Pheulpin, L., Bertrand, N., and Bacchi, V.: Uncertainty quantification and global sensitivity analysis with
dependent inputs parameters: Application to a basic 2D-hydraulic model, LHB, 108(1), 2015265.
doi:10.1080/27678490.2021.2015265, 2022.
Raíces-Cruz, I., Troffaes, M. C., and Sahlin, U.: A suggestion for the quantification of precise and bounded
probability to quantify epistemic uncertainty in scientific assessments, Risk Anal., 42(2), 239-253.
655
doi:10.1111/risa.13871, 2022.
Roy, C. J., and Oberkampf, W. L.: A comprehensive framework for verification, validation, and uncertainty
quantification in scientific computing, Comput. Methods Appl. Mech. Eng., 200(25-28), 2131-2144.
doi:10.1016/j.cma.2011.03.016, 2011.
Rodríguez-Ochoa, R., Nadim, F., Cepeda, J. M., Hicks, M. A., and Liu, Z.: Hazard analysis of seismic submarine
660
slope instability, Georisk, 9(3), 128-147. doi:10.1080/17499518.2015.1051546, 2015.
Sankararaman, S., and Mahadevan, S.: Integration of model verification, validation, and calibration for uncertainty
quantification in engineering systems, Reliab. Eng. Syst. Saf., 138, 194-209. doi:10.1016/j.ress.2015.01.023, 2015.
Shafer, G.: A mathematical theory of evidence. In A mathematical theory of evidence. Princeton university press.
197.
665
Shortridge, J., Aven, T., and Guikema, S.: Risk assessment under deep uncertainty: A methodological comparison,
Reliab. Eng. Syst. Saf., 159, 12-23. doi:10.1016/j.ress.2016.10.017, 2017.
Society for Risk Analysis Society for Risk Analysis Glossary Available from: SRA-Glossary-FINALpdf Accessed
on 25 March 2021, 2018.
Sun, X., Zeng, X., Wu, J., and Wang, D.: A Twostage Bayesian data‐driven method to improve model prediction,
670
Water Resour. Res., e2021WR030436. doi:10.1029/2021WR030436, 2021b.
Sun, X., Zeng, P., Li, T., Wang, S., Jimenez, R., Feng, X., and Xu, Q.: From probabilistic back analyses to
probabilistic run-out predictions of landslides: A case study of Heifangtai terrace, Gansu Province, China, Eng.
Geol, 280, 105950. doi:10.1016/j.enggeo.2020.105950, 2021a.
Tacher, L., Pomian-Srzednicki, I., Parriaux, A.: Geological uncertainties associated with 3-D subsurface models,
675
Comput. Geosci., 32(2), 212-221. doi:10.1016/j.cageo.2005.06.010, 2006.
Tang, X. S., Wang, M. X., and Li, D. Q.: Modeling multivariate cross-correlated geotechnical random fields using
vine copulas for slope reliability analysis, Comput. Geotech., 127, 103784. doi:10.1016/j.compgeo.2020.103784,
2020.
Thiele, S.T., Jessell, M.W., Lindsay, M., Wellmann, J.F., Pakyuz-Charrier, E.: The topology of geology 2:
680
Topological uncertainty, J. Struct. Geol., 91,74-87. doi:10.1016/j.jsg.2016.08.010, 2016.
Ulam, S. M.: Monte Carlo calculations in problems of mathematical physics, Modern Mathematics for the
Engineers, 261-281. McGraw-Hill, New York. 1961.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
23
Uzielli, M., and Lacasse, S.: Scenario-based probabilistic estimation of direct loss for geohazards, Georisk, 1(3),
142-154. doi:10.1080/17499510701636581. 2007.
685
van den Eijnden, A. P., Schweckendiek, T., and Hicks, M. A.: Metamodelling for geotechnical reliability analysis
with noisy and incomplete models, Georisk, 1-18. doi:10.1080/17499518.2021.1952611, 2021.
Vanneste, M., F. Løvholt, D. Issler, Z. Liu, N. Boylan, and Kim. J.: A novel quasi-3D landslide dynamics model:
from theory to applications and risk assessment. Paper presented at the Offshore Technology Conference, May 6
9, in Houston, Texas, OTC-29363-MS. doi:10.4043/29363-MS, 2019.
690
Vanmarcke, E.H.: Random fields: Analysis and synthesis. Cambridge, MA: The MIT Press, 1984.
Wagener, T., Reinecke, R., and Pianosi, F. On the evaluation of climate change impact models. Wiley Interdiscip.
Rev. Clim. Change, e772. doi:10.1002/wcc.772, 2022.
Wellmann, J. F., and Regenauer-Lieb, K.: Uncertainties have a meaning: Information entropy as a quality measure
for 3-D geological models, Tectonophysics, 526, 207-216. doi:10.1016/j.tecto.2011.05.001, 2012.
695
Woo, G.: Downward counterfactual search for extreme events, Front. Earth Sci., 340.
doi:10.3389/feart.2019.00340, 2019.
Yano, J. I.: What is the Maximum Entropy Principle? Comments on “Statistical theory on the functional form of
cloud particle size distributions, J Atmos Sci, 76(12), 3955-3960. doi:10.1175/JAS-D-18-0223.1, 2019.
Zadeh, L. A.: Probability measures of fuzzy events, J. Math. Anal. Appl., 23(2), 421-427. doi:10.1016/0022-
700
247X(68)90078-4, 1968.
Zhao, C., Gong, W., Li, T., Juang, C. H., Tang, H., and Wang, H.: Probabilistic characterization of subsurface
stratigraphic configuration with modified random field approach, Eng. Geol, 288, 106138.
doi:10.1016/j.enggeo.2021.106138, 2021.
https://doi.org/10.5194/gmd-2022-210
Preprint. Discussion started: 17 October 2022
c
Author(s) 2022. CC BY 4.0 License.
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Elementary formulas for propagating information about means and variances through mathematical expressions have long been used by analysts. Yet the precise implications of such information are rarely articulated. This paper explores distribution-free techniques for risk analysis that do not require simulation, sampling or approximation of any kind. We describe best-possible bounds on risks that can be inferred given only information about the range, mean and variance of a random variable. These bounds generalise the classical Chebyshev inequality in an obvious way. We also collect in convenient tables several formulas for propagating range and moment information through calculations involving 7 binary convolutions (addition, subtraction, multiplication, division, powers, minimum, and maximum) and 9 unary transformations (scalar multiplication, scalar translation, exponentiation, natural and common logarithms, reciprocal, square, square root and absolute value) commonly encountered in risk expressions. These formulas are rigorous rather than approximate, and in most cases are either exact or mathematically best-possible. The formulas can be used effectively even when only interval estimates of the moments are available. Although most discussions of moment propagation assume stochastic independence among variables, this paper shows the assumption to be unnecessary and generalises formulas for the case when no assumptions are made about dependence, and when correlations are partially known. Along with partial means and variances, we show how interval covariances may be propagated and tracked through expressions.
Article
Full-text available
Hydraulic models include many uncertainties related to model parameters. Uncertainty Quantification (UQ) and Global Sensitivity Analysis (GSA) allow to quantify these uncertainties and identify the most influent parameters. To use traditional methods of UQ and GSA, the inputs must be independent. Therefore, this research aims to show that the dependence between inputs should not be neglected in uncertainty studies. To illustrate this, we used a simple hydraulic model built with TELEMAC-2D, representing a hypothetical linear river protected by a dyke that can break. The methodology is completed following these steps: the uncertain inputs and the outputs of interest are identified. Then, the inputs margins are defined through a statistical analysis using real dataset. The dependence structure between inputs is figured using copulas. Kriging metamodels are used to increase the number of experiences in a short time period. Finally, UQ and GSA are achieved. Regarding the UQ results, the outputs distribution is different if the inputs are considered dependent or not, whereas regarding the GSA, some parameters, usually considered non-influent in the hypothesis of independent inputs, have a real impact on the outputs. These results suggest that it could be interesting to consider the dependencies between inputs for “real” applications.
Article
Full-text available
The displacement-based back-analysis is an effective approach to determine the values of soil parameters of a slope. However, the applicability of different methods of the displacement-based back-analysis varies. This paper divides the displacement-based back-analysis into the deterministic back-analysis under frequentist inference and the probabilistic back-analysis under Bayesian inference. A framework for the deterministic back-analysis is proposed using the maximum likelihood estimation, which is a typical method of frequentist inference. In the framework for the deterministic back-analysis, the dual annealing (DA) algorithm is applied to search for the globally optimal solution. A framework for the probabilistic back-analysis considering spatial variability is proposed based on the Bayesian theory, random field theory, and truncated Karhunen-Loève expansion (KLE), which overcome the "curse of dimensionality." It is explained that this framework will also work in the probabilistic back-analysis based on the random variable model. In the two frameworks, metamodels are constructed by gradient boosting decision trees (GBDT). This algorithm perfectly fits the relationship between parameters and displacements of a slope, which replaces the time-consuming finite-element simulations. The adaptability of the two proposed frameworks is illustrated with a real case study of a highway slope. The case study also proves that the fitting accuracy of metamodels constructed by the GBDT algorithm is higher than those constructed by the neural network (NN) and the random forests (RF). Processes of frequentist inference and Bayesian inference show that the difference in their results originates from the different perceptions of the parameter space. However, the two different back-analysis results are not in competition but should be complementary. They both play an important role in slope parameter determination and slope stability analysis .
Article
Full-text available
We present a case study for Bayesian analysis and proper representation of distributions and dependence among parameters when calibrating process-oriented environmental models. A simple water quality model for the Elbe River (Germany) is referred to as an example, but the approach is applicable to a wide range of environmental models with time-series output. Model parameters are estimated by Bayesian inference via Markov Chain Monte Carlo (MCMC) sampling. While the best-fit solution matches usual least-squares model calibration (with a penalty term for excessive parameter values), the Bayesian approach has the advantage of yielding a joint probability distribution for parameters. This posterior distribution encompasses all possible parameter combinations that produce a simulation output that fits observed data within measurement and modeling uncertainty. Bayesian inference further permits the introduction of prior knowledge, e.g., positivity of certain parameters. The estimated distribution shows to which extent model parameters are controlled by observations through the process of inference, highlighting issues that cannot be settled unless more information becomes available. An interactive interface enables tracking for how ranges of parameter values that are consistent with observations change during the process of a step-by-step assignment of fixed parameter values. Based on an initial analysis of the posterior via an undirected Gaussian graphical model, a directed Bayesian network (BN) is constructed. The BN transparently conveys information on the interdependence of parameters after calibration. Finally, a strategy to reduce the number of expensive model runs in MCMC sampling for the presented purpose is introduced based on a newly developed variant of delayed acceptance sampling with a Gaussian process surrogate and linear dimensionality reduction to support function-valued outputs.
Article
Full-text available
In‐depth understanding of the potential implications of climate change is required to guide decision‐ and policy‐makers when developing adaptation strategies and designing infrastructure suitable for future conditions. Impact models that translate potential future climate conditions into variables of interest are needed to create the causal connection between a changing climate and its impact for different sectors. Recent surveys suggest that the primary strategy for validating such models (and hence for justifying their use) heavily relies on assessing the accuracy of model simulations by comparing them against historical observations. We argue that such a comparison is necessary and valuable, but not sufficient to achieve a comprehensive evaluation of climate change impact models. We believe that a complementary, largely observation‐independent, step of model evaluation is needed to ensure more transparency of model behavior and greater robustness of scenario‐based analyses. This step should address the following four questions: (1) Do modeled dominant process controls match our system perception? (2) Is my model's sensitivity to changing forcing as expected? (3) Do modeled decision levers show adequate influence? (4) Can we attribute uncertainty sources throughout the projection horizon? We believe that global sensitivity analysis, with its ability to investigate a model's response to joint variations of multiple inputs in a structured way, offers a coherent approach to address all four questions comprehensively. Such additional model evaluation would strengthen stakeholder confidence in model projections and, therefore, into the adaptation strategies derived with the help of impact models. This article is categorized under: Climate Models and Modeling > Knowledge Generation with Models Assessing Impacts of Climate Change > Evaluating Future Impacts of Climate Change
Article
Full-text available
The societal importance of geothermal energy is significantly increasing because of its low carbon-dioxide footprint. However, geothermal exploration is also subject to high risks. For a better assessment of these risks, extensive parameter studies are required that improve the understanding of the subsurface. This yields computationally demanding analyses. Often, this is compensated by constructing models with a small vertical extent. This paper demonstrates that this leads to entirely boundary-dominated and hence uninformative models. It demonstrates the indispensable requirement to construct models with a large vertical extent to obtain informative models with respect to the model parameters. For this quantitative investigation, global sensitivity studies are essential since they also consider parameter correlations. To compensate for the computationally demanding nature of the analyses, a physics-based machine learning approach is employed, namely the reduced basis method, instead of reducing the physical dimensionality of the model. The reduced basis method yields a significant cost reduction while preserving the physics and a high accuracy, thus providing a more efficient alternative to considering, for instance, a small vertical extent. The reduction of the mathematical instead of physical space leads to less restrictive models and, hence, maintains the model prediction capabilities. The combination of methods is used for a detailed investigation of the influence of model boundary settings in typical regional-scale geothermal simulations and highlights potential problems.
Article
Full-text available
An honest communication of uncertainty about quantities of interest enhances transparency in scientific assessments. To support this communication, risk assessors should choose appropriate ways to evaluate and characterize epistemic uncertainty. A full treatment of uncertainty requires methods that distinguish aleatory from epistemic uncertainty. Quantitative expressions for epistemic uncertainty are advantageous in scientific assessments because they are nonambiguous and enable individual uncertainties to be characterized and combined in a systematic way. Since 2019, the European Food Safety Authority (EFSA) recommends assessors to express epistemic uncertainty in conclusions of scientific assessments quantitatively by subjective probability. A subjective probability can be used to represent an expert judgment, which may or may not be updated using Bayes's rule to integrate evidence available for the assessment and could be either precise or approximate. Approximate (or bounded) probabilities may be enough for decision making and allow experts to reach agreement on certainty when they struggle to specify precise subjective probabilities. The difference between the lower and upper bound on a subjective probability can also be used to reflect someone's strength of knowledge. In this article, we demonstrate how to quantify uncertainty by bounded probability, and explicitly distinguish between epistemic and aleatory uncertainty, by means of robust Bayesian analysis, including standard Bayesian analysis through precise probability as a special case. For illustration, the two analyses are applied to an intake assessment.
Article
Full-text available
Because of the earth system complexity, groundwater/hydrology models are always built with structural errors, which may lead to systematic errors in model predictions. Bayesian data‐driven methods (DDMs) provide a feasible way to correct systematic model errors statistically. Generally, the physical and statistical model parameters, namely physical parameters and hyperparameters, are assumed to be independent and jointly calibrated. However, this assumption may be unreasonable and lead to over‐adjusted parameter estimation and biased model prediction. This study proposes a two‐stage DDM to calibrate physical parameters and hyperparameters separately, which does not make the independence assumption. Three case studies, including a groundwater solute transport analytical model, a three‐dimensional groundwater flow model, and a real‐world snowmelt runoff model, were used to evaluate the predictive performance of this two‐stage DDM. Based on the three case studies, we found that the independence assumption of physical parameters and hyperparameters could lead to the over‐fitting of parameter estimation and deviations in model predictions. Two‐stage DDM can constrain the systematic error model calibration; that is, physical parameters are first calibrated in the entire hyperparameter prior probability space, and then hyperparameters are calibrated in the posterior probability space of physical parameters obtained previously. As a result, compared with traditional joint calibration‐based DDM, two‐stage DDM can alleviate parameter over‐fitting and improve model predictive performance.
Article
Full-text available
In this paper we present a framework for addressing a variety of engineering design challenges with limited empirical data and partial information. This framework includes guidance on the characterisation of a mixture of uncertainties, efficient methodologies to integrate data into design decisions, and to conduct reliability analysis, and risk/reliability based design optimisation. To demonstrate its efficacy, the framework has been applied to the NASA 2020 uncertainty quantification challenge. The results and discussion in the paper are with respect to this application.
Article
A new methodology for rigorous Bayesian learning of high-dimensional stochastic dynamical models is developed. The methodology performs parallelized computation of marginal likelihoods for multiple candidate models, integrating over all state variable and parameter values, and enabling a principled Bayesian update of model distributions. This is accomplished by leveraging the dynamically orthogonal (DO) evolution equations for uncertainty prediction in a dynamic stochastic subspace and the Gaussian Mixture Model-DO filter for inference of nonlinear state variables and parameters, using reduced-dimension state augmentation to accommodate models featuring uncertain parameters. Overall, the joint Bayesian inference of the state, model equations, geometry, boundary conditions, and initial conditions is performed. Results are exemplified using two high-dimensional, nonlinear simulated fluid and ocean systems. For the first, limited measurements of fluid flow downstream of an obstacle are used to perform joint inference of the obstacle’s shape, the Reynolds number, and the O(105) fluid velocity state variables. For the second, limited measurements of the concentration of a microorganism advected by an uncertain flow are used to perform joint inference of the microorganism’s reaction equation and the O(105) microorganism concentration and ocean velocity state variables. When the observations are sufficiently informative about the learning objectives, we find that our posterior model probabilities correctly identify either the true model or the most plausible models, even in cases where a human would be challenged to do the same.