modern statistical methods in oceanography: a …arxiv:1312.5904v1 [stat.me] 20 dec 2013 statistical...

22
arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical Statistics, 2013 Modern Statistical Methods in Oceanography: A Hierarchical Perspective Christopher K. Wikle, Ralph F. Milliff, Radu Herbei and William B. Leeds Abstract. Processes in ocean physics, air–sea interaction and ocean biogeochemistry span enormous ranges in spatial and temporal scales, that is, from molecular to planetary and from seconds to millennia. Identifying and implementing sustainable human practices depend crit- ically on our understandings of key aspects of ocean physics and ecology within these scale ranges. The set of all ocean data is distorted such that three- and four-dimensional (i.e., time-dependent) in situ data are very sparse, while observations of surface and upper ocean properties from space-borne platforms have become abundant in the past few decades. Precisions in observations of all types vary as well. In the face of these challenges, the interface between Statistics and Oceanography has proven to be a fruitful area for research and the development of useful models. With the recognition of the key importance of identify- ing, quantifying and managing uncertainty in data and models of ocean processes, a hierarchical perspective has become increasingly produc- tive. As examples, we review a heterogeneous mix of studies from our own work demonstrating Bayesian hierarchical model applications in ocean physics, air–sea interaction, ocean forecasting and ocean ecosys- tem models. This review is by no means exhaustive and we have endeav- ored to identify hierarchical modeling work reported by others across the broad range of ocean-related topics reported in the statistical liter- ature. We conclude by noting relevant ocean-statistics problems on the immediate research horizon, and some technical challenges they pose, for example, in terms of nonlinearity, dimensionality and computing. Key words and phrases: Bayesian, biogeochemical, ecosystem, ocean vector winds, quadratic nonlinearity, spatio-temporal, state–space, sea surface temperature. Christopher K. Wikle is Professor, Department of Statistics, University of Missouri,146 Middlebush Hall, Columbia, Missouri 65203, USA e-mail: [email protected]. Ralph F. Milliff is Senior Research Associate, CIRES, University of Colorado, Boulder, Colorado, USA e-mail: Ralph.Milliff@colorado.edu . Radu Herbei is Assistant Professor, Department of Statistics, The Ohio State University, Columbus, Ohio, USA e-mail: [email protected]. William B. Leeds is Postdoctoral Researcher, Department of Statistics and Department of Geophysical Sciences, University of Chicago, Chicago, Illinois, USA e-mail: [email protected]. 1. INTRODUCTION The global ocean dominates the iconic image of Earth viewed from space, leading to the now fa- mous “blue marble” descriptor for our planet. The ocean covers more than 70% of the planetary surface and ocean processes are critical to life-sustaining This is an electronic reprint of the original article published by the Institute of Mathematical Statistics in Statistical Science, 2013, Vol. 28, No. 4, 466–486. This reprint differs from the original in pagination and typographic detail. 1

Upload: others

Post on 01-Jun-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

arX

iv:1

312.

5904

v1 [

stat

.ME

] 2

0 D

ec 2

013

Statistical Science

2013, Vol. 28, No. 4, 466–486DOI: 10.1214/13-STS436c© Institute of Mathematical Statistics, 2013

Modern Statistical Methods inOceanography: A Hierarchical PerspectiveChristopher K. Wikle, Ralph F. Milliff, Radu Herbei and William B. Leeds

Abstract. Processes in ocean physics, air–sea interaction and oceanbiogeochemistry span enormous ranges in spatial and temporal scales,that is, from molecular to planetary and from seconds to millennia.Identifying and implementing sustainable human practices depend crit-ically on our understandings of key aspects of ocean physics and ecologywithin these scale ranges. The set of all ocean data is distorted suchthat three- and four-dimensional (i.e., time-dependent) in situ data arevery sparse, while observations of surface and upper ocean propertiesfrom space-borne platforms have become abundant in the past fewdecades. Precisions in observations of all types vary as well. In the faceof these challenges, the interface between Statistics and Oceanographyhas proven to be a fruitful area for research and the development ofuseful models. With the recognition of the key importance of identify-ing, quantifying and managing uncertainty in data and models of oceanprocesses, a hierarchical perspective has become increasingly produc-tive. As examples, we review a heterogeneous mix of studies from ourown work demonstrating Bayesian hierarchical model applications inocean physics, air–sea interaction, ocean forecasting and ocean ecosys-tem models. This review is by no means exhaustive and we have endeav-ored to identify hierarchical modeling work reported by others acrossthe broad range of ocean-related topics reported in the statistical liter-ature. We conclude by noting relevant ocean-statistics problems on theimmediate research horizon, and some technical challenges they pose,for example, in terms of nonlinearity, dimensionality and computing.

Key words and phrases: Bayesian, biogeochemical, ecosystem, oceanvector winds, quadratic nonlinearity, spatio-temporal, state–space, seasurface temperature.

Christopher K. Wikle is Professor, Department of

Statistics, University of Missouri,146 Middlebush Hall,

Columbia, Missouri 65203, USA e-mail:

[email protected]. Ralph F. Milliff is Senior

Research Associate, CIRES, University of Colorado,

Boulder, Colorado, USA e-mail:

[email protected]. Radu Herbei is Assistant

Professor, Department of Statistics, The Ohio State

University, Columbus, Ohio, USA e-mail:

[email protected]. William B. Leeds is Postdoctoral

Researcher, Department of Statistics and Department of

Geophysical Sciences, University of Chicago, Chicago,

Illinois, USA e-mail: [email protected].

1. INTRODUCTION

The global ocean dominates the iconic image ofEarth viewed from space, leading to the now fa-mous “blue marble” descriptor for our planet. Theocean covers more than 70% of the planetary surfaceand ocean processes are critical to life-sustaining

This is an electronic reprint of the original articlepublished by the Institute of Mathematical Statistics inStatistical Science, 2013, Vol. 28, No. 4, 466–486. Thisreprint differs from the original in pagination andtypographic detail.

1

Page 2: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

2 WIKLE, MILLIFF, HERBEI AND LEEDS

(and life-challenging) events and processes occur-ring across broad ranges of temporal and spatialscales. Understanding issues of ocean resource con-sumption (e.g., fisheries, coastal pollution, etc.) leadto foci on ocean ecosystem dynamics and their cou-pling to physical processes (e.g., mixing, transports,upwelling, ice dynamics, etc.). Understanding theocean role in climate dynamics (e.g., sequestrationof atmospheric CO2, absorption of atmospheric heat,impacts on the hydrologic cycle, teleconnections,etc.) lead to foci on massive and complex simula-tions and forecasts based on equations of geophysi-cal fluid dynamics. In all instances, the broad rangeof scales, the energetic exchanges across them, andthe associated uncertainties drive innovations thatinvolve methods of modern statistical modeling.Oceanography has historically been a “data poor”

science and the need to use advanced statistical meth-odology to perform inference and prediction has beenparamount throughout its development. Althoughit is the case that there are too few in situ obser-vations of the ocean to characterize its evolutionand its interaction with marine ecosystems, in anironic twist, the discipline also suffers from havingan abundance of particular data types when one fac-tors in the satellite observations that have becomeavailable in the last couple of decades. The needfor statistical collaboration in Oceanography resultsfrom both the situation of not having enough obser-vations in some parts of the system and having hugeamounts of data in other parts of the system.

1.1 The Physical Ocean

The physical ocean is governed by basic laws ofphysics (see, e.g., Vallis (2006)). The primitive equa-tions consist of the following: three equations corre-sponding to the conservation of momentum (for thetwo horizontal and one vertical components of veloc-ity), a continuity equation representing the conser-vation of mass, an equation of state (relating density,pressure, temperature and salinity), and equationscorresponding to the conservation of temperatureand salinity. There are seven state variables (threevelocity components, density, pressure, temperatureand salinity). This system of equations is nonlinearand exhibits a huge range of spatial and temporalscales of variability. Given that many of these scalesof variability are not resolved in data or in determin-istic or “forward” ocean models, the equations aretypically simplified by scale analysis arguments and

the small scale (turbulent) structures are param-eterized. These parameterizations involve relation-ships between the mean of the state variables andtheir gradients. In this way the eddy viscosity anddiffusivity terms serve as so-called “sub-grid scale”parameterizations, representing the unresolved pro-cesses that are sinks of momentum and heat at thegrid scale of a given model, for example, O(10) km inglobal ocean models and O(1) km in regional oceanmodels.The ocean system is nonlinearly coupled to the

atmosphere across a broad range of scales. At thelargest scales, air–sea fluxes of heat and fresh waterdrive mostly vertical or “thermohaline” circulationswhile the surface shear stress and wind stress curldrive mostly horizontal or wind-driven gyre circula-tions (e.g., Pedlosky (1998)). At smaller scales, thevertical-horizontal separation breaks down and theocean response to external forcing and internal in-stabilities results in a broadband spectrum of vigor-ous eddy circulations (e.g., McWilliams (2006)). Onall scales, the circulation provides the context forocean processes affecting other components of theocean system such as those related to ocean biologyand chemistry. There are significant nonlinear inter-actions between the ocean chemistry, biology and itsphysical state.

1.2 Ocean Biogeochemistry

Ocean biogeochemistry is concerned with the in-teraction of the biology, chemistry and geology ofthe ocean (e.g., Miller (2004)). This is a very com-plex system that contains many interactions acrossa variety of scales. The system can be simply illus-trated by thinking about the interactions of broadclasses of its components. For example, the presenceof nutrients near the ocean surface, where there islight, allows for the growth of phytoplankton, whichdeplete the nutrients as their population expands.The increased abundance of phytoplankton then pro-vides a food source for zooplankton, which leadsto growth in the zooplankton population. The con-sumption of phytoplankton leads to waste productsfrom the zooplankton that settles as detritus to theocean floor. As the zooplankton deplete the phyto-plankton, the zooplankton population decreases dueto the lack of a sufficient food source. Eventually,the detritus at depth is transferred to the surfacethrough upwelling and mixing, providing the nutri-ents that lead to another bloom in phytoplankton,etc.

Page 3: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 3

This simple four component system is a vast over-simplification, as there are many different species in-teracting at any one time. More critically, this lowertrophic ecosystem is also coupled to higher levels ofthe food web, for example, with foraging fish predat-ing the zooplankton, which are in turn predated byhigher trophic fish, marine mammals, commercialfishing, etc. The chemical component of the cycleis critical in several respects. It provides a way forcarbon to be removed from the atmosphere, as thephytoplankton remove carbon from the ocean waterand are consumed by the zooplankton. Some of thatcarbon is contained in the detritus that sinks to theocean floor, becoming buried in the sediment andleading to a (temporary) carbon sink in the globalcarbon cycle, which is very important in the contextof sequestration of carbon relative to potential cli-mate change from greenhouse gases. In addition, thebiological cycle in the ocean is closely tied to the dis-tribution of dissolved oxygen in the water and alsoinfluences the distribution of other chemicals, suchas silicon, nitrates and phosphates, for example, inthe shells of diatoms.

1.3 Uncertainty

Given the complexities in the ocean system, itis not surprising that there are numerous sourcesof uncertainty. First, although the large-scale equa-tions of motion are in some sense deterministic, thescale issues that lead to eddy viscosity/diffusivityparameterizations are inherently uncertain. Further-more, the forms of the linkages between system com-ponents (e.g., wind stress, heat and moisture fluxesbetween atmosphere and ocean) are not known withcertainty. The components of traditional biogeochem-ical models are even more uncertain, both in termsof the functional forms and parameters.The process and parameter uncertainty is com-

pounded by the inherent data issues in the oceansystem. In situ observations of the ocean are quitelimited in terms of spatial and temporal resolution,and in terms of the variables measured. For example,it is a painstaking process to measure zooplanktonabundance, often requiring a scientist or technicianto literally count critters through a microscope. For-tunately, many surface variables can be observedremotely, particularly through satellite proxies. Insome cases, for example, near surface winds fromscatterometers, sea surface height from altimetersand sea surface temperature (SST) from radiome-ters, the satellite observations are typically quite

precise, albeit with gaps corresponding to orbital ge-ometries, swath widths and fields of view. In othercases, for example, ocean color as a proxy for phy-toplankton, the observational representation of theprocess is more uncertain, at least on fairly shorttime scales. Thus, a key issue in state prediction,parameter estimation and inference is to deal withincomplete observations that vary in precision andin spatial and temporal support. This is particularlyimportant when one considers ocean “data assimi-lation,” that is, the blending of prior information(e.g., a numerical solution of the deterministic rep-resentation of the ocean state) with observations.Another traditionally important component of un-

certainty in ocean process modeling corresponds tothe selection of reduced-dimensional representationsof the process. Given the assumption that much ofthe larger scale processes in the ocean can be repre-sented in a lower dimensional manifold, with smallerscales corresponding to turbulent scales (that cer-tainly may interact with the larger scale modes, orat least suggest the form of parameterizations orstochastic noise terms), there has been considerableattention given to different approaches to obtain thereduced-dimension basis functions. The choices varydepending on the part of the system being consid-ered as well as whether one is looking at the systemdiagnostically or predictively.In the context of statistical models used to de-

scribe or predict portions of the ocean system, thenature of the error structures is important. Giventhe nonlinearity that is inherent in the system, manyprocess distributions are not well represented byGaussian errors. In addition, in some cases (e.g., bi-ological abundance variables) the distributions canonly have positive support.

1.4 Statistical Methods

The ocean and atmospheric sciences have benefit-ted from a strong tradition in applying fairly com-plex statistical methods to deal with many of theuncertainty issues described above. In particular,general monographs such as Emery and Thomson(2001), Von Storch and Zwiers (2002), and Wilks(2011) provide comprehensive descriptions of tradi-tional methods used to analyze such data. In addi-tion to overviews of basic statistical concepts, thesebooks describe multivariate methods (e.g., principalcomponents—empirical orthogonal functions, canon-ical correlation analysis, discriminant analysis, etc.),spectral methods (e.g., cross-spectral analysis) and

Page 4: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

4 WIKLE, MILLIFF, HERBEI AND LEEDS

dynamically-based reduction methods (e.g., princi-pal oscillation patterns) to facilitate analysis of high-dimensional data that has inherent dependence intime and space. This is in addition to more focusedmonographs such as Preisendorfer and Mobley(1988), which gives a comprehensive overview ofeigen-decomposition methods, and several books andreview papers devoted to various aspects of data as-similation (see Section 3.1). Statistical presentationsof many of these methods can be found in Jolliffe(2002) and Cressie and Wikle (2011).Recognizing the challenges related to uncertainty

in the ocean system and the need to foster morecollaborative research between oceanographers andstatisticians, the U.S. National Research Council(NRC) commissioned a panel to write a report on“Statistics and Physical Oceanography” (NRC, 1994);see also the accompanying article by Chelton (1994)and published comments. This report contains a verynice review of physical oceanography for nonoceanog-raphers and outlined the need for research in severalkey areas, including the change of support prob-lem and the indirect nature of satellite observations,non-Gaussian random fields, the incorporation ofLagrangian and Eulerian data, data assimilation,inverse modeling, model/data comparison and fea-ture identification, to name some of the most promi-nent. The report focused on the physical componentof the ocean and did not address biogeochemistrynor many issues of current interest, such as climatechange and reduced-dimensional representations.

1.5 Paper Outline

Our goal with this review is to provide an overviewof some of the advancements that have occurred atthe interface of Statistics and Oceanography sincethe NRC (1994) report. In particular, we believestrongly that the hierarchical statistical perspectivehas played a significant role in this developmentand will focus our review from that perspective.Section 2 presents a brief discussion of hierarchicalmodeling, both empirical and Bayesian, with somediscussion of the need for computational tools. Wenote that although this paper is in a Statistics jour-nal, we hope that it will generate interest from bothstatisticians and oceanographers. For the same rea-son that we gave a brief and general overview ofOceanography above, we will also give a brief andgeneral overview of hierarchical modeling for thosereaders with little exposure to these ideas. In Sec-tion 3 we focus in more depth on examples related

to data assimilation and inverse modeling, long leadforecasting and uncertainty quantification in biogeo-chemical models. We will follow this review with abrief discussion of current and future challenges inSection 4.

2. THE HIERARCHICAL MODELING

PARADIGM

The idea of hierarchical modeling of scientific pro-cesses arose largely out of Berliner (1996), whenMark Berliner was the director of the GeophysicalStatistics Project at the National Center for Atmo-spheric Research. The idea, although fundamentallyquite simple, was revolutionary in that it provideda probabilistically consistent way to partition un-certainty in systems with complicated data, processand parameter relationships, and coincided with thedevelopment and popularization of Markov ChainMonte Carlo (MCMC) methods in Bayesian statis-tics. As described below, the key idea is to considerthe joint model of data, process and parameters asthree general linked model components, that is, thedata conditioned on the process and parameters,the process conditioned on parameters, and the pa-rameters. These ideas quickly spread into Statisticsand subject matter journals in Climatology, Meteo-rology, Oceanography and Ecology, to name a few.This particular perspective on hierarchical model-ing is summarized in several books, including Clark(2007), Royle and Dorazio (2008) and Cressie andWikle (2011). More traditional Bayesian presenta-tions of hierarchical models can be found in manybooks on Bayesian statistics (e.g., Gelman et al.(2004); Banerjee, Carlin and Gelfand (2003)).Statistical modeling and analysis are about the

synthesis of information. This information may comefrom expert opinion, physical laws, previous empiri-cal results or various observations—both direct andindirect. Consider the case where we have a scientificprocess of interest, denoted by Y . As an example,say that this process corresponds to the near-surfacenorth/south and east/west wind components overa portion of the ocean (i.e., a multivariate spatio-temporal process). We also have observed data as-sociated with this process, say Z, which might comefrom a satellite-based scatterometer (i.e., wind com-ponent observations derived from speed and direc-tion that are incomplete in space and time). Weassume that we have parameters associated withthe measurement process, say θZ , that might rep-resent differences in support and representativeness

Page 5: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 5

between the satellite observations and the true windprocess at the resolution of interest. In addition, weassume that there are some parameters, say θY , thatdescribe the underlying wind process dynamics (i.e.,the evolution operator and innovation covariancesthat propagate the joint spatial fields of the windcomponents through time). Thus, using the total lawof probability, it is natural to write the decomposi-tion of the joint distribution of the data and processconditioned on the parameters as

[Z,Y |θZ , θY ] = [Z|Y, θZ ][Y |θY ],(1)

where [Z|Y, θZ ] is the “data distribution” (or “datamodel”) and [Y |θY ] is the “process distribution” (or“process model”), and we have assumed conditionalindependence of the parameters on the distributionsof the right-hand side (RHS) of (1). Note, we areusing brackets “[ ]” to refer to a distribution andthe vertical bar “|” to denote “conditioned upon.”Clearly, we could also consider an alternative decom-position in which Y is conditioned on Z followed bythe marginal distribution of Z. However, such a de-composition is less scientific as described below.In traditional statistics, one might think about the

data Z given some specified distributional form andsome associated parameters, θ (e.g., correspondingto a spatio-temporal mean and associated variancesand covariances, or their parameterization). In thecontext of (1), such a distribution arises from in-tegrating out the random Y process, yielding thedistribution [Z|θ ≡ θZ , θY ]. We are then typicallyinterested in estimating these parameters given thedata. Such estimation (e.g., maximum likelihood es-timation) does not include an explicit representationof a model for the underlying dynamical wind pro-cess, Y , but rather includes it implicitly through thefirst and second moments (as a consequence of theintegration). In addition, this distribution accountsfor the uncertainty that is due to sampling and mea-surement.The question is then why might we be interested

in Y ? First, in many such applications, one is ac-tually interested in predicting the true, but unob-served process, Y , rather than just accounting forits (co)variability. Second, given the complexity ofmost ocean and atmospheric processes, the multi-variate spatiotemporal dependence structures asso-ciated with Y can be very complicated (e.g., nonlin-ear in time, nonstationary in space and/or time) andpotentially very high-dimensional. This seriouslycomplicates the likelihood-based inference, as it puts

much of the modeling burden on the realistic specifi-cation of complicated dependence structures. Rather,by focusing attention on the process Y directly, onecan incorporate scientific insight (e.g., Markovianapproximations to mathematical representations ofthe process, spatially or time-varying parameters,etc.) and, critically, disentangle the measurementuncertainty (which can also be quite complicated)and the process (co)variability and uncertainty. Thatis, marginal means and covariances contain poten-tially complicated functions of θZ and θY , which canbe difficult to specify without explicitly doing theintegration of the random process. Thus, one is ef-fectively trading the complexity of specifying verycomplicated marginal dependence structures with amore scientific specification of the conditional meanas a random process at the next level of the hierar-chy. This also allows one to focus effort on modelingthe conditional error structure in the data stage,without having to try to disentangle the measure-ment uncertainty with the process (co)variability.Statisticians will recognize this as just a manifesta-tion of the issue that one faces in traditional mixed-model analysis where one has a choice of consideringa marginal model, whereby the random effects areintegrated out, or a conditional model, whereby therandom effects are predicted and the conditional co-variance structure in the data model is simpler (e.g.,Verbeke and Molenberghs (2009)).Applying Bayes’ rule to the hierarchical decompo-

sition in (1) gives

[Y |Z,θZ , θY ]∝ [Z|Y, θZ ][Y |θY ],(2)

with the normalizing constant obtained by the inte-gral of (1) with respect to the process Y . Note thatthis implies that we are updating our knowledge ofthe process of interest given the data we have ob-served. This is the goal of prediction and confirmsthat the hierarchical decomposition on the RHS of(2) is the more plausible scientific decomposition ofthe joint distribution of data and process describedabove in (1). Clearly, (2) assumes that the parame-ters are known without uncertainty. This is seldomthe case in reality, although one may have estimatesof these parameters from some source and be com-fortable with substituting them into (2), leading towhat Cressie and Wikle (2011) refer to as an empir-ical hierarchical model (EHM). Alternatively, andmore realistically for most ocean processes, at leastsome of the parameters are typically not known, andwe are interested in learning more about them or,

Page 6: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

6 WIKLE, MILLIFF, HERBEI AND LEEDS

at least, would like to account for their uncertainty.In this case, we consider the fully-hierarchical orBayesian hierarchical model (BHM):

[Y, θY , θZ |Z]∝ [Z|Y, θZ ][Y |θY ][θZ , θY ],(3)

where one must specify a prior distribution for theparameters [θZ , θY ] and we note that the normaliz-ing constant integrates over the parameters in ad-dition to the process. It is often the case that wehave information to inform these parameter distri-butions. For example, in the case of the wind processdescribed above, we have knowledge about the qual-ity of scatterometer observations of wind compo-nents, and recognize that certain turbulent scalinglaws must be followed, which can be incorporatedthrough restrictions and informative prior distribu-tions (e.g., Wikle et al. (2001)).Schematically, it is helpful to think about the RHS

of (3) by using the following representation of Berli-ner (1996):

[data,process,parameters]

= [data|process,parameters](4)

× [process|parameters]× [parameters].

The strength of this hierarchical representation isthat it provides a probabilistically consistent way tothink about modeling complex systems while quan-tifying uncertainty. Critically, each of the stages onthe RHS of (3) or (4) can be split into sub-models.For example, multiple data sets with differing sup-ports can be accommodated with a model such as

[Z(1),Z(2)|Y, θZ(1) , θZ(2)](5)

= [Z(1)|Y, θZ(1) ][Z(2)|Y, θZ(2) ],

where Z(1) and Z(2) correspond to data sets (1)and (2), respectively, with associated parametersθZ(1) and θZ(2) . We note that (5) makes the assump-tion that, conditioned on the true process, both datasets are independent. This is often a reasonable sim-plifying assumption and greatly facilitates the com-bination of differing data sets (although this assump-tion should be verified in specific applications). Theparameters and distributional form for the two dis-tributions on the RHS of (5) can be quite different,perhaps accommodating differing types of spatial ortemporal support, and/or measurement and sam-pling errors. Examples of this in the context of thewind example discussed previously are data fromsatellite scatterometers as well as data from ocean

buoys or even weather center analysis winds (e.g.,Wikle et al. (2001)).It is also important to note that the process model

component on the RHS of (3) or (4) can also bedecomposed into subcomponents. This could corre-spond to a hierarchical Markov decomposition intime, for example, [Y ] =

∏Tt=1[Yt|Yt−1][Y0], where

Y = Y0, Y1, . . . , YT . Or, it could correspond to amultivariate decomposition, [Y ] = [Y (2)|Y (1)][Y (1)],where Y = Y (1), Y (2). In both cases, there wouldalso be process model parameters. Clearly, combina-tions of these types of distributions and other typesof dependencies (e.g., spatial, spatiotemporal, etc.)could be considered. In the context of the wind ex-ample, it would be natural for the wind componentsto be represented by Y (2) and these could be condi-tioned on near surface atmospheric pressure fields,say Y (1), both of which would be spatiotemporalprocesses with further conditioning possible.Last, we recognize that a huge advantage of hier-

archical models in complicated systems is that theparameters are themselves endowed with potentiallyquite complex distributions. That is, they exhibitmultivariate dependence between parameters or interms of space and time, or may themselves be de-pendent on exogenous information. Such complexdependencies in parameters are very difficult to ac-commodate in the classical paradigm. As discussedabove, in the wind example, it is critical to specifyparameter distributions that adhere to known em-pirical and theoretical turbulent scaling properties(e.g., Wikle et al. (2001)).There is no free lunch! Although the hierarchical

modeling paradigm is extremely powerful, it oftencomes at a substantial computational cost. In par-ticular, the normalizing constant in (3) involves theintegration over all random quantities in the model,which can correspond to a very high-dimensionalintegration in many applications. Given that ana-lytical solutions to these integrals are almost neveravailable in complex models, one has to resort to nu-merical methods. In the fully Bayesian context thisis typically some type of Markov Chain Monte Carloalgorithm (e.g., Robert and Casella (2004)). In theEHM case represented schematically in (2), one maybe able to work out alternative computational solu-tions [e.g., expectation-maximization (E–M) or nu-merical maximum likelihood or method-of-momentsestimation; e.g., see Chapter 7 of Cressie and Wikle(2011)]. It is critical to note that the complexityof these calculations often leads to modifications inmodel structure to facilitate practical computation.

Page 7: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 7

3. HIERARCHICAL MODELING AND

OCEANOGRAPHY

There have been many examples of hierarchicalmodeling in the ocean sciences since the late 1990s.In this section we briefly review some of that workat the interface of Statistics and Oceanography. Wefocus most attention on two data assimilation ex-amples (ocean vector winds and ocean tracer stateestimation), long lead forecasting of SST, and un-certainty quantification and assimilation of biogeo-chemical models. For these topics, we present vi-gnettes to illustrate the power of hierarchical mod-eling from problems we have worked on. We alsobriefly describe some of the work at the interfacerelated to other important oceanographic problems.

3.1 Data Assimilation and Inverse Modeling

Wikle and Berliner (2007) summarize data assim-ilation (DA) as an approach for optimally blendingobservations with prior information concerning thesystem (i.e., mathematical representations of mech-anistic relationships, model output, etc.) to obtaina distributional characterization of the true state ofthe system. Relevant to this overview paper, theseconcepts originated in the atmospheric and oceansciences and there is an extensive literature describ-ing various methodologies (Daley (1991); Ghil andMalanotte-Rizzoli (1991); Bennett (2002); Kalnay(2003); Lorenc (1986); Tarantola (1987); Wikle andBerliner (2007)). In essence, DA can be thought of asan inverse problem, and one can derive algorithmsfrom numerous perspectives, including optimal es-timation theory, variational analysis and Bayesianstatistics. The Bayesian approach to DA (e.g., Lorenc(1986); Tarantola (1987)) is helpful because the prob-lem is inherently hierarchical and, thus, it provides acoherent probabilistic approach that can be used todescribe the various approaches. In addition, if one iswilling to consider the BHM perspective, one can of-ten obtain a more realistic characterization of uncer-tainty in various components of the data and mech-anistic models (e.g., Wikle and Berliner (2007)).Most oceanographic processes of interest in the

DA context concern spatial fields that vary withtime. A general dynamical spatiotemporal model(DSTM) can be written in the BHM paradigm (e.g.,Cressie and Wikle (2011)). The data model is givenby

Zt(·) =H(Yt(·),θd(t),εt), t= 1, . . . , T,

where Zt(·) represents the data at time t, Yt(·) theassociated process, where the mapping function, H,

may be linear or nonlinear, the error process, εt,may be additive or multiplicative, and the modeldepends on parameters given by θd(t) that may bespatial or time-varying. The process model is givenby

Yt(·) =M(Yt−1(·),θp(t),ηt), t= 1,2, . . . ,

where the evolution operator M may be linear ornonlinear for the oceanographic process of interest,the error process, ηt, may be additive or multiplica-tive, and the parameters, θp(t), may be spatial ortime-varying. Note that this model is assumed to bevalid beyond the maximum data time-period (T ), sothat forecasting is appropriate. Finally, the model iscompleted by the specification of distributions forthe parameters in the previous stages, [θd(t)][θp(t)],where we have assumed that the parameters fromthe two stages are independent for convenience. Note,this hierarchical framework would also include a spec-ification of the initial process distribution, [Y0|θ0].The sequential modeling perspective is a naturalframework in which to consider the DA problem, asone can interpret the prior distribution from aboveas a forecast distribution, which gets updated givennew observations (e.g., see Cressie andWikle (2011)).

3.1.1 Ocean surface vector winds Surface windsdirectly transfer momentum to the ocean and sur-face wind speed modulates the exchanges of heatand fresh water to and from the upper ocean as mod-eled by bulk transfer formulae (e.g., Large (2006)).The advent of space-borne scatterometer instrumentsin the 1990s provided the first global wind fields,on daily timescales, from observations. Prior to thescatterometer era, ocean winds were inferred fromglobal weather forecast models and reanalyses (e.g.,Hellerman and Rosenstein (1983)) that dependedupon a very sparse global network of in situ windobservations from buoys and ships of opportunity.Reliable resolutions in the pre-scatterometer windfields were limited to ocean basin scale features (e.g.,associated with large-scale high and low pressuresystems) and monthly averages.The wind fields retrieved from scatterometer ob-

servations are not direct measures of the wind, butrather the observations are of the roughness im-parted on the ocean by surface capillary waves inresponse to the shear stress vector at the air–seainterface. The amplitudes and orientations of sur-face capillary waves are in equilibrium with the sur-face shear stress, and these amplitudes and orien-tations are retrievable from measures of how the

Page 8: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

8 WIKLE, MILLIFF, HERBEI AND LEEDS

capillary waves scatter impinging radar pulses ofknown frequency and polarization, that is, so-calledmicrowave backscatter cross-sections; for details, seeFreilich (1996). Radar backscatter cross-sections arespatially averaged over wind-vector cells and relatedto a surface vector wind (SVW) via a geophysi-cal model function. The SVW retrievals from scat-terometers are accurate to within at least 2 ms−1

in speed and 30 in direction. Depending on thesensor system and agency providing the data, res-olutions are on the order of 12.5–50 km for up to90% global coverage on daily timescales. The SVWretrievals occur in swaths along the polar-orbittingsatellite ground track. Swath widths vary by sys-tem from 600–1800 km, that is, between about 20and 100 wind-vector cells across a given swath de-pending on resolution. Because of the polar orbit(about 14 polar orbits per day) the swaths overlapat high latitudes and are separated by gaps in cov-erage at low latitudes, with the largest swath gapsoccurring at the equator. Again, depending on thesystem, the swath gaps at the equator are on the or-der of 10 degrees longitude and can take up to threedays to fill. The SVW from scatterometers resolvefeatures of the synoptic storms (e.g., fronts, conver-gences in rain bands, closed circulations, etc.) thatform, propagate and dissipate every day, over theworld ocean.To fill the gaps in the scatterometer winds, one

could make use of the complete (yet lower resolu-tion) wind fields from the operational meteorologi-cal centers. However, these wind fields have differ-ent properties than the scatterometer observations.Differences in the true resolutions of weather-centeranalyses and scatterometer winds are efficiently de-scribed in terms of surface wind kinetic energy spec-tra, that is, kinetic energy as a function of spatialwavenumber. Wikle, Milliff and Large (1999) showeda power-law relation for surface winds from multipledata sources in the tropical Pacific and the powerlaw behavior in surface winds from scatterometryhas been documented for the globe (Milliff et al.,1999). Milliff et al. (2011) compare kinetic energybetween weather center analyses and SVW retrievalsand find that the kinetic energy drops off unrealis-tically in weather center analyses at smaller scales.For example, at spatial scales on the order of 102 km,the weather center kinetic energy content is morethan an order of magnitude weaker than the SVWobservations. The goal of a statistical data assim-ilation is then to blend the complete, but energy-deficient, weather center analyses with the incom-

plete, yet energy-realistic, SVW in order to providespatially complete wind fields at sub-daily intervalswhile managing the uncertainties associated withthe different data sources and the blending proce-dure.Wikle et al. (2001) implemented a spatiotemporal

BHM for tropical winds in a region of the equato-rial western Pacific. As noted above, the inter-swathgaps in scatterometer coverage on daily timescalesare largest in the tropics. A BHM formulation toblend weather-center analyses and scatterometerwinds requires space–time properties that accountfor the greater intermittency in the scatterometerdata. This is achieved in the process model stage inWikle et al. (2001), which is posed in terms of threescientifically-motivated components as

Yt = µu +Φ(1)αut +Φ(2)βu

t ,(6)

whereYt is a vector of spatially-indexed zonal (east–west) wind velocity components (typically denotedby u). Analogous terms apply for the meridional ve-locity (v). Here, µu is the mean zonal wind processthat accounts for the prevailing wind and its vari-ance in the tropical domain, Φ(1) is a basis func-tion matrix corresponding to the large-scale modesof the equatorial beta-plane approximation (i.e., alinear, thin fluid approximation) to the momentum

equations, and Φ(2) are nested wavelet basis func-tions used to model fine-scale winds according toobserved kinetic energy-wavenumber properties forthe region. The use of scientifically-motivated ba-sis functions is crucial here, as we have informa-tion from the thin-fluid analytical approximation tothe wind field that is appropriate over the tropics.In particular, the leading equatorial normal modes[Φ(1)] are large-scale wave signals in the tropics, thatis, westward propagating Rossby waves, eastwardKelvin waves and mixed Rossby–Gravity waves (e.g.,Matsuno (1966)). Based on data analyses (e.g.,Wheeler and Kiladis (1999)), the equatorial normalmode basis functions were limited to the leadingmodes defined by two zonal wavenumbers and fourwavenumbers in the meridional direction (Wikle et al.,2001). These are sufficient to support the impor-tant basin-scale Kelvin, Rossby and mixed Rossby–Gravity waves that can be used to describe muchof the tropical dynamics within the Pacific basin.The amplitude coefficients, αu

t and βut , are treated

as random variables in the BHM, with first-orderMarkov models. The priors on the parameters asso-ciated with these models correspond to time series of

Page 9: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 9

Fig. 1. Tropical surface wind BHM after Wikle et al. (2001).Top panel is the NCEP reanalysis at the time tropical cy-clone Dale occupied the BHM domain. Color contours denoteconvergence (blues) and divergence (reds) in the surface windfield. Bottom panel is the posterior mean wind field (vectors)and convergence/divergence (colors) map for tropical cycloneDale from the tropical wind BHM. Strong convergences inrainband structures spiraling into the tropical cyclone centercoincide with coldest cloud-top temperatures from independentsatellite observations.

the theoretical amplitudes for each mode for the αut

coefficients and the slopes of the kinetic energy spec-trum for the βu

t coefficients as obtained from Wikle,Milliff and Large (1999). Data stage inputs were ob-tained from NSCAT SVW retrievals and from sealevel pressure and surface winds taken from reanal-yses of the National Centers for Environmental Pro-tection (NCEP).The posterior mean surface wind and surface con-

vergence/divergence exhibited variability on scalesnot achievable in the NCEP reanalyses. This is par-ticularly evident in a comparison of the NCEP re-analysis wind field with the posterior mean windsfrom the tropical wind BHM for the period whentropical cyclone Dale crossed the study area (Fig-ure 1). The posterior mean winds are organized intostrong convergence regions that spiral into the cen-ter of the tropical cyclone. These are consistent withrain band features of tropical cyclones and this wasconfirmed by comparing the posterior mean windswith cloud-top temperatures from an independentsatellite observation nearly coincident with the snap-

shot from the posterior distribution of the BHM (seeWikle et al. (2001)).Uncertainty management properties of SVWBHM

based on mechanisitic models (i.e., leading orderterms and/or approximations of the primitive equa-tions) are particularly relevant in ocean forecast set-tings. The Mediterranean Forecast System (MFS)is an operational system producing 10-day forecastsfor upper ocean fields every day. The MFS oceanforecast models resolve synoptic variability in theupper ocean. The uncertain parts of the forecastfields are at ocean mesoscales, that is, hourly and10–50 km scales. These are the scales of upper oceanhydrodynamic instabilities driven by the surface wind.So, modeling uncertainty in the surface wind fieldis a useful means of quantifying uncertainty in theMFS ocean forecasts on the scales that are most im-portant to daily users.Milliff et al. (2011) describe the details of the

SVW BHM for MFS, and Pinardi et al. (2011) re-view the impacts of BHM SVW fields in an ensem-ble forecast methodology built around realizationsfrom the posterior distribution for SVW from theBHM. The mechanistic process model in Milliff et al.(2011) involves the leading-order terms in a RayleighFriction equation approximation at synoptic scalesand, again, a nested wavelet basis model to representturbulent closure at the finest spatial scales. Thus,again it is critical to incorporate scientific informa-tion into the specification of the process rather thantry to model such behavior through the marginalcovariance structure. This then also allows specifi-cation of the measurement uncertainties associatedwith the data conditioned upon the true process.Specifically, data stage inputs are obtained from theQuikSCAT scatterometer SVW retrievals and fromweather center sea level pressure and surface windfields.Ten realizations of the posterior distribution for

SVW were used to drive ensemble data assimilationand ensemble forecasts in the MFS. Figure 2 depictsthe data assimilation system response in sea-surfaceheight (SSH) at the forecast initial condition time.The mean SSH initial condition (Figure 2, top) re-flects the accurate spatial scales for MFS forecastswhere synoptic variability overlies the general cir-culation patterns for the Mediterranean Sea. Sub-basin scale cyclonic gyres are represented by blue(depressed) SSH and anticyclones by reds. The gra-dients between blue and red signals are proportionalto surface current speeds. The spread in the SSH

Page 10: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

10 WIKLE, MILLIFF, HERBEI AND LEEDS

Fig. 2. Forecast initial condition mean (top) and spread(bottom) in the sea-surface height field computed from 10 en-semble members, each driven by a realization from the pos-terior distribution for the SVW BHM (Milliff et al., 2011).Ensemble spread is usefully localized in the most uncertainregions of the domain where wind forcing is driving hydro-dynamic instabilities and a local pulse in the mesoscale eddyfield (Pinardi et al., 2011).

initial condition (Figure 2, bottom) demonstratesthe value of managing uncertainty in the SVW. Thelargest amplitude signals in the spread are localizedin a few places only—where surface wind is drivinghydrodynamic instabilities and pulsing the energyin ocean mesoscale response.

3.1.2 Inversion of oceanographic tracer data In-ferring the structure of the circulation in the world’soceans is a significant part of our quest for under-

standing the climate. Direct measurements of thecirculation are difficult to obtain; however, in thepast decades, scientists have collected large amountsof hydrographic data (hydrostatic pressure, temper-ature, salinity, oxygen concentration, etc.). Such datahave been used to produce climatological maps thatexhibit the large scale structure (Lozier, Owens andCurry 1995). Based on these maps and data, onecan infer some constituents of the ocean circula-tion, such as boundary currents, wind-driven gyresand abyssal interior flow. Yet, hydrographic dataare typically available at very sparse locations inspace and time. For example, the World Ocean Cir-culation Experiment lasted from 1990 until 1998. InFigure 3 we illustrate these data that were availablefor the South Atlantic Ocean. Consequently, statis-tical smoothing methods are required to provide es-timates at space–time locations of interest. To thatend, one would typically combine data sets collectedat different times. This seems reasonable only underthe assumption that the distribution of dynamically-passive tracers is representative of a mean circula-tion and that such tracers are sufficiently stable ona fixed neutral density layer.A classical approach to inverting hydrographic data

into circulation structure is the box inverse method

(Wunsch, 1996). Starting with the thermal wind equa-tions, the problem reduces to estimation of a refer-ence velocity field. This is achieved via a large sys-tem of linear equations expressing conservation ofmass within a collection of connected boxes. Dueto the size of the problem, a high-dimensional re-gression approach is subsequently employed. Tra-ditionally, ridge regression estimators have proved

Fig. 3. Left: World Ocean Circulation Experiment era cruise tracks in the South Atlantic Ocean. Right: Interpolated oxygenmeasurements in a deep neutral layer of the South Atlantic Ocean.

Page 11: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 11

useful, however, modern statistical/machine learn-ing methods can provide more effective alternatives.Other approaches are based on more complex in-verse modeling and nonlinear optimization (Pailletand Mercier (1997); Zika, McDougall and Sloyan(2010); Zhang and Hogg (1992); Wunsch (1994)).The inverse problem described here is ill-posed andsome type of regularization is required (Kirsch, 1996).We describe the Bayesian inversion approach of McK-eague et al. (2005) (see also Herbei, McKeague andSpeer (2008)), which connects the data (tracer con-centrations) to parameters (ocean circulation) us-ing a system of partial differential equations. In thiscase, regularization is imposed through a prior dis-tribution over the parameters. The solution is theposterior distribution. It can be used to select “rep-resentative” values and the associated uncertaintyfor any aspect of the circulation.Formally, let C = C(x, y) denote the concentra-

tion of a tracer of interest at a location (x, y) in arectangular domain Ω ⊂ R

2 representing the layerof the ocean being studied. The problem is to esti-mate the (horizontal) water velocities and diffusioncoefficients based on noisy measurements of C at asparse set of locations in Ω. In the right panel ofFigure 3 we display an interpolated map of oxygenconcentration for a 2000 m deep layer in the SouthAtlantic Ocean. The connection between the tracerconcentration C and the velocities and diffusion co-efficients is modeled by the steady-state advection-diffusion equation

u · ∇C =∇ · (K∇C) +QC , (x, y) ∈Ω,(7)

where u= (u, v) is the horizontal water velocity, thediagonal diffusivity matrix K = diag(κ(x), κ(y)) doesnot vary with location, and the sink term QC =−λC is present only when C represents oxygen con-centration. Equation (7) is augmented with Dirich-let boundary conditions C =C∂Ω when (x, y) ∈ ∂Ω,where ∂Ω denotes the boundary of Ω.The statistical model assumes additive observa-

tional error

Cobsj (xi, yi) =Cj(xi, yi) + εj(xi, yi).

Here, j = 1, . . . , nC indexes a particular tracer andi = 1, . . . ,Nj indexes a spatial location where datafor tracer j are available. The underlying tracer con-centration C(x, y) is obtained as a solution of the ad-vection diffusion equation (7). As this solution is notavailable in closed form, one uses a numerical (grid-based) approximation. We collect all quantities of

interest (velocities, diffusion coefficients, boundaryvalues) in a high-dimensional vector γ. Under theassumption that the measurement errors ε(·) are un-biased Gaussian variables with constant (yet tracer-dependent) variance, the posterior distribution iswritten as

π(γ|Cobs)

nC∏

j=1

Nj∏

i=1

exp

−(C(γ;xi, yi)−Cobs(xi, yi))

2

2σ2j

(8)

· π(γ),

where π(γ) represents the selected prior distribu-tion. The posterior probability model (8) is exploredvia Monte Carlo methods (e.g., Robert and Casella(2004)).The ill-posedness mentioned above is resolved by

specifying a proper prior distribution π(γ). This im-plies that the posterior distribution is proper. How-ever, an efficient MCMC approach is still required.In addition, one can design specific sampling strate-gies as described in McKeague et al. (2005) andHerbei, McKeague and Speer (2008). It is impor-tant to understand that each component describedabove (physical model, prior distribution, likelihoodfunction, data) plays a crucial role in determiningthe solution. Under the Bayesian approach, a sen-sitivity analysis, although possible, is hampered bythe immense computational cost associated with theMCMC sampler. In addition, the under-determina-tion problem is present in this case. There are(roughly) 300 data points, while there are thousandsof parameters to estimate (velocities, diffusions,boundary values). In the results given in McKeagueet al. (2005) and Herbei, McKeague and Speer (2008),not all estimated velocities are significantly differentfrom zero, which is the prior mean. However, thedata are informative about the large-scale features(zonal jets, gyres). The posterior mean velocitiesare compared with velocities determined from floatdata (Hogg and Owens, 1999), showing a comfort-ing consistency (McKeague et al. (2005) and Herbei,McKeague and Speer (2008)).

3.2 Long Lead Forecasting: SST

Tropical Pacific SST exhibits some of the mostimportant variability on inter-annual time scales forthe ocean/atmosphere system (e.g., see the overviewin Philander (1990)). This variation arises from com-plicated interactions, across a large range of spa-tiotemporal scales, between the ocean and the at-mosphere. The most prominent signal on these time

Page 12: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

12 WIKLE, MILLIFF, HERBEI AND LEEDS

scales is the El Nino-Southern Oscillation (ENSO)phenomenon. This is the well-known quasi-periodic(3–5 year period) warming (El Nino) and cooling (LaNina) in the tropical Pacific. These warming andcooling events lead to dramatic effects in weatheracross the globe due to teleconnections with theglobal atmospheric circulation. Because of these sig-nificant weather related impacts (e.g., droughts,floods, etc.), it is critical to be able to forecast sev-eral months in advance the possible developmentand transition of these events. Increasingly, such“long lead” forecasts have shown useful skill andcorrespond to one of the few situations in oceanscience where a purely statistical forecast method-ology is competitive with, and in many cases bet-ter than, equivalent deterministic model forecasts(Barnston, He and Glantz (1999); van Oldenborghet al. (2005)).At typical spatial resolutions, there can easily be

several thousand gridded spatial locations corre-sponding to the tropical Pacific region of forecast-ing interest. Complicated spatiotemporal statisticalmodels are difficult or impossible to implement atthese dimensions. For this reason, statistically-basedmodels for tropical SST have been “spectral” in thesense that they are based on coefficients associatedwith a projection of the SST fields on spatial ba-sis functions. The associated projection coefficientsthat are evolved are typically a reduced set, usu-ally corresponding to larger modes. That is, let Yt

correspond to an n-dimensional vector of the trueSST process at n spatial locations and time t andconsider the decomposition of this process vector

Yt =Φ(1)αt +Φ(2)βt,(9)

where Φ(i), i = 1,2, is an n × pi matrix of spatialbasis functions, and αt and βt are the associatedexpansion coefficients. A first-order linear Markovassumption on the evolution of the coefficients αt,for example, αt+τ = Mαt + ηt+τ , for appropriatetime increment, τ , and with Gaussian errors, ηt ∼Gau(0,Q), was shown in the early 1990s to be amodel with reasonable skill at long-lead forecasting(e.g., Penland and Magorian (1993)). As with thetropical wind case presented previously, the specifi-cation of the process in terms of two basis sets is crit-ical from a scientific perspective. In particular, it isthought that the active dynamical process is drivenby a lower dimensional manifold (corresponding tothe first set of basis functions) and, yet, the resid-ual spatially-dependent structures captured by the

second set of basis functions remain an importantsource of variability.Although dimension-reduced linear Markov mod-

els showed reasonable skill, the ENSO phenomenonis better characterized as a nonlinear process (e.g.,Hoerling, Kumar and Zhong (1997); Burgers andStephenson (1999)). Nonlinear statistical models typ-ically are better at capturing the magnitude of thepredicted El Nino and La Nina events (Tanganget al. (1998); Berliner, Wikle and Cressie (2000);Tang et al. (2000); Timmermann, Voss and Pasman-ter (2001); Kondrashov et al. (2005)). Most nonlin-ear stochastic methods have not successfully char-acterized the uncertainty in the forecasts. An ex-ception to this is the model of Berliner, Wikle andCressie (2000), who use a dimension-reduced thresh-old Markov model with certain components gov-erned by the onset of intra-seasonal oscillations (so-called “westerly wind bursts”).There are several key components of the model

in Berliner, Wikle and Cressie (2000) that illustratethe power of hierarchical modeling. First, the datamodel is constructed

Zt =Φ(1)αt + εt, εt ∼Gau(0,R),(10)

where the data vector, Zt, is very high-dimensional(say, n× 1), and Φ(1) is a reduced-dimension set ofspatial basis functions as described above (an n×p1matrix). It is important to note that the remain-

ing basis functions (e.g., Φ(2) from above) associ-ated with the basis expansion are used to param-eterize the observational spatial covariance matrix,R, which now accounts for both observation errorand the truncation. The assumption is that the ac-tive dynamics exist on the lower dimensional mani-fold represented by Φ(1) and, like in typical turbu-lence parameterizations, the small scale structuresaccounted for by the Φ(2) portion of the expansionare associated with nonpredictive covariability.Critically, the evolution of the dynamical process

is nonlinear, but conditionally linear. Specifically,

αt+τ =µt+Mtαt+ηt+τ , ηt ∼Gau(0,Q),(11)

where prior distributions are also specified for αt :t= 1, . . . , τ. The important modeling assumption isthat Mt and µt depend on both the current (timet) and future (time t + τ ) climate “regimes,” thatis, Mt =M(It, Jt) and µt = µ(It, Jt), where It clas-sifies the current SST regime as “cool,” “normal”or “warm,” and Jt anticipates one of these threeregimes at time t + τ . Specifically, It is based on

Page 13: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 13

the Southern Oscillation Index and Jt is based ona latent variable that is modeled, hierarchically, interms of the east-west component of the near sur-face wind in the Western Pacific ocean (for details,see Berliner, Wikle and Cressie (2000)). The pointis that this leads to nine potential mean states andnine potential dynamical operators, accommodatingthe nonlinear transition between ENSO phases.Because this model was developed in a Bayesian

hierarchical framework, it was able to capture theuncertainty characterized by the data, process andparameters. However, despite the fact that this modelused physical notions (i.e., westerly wind bursts) inthe lower stages of the model hierarchy, it was un-able to directly account for quadratic interactionsacross spatial scales. As mentioned previously, non-linearity is important in most atmosphere and oceanprocesses. Kondrashov et al. (2005) demonstrate theeffectiveness of a quadratic nonlinear model for long-lead prediction of ENSO from a classical regressionperspective. Wikle and Hooten (2010) demonstratethe implementation of a quadratic nonlinear modelin the context of a Bayesian hierarchical frameworkin which the random nonlinear process is “hidden”and parameters are random as well. They illustratethat simple arguments from turbulence scaling sup-port the form of this model and further suggest adimension reduction strategy.As defined in Wikle and Hooten (2010), general

quadratic nonlinearity (GQN) with respect to theαt process can be written

αt+τ (i)

=

p∑

j=1

mLijαt(j)(12)

+

p∑

k=1

p∑

l=1

mQi,klαt(k)g(αt(l);θg) + ηt+τ (i)

for i= 1, . . . , p, where g(·) is some transformation ofαt that depends on parameters θg and gives the pro-cess more generality than the simple dyadic interac-tions alone. This framework is exceptionally flexi-ble in accommodating many real-world mechanisticprocesses, but it comes at the cost of a curse of di-mensionality in the parameter space; that is, thereare O(p3) parameters corresponding to the nonlin-

ear coefficients (mQi,kl), in addition to the linear co-

efficients (mLij) and θg.

One can use scale analysis to help with the dimen-sionality concerns, and also to motivate components

of the hierarchical model. Specifically, as before, as-sume we can decompose the spectral coefficients intolarge-scale components (αt) and small-scale coeffi-cients (βt) corresponding to the spatial basis func-

tions Φ(1) and Φ(2) discussed above. Consider allof the possible dyadic interactions of the elementsof this vector—that is, there are small-scale/small-scale, small-scale/large-scale and large-scale/large-scale interactions. Loosely motivated by “Reynoldsaveraging” in turbulence theory (e.g., Holton (2004)),Wikle and Hooten (2010) suggest considering thelarge-scale/large-scale pairwise interactions explic-itly, with the small-scale/small-scale interactions con-tributing to a correlated dependence structure in theadditive error, and the small-scale/large-scale inter-actions corresponding to the linear term in the large-scale coefficients with random parameters (i.e., thesmall-scale coefficients play the role of random coef-ficients in the interaction). Such an argument leadsto a quadratic nonlinear model on the large-scalecoefficients, which can be written in matrix form as

αt+τ =MLαt + (Ip1 ⊗α′

t)MQαt + ηt+τ ,(13)

ηt ∼Gau(0,Q).

Prior distributions are then given to the parametersin ML and MQ as well as the covariance matrix Q

(e.g., Wikle and Hooten (2010)). This approach wasshown to quite reasonably account for the uncer-tainties so that the prediction error bounds coveredthe extreme ENSO events, even when the forecastsdid not adequately capture the true magnitude.Critically, these quadratic nonlinear implementa-

tions still suffer from a fairly high curse of dimen-sionality in the parameters. In this sense, variousmodel reduction approaches have to be implementedthat necessarily fail to account for as much modeluncertainty as is likely present. It is difficult to knowa priori which quadratic interactions are important.To alleviate this curse of dimensionality and to pro-vide a natural framework in which to “average” acrossthe various model specifications, Wikle and Holan(2011) employed a stochastic search variable selec-tion methodology (e.g., George and McCulloch, 1993,1997).In general, although the GQN statistical models

are very flexible in representing oceanographic pro-cesses, these models can easily experience finite-timeblow up (i.e., explosive growth) when fit to data(Majda and Yuan, 2012). This is seldom an issuewhen one is using these models for data assimila-tion, given the presence of observations to act as

Page 14: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

14 WIKLE, MILLIFF, HERBEI AND LEEDS

a control (e.g., Leeds et al. (2013)), nor is it typi-cally a problem when one is only forecasting out onetime step (e.g., as in the SST example). It can be aproblem when multiple time steps are forecast thatrequire some form of constraint, either statistical orphysical (Majda and Harlim, 2013).

3.3 Biogeochemical Models

Analysis of marine ecosystem dynamics involvesvarious sources of uncertainty in the observations,the underlying scientific process and the parame-ters that describe the process dynamics. Critically,it is often the case that the observation errors arenon-Gaussian and that the process being modeled isnonlinear, that is, there is an explicit system of cou-pled nonlinear differential equations that describethe complex ecosystem dynamics. As a result of theseuncertainties and complexities, the BHM frameworkis natural, but can be difficult to implement.Soon after the introduction of MCMC methods

into Bayesian computation, the BHM approach wasused in data assimilation for marine ecosystem mod-els. Anticipating the forthcoming increase in re-motely-sensed ocean color observations, Harmon andChallenor (1997) implemented an MCMC-based sam-pling protocol that explored the ability to recovervarious model parameters in a seven-compartmentmarine ecosystem model with and without addedmodel noise. They noted the computational diffi-culties required to adequately account for correla-tion in these parameters. While identifying corre-lation in only ten parameters may no longer be acomputational issue, there are still issues related tothe sampling methodology that adequately gener-ates reasonable block proposals for the entire pa-rameter vector. Adaptive Metropolis–Hastings al-gorithms, which update the covariance of the pro-posal distribution, may be useful when trying togenerate adequate proposals for a nonlinear marineecosystem model (see, e.g., Haario, Saksman andTamminen (2001)). Current research has taken intoaccount advances in sampling methodology, includ-ing the following: sequential importance resampling,particle filters, ensemble Kalman filters and state-augmentation approaches (Evensen, 1994, 2009;Dowd, 2006, 2007, 2011; Parslow et al. (2013); Stroudet al. (2010)).An innovative practice in statistical modeling of

ocean ecosystems is the inclusion of information fromdeterministic, mechanistic models into the statis-tical framework, typically in the process stage of

a BHM. However, the necessary estimation proce-dures often require iteratively running the mecha-nistic model, which poses a problem when the modelis computationally expensive and can only be run avery limited number of times. In certain situationswhere the computer model is too computationallyexpensive to run a sufficient number of times forthe desired analysis, statistical surrogates (i.e., em-ulators) are used. In its simplest form, an emulatoris simply the resulting estimated statistical modelwhen computer model output is used as data. Then,this model is used to predict the output of the com-puter model under untried input settings (e.g., ini-tial conditions, parameter values, forcings).Traditionally, so-called “computer model” emula-

tion has been done using Gaussian Process (GP)models (e.g., Sacks et al. (1989); Currin et al. (1991)).In our case, these “computer models” are determin-istic ocean forward models. GP emulators are re-lated to spatial GPs, which use a correlation func-tion such that the model output is more highly cor-related for inputs that are “nearer” to one anotherin a given sense than those that are “farther apart.”However, because ocean ecosystem models are non-linear, a GP emulator of the joint distribution of theoutput may be inappropriate (as a nonlinear processcannot be specified by only two moments). In thiscase, one could consider the use of a dynamic GPemulator, which considers the value of the processat the current time step (i.e., the initial conditions)as an input to the ocean forward model (and GPemulator). This offers several benefits over the tra-ditional GP emulation approach (e.g., Conti et al.(2009); Liu and West (2009)).Rather than modeling the dynamics through a co-

variance function in a GP, it may be more appro-priate to model the output using first-order char-acteristics (as the computer models themselves arewritten). The use of so-called “first-order emulators”has appeared in several applications, for example,van der Merwe et al. (2007) and Frolov et al. (2009)use neural networks and Hooten et al. (2011) userandom forests, a machine learning algorithm. Thereis also the potential to use other nonlinear statisti-cal models, polynomial chaos expansions (Mattern,Fennel and Dowd, 2012) or GP models in a dynam-ical context (Margvelashvili and Campbell, 2012).These approaches use dimension-reduction techniquesto overcome the curse of dimensionality, with theemulator describing the dynamics of the reduced-dimensional process. Most of these implementations

Page 15: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 15

use nonparametric approaches. Alternatively, Leeds,Wikle and Fiechter (2013) use an emulator basedon a reduced-rank multivariate quadratic nonlinearstatistical model as described above.Related to the use of emulators for dynamic mod-

els, there is also a growing body of work consider-ing the use of emulators for spatial or spatiotempo-ral forward model output. Marine ecosystem mod-els can vary from nonspatial, 0-D models, to modelsthat have 3-D spatial structure. Leeds et al. (2013)used a 1-D (in the vertical) four-component modelthat included nutrients, zooplankton, phytoplank-ton and detritus (e.g., a “1-D NZPD” model) withiron limitation (i.e., “1-D NPZDFe”) to model a 3-Dprocess by creating a forest of 1-D models (morespecifically, a forest of emulators of the 1-D model).These 1-D models resolve vertical processes, but donot account for horizontal diffusion and advection.Variability resulting from horizontal advection anddiffusion is accommodated by putting a spatial GPmodel on the parameters (inputs) to the 1-DNPZDFe model.In certain circumstances, the forward model emu-

lator is developed simply to learn about the behaviorof the forward model itself. However, in other situa-tions, the forward model emulator may be useful ina data assimilation context. In those circumstances,the emulator may be used in a two-stage approach,where the emulator is fit “off-line,” and then is usedinside a Bayesian hierarchical model in the place ofthe forward model itself. Calder et al. (2011) andLeeds, Wikle and Fiechter (2013) developed BHMsusing forward model output as data.Leeds, Wikle and Fiechter (2013) performed data

assimilation using SeaWiFS ocean color observations,as well as sea surface height and SST output fromthe Regional Ocean Model System (ROMS) forwardmodel coupled with the NPZDFe model (Fiechterand Moore (2012)). They consider a reduced-dimen-sion quadratic nonlinear process model similar to(9) and (13), but where Yt = (Y′

1,t,Y′

2,t,Y′

3,t)′, rep-

resenting the three state variables of interest. Basisfunctions Φ(1) and Φ(2) were based on output fromthe ROMS-NPZDFe model for a different time pe-riod using a singular value decomposition and thepriors for ML and MQ were developed using theremaining right singular vectors. Leeds, Wikle andFiechter (2013) show that this approach was suffi-cient to accommodate nonlinear dynamics even inthe absence of observations over a substantial por-tion of the domain as shown in Figure 4.

3.4 Other Important Problems

Space limitations prevent us from giving completeoverviews of all hierarchical statistical models inoceanography. However, we mention here some no-table and important papers that have been pub-lished in recent years in areas outside of those men-tioned above. By necessity, this list is not exhaus-tive, neither in terms of topics nor citations withintopics, but it does give some sense of the breadthand depth of work being done in the area.As mentioned above, the amount of in situ data

for ocean state variables is fairly limited and thereis a need to construct complete spatial fields rep-resenting climatological features of the ocean state,as well as the uncertainty in those states using spa-tial and spatiotemporal models. There are many no-table examples from a BHM perspective (e.g., Hig-don (1998); Lavine and Lozier (1999); Lemosand Sanso, 2009, 2012; Lemos, Sanso and Santos(2009)).It has long been known that there are strong tele-

connections between the ocean and the atmosphereand this has led to useful statistical models for pre-diction (e.g., Davis (1976); Barnett (1981); Barnettand Preisendorfer (1987)). There is an increasinglylarge number of papers that have used the BHMperspective to model teleconnections as well as theirimpacts. For example, Wikle and Anderson (2003)used such a model when evaluating the impacts ofSST on tornado report counts in the eastern two-thirds of the US; Elsner, Murnane and Jagger (2006)(and references therein) have shown great success inmodeling hurricane activity and forecasts (see alsoFlay and Nott (2007)); Lima and Lall (2009) use aBHM to model precipitation considering ocean con-ditions. In addition, there are recent studies show-ing linkages between the physical ocean and ecolog-ical impacts (e.g., Cloern et al. (2010); Ruiz et al.(2009)).Bayesian hierarchical models have long been used

to model the higher trophic levels of the marineecosystem. For example, stock-recruitment modelshave been an important tool in marine fisheries man-agement (e.g., Thompson (1992); Hilborn, Pikitchand McAllister (1994); McAllister and Kirkwood(1998); Dorn (2002); Michielsens and McAllister(2004); Hirst et al. (2005)). In addition, BHMs havebeen used extensively in recent years to model dis-tributions and movement of marine mammals (e.g.,Jonsen, Myers and James (2007); Ver Hoef and

Page 16: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

16 WIKLE, MILLIFF, HERBEI AND LEEDS

Fig. 4. A plot of SeaWiFS ocean color observations (top row), ROMS-NPZDFe output for phytoplankton (second row), theposterior mean for phytoplankton (third row) and the posterior standard deviation (bottom row), for three consecutive eight-daytime periods, corresponding to May 16, May 24 and June 1, 2002, respectively. Note that observations were log-transformedand that the MCMC was run using these log-transformed observations.

Jansen (2007); Johnson et al. (2008); Cressie et al.(2009); Hanks et al. (2011); Moore and Barlow (2011);Conn et al. (2012); Hiruki-Raring et al. (2012); Mc-Clintock et al. (2012)).Clearly, a critical problem of vast societal interest

concerns climate change. Given the role of the oceanin the climate system and its direct connections toweather events and ecological impacts, oceanic cli-

mate change and its uncertainty characterization areextremely important. This is a vast research areawith many papers that have taken a hierarchicalBayesian approach. This topic is beyond the scopeof this review, but a few notable examples includeTebaldi et al. (2005), Furrer et al. (2007), Tebaldiand Sanso (2009), Aldrin et al. (2012) and Satterth-waite et al. (2012).

Page 17: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 17

4. CURRENT AND FUTURE CHALLENGES

As demonstrated throughout this review, the oceansystem is quite complex with numerous interact-ing subcomponents and external processes, and withsubstantial uncertainty in terms of knowledge anddata. Statistical methods have been critically im-portant to improve our understanding of this sys-tem and to characterize uncertainty. In recent years,the hierarchical statistical modeling paradigm hasproven to be an exceptionally useful tool to man-age the various sources of uncertainty. We have onlyjust scratched the surface in terms of our presenta-tion of the work that has been done in this area,but it is clear that the use of hierarchical modelingin oceanography has blossomed. That being said, inaddition to continued development in the areas men-tioned above, there are still many important prob-lems at the interface of Statistics and Oceanographyto be considered and methodologies to be explored.One of the biggest challenges for statisticians work-

ing in Oceanography is to develop statistical mod-els that can effectively parameterize the complexnonlinear interactions that are associated with theocean system. In particular, models must be de-veloped to account for potentially high-dimensionalstate-processes, and yet effectively manage uncer-tainty with varying data quality and non-Gaussianerrors. In general, to be effective in this context, di-mension reduction should not be independent of thephysical and biological environment of the problemunder consideration.One component of nonlinearity that is important

concerns the coupling of subsystems, whether thatbe the atmosphere/ocean interface, the ocean/ice in-terface, the physical/biological interface or interac-tions between trophic levels in the ocean. Many ofthese couplings require parameterizations for fluxesacross boundaries, and there is often only limited ob-servational data available to help inform the process.This presents an extreme challenge and one in whichwe must use the large amounts of remotely sensedobservations along with the sparse in situ data tobuild these nonlinear relationships.Not unrelated is the notion of accounting for and

describing model error. The precise characterizationof uncertainty that is fundamental in BHM can beused to help identify and characterize model errorin deterministic models of the ocean system, for ex-ample, ocean forecast models. As with all modelabstractions, deterministic approaches accept trade-offs in resolution and approximations of the ocean

variability to gain affordability in simulations andforecasts. Because the ocean system and the mod-els are inherently nonlinear, the errors introducedby acceptable approximations can grow and lead tomodel error properties that are difficult to diagnose.Parameters of posterior distributions from indepen-dent BHM analyses for ocean model response orcontrol variables can be used as a standard againstwhich deterministic model error can be quantified.For example, if the incremental adjustment of thesurface momentum flux control vector in a varia-tional data assimilation procedure pushes the mo-mentum flux at a point outside the reasonable boundsof a posterior distribution for momentum flux froma BHM, the variational procedure is probably com-pensating for forecast model error as opposed to un-certainty in the control vector.To gain understanding of the impact of physical

processes on biota as well as the interaction of bi-ological organisms, there is increasing reliance onindividual (or agent-based) models. These modelshave become more prevalent in the ecological realmand are starting to be considered from the BHM per-spective (e.g., Hooten and Wikle (2010)). The use ofthese models in a statistical context across the spec-trum of scales and processes involved in oceanog-raphy is a growing area of interest. For example,Megrey et al. (2007) link physical ocean models toindividual-based bioenergetic models for fish.The issues described above will almost certainly

require new computational strategies, particularlyin the Bayesian paradigm. As the models get moreand more complex and larger data sets become avail-able, standard MCMC methods begin to fail to pro-vide meaningful results. Among the many challengesassociated with Bayesian computing, two stand out:(1) there are very high-dimensional distributions tobe explored, and (2) the models are complex, whichleads to inexact and sometimes impossible likeli-hood evaluations. Novel MCMC methodology willaddress these issues, while maintaining feasibility.For example, particle MCMC (PMCMC) methodsare designed to address the first issue. For high-dimensional posterior distributions (thousands ofstate variables and parameters), it is nearly impos-sible to design good Metropolis–Hastings proposaldistributions and in this case, even adaptive MCMCmay fail. Andrieu, Doucet and Holenstein (2010)propose to use sequential Monte Carlo (SMC) meth-ods combined with importance sampling to designnear optimal proposal distributions. The resulting

Page 18: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

18 WIKLE, MILLIFF, HERBEI AND LEEDS

algorithm, which has similar features to a particlefilter, will update the entire collection of state vari-ables at once and can be extremely useful for space–time models. Parslow et al. (2013) use PMCMC forstate and parameter estimation as well as state fore-casting in the context of a marine biogeochemicalmodel. In addition, Hamiltonian MCMC method-ologies may provide an attractive alternative for sit-uations for which samples from complicated high-dimensional processes and parameter distributionsare required (e.g., Beskos, Kalogeropoulos and Pa-zos (2013)).Approximate Bayes Computing (ABC) methods

are a new and emerging class of likelihood-freeMCMC algorithms. They address the second issueabove—that is, cases when evaluation of the like-lihood function involves integrals over very largespaces that are impossible to calculate (e.g., Beau-mont, Zhang and Balding (2002)). ABC relies onthe ability to simulate from the selected model with-out much computing effort. Consequently, the useris forced to select a relevant (multivariate) statisticand “posterior samples” are defined as parametervalues that result in a statistic similar to the oneobserved. This algorithm, in fact, explores the distri-bution of the variables of interest conditional on the

selected statistic, not the data. Although this mayraise some concern regarding the interpretation ofthe results, Fearnhead and Prangle (2012) describea semi-automatic way of selecting good summarystatistics in a general setting.In general, it seems likely that for inference and

prediction of complicated oceanographic processesin the foreseeable future, one will have to continue tomake compromises between model complexity andcomputational feasibility. Indeed, one must be will-ing to accept some reasonable lack of “optimality”in the solution in order to make headway on manyof these problems. This is true regardless of whetherone takes a Bayesian or frequentist perspective. Thehierarchical paradigm helps to some extent as it al-lows one to consider trade-offs between approximatecomputational strategies, incorporation of scientificinformation and model specification for the differentmodel components (data, process and parameters)separately.In conclusion, there is a long history of activity at

the interface of Statistics and Oceanography. In re-cent years, the hierarchical statistical paradigm hasproven to be very helpful for managing the uncer-tainty associated with data, process and parame-ters in modeling the ocean and its related systems.

There are increasingly more oceanographers withadvanced statistical training and more statisticianswith oceanographic backgrounds and this is sure tolead to even more innovative and exciting method-ological developments in years to come.

ACKNOWLEDGMENTS

The authors wish to thank the Editor, AE and re-viewer who provided substantial suggestions to im-prove the manuscript. In addition, Wikle acknowl-edges the support of National Science Foundation(NSF) Grants DMS-10-49093, OCE-0814934 and Of-fice of Naval Research (ONR) Grant ONR-N00014-10-0518. Herbei acknowledges support from NSFGrant DMS-12-09142 and ONR Grant N00014-10-1-0488. Milliff acknowledges early support for BHMresearch from the Physical Oceanography Program,ONR; from the NASA, International Ocean Vec-tor Winds Science Team, and recent support fromUS Globec under Grant number NSF OCE-0815030.It is a pleasure to acknowledge our BHM mentorL. Mark Berliner and the inspirations he providesdating back to his days as Director, GeophysicalStatistics Project at NCAR through the present.We are also pleased to acknowledge our colleaguesand collaborators who participate in vigorous cross-disciplinary discussions at summer Confabs in Boul-der, CO. In recent years, these include, ProfessorsBerliner, Mevin B. Hooten, Andrew M. Moore, Na-dia Pinardi and Thomas M. Powell, and DoctorsJeremiah Brown, Manny Fiadeiro and Jerome Fiech-ter.

REFERENCES

Aldrin, M., Holden, M., Guttorp, P., Skeie, R. B.,Myhre, G. and Berntsen, T. K. (2012). Bayesian es-timation of climate sensitivity based on a simple climatemodel fitted to observations of hemispheric temperaturesand global ocean heat content. Environmetrics 23 253–271.MR2914207

Andrieu, C., Doucet, A. and Holenstein, R. (2010). Par-ticle Markov chain Monte Carlo methods. J. R. Stat. Soc.Ser. B Stat. Methodol. 72 269–342. MR2758115

Banerjee, S., Carlin, B. and Gelfand, A. (2003). Hierar-chical Modeling and Analysis for Spatial Data 101. Chap-man & Hall/CRC, Boca Raton.

Barnett, T. (1981). Statistical prediction of North Amer-ican air temperatures from Pacific predictors. MonthlyWeather Review 109 1021–1041.

Barnett, T. and Preisendorfer, R. (1987). Origins andlevels of monthly and seasonal forecast skill for UnitedStates surface air temperatures determined by canonicalcorrelation analysis. Monthly Weather Review 115 1825–1850.

Page 19: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 19

Barnston, A., He, Y. and Glantz, M. (1999). Predictiveskill of statistical and dynamical climate models in SSTforecasts during the 1997–98 El Nino episode and the 1998La Nina onset. Bulletin of the American Meteorological So-ciety 80 217–243.

Beaumont, M. A., Zhang, W. and Balding, D. J. (2002).Approximate Bayesian computation in population genet-ics. Genetics 162 2025–2035.

Bennett, A. F. (2002). Inverse Modeling of the Oceanand Atmosphere. Cambridge Univ. Press, Cambridge.MR1920432

Berliner, L. M. (1996). Hierarchical Bayesian time se-ries models. In Maximum Entropy and Bayesian Methods(Santa Fe, NM, 1995). Fund. Theories Phys. 79 15–22.Kluwer Academic, Dordrecht. MR1446713

Berliner, L., Wikle, C. and Cressie, N. (2000). Long-leadprediction of Pacific SSTs via Bayesian dynamic modeling.Journal of Climate 13 3953–3968.

Beskos, A., Kalogeropoulos, K. and Pazos, E. (2013).Advanced MCMC methods for sampling on diffusionpathspace. Stochastic Process. Appl. 123 1415–1453.MR3016228

Burgers, G. and Stephenson, D. (1999). The “normality”of El Nino. Geophys. Res. Lett. 26 1027–1030.

Calder, C., Berrett, C., Shi, T., Xiao, N. andMunroe, D. (2011). Modeling space–time dynamics ofaerosols using satellite data and atmospheric transportmodel output. J. Agric. Biol. Environ. Stat. 16 495–512.

Chelton, D. (1994). Physical oceanography: A brief overviewfor statisticians. Statist. Sci. 9 150–166.

Clark, J. S. (2007). Models for Ecological Data: An Intro-duction. Princeton Univ. Press, Princeton, NJ. MR2292764

Cloern, J., Hieb, K., Jacobson, T., Sanso, B.,Di Lorenzo, E., Stacey, M., Largier, J., Meiring, W.,Peterson, W. and Powell, T. et al. (2010). Biologicalcommunities in San Francisco Bay track large-scale cli-mate forcing over the North Pacific. Geophys. Res. Lett.37 L21602.

Conn, P., Johnson, D., London, J. and Boveng, P.

(2012). Accounting for missing data when assessing avail-ability in animal population surveys: An application to ice-associated seals in the Bering Sea. Methods in Ecology andEvolution 3 1039–1046.

Conti, S., Gosling, J. P., Oakley, J. E. and O’Hagan, A.

(2009). Gaussian process emulation of dynamic computercodes. Biometrika 96 663–676. MR2538764

Cressie, N. and Wikle, C. K. (2011). Statistics for Spatio-temporal Data. Wiley, Hoboken, NJ. MR2848400

Cressie, N., Calder, C., Clark, J., Hoef, J. andWikle, C. (2009). Accounting for uncertainty in ecolog-ical analysis: The strengths and limitations of hierarchicalstatistical modeling. Ecological Applications 19 553–570.

Currin, C., Mitchell, T., Morris, M. and Ylvisaker, D.

(1991). Bayesian prediction of deterministic functions, withapplications to the design and analysis of computer exper-iments. J. Amer. Statist. Assoc. 86 953–963. MR1146343

Daley, R. (1991). Atmospheric Data Analysis. CambridgeAtmospheric and Space Science Series 4 57. CambridgeUniv. Press, Cambridge.

Davis, R. (1976). Predictability of sea surface temperature

and sea level pressure anomalies over the North Pacific

Ocean. Journal of Physical Oceanography 6 249–266.

Dorn, M. (2002). Advice on West Coast rockfish harvest

rates from Bayesian meta-analysis of stock-recruit relation-

ships. North American Journal of Fisheries Management

22 280–300.

Dowd, M. (2006). A sequential Monte Carlo approach for

marine ecological prediction. Environmetrics 17 435–455.

MR2240936

Dowd, M. (2007). Bayesian statistical data assimilation for

ecosystem models using Markov Chain Monte Carlo. Jour-

nal of Marine Systems 68 439–456.

Dowd, M. (2011). Estimating parameters for a stochastic

dynamic marine ecological system. Environmetrics 22 501–

515. MR2843404

Elsner, J., Murnane, R. and Jagger, T. (2006). Forecast-

ing US hurricanes 6 months in advance. Geophys. Res. Lett.

33 L10704.

Emery, W. and Thomson, R. (2001). Data Analysis Methods

in Physical Oceanography. Elsevier, Amsterdam.

Evensen, G. (1994). Sequential data assimilation with a non-

linear quasi-geostrophic model using Monte Carlo methods

to forecast error statistics. Journal of Geophysical Research

99 10.

Evensen, G. (2009). Data Assimilation: The Ensemble

Kalman Filter, 2nd ed. Springer, Berlin. MR2555209

Fearnhead, P. and Prangle, D. (2012). Constructing sum-

mary statistics for approximate Bayesian computation:

Semi-automatic approximate Bayesian computation. J. R.

Stat. Soc. Ser. B Stat. Methodol. 74 419–474. MR2925370

Fiechter, J. and Moore, A. M. (2012). Iron limitation im-

pact on eddy–induced ecosystem variability in the coastal

Gulf of Alaska. Journal of Marine Systems 92 1–15.

Flay, S. and Nott, J. (2007). Effect of ENSO on Queensland

seasonal landfalling tropical cyclone activity. International

Journal of Climatology 27 1327–1334.

Freilich, M. (1996). Sea winds algorithm theoretical basis

document. Jet Propulsion Laboratory, Pasadena, CA.

Frolov, S., Baptista, A., Leen, T., Lu, Z. and van der

Merwe, R. (2009). Fast data assimilation using a nonlin-

ear Kalman filter and a model surrogate: An application

to the Columbia River estuary. Dynamics of Atmospheres

and Oceans 48 16–45.

Furrer, R., Sain, S. R., Nychka, D. and Meehl, G. A.

(2007). Multivariate Bayesian analysis of atmosphere-

ocean general circulation models. Environ. Ecol. Stat. 14

249–266. MR2405329

Gelman, A., Carlin, J. B., Stern, H. S. and Rubin, D. B.

(2004). Bayesian Data Analysis, 2nd ed. Chapman &

Hall/CRC, Boca Raton, FL. MR1385925

George, E. and McCulloch, R. (1993). Variable selection

via Gibbs sampling. J. Amer. Statist. Assoc. 88 881–889.

George, E. and McCulloch, R. (1997). Approaches for

Bayesian variable selection. Statist. Sinica 7 339–374.

Ghil, M. and Malanotte-Rizzoli, P. (1991). Data assimi-

lation in meteorology and oceanography. Adv. Geophys 33

141–266.

Page 20: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

20 WIKLE, MILLIFF, HERBEI AND LEEDS

Haario, H., Saksman, E. and Tamminen, J. (2001).An adaptive Metropolis algorithm. Bernoulli 7 223–242.MR1828504

Hanks, E., Hooten, M., Johnson, D. and Sterling, J.

(2011). Velocity-based movement modeling for individualand population level inference. PloS One 6 e22795.

Harmon, R. and Challenor, P. (1997). A Markov chainMonte Carlo method for estimation and assimilation intomodels. Ecological Modelling 101 41–59.

Hellerman, S. and Rosenstein, M. (1983). Normalmonthly wind stress over the world ocean with errorestimates. Journal of Physical Oceanography 13 1093–1104.

Herbei, R., McKeague, I. W. and Speer, K. G (2008).Gyres and jets: Inversion of tracer data for ocean circula-tion structure. J. Phys. Ocean. 38 1180–1202.

Higdon, D. (1998). A process-convolution approach to mod-elling temperatures in the North Atlantic Ocean. Environ.Ecol. Stat. 5 173–190.

Hilborn, R., Pikitch, E. and McAllister, M. (1994).A Bayesian estimation and decision analysis for an age-structured model using biomass survey data. Fisheries Re-search 19 17–30.

Hirst, D., Storvik, G., Aldrin, M., Aanes, S. andHuseby, R. (2005). Estimating catch-at-age by combiningdata from different sources. Canadian Journal of Fisheriesand Aquatic Sciences 62 1377–1385.

Hiruki-Raring, L., Hoef, J., Boveng, P. and Bengt-

son, J. (2012). A Bayesian hierarchical model of Antarcticfur seal foraging and pup growth related to sea ice and preyabundance. Ecological Applications 22 668–684.

Hoerling, M., Kumar, A. and Zhong, M. (1997). El Nino,La Nina, and the nonlinearity of their teleconnections.Journal of Climate 10 1769–1786.

Hogg, N. G. and Owens, W. B. (1999). Direct measurementof the deep circulation within the Brazil Basin. Deep-SeaRes. 46 335–353.

Holton, J. (2004). An Introduction to Dynamic Meteorology,4th ed. Elsevier Academic Press, Burlington, MA.

Hooten, M. B. and Wikle, C. K. (2010). Statistical agent-based models for discrete spatio-temporal systems. J.Amer. Statist. Assoc. 105 236–248. MR2757201

Hooten, M. B., Leeds, W. B., Fiechter, J. andWikle, C. K. (2011). Assessing first-order emulator infer-ence for physical parameters in nonlinear mechanistic mod-els. J. Agric. Biol. Environ. Stat. 16 475–494. MR2862294

Johnson, D. S., London, J. M., Lea, M.-A. and Dur-

ban, J. W. (2008). Continuous-time correlated randomwalk model for animal telemetry data. Ecology 89 1208–1215.

Jolliffe, I. (2002). Principal Component Analysis, 2nd ed.Springer, New York.

Jonsen, I., Myers, R. and James, M. (2007). Identifyingleatherback turtle foraging behaviour from satellite teleme-try using a switching state-space model. Marine EcologyProgress Series 337 255–264.

Kalnay, E. (2003). Atmospheric Modeling, Data Assimi-lation and Predictability. Cambridge Univ. Press, Cam-bridge.

Kirsch, A. (1996). An Introduction to the Mathematical The-ory of Inverse Problems. Applied Mathematical Sciences120. Springer, New York. MR1479408

Kondrashov, D., Kravtsov, S., Robertson, A. andGhil, M. (2005). A hierarchy of data-based ENSO models.Journal of Climate 18 4425–4444.

Large, W. (2006). Surface fluxes for practitioners of globaldata assimilation. In Ocean Weather Forecasting (E. Chas-

signet and J. Verron, eds.) 229–270. Springer, Dor-drecht.

Lavine, M. and Lozier, S. (1999). A Markov random fieldspatio-temporal analysis of ocean temperature. Environ.Ecol. Stat. 6 249–273.

Leeds, W., Wikle, C. and Fiechter, J. (2013).Emulator-assisted reduced-rank ecological data as-similation for nonlinear multivariate dynamicalspatio-temporal processes. Stat. Methodol. To appear.DOI:10.1016/j.statmet.2012.11.004.

Leeds, W., Wikle, C., Fiechter, J., Brown, J. and Mil-

liff, R. (2013). Modeling 3-D spatio-temporal biogeo-chemical processes with a forest of 1-D statistical emula-tors. Environmetrics 24 1–12.

Lemos, R. T. and Sanso, B. (2009). A spatio-temporalmodel for mean, anomaly, and trend fields of North At-lantic sea surface temperature. J. Amer. Statist. Assoc. 1045–18. MR2662306

Lemos, R., Sanso, B. and Santos, F. (2009). HierarchicalBayesian modelling of wind and sea surface temperaturefrom the Portuguese coast. International Journal of Cli-matology 30 1423–1430.

Lemos, R. T. and Sanso, B. (2012). Conditionally linearmodels for non-homogeneous spatial random fields. Stat.Methodol. 9 275–284. MR2863614

Lima, C. and Lall, U. (2009). Hierarchical Bayesian model-ing of multisite daily rainfall occurrence: Rainy season on-set, peak, and end. Water Resources Research 45 W07422.

Liu, F. and West, M. (2009). A dynamic modelling strategyfor Bayesian computer model emulation. Bayesian Anal. 4393–411. MR2507369

Lorenc, A. (1986). Analysis methods for numerical weatherprediction. Quarterly Journal of the Royal MeteorologicalSociety 112 1177–1194.

Lozier, M. S., Owens, W. and Curry, R. (1995). The cli-matology of the North Atlantic. Prog. Oceanogr. 36 1–44.

Majda, A. J. and Harlim, J. (2013). Physics constrainednonlinear regression models for time series. Nonlinearity26 201–217. MR3001768

Majda, A. J. and Yuan, Y. (2012). Fundamental limitationsof ad hoc linear and quadratic multi-level regression modelsfor physical systems. Discrete Contin. Dyn. Syst. 4 1333–1363.

Margvelashvili, N. and Campbell, E. (2012). Sequen-tial data assimilation in fine-resolution models using error-subspace emulators: Theory and preliminary evaluation.Journal of Marine Systems 90 13–22.

Matsuno, T. (1966). Quasi-geostrophic motions in the equa-torial area. J. Meteor. Soc. Japan 44 25–43.

Mattern, J., Fennel, K. andDowd, M. (2012). Estimatingtime-dependent parameters for a biological ocean model

Page 21: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

HIERARCHICAL STATISTICAL MODELS: OCEANOGRAPHY 21

using an emulator approach. Journal of Marine Systems96 32–47.

McAllister, M. and Kirkwood, G. (1998). Bayesian stockassessment: A review and example application using thelogistic model. ICES Journal of Marine Science: Journaldu Conseil 55 1031–1060.

McClintock, B., King, R., Thomas, L., Matthiopou-

los, J., McConnell, B. and Morales, J. (2012). A gen-eral discrete-time modeling framework for animal move-ment using multistate random walks. Ecological Mono-graphs 82 335–349.

McKeague, I. W., Nicholls, G. K., Speer, K. G. andHerbei, R. (2005). Statistical inversion of South Atlanticcirculation in an abyssal neutral density layer. J. Mar. Res.63 683–704.

McWilliams, J. (2006). Fundamentals of Geophysical FluidDynamics. Cambridge Univ. Press, Cambridge.

Megrey, B., Rose, K., Klumb, R., Hay, D., Werner, F.,Eslinger, D. and Smith, S. (2007). A bioenergetics-based population dynamics model of Pacific herring (Clu-pea harengus pallasi) coupled to a lower trophic levelnutrient–phytoplankton–zooplankton model: Description,calibration, and sensitivity analysis. Ecological Modelling202 144–164.

Michielsens, C. and McAllister, M. (2004). A Bayesianhierarchical analysis of stock recruit data: Quantifyingstructural and parameter uncertainties. Canadian Journalof Fisheries and Aquatic Sciences 61 1032–1047.

Miller, C. (2004). Biological Oceanography. Blackwell, Ox-ford.

Milliff, R., Large, W., Morzel, J., Danabasoglu, G.

and Chin, T. (1999). Ocean general circulation model sen-sitivity to forcing from scatterometer winds. Journal ofGeophysical Research 104 11337–11411.

Milliff, R., Bonazzi, A., Wikle, C., Pinardi, N. andBerliner, L. (2011). Ocean ensemble forecasting. Part I:Ensemble Mediterranean winds from a Bayesian hierarchi-cal model. Quarterly Journal of the Royal MeteorologicalSociety 137 858–878.

Moore, J. and Barlow, J. (2011). Bayesian state-spacemodel of fin whale abundance trends from a 1991–2008time series of line-transect surveys in the California Cur-rent. Journal of Applied Ecology 48 1195–1205.

NRC (1994). Report on statistics and physical oceanography.Statist. Sci. 9 167–201.

Paillet, J. andMercier, H. (1997). An inverse model of theeastern North Atlantic general circulation and thermoclineventilation. Deep-Sea Res. 44 1293–1328.

Parslow, J., Cressie, N., Campbell, E., Jones, E. andMurray, L. (2013). Bayesian learning and predictabilityin a stochastic nonlinear dynamical model. Ecological Ap-plications 23 679–698.

Pedlosky, J. (1998). Ocean Circulation Theory. Springer,Berlin.

Penland, C. and Magorian, T. (1993). Prediction of Nino3 sea surface temperatures using linear inverse modeling.Journal of Climate 6 1067–1076.

Philander, S. (1990). El Nino, La Nina, and the SouthernOscillation 46. Academic Press, San Diego, CA.

Pinardi, N., Bonazzi, A., Dobricic, S., Milliff, R.,Wikle, C. and Berliner, L. (2011). Ocean ensemble fore-casting. Part II: Mediterranean Forecast System response.Quarterly Journal of the Royal Meteorological Society 137

879–893.Preisendorfer, R. andMobley, C. (1988). Principal Com-

ponent Analysis in Meteorology and Oceanography 425. El-sevier, New York.

Robert, C. P. and Casella, G. (2004). Monte Carlo Sta-tistical Methods, 2nd ed. Springer, New York. MR2080278

Royle, J. and Dorazio, R. (2008). Hierarchical Model-ing and Inference in Ecology: The Analysis of Data fromPopulations, Metapopulations and Communities. AcademicPress, San Diego, CA.

Ruiz, J., Gonzalez-Quiros, R., Prieto, L. andNavarro, G. (2009). A Bayesian model for anchovy(Engraulis encrasicolus): The combined forcing of manand environment. Fisheries Oceanography 18 62–76.

Sacks, J., Welch, W. J., Mitchell, T. J. andWynn, H. P. (1989). Design and analysis of computer ex-periments. Statist. Sci. 4 409–435. MR1041765

Satterthwaite, W., Mohr, M., O’Farrell, M.,Wells, B. and Walters, C. (2012). A Bayesianhierarchical model of size-at-age in ocean-harvestedstocks—quantifying effects of climate and temporalvariability. Canadian Journal of Fisheries and AquaticSciences 69 942–954.

Stroud, J. R., Stein, M. L., Lesht, B. M., Schwab, D. J.

and Beletsky, D. (2010). An ensemble Kalman filter andsmoother for satellite data assimilation. J. Amer. Statist.Assoc. 105 978–990. MR2752594

Tang, B., Hsieh, W., Monahan, A. and Tangang, F.

(2000). Skill comparisons between neural networks andcanonical correlation analysis in predicting the equatorialPacific sea surface temperatures. Journal of Climate 13

287–293.Tangang, F., Tang, B., Monahan, A. and Hsieh, W.

(1998). Forecasting ENSO events: A neural network-extended EOF approach. Journal of Climate 11 29–41.

Tarantola, A. (1987). Inverse Problem Theory: Methodsfor Data Fitting and Model Parameter Estimation. Else-vier Science Publishers B.V., Amsterdam. MR0930881

Tebaldi, C. and Sanso, B. (2009). Joint projections of tem-perature and precipitation change from multiple climatemodels: A hierarchical Bayesian approach. J. Roy. Statist.Soc. Ser. A 172 83–106. MR2655606

Tebaldi, C., Smith, R., Nychka, D. and Mearns, L.

(2005). Quantifying uncertainty in projections of regionalclimate change: A Bayesian approach to the analysis ofmultimodel ensembles. Journal of Climate 18 1524–1540.

Thompson, G. (1992). A Bayesian approach to managementadvice when stock-recruitment parameters are uncertain.Fishery Bulletin 90 561–573.

Timmermann, A., Voss, H. and Pasmanter, R. (2001).Empirical dynamical system modeling of ENSO using non-linear inverse techniques. Journal of Physical Oceanography31 1579–1598.

Vallis, G. (2006). Atmospheric and Oceanic Fluid Dynam-ics: Fundamentals and Large-Scale Circulation. CambridgeUniv. Press, Cambridge.

Page 22: Modern Statistical Methods in Oceanography: A …arXiv:1312.5904v1 [stat.ME] 20 Dec 2013 Statistical Science 2013, Vol. 28, No. 4, 466–486 DOI: 10.1214/13-STS436 c Institute of Mathematical

22 WIKLE, MILLIFF, HERBEI AND LEEDS

van der Merwe, R., Leen, T. K., Lu, Z., Frolov, S. andBaptista, A. M. (2007). Fast neural network surrogatesfor very high dimensional physics-based models in compu-tational oceanography. Neural Netw. 20 462–478.

van Oldenborgh, G. J., Balmaseda, M., Ferranti, L.,Stockdale, T. and Anderson, D. (2005). Did theECMWF seasonal forecast model outperform statisticalENSO forecast models over the last 15 years? Journal ofClimate 18 3240–3249.

Ver Hoef, J. M. and Jansen, J. K. (2007). Space–timezero-inflated count models of harbor seals. Environmetrics18 697–712. MR2408939

Verbeke, G. and Molenberghs, G. (2009). Linear MixedModels for Longitudinal Data. Springer, New York.MR2723365

Von Storch, H. and Zwiers, F. (2002). Statistical Analysisin Climate Research. Cambridge Univ. Press, Cambridge.

Wheeler, M. and Kiladis, G. N. (1999). Convectively cou-pled equatorial waves: Analysis of clouds and temperaturein the wavenumber-frequency domain. J. Atmospheric Sci.56 374–399.

Wikle, C. and Anderson, C. (2003). Climatological analy-sis of tornado report counts using a hierarchical Bayesianspatiotemporal model. J. Geophys. Res 108 9005.

Wikle, C. K. and Berliner, L. M. (2007). A Bayesian tu-torial for data assimilation. Phys. D 230 1–16. MR2345198

Wikle, C. K. and Holan, S. H. (2011). Polynomial nonlin-ear spatio-temporal integro-difference equation models. J.Time Series Anal. 32 339–350. MR2841788

Wikle, C. K. and Hooten, M. B. (2010). A general science-based framework for dynamical spatio-temporal models.TEST 19 417–451. MR2745992

Wikle, C.,Milliff, R. and Large, W. (1999). Surface windvariability on spatial scales from 1 to 1000 km observedduring TOGA COARE. J. Atmospheric Sci. 56 2222–2231.

Wikle, C. K., Milliff, R. F., Nychka, D. andBerliner, L. M. (2001). Spatiotemporal hierarchicalBayesian modeling: Tropical ocean surface winds. J. Amer.Statist. Assoc. 96 382–397. MR1939342

Wilks, D. (2011). Statistical Methods in the Atmospheric Sci-ences 100. Academic Press, San Diego, CA.

Wunsch, C. (1994). Dynamically consistent hydrographyand absolute velocity in the eastern North Atlantic Ocean.J. Geophys. Res. 99 14071–14090.

Wunsch, C. (1996). The Ocean Circulation Inverse Problem.Cambridge Univ. Press, Cambridge. MR1410266

Zhang, H. M. and Hogg, N. (1992). Circulation and watermass balance in the Brazil Basin. J. Mar. Res. 50 385–420.

Zika, J., McDougall, T. and Sloyan, B. (2010). A tracer-contour inverse method for estimating ocean circulationand mixing. J. Phys. Ocean. 40 26–47.