-
-
Notifications
You must be signed in to change notification settings - Fork 1
/
Copy pathbayesian_plate_reconstruction.tex
1139 lines (993 loc) · 75.3 KB
/
bayesian_plate_reconstruction.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
%%
%% Copyright 2007, 2008, 2009 Elsevier Ltd
%%
%% This file is part of the 'Elsarticle Bundle'.
%% ---------------------------------------------
%%
%% It may be distributed under the conditions of the LaTeX Project Public
%% License, either version 1.2 of this license or (at your option) any
%% later version. The latest version of this license is in
%% http://www.latex-project.org/lppl.txt
%% and version 1.2 or later is part of all distributions of LaTeX
%% version 1999/12/01 or later.
%%
%% The list of all files belonging to the 'Elsarticle Bundle' is
%% given in the file `manifest.txt'.
%%
%% Template article for Elsevier's document class `elsarticle'
%% with harvard style bibliographic references
%% SP 2008/03/01
\documentclass[preprint,12pt,authoryear]{elsarticle}
%% Use the option review to obtain double line spacing
%% \documentclass[authoryear,preprint,review,12pt]{elsarticle}
%% Use the options 1p,twocolumn; 3p; 3p,twocolumn; 5p; or 5p,twocolumn
%% for a journal layout:
%% \documentclass[final,1p,times,authoryear]{elsarticle}
%% \documentclass[final,1p,times,twocolumn,authoryear]{elsarticle}
%% \documentclass[final,3p,times,authoryear]{elsarticle}
%% \documentclass[final,3p,times,twocolumn,authoryear]{elsarticle}
%% \documentclass[final,5p,times,authoryear]{elsarticle}
%% \documentclass[final,5p,times,twocolumn,authoryear]{elsarticle}
%% For including figures, graphicx.sty has been loaded in
%% elsarticle.cls. If you prefer to use the old commands
%% please give \usepackage{epsfig}
%% The amssymb package provides various useful mathematical symbols
\usepackage{amssymb}
\usepackage{amsmath}
\usepackage{color, soul}
\usepackage{url}
\usepackage{booktabs}
\usepackage{longtable}
\usepackage{pdflscape}
%% The amsthm package provides extended theorem environments
%% \usepackage{amsthm}
%% The lineno packages adds line numbers. Start line numbering with
%% \begin{linenumbers}, end it with \end{linenumbers}. Or switch it on
%% for the whole article with \linenumbers.
\usepackage{lineno}
%% Block of code for fixing corresponding author bug
%% in elsarticle template... Don't really understand it
\makeatletter
\def\@author#1{\g@addto@macro\elsauthors{\normalsize%
\def\baselinestretch{1}%
\upshape\authorsep#1\unskip\textsuperscript{%
\ifx\@fnmark\@empty\else\unskip\sep\@fnmark\let\sep=,\fi
\ifx\@corref\@empty\else\unskip\sep\@corref\let\sep=,\fi
}%
\def\authorsep{\unskip,\space}%
\global\let\@fnmark\@empty
\global\let\@corref\@empty %% Added
\global\let\sep\@empty}%
\@eadauthor={#1}
}
\makeatother
%% End block of weird code
%% Command for bold Greek symbols
\newcommand{\mitbf}[1]{\hbox{\mathversion{bold}$#1$}}
\journal{Earth and Planetary Science Letters}
\begin{document}
\begin{frontmatter}
%% Title, authors and addresses
%% use the tnoteref command within \title for footnotes;
%% use the tnotetext command for theassociated footnote;
%% use the fnref command within \author or \address for footnotes;
%% use the fntext command for theassociated footnote;
%% use the corref command within \author for corresponding author footnotes;
%% use the cortext command for theassociated footnote;
%% use the ead command for the email address,
%% and the form \ead[url] for the home page:
%% \title{Title\tnoteref{label1}}
%% \tnotetext[label1]{}
%% \author{Name\corref{cor1}\fnref{label2}}
%% \ead{email address}
%% \ead[url]{home page}
%% \fntext[label2]{}
%% \cortext[cor1]{}
%% \address{Address\fnref{label3}}
%% \fntext[label3]{}
\title{Bayesian inversion for paleomagnetic reconstruction and plate kinematics}
%% use optional labels to link authors explicitly to addresses:
%% \author[label1,label2]{}
%% \address[label1]{}
%% \address[label2]{}
\author{Ian Rose\corref{cor1}\fnref{ref1}}
\author{Bruce Buffett\fnref{ref1}}
\author{Nicholas Swanson-Hysell\fnref{ref1}}
\fntext[ref1]{University of California, Berkeley}
\cortext[cor1]{Corresponding author, \url{[email protected]}}
\address{}
\begin{abstract}
Apparent polar wander (APW) paths from paleomagnetic poles provide the most direct data
for reconstructing past paleogeography and plate motions for times earlier than $\sim$200 Ma.
However, it can be difficult to interpret APW paths in the presence of errors on pole positions,
age uncertainties, and the lack of paleolongitude control in traditional paleomagnetic analysis.
Approaches for dealing with the uncertainties and compiling paleomagnetic poles
into a single APW path include spline fits and running means.
Here we propose a new approach for interpretation of APW paths.
It extends the paleomagnetic Euler pole analysis of \citet{gordon1984paleomagnetic}
by placing it within the framework of a Bayesian inverse problem.
This approach allows the incorporation of uncertainties in both pole position and age.
The paleomagnetic Euler poles resulting from the inversion provide estimates for the total
plate motions (not just the latitudinal components) as well as their uncertainties.
In addition to estimating a single Euler pole, the method allows for changepoints
from one Euler pole to another to be solved for as part of the inversion.
The inversion path be restricted to be a great circle therefore allowing inversion for true
polar wander or true polar wander in combination with small circle paths.
We show several example inversions on simple synthetic data to demonstrate the capabilities of the method.
We apply this method to the Cenozoic APW path of Australia and to the
Mesoproterozoic Keweenawan Track of cratonic North America (Laurentia).
\end{abstract}
\begin{keyword}
%% keywords here, in the form: keyword \sep keyword
%% PACS codes here, in the form: \PACS code \sep code
%% MSC codes here, in the form: \MSC code \sep code
%% or \MSC[2008] code \sep code (2000 is the default)
\end{keyword}
\end{frontmatter}
\linenumbers
%% main text
\section{Introduction}
\label{sec:introduction}
Plate tectonics is the motion of near-rigid blocks of lithosphere across the surface of Earth,
separated by narrow regions of deformation in spreading centers, transform faults, and subduction zones.
The rigidity of plates means that the motion of most of Earth's surface
can be described by a set of Euler poles which specify the position and magnitude of the rotation axis for a
given plate \citep[cf.][]{cox2009plate}. Individual points on a plate undergoing
rigid rotation are described by small circle paths.
Euler poles are ubiquitous in describing current plate motions
\citep[e.g.][]{demets1990current, argus2011geologically} due to their simplicity and compactness.
Furthermore, there are good reasons to think that plate motions remain constant, or approximately
so, over millions to tens of millions of years. This consistency of motion is most dramatically seen in the shape
of oceanic fracture zones and in hotspot tracks across the lithosphere. These features
form gently curving arcs over large portions of Earth's surface which are well described by small
circles, consistent with finite Euler rotations of the plate for an extended period of time.
As such, the combination of an Euler pole plus a time interval for which it is active
(often called a ``stage pole'') is a convenient description of plate motions through Earth history.
The stage pole description of plate motions is therefore a convenient way of reconstructing
plate tectonic history, and is widely used in both continental reconstruction
\citep[e.g.][]{boyden2011next} and in geodynamical
modeling \citep[e.g.][]{mcnamara2005thermochemical, bull2014effect, rudolph2014history}.
Most reconstructions of plate motions over the past 200 million years rely heavily on fitting Euler pole rotations
to oceanic fracture zones, hotspot tracks, seafloor magnetic isochrons,
and, to a lesser extent, paleomagnetic data \citep{muller1993revised, seton2012global}.
However, as we look further back in Earth history, many of the records on which these
plate tectonic reconstructions rely largely disappear due to the
subduction of oceanic lithosphere. Before $\sim$200 Ma there is no record of oceanic crust,
and the paleomagnetic record from continental rocks is the dominant remaining evidence.
These paleomagnetic data are used in conjunction with geological data that provides information on the tectonic setting of plate margins, as well as the correlation of geologic units, to develop paleogeographic reconstructions.
It is more challenging to reconstruct past plate motions from the paleomagnetic
record than from data derived from oceanic lithosphere for a number of reasons, including
(1) the data are often sparser,
(2) traditional paleomagnetic analysis constraints paleolatitude and the orientation
of the continental block but does not constrain paleolongitude without additional assumptions, and
(3) some paleomagnetic poles have poor age control.
\citet{gordon1984paleomagnetic} noted that apparent polar wander paths (APW paths) have
arcing trajectories similar to fracture zones and hotspot tracks, which is
to be expected if similar tectonic processes are responsible for creating them.
They therefore suggested fitting small circles to paleomagnetic poles tracks,
which would furnish Euler poles for the plate in question for that time period.
This model for understanding APW paths, called paleomagnetic Euler pole (PEP) analysis,
has the attractive feature of providing a complete description of the plate motion,
including paleolongitudinal changes and speeds.
However, it has the drawback of being somewhat difficult to estimate,
uncertainties in the fit were not easily computed,
and it did not incorporate age uncertainties.
A rigorous treatment of the uncertainties requires significant computational effort.
With a few exceptions \citep[e.g.][]{beck1989paleomagnetism, tarling1996palaeomagnetic, bryan1986rotation, beck2003absolute, smirnov2010co},
PEP analysis has not seen wide adoption.
Herein we extend paleomagnetic Euler pole (PEP) analysis by placing it within
a Bayesian statistical framework, and demonstrate how to invert for PEPs
using Markov Chain Monte Carlo (MCMC) methods. This framework has the advantage
of naturally incorporating uncertainties in the paleomagnetic pole positions,
as well as widely disparate age uncertainties that commonly occur in APW paths.
The resulting stage poles for which we invert are not a single answer, but are instead
a distribution of possible answers, furnishing uncertainties as part of the solution process.
\citet{iaffaldano2012reconstructing} employed a similar Bayesian approach to inversions for finite plate rotations.
They used seafloor data to reconstruct India's Cenozoic convergence with Asia, incorporating
uncertainties into the inversion.
The paper is organized as follows: in Section~\ref{sec:apwp} we review different
approaches for interpreting APW paths. In Section~\ref{sec:bayesian_inversion} we
describe the formalism of Bayesian inversions and Markov Chain Monte Carlo methods.
In Section~\ref{sec:model} we describe the statistical model which we will be inverting.
In Section~\ref{sec:example_inversion} we demonstrate the inversion on several
synthetic data sets. In Sections~\ref{sec:australia} and~\ref{sec:keweenawan}
we show examples using paleomagnetic data from Australia and Laurentia,
including interpretations of plate speeds.
\section{Interpretation of APW paths}
\label{sec:apwp}
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/cartoon/paleomagnetic_euler_pole.png}
\caption[Conceptual model for a paleomagnetic Euler pole.]{Conceptual model for a paleomagnetic Euler pole. A finite rotation of the plate around an Euler pole (red circle) results in long, arcuate oceanic fracture zones and hotspot tracks which describe small circles on the globe. The same finite rotation produces a small circle in the APW path, which is illustrated by blue paleomagnetic poles. By fitting a small circle to the APW path we may recover the Euler pole that produced the rotation. Adapted from \citet{gordon1984paleomagnetic}}
\label{fig:pep}
\end{figure*}
A sequence of paleomagnetic poles from the same continental block form an APW path,
which can then be used to develop plate tectonic reconstructions and models
of plate speeds through time. Interpretation of these paths becomes difficult in the
case of limited, highly uncertain, or conflicting data, and when the age of paleomagnetic poles are poorly
known. A number of approaches to dealing with uncertainty in APW paths have been developed,
which we briefly review here by.
\subsection{Latitudinal drift}
\label{sec:latitudinal_drift}
Due to the rotational symmetry of Earth's dominantly dipolar time-averaged magnetic field, paleomagnetic poles do not
directly constrain the paleolongitude of the continental block in question \citep{butler1992paleomagnetism}.
The simplest analysis of an APW path is thus to compare the paleolatitudes implied by successive poles for a point on the respective block.
The difference in paleolatitudes gives a minimum angular distance over which the block has traveled.
If the two poles are well-dated, this difference also furnishes a rate of latitudinal motion.
It is also possible to estimate confidence bounds on the rate of latitudinal drift by
bootstrap resampling \citep[e.g.][]{tarduno1990fast} or by taking a Monte Carlo approach.
\citet{swanson2014confirmation} developed a Monte Carlo sampling method and applied the method to a pair of poles from the Proterozoic Midcontinent Rift of North America to estimate the range
of implied latitudinal drift. They also sampled from the uncertainties of radiometric dates associated with the poles, assuming
Gaussian distributions, in order to incorporate age uncertainties into the analysis.
With samples of pole position and ages, they were able to estimate the 95\%
confidence estimates on the rate of latitudinal drift.
Whether using point estimates of the latitudinal drift rate or using Monte Carlo estimates,
the latitudinal drift interpretation of APW paths remains limited as it represents a minimum estimate.
It does not resolve longitudinal drift rate,
nor does it naturally extend to APW paths with more than two poles,
especially if two coeval poles are not in agreement.
\subsection{Spherical splines}
\label{sec:spherical_splines}
When considering APW paths with many poles, it becomes more difficult to perform
latitudinal comparisons between pairs of poles. It is not always clear which pairs of
poles to compare in cases where there are many overlapping paleomagnetic poles,
that can have unclear age progression and variable reliabilities.
One approach to deal with these uncertainties is to fit a spline through the
set of paleomagnetic poles, constraining the path to lie on the surface of a sphere.
This approach was pioneered by \citet{torsvik1992baltica} using the spherical spline
algorithm developed by \citet{jupp1987fitting}.
This approach has the advantage of allowing the weighting of the data by their
uncertainties. The uncertainty assigned to a paleomagnetic pole can
be the 95\% confidence interval on the pole, but it can also be augmented
by various quality screening factors, such as the quality (``Q'') factor of \citet{van1990reliability} \citep{torsvik1992baltica}.
Even with the weighting of the paleomagnetic poles by uncertainty there
can be unrealistic loops in the APW path generated by the spline fit.
To combat this, the spline can also be computed under tension, penalizing
curvature and producing a smoother path \citep{torsvik1996continental}.
The spherical spline approach to interpreting APW paths has several attractive features.
It produces a smooth path through the data, incorporates spatial uncertainties
in the data, and may be efficiently computed.
However, it does have some drawbacks.
It is not easy to determine the appropriate uncertainty weighting and spline
tension parameters for the fit, and what effect those choices have on the result.
Furthermore, the resulting fit does not have an uncertainty with a physically
interpretable meaning \citep{torsvik1996continental}.
It also does not have a simple way of incorporating age uncertainties of the paleomagnetic poles.
Finally, by their very nature, splines cannot represent the sharp hairpin cusps
that characterize the abrupt shifts in motion that plates sometimes undergo \citep{irving1972hairpins, gordon1984paleomagnetic}.
\subsection{Running means}
\label{sec:running_means}
An alternative method for developing APW paths is to perform a running Fisher
mean on the poles with a moving window \citep{Irving1977a, van2001evidence, torsvik2008global}.
In such an analysis, paleomagnetic poles in a compilation are averaged in defined steps (typically 1-10 Myr) with a defined window duration (typically 10-30 Myr). Like spherical splines, the running mean approach has the ability
to effectively damp the effect of outlier poles that could lead to spurious loops in the APW path if there are sufficient data. The duration
of the moving window controls the amount of smoothing.
Furthermore, this moving window method enforces an age progression in the averaged poles.
\citet{torsvik2008global} also investigated the effects of combining running means
with spherical splines, by first computing a set of mean poles and then
fitting a spline through those means.
The simplicity of using the running mean approach to develop an APW, as well as its ability to suppress potentially spurious poles, has led to it being widely adopted.
The running mean approach shares many of the drawbacks of the spline approach.
It is not obvious how to best choose the window duration, and different window durations are
likely appropriate for different data sets.
It is also unclear how to interpret the resulting uncertainties in the path
that are reported as the Fisher $A_{95}$ ellipse of the mean of the poles.
This approach, along with the spherical spline method, does not easily incorporate age uncertainties in the poles, nor the uncertainty in pole positions.
\subsection{Paleomagnetic Euler poles}
Paleomagnetic Euler poles (PEP, introduced in Section~\ref{sec:introduction},
also known as the ``small circle'' method, were first described by \citet{gordon1984paleomagnetic}.
The model rests in recognizing that plate motions are well described by finite
rotations around Euler poles which are approximately steady for millions or
tens of millions of years. As a result, the APW path of the plate
can also be described by Euler rotations, which
produce small circles on Earth's surface.
By fitting a sequence of Euler poles to a small circle path, one specifies
the position of the Euler pole which produces that circle.
PEP analysis has the feature that it closely conforms to our model for how plates move.
Since it specifies the Euler pole which produces a given small circle,
this allows for an estimate of the full motion of a given plate,
as well as the total plate speed (instead of just the latitudinal component of the speed).
On the other hand, PEP analysis has many of the same
deficiencies that spline fits and running means have: it is not easy to compute
uncertainties, especially in the presence of unknown ages of poles.
Furthermore, one has the additional challenge and related uncertainty of deciding how many PEPs to
include for a given sequence of paleomagnetic poles.
In the following sections, we develop a Bayesian statistical approach to
PEP analysis which attempts to address some these deficiencies.
\section{Bayesian inversion}
\label{sec:bayesian_inversion}
\subsection{A general description of inverse problems}
\label{sec:intro_inverse_problems}
The central question motivating inverse problems is ``How probable is a particular model, given my observations?''.
We represent a vector of individual observations by the data vector $\mathbf{d}$, and a model
by the vector of model parameters $\mathbf{m}$, so the above question can be expressed as the function $P(\mathbf{m} \vert \mathbf{d})$.
Traditional frequentist approaches to an inverse problem often proceed by maximizing the likelihood function,
defined by the probability of the data given a particular model \citep[e.g][]{aster2005parameter}:
\begin{equation}
\mathcal{L} ( \mathbf{m} \vert \mathbf{d} ) \equiv P( \mathbf{d} \vert \mathbf{m} ).
\label{eq:likelihood}
\end{equation}
The likelihood function replaces something that is difficult to compute (namely, $P(\mathbf{m} \vert \mathbf{d})$)
with something that is less difficult to compute.
To compute $\mathcal{L}(\mathbf{m}, \mathbf{d})$ we need to have two things: a statistical model for
uncertainties in the observations $\mathbf{d}$ and forward model that allows us to compute
predictions. We denote the forward model by $\mathbf{g}$:
\begin{equation}
\mathbf{d}^p = \mathbf{g}(\mathbf{m}),
\label{eq:forward}
\end{equation}
where the superscript ``$p$'' denotes a predicted value.
If each of the observed data $d_i$ are described by Gaussian random variables
with standard deviations $\sigma_i$, the likelihood function is given by the product
of the individual likelihoods of the observations:
\begin{equation}
\mathcal{L}(\mathbf{d} | \mathbf{m} ) = \displaystyle\prod_i \exp\left({-\frac{(d_i - d_{i}^p)^2}{2 \sigma_i^2}}\right).
\label{eq:example_likelihood}
\end{equation}
The likelihood function $\mathcal{L}$ is maximized by searching over the model parameter space.
If the uncertainties in the observations are Gaussian, then maximizing the likelihood function is
equivalent to the least squares solution \citep{aster2005parameter}.
A standard maximum likelihood fit will frequently overfit the observations, resulting
in unrealistic solutions. In the context of APW paths, these overfit solutions may
pass through every paleomagnetic pole, including less reliable ones, resulting in
loopy or jerky paths. In order to address such overfitting, some form of regularization is usually
included in the solution of the inverse problem, such as penalizing the magnitude or
curvature of the solution. Both the running-mean and the spline under tension approaches
to APW paths are a form of regularization on the problem.
\subsection{Bayesian approach}
The Bayesian approach to inverse problems takes a different strategy from the frequentist one.
Rather than finding point estimates of a model fit, it treats the underlying model
as a set of random variables with individual probability distributions.
The probability distribution of the model $P(\mathbf{m} \vert \mathbf{d})$ given the data
is then found by an application of Bayes theorem \citep[cf.][]{sivia2006data}:
\begin{equation}
P\left(\mathbf{m} \vert \mathbf{d} \right) = \frac{ P \left(\mathbf{d}\vert \mathbf{m} \right) P \left( \mathbf{m} \right) }{P \left( \mathbf{d}\right).}
\label{eq:bayes}
\end{equation}
It is often unnecessary to calculate the denominator of Equation~\eqref{eq:bayes}, which is a normalization constant, leaving us with
\begin{equation}
P\left(\mathbf{m} \vert \mathbf{d} \right) \propto P \left( \mathbf{d} \vert \mathbf{m} \right) P \left( \mathbf{m} \right).
\label{eq:propbayes}
\end{equation}
The quantity $P(\mathbf{m} \vert \mathbf{d})$ is known as the posterior probability,
and it represents our desired knowledge about the distributions of the parameters $\mathbf{m}$.
The first factor on the right-hand-side of Equation~\eqref{eq:propbayes} is identical to the likelihood
function described in Section~\ref{sec:intro_inverse_problems}, and the second factor
is known as the prior probability of the model.
The prior probability reflects the state of our knowledge and beliefs of the values
of the model parameters prior to the consideration of our data. It also allows us
to incorporate constraints that are not otherwise included in the forward model.
In contrast with the classical statistical approach of regularization, the Bayesian
inverse problem can (in effect) regularize the problem through the choice of prior distribution, by making
choices of probability distributions that have less probability density in
regions with less realistic values \citep{minson2013bayesian, sambridge2013transdimensional}.
\subsection{Markov chain Monte Carlo methods}
It is usually impossible to calculate the posterior
probability distribution in Equation~\eqref{eq:bayes} directly \citep{davidson2015bayesian}.
It is much more tractable to generate a Markov chain which, upon convergence,
generates samples from the desired posterior \citep{gelman2014bayesian}.
This approach defines a class of methods known as Markov chain Monte Carlo (MCMC) methods.
The literature on MCMC methods is extensive and we do not cover it here, but
the interested reader can refer to \citet{gelman1996markov}, \citet{sambridge2013transdimensional},
and \citet{davidson2015bayesian}. A number of high-quality open source software packages
for implementing MCMC models exist, including WinBUGS \citep{lunn2000winbugs},
PyMC \citep{patil2010pymc}, and Stan \citep{carpenter2016stan}.
We make extensive use of PyMC in this work.
\subsection{Distributions on a sphere}
In order to proceed with a Bayesian description of the problem, every parameter
in the model should be described by some statistical distribution that determines
the probability that the parameter takes a specific value.
Parameters like pole ages can be described by familiar 1D probability distributions
(such as uniform or normal distributions), whereas Euler pole locations are
described by 2D distributions of directional data on the surface of a sphere.
We review several of these distributions here.
For a comprehensive discussion of spherical probability distributions,
see \citet{fisher1987statistical}. Plots of the following distributions, as well as
samples drawn from them, are shown in Figure~\ref{fig:distributions}.
\subsubsection{Uniform distribution}
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/cartoon/distributions.pdf}
\caption[Spherical probability distributions.]{Probability densities for distributions of directional data, as well as samples drawn from them. All are plotted using an orthographic projection. (a) Uniform distribution. (b) Fisher distribution. The center of the distribution is at $30^\circ$N, $30^\circ$E, with concentration parameter of $\kappa=20$. (c) Watson girdle distribution. The pole of symmetry is at $70^\circ$N, $90^\circ$E, with a concentration parameter of $\kappa=-5$.}
\label{fig:distributions}
\end{figure*}
The simplest probability distribution on a sphere is the spherical uniform distribution.
It has a probability density given by
\begin{equation}
\rho_U(\phi, \psi) = \frac{1}{4 \pi},
\end{equation}
where $\rho_U$ is the probability density, $\phi$ is the longitude, and $\psi$ is the latitude
(we will also refer to the Cartesian unit vector $\hat{\mathbf{x}}$ as a concise representation of $\phi$ and $\psi$).
Most non-uniform distributions on a sphere reduce to the uniform distribution in some limit.
We will use this distribution when we want to specify an uninformative prior distribution for directional parameters.
\subsubsection{Fisher distribution}
The Fisher distribution (also called the von Mises-Fisher distribution) is the analogue
of a 2D normal distribution on a sphere.
The probability density $\rho_F$ at a point $\hat{\mathbf{x}}$ is given by
\begin{equation}
\begin{aligned}
\rho_F(\phi, \psi ; \kappa_F, \hat{\mitbf{\mu}})
&= \frac{1}{C_F} \exp \left( \kappa_F \hat{\mathbf{x}}^T \hat{\mitbf{\mu}} \right) \\
&= \frac{1}{C_F} \exp \left( \kappa_F \cos \theta \right),
\end{aligned}
\end{equation}
where $\kappa_F$ is the concentration of the distribution,
$\hat{\mitbf{\mu}}$ the unit vector for the mean direction of the distribution,
and $C_F$ is a normalization coefficient. It can be alternatively
parameterized using $\theta$, which is the angle between $\hat{\mathbf{x}}$ and $\hat{\mitbf{\mu}}$.
The normalization factor is given by
\begin{equation}
C_F = \frac{\kappa_F}{4 \pi \sinh{\kappa_F}}.
\end{equation}
When $\kappa_F$ goes to zero, the Fisher distribution is equivalent to the spherical uniform distribution.
The uncertainty ellipses for most paleomagnetic poles are calculated assuming a Fisher distribution,
and we will use this distribution to calculate the likelihood function for pole positions in the model.
\subsubsection{Watson girdle distribution}
Whereas the Fisher distribution concentrates probability density near around a pole on
the surface of the sphere, the Watson girdle probability distribution is concentrated
in a belt orthogonal to the pole. It is useful for characterizing planar data, and is given by
\begin{equation}
\begin{aligned}
\rho_W(\phi, \psi; \kappa_W, \hat{\mitbf{\mu}})
&= \frac{1}{C_W} \exp \left( \kappa_W (\hat{\mathbf{x}}^T \hat{\mitbf{\mu}})^2 \right) \\
&= \frac{1}{C_W} \exp \left( \kappa_W \cos^2 \theta \right),
\end{aligned}
\label{eq:watson}
\end{equation}
where $\kappa_W$ is the concentration of the girdle, $C_W$ is a normalization coefficient,
and the other parameters are identical to those in the Fisher distribution.
The Watson distribution is girdle-shaped only when $\kappa_W$ is a negative number,
which is the only case we consider here.
The normalization factor is given by
\begin{equation}
C_W = \left[ {}_1 F_1 \left( \frac{1}{2}, \frac{3}{2}, \kappa_W \right) \right]^{-1},
\end{equation}
where ${}_1 F_1()$ is Kummer's confluent hypergeometric function, which is available
in most software libraries of special mathematical functions.
As with the Fisher distribution, when $\kappa_W$ goes to zero,
the Watson distribution is equivalent to the spherical uniform distribution.
\section{A model for PEP inversion}
\label{sec:model}
\subsection{Forward model}
\label{sec:forward_model}
A forward model describes how we generate predicted observations, given a set of model parameters (Equation~\eqref{eq:forward}).
The forward model for PEP analysis in this study is essentially unchanged from that of \citet{gordon1984paleomagnetic}.
We describe plate motions (and hence paleomagnetic pole motions) with a series of Euler poles.
Each Euler pole has three parameters: a latitude, a longitude, and a rotation rate.
We also must specify the ages where we switch from
one Euler pole to the next (the cusps, or ``hairpins'' of \citet{irving1972hairpins}).
In the context of parameter inversion these are often known as ``changepoints.''
Finally, we need a starting position on the globe, which, in practice, can be sampled
from the Fisher distribution of the oldest paleomagnetic pole in the dataset.
The starting point contributes two parameters (a latitude and a longitude).
Therefore, a model with $n_e$ Euler rotations will
have $3 n_e$ parameters for the poles, $(n_e-1)$ parameters for the changepoints,
and 2 parameters for the starting location.
The number parameters for which we are inverting is then given by
\begin{equation}
\begin{aligned}
N &= 3 n_e + (n_e -1) + 2 \\
&= 4 n_e + 1.
\end{aligned}
\label{eq:n_parameters}
\end{equation}
For each Euler pole $\mitbf{\omega}_i$ the velocity $\mathbf{v}$ of a point
$\mathbf{p}$ on the surface of the globe is given by
\begin{equation}
\mathbf{v} = \mitbf{\omega}_i \times \mathbf{p}.
\label{eq:rigid_rotation}
\end{equation}
Finite rotations can be performed by constructing Euler angle rotation matrices \citep[cf.][]{goldstein1965classical}.
We generate synthetic paleomagnetic pole positions from the forward model by stringing together
finite rotations through the stage poles until the age of the paleomagnetic pole
is reached. These positions can then be compared to the actual paleomagnetic poles in our dataset.
\subsection{Choice of prior distributions}
\label{sec:priors}
Bayesian analysis requires us to specify prior probability distributions
for each of the model parameters in the inverse problem.
These distributions reflect our state of belief about the values of the parameters before we begin,
and allow us the option of incorporating information otherwise not captured by the model.
To avoid biasing the results of the model towards a specific posterior distribution,
we usually try to choose prior distributions that are as uninformative as possible.
Depending upon the context, and the type of parameter, that choice may vary.
The central parameters in the paleomagnetic Euler pole problem are the Euler pole positions,
the Euler pole magnitudes, the changepoints, the starting point, and the paleomagnetic pole ages, which we treat in turn.
We use the notation $x \sim y$ to indicate that the parameter $x$ is drawn from distribution $y$.
\textbf{Euler pole directions:}
The first parameter we consider is the position of the Euler poles, which should be drawn
from a spherical probability distribution.
The least informative prior distribution for the i'th Euler pole is the uniform spherical distribution:
\begin{equation}
\hat{\mitbf{\omega}}_i \sim \rho_U.
\end{equation}
essentially allowing the Euler pole to be anywhere on the globe with equal probability.
An interesting alternative choice is to inform our prior distribution for Euler pole position based
on current plate motions. It has long been observed that, to first order, plate motions
are well explained by slab-pull torques acting along subduction zones, and to a lesser
extent, ridge push and mantle traction effects \citep{forsyth1975relative, gordon1978absolute, richardson1992ridge}.
This observation is explained by the fact that plate tectonics is the
surface expression of Earth's convection.
We can then ask the question of whether the Euler pole for a given plate is more likely
to be on top of the plate (corresponding to a spinning motion for that plate) or far away
from that plate (corresponding to motion across the surface of Earth).
Given that tectonic plates are the surface expression of convection,
we can hypothesize that the second possibility is more likely because a spinning plate has no divergence
(i.e., spreading centers and subduction zones, \citep{forte1987plate, gable1991convection}).
Without divergence the plate motion does not contribute to convection.
To test this hypothesis, we generated position samples on the surface of Earth and computed
the angular distance between that point and the Euler pole for the plate in which that point resides.
We used the NNR-MORVEL56 model for current plate motions \cite{argus2011geologically}
and restricted our analysis to the fourteen largest plates.
We then fit those angular distance samples to a Watson girdle distribution (Equation~\eqref{eq:watson}),
inverting for the concentration parameter $\kappa_W$.
If an Euler pole position has no preference for being a particular angular distance from
a point on a plate, then $\kappa_W$ should be close to zero, corresponding to a uniform distribution.
We find that the distribution is best fit with $\kappa_W \approx -0.8$, which corresponds
to the Euler pole probability density being roughly twice as large $90^\circ$
away from a given point than on top of the point (see Figure~\ref{fig:euler_pole_prior}).
\textbf{Euler pole magnitudes:}
The magnitude of each Euler pole is a strictly positive number, specifying the
rotation rate of that pole (negative rotations can be accommodated by flipping an Euler
pole to the antipode).
There are several possibilities for the prior distribution for the rates.
In order to not bias the inversion towards a particular
rate, we can choose a uniform prior distribution with large support:
\begin{equation}
\vert \mitbf{\omega}_i \vert \sim U(0, 4),
\end{equation}
where $U(\cdot, \cdot)$ is a uniform distribution between two values, and is specified
in degrees per Myr. Typical rotation rates for present day plate motions
are under $1^\circ$/Myr \citep{argus2011geologically}, which corresponds to rates
of about 11 cm/yr at a position $90^\circ$ from the pole.
Another option is to choose a weakly informative prior distribution for the Euler pole magnitudes
informed by recent plate motions (similar to the Watson girdle prior distribution for the directions).
\citet{zahirovic2015tectonic} found, based on Cenozoic and Mesozoic plate reconstructions,
that plate speeds much higher than 15 cm/yr were unlikely.
A reasonable choice of distribution for strictly positive numbers is the exponential distribution,
given by
\begin{equation}
\rho_E(\vert \mitbf{\omega}_i \vert) = \lambda \exp(-\lambda \vert \mitbf{\omega}_i \vert),
\end{equation}
which has higher probability density at lower values, and falls off exponentially.
We sampled the current plate rates on Earth's surface according to NNR-MORVEL56
and fit those to an exponential distribution. The best fitting scale parameter $\lambda$
for current plate rates is $\lambda\approx2.5$.
Making this choice of prior distribution for Euler pole rotation rates can be seen as a form
of regularization on plate speeds.
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/euler_pole_prior/euler_pole_prior.pdf}
\caption[Informative prior distributions for Euler poles]{Informative prior distributions for Euler poles. (a) Prior probabilities for rotation rates. The histogram is the angular rotation rate from one thousand samples from the surface of Earth, using the NNR-MORVEL56 model. A fit to this sample set with an exponential distribution yields a scale parameter of $\lambda \approx 2.5$. We also show the distribution for $\lambda = 1.0$, which imposes less regularization on the rate, and a uniform $U(0,4)$ distribution, which has no preference for slower speeds.
(b) Prior probability density for the position of the Euler poles, with the north pole as the site latitude and longitude.
We again sampled one thousand points on Earth's surface, calculating the angular distance between that point and the Euler pole for its plate.
If we model the probability distribution as being drawn from a Watson distribution, these angular distances correspond to colatitudes, where the pole is the sampled point.
Fitting the resulting angular distribution to a Watson girdle distribution finds $\kappa_W \approx -0.8$.
Since the Watson distribution is rotationally symmetric, longitudes do not contribute to the fit.
For $\kappa_W \approx -0.8$ the probability density is roughly twice as large at the equator as at the pole.}
\label{fig:euler_pole_prior}
\end{figure*}
\textbf{Changepoints:}
Changepoints occur sequentially between the oldest (at age $a_\mathrm{max}$) and
youngest (at age $a_\mathrm{min}$) paleomagnetic poles.
We choose a uniform distribution as a prior for these changepoints:
\begin{equation}
c_i \sim U( a_\mathrm{min}, a_\mathrm{max}),
\end{equation}
where $c_i$ is the i'th changepoint.
\textbf{Starting position:}
Finally, the starting position $\hat{\mathbf{x}}_\mathrm{start}$ for the set of Euler pole rotations needs a prior distribution.
We could choose another uniform distribution, but a more reasonable choice
is to start from near the oldest paleomagnetic pole in the dataset.
We therefore choose the Fisher distribution of the oldest paleomagnetic pole as a reasonable prior distribution for a start point:
\begin{equation}
\hat{\mathbf{x}}_\mathrm{start} \sim \rho_F(\kappa_{F0}, \hat{\mitbf{\mu_0}}),
\end{equation}
where $\kappa_{F0}$ and $\hat{\mitbf{\mu}}_0$ are the concentration parameter and mean direction
of the oldest paleomagnetic pole in the dataset.
\textbf{Pole ages:}
One of the major advantages of Bayesian analysis is the ability to naturally incorporate
uncertainties in as many parameters as needed. Previous approaches to
modeling APW paths have the drawback that they do not easily account for uncertainties in
the age of paleomagnetic poles. In our approach, we can include age uncertainty
by including the age of the poles and associated uncertainty as parameters in our model.
There are many different ways to constrain the ages of the geologic units from
which we obtain paleomagnetic poles, including radiometric dating, biostratigraphy,
magnetostratigraphy, cross-cutting relationships, and stratigraphic relationships.
Here we concentrate on poles that are either interpreted to be the age of a single radiometric
date or are interpreted to be bracketed stratigraphically between two dates
(derived radiometrically or by using other age control such as biostratigraphy).
If a geologic unit has been radiometrically dated, we can model the age of
the j'th paleomagnetic pole $a_j$ as a normal distribution with mean $\mu_j$ and standard deviation $\sigma_j$:
\begin{equation}
a_j \sim N(\mu_j, \sigma_j),
\end{equation}
where $N(.,.)$ denotes a normal distribution.
Frequently, however, the geologic unit from which we obtain a paleomagnetic pole
is not well dated, but its age can be constrained to lie between those
of well-dated units stratigraphically above and below it. In this case,
we argue that a uniform distribution between those ages is a reasonable choice
of prior distribution:
\begin{equation}
a_j \sim U(a_\mathrm{young}, a_\mathrm{old}),
\end{equation}
where $a_\mathrm{young}$ and $a_\mathrm{old}$ are the ages of the lower and upper
age constraints, respectively.
To summarize our choices for prior distributions:
\begin{itemize}
\item Euler pole positions: spherical uniform distribution, or a Watson girdle distribution with $\kappa_W \approx -0.8$.
\item Euler pole magnitudes: Uniform distribution between 0$^\circ$ and 4$^\circ$/Myr, or an exponential distribution with $\lambda \approx 2.5$.
\item Changepoints: uniform distribution between $a_\mathrm{min}$ and $a_\mathrm{max}$.
\item Paleomagnetic pole ages: normal or uniform distribution, depending on the type of age control for the geologic unit.
\end{itemize}
\subsection{Likelihood}
\label{sec:likelihood}
In addition to the choice of prior distributions, we need a statistical description of the observations.
This description will allow us to calculate the likelihood function, which, when combined with the prior distributions,
allows us to evaluate Bayes' theorem (Equation~\eqref{eq:propbayes}).
In the case of APW paths, our observations are paleomagnetic poles.
The most common statistical distribution for describing paleomagnetic poles is the Fisher distribution
(although other distributions are sometimes used, such as the Kent or Bingham distributions, c.f. \citet{tauxe2010essentials}).
Given the set of model parameters $\mathbf{m}$ and the forward model $\mathbf{g}(\mathbf{m})$, described
in Section~\ref{sec:forward_model}, we can calculate the predicted paleomagnetic pole unit vectors $\hat{\mathbf{x}}_i^p$.
For a set of $n$ paleomagnetic poles, the likelihood is then given by the product of the probabilities
of each observation:
\begin{equation}
P(\mathbf{d} \vert \mathbf{m}) = \displaystyle\prod_{i=1}^n \frac{1}{C_{F,i}} \exp \left( \kappa_{F,i} \hat{\mathbf{x}}_{i}^{pT} \hat{\mitbf{\mu}}_i \right).
\label{eq:model_likelihood}
\end{equation}
\section{Example inversions}
\label{sec:example_inversion}
Before proceeding with inversions for paleomagnetic Euler poles using real paleomagnetic data,
it is useful to consider a few examples of inversions for synthetic datasets.
We have implemented the forward model described in Section~\ref{sec:forward_model}
in Python, and used the package PyMC \citep{patil2010pymc} to perform the Monte Carlo inversion.
We use a Metropolis-Hastings sampler, starting from a fit to the maximum a posteriori (MAP) probability.
In most cases shown here, we generate $10^6$ samples and discard the first 20\% to avoid
potential bias in the posterior due to a poor starting point.
Our code for the inversions has an open-source GPL license and may be found at \url{https://github.com/ian-r-rose/mcplates}.
In all of the inversions we show here, we choose $\kappa_W=0$ for the Watson concentration parameter
in the Euler pole direction prior distribution (that is, equivalent to a uniform distribution on a sphere),
and $\lambda=2.5$ for the scale parameter in the Euler pole magnitude prior distribution.
\subsection{One Euler rotation}
\label{sec:one_stage_pole}
We begin by trying to recover the Euler pole for a single rotation.
We generate an idealized synthetic APW track of four poles by starting from a pole at $0^\circ$ N, $30^\circ$ E,
and rotating around an Euler pole at $0^\circ$ N, $0^\circ$ E for 180 Myr at a rate of $1^\circ$/Myr.
We produce paleomagnetic poles at 190 Ma, 130 Ma, 70 Ma, and 10 Ma, and prescribe $A_{95}$
of $10^\circ$ to each pole (where $A_{95}$ indicates the 95\% angular confidence interval for the pole position).
The results of the inversion are shown in Figure~\ref{fig:one_euler_pole}.
The Bayesian approach successfully recovers a posterior probability distribution for the position
of the Euler pole that includes the start pole, as well as a rate that is centered
near the true value of $1^\circ$/Myr (Figure~\ref{fig:one_euler_pole}). The posterior distribution for the rate
has a highest posterior density (HPD) credible interval at 95\%
(which we abbreviate from here as a 95\% credible interval)
between $0.8^\circ$/Myr and $1.2^\circ$/Myr, reflecting the resolving power of the inversion
for data with the given uncertainties.
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/synthetic/one_euler_pole.pdf}
\caption[Inversion for a single paleomagnetic Euler pole.]{Inversion for a single paleomagnetic Euler pole. (a) Four paleomagnetic poles are generated during a net $180^\circ$ rotation about an Euler pole at $0^\circ$N, $0^\circ$E over 180 Myr, for a rotation rate of $1^\circ$/Myr. The red distribution is the probability density function recovered by MCMC inversion, and the red lines are a sampling of 200 synthetic APW paths generated by the inversion. (b) Posterior probability density for the rotation rate of the Euler pole recovered by the inversion. The solid line shows the median of the distribution ($1^\circ$/Myr), and the dashed lines show the 95\% credible interval ($0.8^\circ-1.2^\circ$/Myr). }
\label{fig:one_euler_pole}
\end{figure*}
\subsection{Two Euler rotations}
\label{sec:two_stage_poles}
We next consider an inversion for an APW path with two stage poles.
Unlike the example in Section~\ref{sec:one_stage_pole}, this inversion also requires a changepoint.
We generate five paleomagnetic poles from a starting point at $0^\circ$ N, $60^\circ$ W.
The first rotation is around an Euler pole at $41^\circ$ N, $60^\circ$ W, and rotates at $1^\circ$/Myr for 130 Myr.
The rotation rotation is around an Euler pole at $41^\circ$ N, $60^\circ$ E, rotates at the same rate for the same amount of time.
The ``observations'' produced by the two stage model are shown in Figure~\ref{fig:two_euler_poles}.
The inversion successfully recovers the Euler pole positions and rates, and the changepoint is centered near 130 Ma.
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/synthetic/two_euler_poles.pdf}
\caption[Inversion for two successive paleomagnetic Euler poles.]{Inversion for two successive paleomagnetic Euler poles. (a) Five paleomagnetic poles are generated, beginning with a pole at $0^\circ$N, $60^\circ$W. The first Euler pole is located at $41^\circ$N, $60^\circ$W, and rotates at $1^\circ$/Myr for 130 Myr. The second Euler pole is located at $41^\circ$N, $60^\circ$E, and rotates at the same speed and for the same duration. The red and green distributions show the location of the first and second Euler poles (respectively) recovered by the MCMC inversion. The red and green lines are a sampling of 200 synthetic APW paths generated by the inversion. (b) Posterior probability density for the rotation rates of the Euler poles recovered by the inversion. The solid lines show the median values of the distributions ($\sim 0.97^\circ$/Myr, and the dashed lines show the 95\% credible intervals ($\sim 0.61^\circ-1.3^\circ$/Myr). The two distributions are nearly identical, and centered on the true value of the rate.}
\label{fig:two_euler_poles}
\end{figure*}
\subsection{Incorporating age uncertainty}
\label{sec:age_uncertainty}
A major benefit of the Bayesian approach to inverse problems is its generality.
As long as some effect can be described statistically and incorporated into our
forward model, we can include it in the inverse problem.
In this case, we include uncertainties in the ages of the paleomagnetic poles.
We use the same test case as in Section~\ref{sec:one_stage_pole}, but assign uncertain prior
distributions to the ages of the poles. For the first and last poles we assume they
are radiometrically dated with standard deviations of 2 Myr.
However, we assume that the middle two poles have no age control, except that their
respective rock units lie stratigraphically between the first and last poles.
We thus assign Gaussian prior distributions to the first and last poles and uniform prior distributions
to the middle two.
The primary effect of adding uncertainties to the ages of the poles is that they can
help to constrain the location of the APW path without providing an unwanted constraint
on the timing of the path.
Figure~\ref{fig:age_uncertainty_samples} shows the prior and posterior distributions for the
ages of the poles. We can see from the posterior distribution that the inversion
successfully places the ages of the middle two poles at $\sim70$ Ma and $\sim130$ Ma.
However, the 95\% credible intervals on these age estimates are relatively wide.
The posterior distributions for the Euler pole position and magnitude
which we recover from this inversion are visually identical to those in Figure~\ref{fig:one_euler_pole}.
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/synthetic/age_uncertainty_samples.pdf}
\caption[Probability distributions for ages of the paleomagnetic poles in the one-Euler pole inversion test.]{Probability distributions for the ages of the paleomagnetic poles in the one-Euler pole inversion test. Top: prior distributions. We presume the first and last poles to be radiometrically dated with one-sigma uncertainties of 2 Myr, and are assigned Gaussian prior distributions. The middle two poles are undated, and are only stratigraphically constrained to be between the first and last poles. Bottom: posterior distributions after $10^5$ MCMC samples. The distributions of the first and last samples are largely unchanged, but the distributions for the middle two poles are centered on their true values of 70 Ma and 130 Ma. Importantly, the middle poles help to constrain the location of the Euler pole, but the wide prior distributions on their ages do little to constrain the rotation rates.}
\label{fig:age_uncertainty_samples}
\end{figure*}
\subsection{Reporting the apparent polar wander path}
\label{sec:age_uncertainty}
A difficulty with the Bayesian approach is that the credible interval of Euler pole solutions are not easily reported as they do not neatly correspond to a parametric statistical distribution.
We visualize the solutions with spatial histograms for the Euler pole positions and by showing example inverted paths.
A typical product that one is seeking with such an inversion is an apparent polar wander path reported as interpreted pole position at given intervals.
An approach that can be taken is to calculate the predicted pole position implied by inverted Euler pole positions at a given time and then to calculate the Fisher mean of these inverted pole positions.
This approach provides the mean pole position on the apparent polar wander path.
We also wish to report the uncertainty on that estimated position.
The spread in the position of these inverted pole positions has real meaning related to the certainty of the position within the inversion that corresponds both to the best-fit in space through the poles and that in time which affects the rate of the rotation about the Euler pole.
The spread in pole positions can be approximated as a Fisher distribution from which the angle from the mean that contains 95$\%$ of the solutions ($\theta_{95}$) can be calculated:
\begin{equation}
\theta_{95}=\frac{140º}{\sqrt{\kappa}}
\label{eq:angular_deviation}
\end{equation}
This angle is equivalent to a 2$\sigma$ uncertainty in Gaussian statistics.
Implied pole positions from these inversions are more tightly clustered than the inverted Euler pole positions themselves and typically are consistent with being drawn from a Fisher distribution such that reporting the 95$\%$ confidence of angular deviation is appropriate.
\section{Application to Cenozoic Australian APW path}
\label{sec:australia}
A first application of our model for PEP inversion is with the case of Cenozoic Australia.
Australia's Cenozoic plate motions are relatively well constrained by oceanic fracture zones
and hotspot volcanism \citep{muller1993revised, seton2012global}.
This gives us the opportunity to compare the results of our model to a one derived from an independent dataset.
\begin{table}
\scriptsize
\input{tables/australia_poles}
\caption[Paleomagnetic poles used for the Australia inversion.]{Paleomagnetic poles used for the Australia inversion, as well as references.
$\psi_p$ and $\phi_p$ give the latitude and longitude of the mean pole position, and $A_{95}$ gives the 95\% angular confidence interval for that position.
The paleomagnetic poles come primarily from sedimentary successions with biostratigraphic age control
(with the exception of the Newer Volcanics, which erupted from $\sim2-4$ Ma).
Upper and lower age bounds come from estimates from \citet{idnurm1985lateI} and version 4.6 of the Global paleomagnetic database (GPMDB). The upper and lower age bounds for the Browns Creek Formation come from the magnetostratigraphic/biostratigraphic correlation of \citep{Shafik1997a} which are assigned ages based on the 2012 Geological Timescale.}
\label{tab:australia_poles}
\end{table}
The most reliable Cenozoic Australian paleomagnetic data comes from \citet{idnurm1985lateI} and \citet{idnurm1994new}.
While some Cenozoic paleomagnetic poles from Australia were developed prior to Idnurm's work, they show considerably more scatter,
and there are concerns that they do not adequately average secular variations
\citep{idnurm1985lateI, klootwijk2009sedimentary}.
\citet{idnurm1985lateII} found that the latitude-age progression of the paleomagnetic
data was significantly faster than that given by hotspot tracks.
He proposed that the best explanation for this discrepancy was a long-lived
departure of Earth's magnetic field from a geocentric axial dipole (GAD).
Here we reanalyze the available paleomagnetic data with our statistical model,
which allows us to incorporate more recent models for Australia's plate motions
and perform an accounting of the uncertainties within our Bayesian framework.
The pole list we use is given in Table~\ref{tab:australia_poles}.
We fit the paleomagnetic pole list to one and two PEPs, and compare them
to the global plate motion model from \citet{seton2012global}.
Figure~\ref{fig:australia_paths_1} shows the results for a single Euler pole
plotted on the globe, as well as modeled paleomagnetic poles drawn from the posterior distribution.
Figure~\ref{fig:australia_speeds_1}b shows the single Euler pole result plotted as latitude vs. age.
Both the paleomagnetic data and the \citet{seton2012global} model show a speed-up in
Australia's motion as it approaches the present day, though the paleomagnetic data
has systematically higher recent latitudes, implying faster plate speeds.
While the single Euler pole fit does a reasonable job of fitting the azimuth of the plate motion model
(Figure~\ref{fig:australia_paths_1}), it cannot fit a change in speed,
and does not pass through all the paleomagnetic poles in Figure~\ref{fig:australia_speeds_1}b.
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/australia/australia_1_false_paths.pdf}
\caption[Australian Cenozoic APW path fit to one Euler rotation.]{Australian Cenozoic APW path fit to one Euler rotation.
(a) Euler pole position and sample tracks. The blue distribution shows the posterior distribution of Euler pole positions, and the blue paths show a sampling of paths drawn from the posterior. The red line shows the polar wander path derived from the global plate motion model of \citet{seton2012global}, and the red star shows the average Euler pole from 0-60 Ma defined by the \citet{seton2012global} path. We calculate plate speeds referenced to Uluru, shown by the black star.
(b) Paleomagnetic pole positions for draws from the posterior distribution of the inversion superimposed on observed pole positions and their uncertainty.}
\label{fig:australia_paths_1}
\end{figure*}
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/australia/australia_1_false_speeds.pdf}
\caption[Results for Australian Cenozoic APW path with one Euler rotation.]{Results for Australian Cenozoic APW path with one Euler rotation.
(a) Posterior distribution for plate speed. The solid line shows the median plate speed (6.2 cm/yr) and the dashed lines show the 95\% credible interval (5.4-8.1 cm/yr).
(b) Latitude vs. age. The paleomagnetic poles are shown as data points, with age and latitude uncertainty. The blue lines show a sampling of the posterior path distribution, and the red line shows the \citet{seton2012global} plate motion model. Note that the paleomagnetic poles are systematically above the \citet{seton2012global} model, and that the single Euler rotation path is not able to pass through all the data points.
(c) Prior distributions for the ages of paleomagnetic poles.
(d) Posterior distributions for the ages of paleomagnetic poles.}
\label{fig:australia_speeds_1}
\end{figure*}
The two-Euler pole inversion, shown in Figures~\ref{fig:australia_paths_2}-\ref{fig:australia_speeds_2} does a better job of
fitting the paleomagnetic poles, preferring a change in speed at $\sim 23$ Ma.
The path shares some features with the \citet{seton2012global} model, including the overall azimuth and an
increase in plate speed midway through the track.
However, the $\sim23-0$ Ma speeds are significantly faster than those from the global plate motion model,
consistent with the conclusions of \citet{idnurm1985lateII}.
We conclude that the discrepancy between the Australian paleomagnetic data and
Cenozoic plate motion models remains.
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/australia/australia_2_false_paths.pdf}
\caption[Australian Cenozoic APW path fit to two Euler rotations.]{Australian Cenozoic APW path fit to two Euler rotations.
(a) Euler pole locations and sample tracks. The blue distribution shows the posterior of positions for the first Euler pole, and the red distribution shows the posterior for the second Euler pole. The distribution of the first Euler pole is quite broad, reflecting the short length of the portion of the path that it is trying to fit (shorter paths provide less of a constraint on the position). The blue paths show a sampling of paths drawn from the posterior. As in Figure~\ref{fig:australia_paths_1}, the red line shows the polar wander path derived from the global plate motion model of \citet{seton2012global}, and the red star shows the average Euler pole from 0-60 Ma defined by the \citet{seton2012global} path.
(b) Paleomagnetic pole positions for draws from the posterior distribution of the inversion superimposed on observed pole positions and their uncertainty.}
\label{fig:australia_paths_2}
\end{figure*}
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/australia/australia_2_false_speeds.pdf}
\caption[Results for Australian Cenozoic APW path with two Euler rotations.]{Results for Australian Cenozoic APW path with two Euler rotations.
(a) Posterior distributions for plate speeds. The blue distribution shows the plate speeds for the first rotation, which occurs from approximately 60-23 Ma. The red distribution shows the plate speeds for the second rotation, which occurs from approximately 23-0 Ma. The changepoint is approximate, since it is also a random variable with its own posterior distribution. This inversion prefers a slow first rotation (2.2 cm/yr with a 95\% credible interval of 0-7.3 cm/yr) and a fast second rotation (11.8 cm/yr with a 95\% credible interval of 7.1-19 cm/yr).
(b) Latitude vs. age. The paleomagnetic poles are shown as data points, with age and latitude uncertainty. The blue lines show a sampling of the posterior path distribution, and the red line shows the \citet{seton2012global} plate motion model. Compared to the single Euler pole fits (Figure~\ref{fig:australia_speeds_1}b), the two Euler pole fit does a much better job of passing through the paleomagnetic poles. However, while the modeled APW paths are qualitatively similar to the \citet{seton2012global} path in shape, the 0-23 Ma segment rotates significantly faster
(c) Prior distributions for the ages of paleomagnetic poles.
(d) Posterior distributions for the ages of paleomagnetic poles.}
\label{fig:australia_speeds_2}
\end{figure*}
\section{Application to the Keweenawan track}
\label{sec:keweenawan}
\subsection{Geologic context}
The Keweenawan province is a Mesoproterozoic succession of volcanic and sedimentary rocks
which outcrops in the region around Lake Superior, including the Keweenaw Peninsula.
The province represents the products of a failed rift that developed in the interior of Laurentia and was active
from about 1110 to 1080 Ma \citep{Davis1985a,swanson2014magmatic, fairchild2016end}.
Geochronology and paleomagnetism from the Keweenawan province has been central to reconstructions
of paleogeography and dynamics of the Mesoproterozoic \citep[e.g.][]{li2008assembly, evans2009palaeomagnetically}.
Paleomagnetic studies of the Keweenawan province have long shown asymmetry between normal and
reverse polarities of the magnetization, which has been interpreted to be a result of significant
non-dipole behavior of Earth's magnetic field during the Mesoproterozoic \citep{pesonen1981late, nevanlinna1983late, pesonen1983geomagnetic}.
More recent high-resolution paleomagnetism has shown that the apparent reversal asymmetry
is an artifact of temporal gaps in sampling \citep{swanson2009no, kulakov2013paleomagnetism}.
Instead, the paleomagnetic poles are symmetric with respect to polarity,
but the paleomagnetic directions recorded in Midcontinent rift volcanics and intrusions rapidly shallow over the course of the rifting period,
corresponding to equatorward motion of Laurentia.
A consequence of this reinterpretation of the paleomagnetic data is that Laurentia's apparent
polar wander rate, and thus its implied plate motion, is quite fast.
A Monte Carlo approach (described in Section~\ref{sec:latitudinal_drift}) by \citet{swanson2014confirmation}
found an implied latitudinal drift rate of $\sim$24 cm/yr, with a 95\% confidence interval of 15.2-44.4 cm/yr.
At the high end, these rates are significantly faster than the fastest Cenozoic plate speeds \citep{zahirovic2015tectonic}.
Interpretations of plate kinematic models for the past 200 million years, led \cite{zahirovic2015tectonic} to conclude that motion of plates with substantial continental lithosphere above 15 cm/yr requires exceptional explanation.
Possible explanations for such fast rates recorded in the Midcontinent Rift include faster plate speeds in the Proterozoic (possibly due to decreased
mantle viscosity in a hotter, younger Earth), increased convective vigor associated with mantle avalanching or true polar wander \citep{swanson2009no}.
Combined paleomagnetic and geochonologic studies have also been interpreted to indicate that
the APW rate during the Keweenawan rifting slowed down in the later stages of rift development.
\citet{davis1997geochronology} inferred a change in minimum plate speed from $\sim$22 cm/yr
to $\sim$8 cm/yr at around 1095 Ma. Similar conclusions were reached by \citet{swanson2009no}.
However, more recent geochronology and paleomagnetism from the late stage volcanics
of the Midcontinent Rift. \citet{fairchild2016end} have suggested Laurentia's APW path had high rates throughout rifting.
\subsection{Inversion for paleomagnetic Euler poles}
We apply our Bayesian PEP analysis to the $\sim$1110-1080 Keweenawan paleomagnetic track.
We would like to determine allowable plate speeds (not just the latitudinal components), while
incorporating the variable uncertainties in the ages of the paleomagnetic poles. While some poles are tightly constrained with high-precision U-Pb dates, others are more loosely constrained in time.
Furthermore, we want to test whether an abrupt slowdown in the APW path is required by the data,
and if so, when it might have happened.
The paleomagnetic poles and age constraints we use, along with uncertainties, are given in Table~\ref{tab:keweenawan_poles}.
\begin{landscape}
\begin{table}
\scriptsize
\input{tables/keweenawan_poles}
\caption[Paleomagnetic poles used for the Keweenawan track inversion.]{Paleomagnetic poles used for the Keweenawan inversion, as well as references for their positions and ages.
$\psi_p$ and $\phi_p$ give the latitude and longitude of the mean pole position, and $A_{95}$ gives the 95\% angular confidence interval for that position.
For poles with an associated radiometric date we give the age with 2$\sigma$ error bars.
For poles with stratigraphic age control we give upper and lower bounds on the age.}
\label{tab:keweenawan_poles}
\end{table}
\end{landscape}
We invert the Keweenawan APW track for one, two, and three paleomagnetic Euler poles.
Adding a third PEP did not improve the fit, and it left one Euler pole completely unconstrained.
We interpret this as meaning that three PEPs is unnecessary given the data. Therefore we focus
on the results with one and two PEPs.
The fit with a single PEP is shown in Figure~\ref{fig:keweenawan_paths_1}.
The posterior distribution of the Euler pole position is shown in in blue,
and a sampling of the small-circle paths generated from the posterior distribution are plotted over the paleomagnetic poles.
Figure~\ref{fig:keweenawan_speeds_1}(c) and~\ref{fig:keweenawan_speeds_1}(d) show the prior and posterior distributions for the ages of the poles,
and Figure~\ref{fig:keweenawan_speeds_1}(a) shows the posterior distribution of the plate speed for Laurentia,
calculated with respect to Duluth, MN (46.8$^\circ$N, 92.1$^\circ$W).
The median plate speed for the one-Euler pole inversion is 26.5 cm/yr, with a 95\% credible interval of 23.1-30.4 cm/yr.
The fit with two PEPs is shown in Figure~\ref{fig:keweenawan_paths_2}.
The posterior distribution of the position for the first Euler pole is shown in blue,
and the posterior distribution of the position for the second Euler pole is shown in red.
Figure~\ref{fig:keweenawan_speeds_2}(c) and~\ref{fig:keweenawan_speeds_2}(d) show the prior and posterior distributions for the ages of the poles,
and Figure~\ref{fig:keweenawan_speeds_2} shows the posterior distribution of the plate speeds for the two rotations.
The inversion places the changepoint between the two Euler rotations at roughly 1098 Ma.
Before the changepoint the inverted plate motion is much faster, with median at 30.0 cm/yr
and a 95\% credible interval of 22.9-39.1 cm/yr. After the changepoint the plate motion has a median
of 13.9 cm/yr with a 95\% credible interval of 9.9-22.5 cm/yr.
\begin{figure*}
\includegraphics[width=0.9\textwidth]{figures/keweenawan/keweenawan_1_false_paths.pdf}
\caption[Inversion of the Keweenawan track for a single Euler rotation]{Inversion of the Keweenawan track for a single Euler rotation. See Figure~\ref{fig:keweenawan_speeds_1}b for pole labels.
(a) Euler pole location and sample tracks. The posterior probability distribution is shown in blue, and a representative sample of the tracks generated by the inversion are drawn over the paleomagnetic poles. Also shown is the outline of Laurentia, with the location of Duluth drawn as a star.
(b) Paleomagnetic pole positions for draws from the posterior distribution of the inversion superimposed on observed pole positions and their uncertainty.}