This repository has been archived by the owner on Nov 13, 2017. It is now read-only.
-
Notifications
You must be signed in to change notification settings - Fork 1
/
literature.tex
1552 lines (1359 loc) · 98.1 KB
/
literature.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
\chapter{A Brief and Opinionated Literature Survey}%
\label{chap:litsurvey}
Teleoperation systems are structurally simple, two connected robotic manipulators, but equally challenging
systems. This is especially true from a system theoretical point of view. As an example,
if we just focus on the local and the remote devices that would be used for manipulation,
we see that they are, whether linear or nonlinear, motion-control systems with well-studied
properties. Hence, one can view the open-loop teleoperation, i.e., standalone devices without any communication
in between, as a system with a block diagonal structure in which each input to this block effects only one
of the devices. However, unlike the typical motion-control systems, these two disjoint
systems must be stabilized simultaneously by the same controller\footnote{We need to emphasize
that the delayed/undelayed local control loops can be seen as the entries of a structured
central controller. Thus, there is no reason to distinguish local control methods at this
point.}, that is performing sufficiently well in order to \enquote{fool} the user such that the
user feels a force feedback as if s/he is actually operating at the remote medium. With this
structure, the outputs of either of these subsystems become exogenous inputs of the other
and these are regulated by the to-be-designed controller. Therefore, it is this controller
that makes a teleoperation system perform adequately or, as in many cases, drive to instability.
For example, in the case of the so-called \emph{free-air motion}, i.e., the remote device is free
to roam in the remote site, the human force input to the local device and/or the position of the
local device should\footnote{Or better, that's what we believe today.}
be tracked by the remote device. In the case of a hard-contact of the remote device with the environment, however,
these inputs should be counteracted if the force vector points into the obstacle. Hence, the force signal is
simultaneously tracked for mimicking the user motion and is defied in case of a resisting force at the remote
site. As if this is not challenging enough, when the user suddenly decides to release the local device, this
resistance should die out as soon as possible, preventing a kickback. To sample the artificial nature of such
a behavior, consider a user who leans to a wall located at position $x_0$ and beyond, applying a horizontal
force and then retreating after some time. It is not expected that the wall continues to push the user even
after the user has the position $x<x_0$. Such behavior would not only be unrealistic but also misleading as
it can be confused with a sticky surface as far as the immersion is concerned. There are many other scenarios
that would further complicate the requirements but, in short, the user and the environment properties are
time-varying and make it difficult to design a straightforward control law such that these and many other details
are handled properly, and more importantly, simultaneously.
With this short motivation, we can safely claim that looking at the overall system as a typical motion control
system is not sufficient in terms of complexity (though necessary). In general, motion tracking specifications
constitute a subset of the general performance requirements of bilateral teleoperation systems.
%\begin{rem}
%A similar line of reasoning has been given in \cite{buergerhogan1}. But for reasons that are
%not known to us, the authors chose to see the classical loop shaping as the sole servo control method. Moreover, modern
%control theory offers a few options to avoid the obstacles that are given as impossibilities in their paper. Therefore,
%we emphasize that the difference here is merely about the problem formulation and is not about the
%limits of performance\footnote{We refer to the elegant book \cite{boydbarratt} for a detailed investigation of the
%limits of performance.}.
%\end{rem}
The inception of the bilateral teleoperation technology is often attributed to the work
of Raymond Goertz in Argonne National Laboratories, \cite{goertz} (in \cite{basanezsuarez},
it is traced back to Nikola Tesla and, in \cite{sheridan89}, even some 16th century tools are accepted as precursors
of the contemporary teleoperation). The main motivation of Goertz' work (similarly later in Europe
by Vertut \cite{vertutcoiffet}) was handling and manipulating nuclear material. Thus the very
first teleoperators were purely mechanical to cope with hostile environment conditions. Though not much
happened in terms of commercial product realizations, the concept of telemanipulation kept its appeal
and a large body of research was reported until the 1980s. In that decade, with the help of the
ever-increasing computational power and the popularity of Virtual Reality (VR), teleoperation
technology received more attention for a possible use in the space-, underwater-, and medical-related
tasks. Together with the advances in control theory and network theory (e.g. \cite{miyazaki,furuta}),
a more systematic control methodology has been adopted. Especially, stability analysis results that can be
related to design guidelines (physical parameter bounds, bandwidth limitations etc.) were utilized
and limits of performance were explored. A particular phenomenon, namely the destabilizing effect
of the delays in the teleoperation, lead the experts of the field to delve more into the systematic
analysis tools and qualitative aspects of teleoperation. Especially, the use of the concepts such as,
\enquote{passivity}, \enquote{scattering transformations}, and \enquote{wave variables} has become the
standard methods of analysis and synthesis (see, e.g., \cite{hannaford89,andersonspong,nieslotine}).
Arguably, this point is where bilateral teleoperation branched off from the general control theory
and became a specialized area of research, specifically dealing with a particular problem that is still
a matter of unresolved debate, as we touch upon later in this chapter.
We start to summarize the advances from this point as this thesis is precisely built on top these systematic
analysis and synthesis results gathered in the last two decades. However, the reader is referred to
\cite{hokayemspong,burdea}, and \cite{sheridan89} for a more detailed overview including other practical
aspects of teleoperation analysis and the hardware developments with a more historical perspective which
will be omitted here.
As we keep on narrowing down our focus to the control theoretical parts of this challenging problem, we have to
note that many parts of the bilateral teleoperation problem can be scrutinized under different frameworks.
Hence, there is no shortage of techniques for which the bilateral teleoperation problem is an ideal
test case. In this plethora of methods, for example, the variation of human and environment properties give
naturally rise to a robust or an adaptive control approach, the hard-contact problem can be analyzed by viewing it
as switched control systems, jump control systems or constrained linear systems etc. Thanks to these advances,
as we elaborate, the main unsolved problem is not a methodological one but a motivational one. In
other words we are lacking not the solution methods but rather a fundamental understanding of the problem in terms of
what the requirements are and what a good device is if compared to another. Let us sample a few important and
successful approaches reported so far together with their shortcomings if any.
We emphasize that the literature
covered here is far from comprehensive and deliberately shaped with pragmatic intentions. Hence a large body of
research is left out. This is certainly not due to their lack of thoroughness or else, but simply due to the irrelevance
for the purpose of this chapter. In general, the methods that are left out either don't define a performance
objective or only focus on the particular detail about bilateral teleoperation instead of the human perception.
The reasoning behind this choice should be more apparent after \Cref{chap:perf}.
\begin{figure}%
%http://chat.stackexchange.com/transcript/message/4857161#4857161
\centering
\begin{tikzpicture}[scale=1,manstyle/.style={line width=4pt,line cap=round,line join=round}]
\node[rectangle,draw,minimum height=3cm,minimum width=3.8cm] at (0.5cm,0) (a) {};
\node[rectangle,draw,minimum height=3cm,minimum width=3.8cm] at (1.5cm,1cm) (b) {};
\foreach \x in {north east,north west,south east,south west} \draw (a.\x) -- (b.\x);
\node[rectangle,draw,minimum height=3cm,minimum width=3.8cm] at (5.5cm,0) (a) {};
\node[rectangle,draw,minimum height=3cm,minimum width=3.8cm] at (6.5cm,1cm) (b) {};
\foreach \x in {north east,north west,south east,south west} \draw (a.\x) -- (b.\x);
\node[fill,circle,inner sep=2.5pt,outer sep=1pt] at (-0.2mm,7.1mm) {};
\draw[manstyle] (0,0.5cm) -- ++(0,-1.2cm);
\draw[manstyle] (-1.5pt,0) -- ++(0,0.5cm) (1.2pt,1pt) --(0,5mm)--++(-45:4mm);
\draw[line width=1mm,fast cap-fast cap] (0.5cm,0.2cm) -- ++(0.5cm,0);
\begin{scope}[xshift=1.3cm,-stealth,black!20]
\draw (0,0,0) -- (0,0,1); \draw (0,0,0) -- (0,1,0);\draw (0,0,0) -- (1,0,0);
\end{scope}
\begin{scope}[xshift=4.7cm,-stealth,black!20]
\draw (0,0,0) -- (0,0,1); \draw (0,0,0) -- (0,1,0);\draw (0,0,0) -- (1,0,0);
\end{scope}
\begin{scope}[shift={(1.4cm,-0.2cm)}]
\draw (0,0) -- (0.9cm,0) (0.45,0) circle (1.5mm);
\path [fill=white,postaction={pattern=north west lines}] (0cm,0cm) rectangle (0.9cm,-0.16cm);%postaction={pattern=north west lines},
\draw[line width=1.5mm,round cap-round cap] (0.475cm,0.05cm) -- ++(110:7mm);
\draw[line width=1mm,round cap-round cap] (0.475cm,0.05cm) ++(110:6.3mm) --++ (220:5mm);
\end{scope}
\begin{scope}[shift={(5.4cm,-0.2cm)},cm={-1,0,0,1,(0,0)},transform shape]
\draw (0,0) -- (0.9cm,0) (0.45cm,0) circle (1.5mm);
\path [fill=white,postaction={pattern=north west lines}] (0cm,0cm) rectangle (0.9cm,-0.16cm);%postaction={pattern=north west lines},
\draw[line width=1.5mm,round cap-round cap] (0.475cm,0.05cm) -- ++(110:7mm);
\draw[line width=1mm,round cap-round cap] (0.475cm,0.05cm) ++(110:6.3mm) --++ (220:5mm);
\end{scope}
\draw[line width=1mm,fast cap-fast cap] (5.8cm,0.2cm) -- ++(0.5cm,0);
\begin{scope}[shift={(6.6cm,0.4cm)}]
\draw[thick]
\foreach \i in {1,2,...,10} {%
[rotate=(\i-1)*36] (0:2mm) arc (0:12:2mm) -- (18:2.4mm) arc (18:30:2.4mm) -- (36:2mm)
};
\node[circle,draw,inner sep=2pt,fill] (merkez) at (0,0) {};
\end{scope}
\begin{scope}[shift={(6.85cm,0cm)},scale=1.2]
\draw[thick]
\foreach \i in {1,2,...,10} {%
[rotate=(\i-1)*36] (0:2mm) arc (0:12:2mm) -- (18:2.4mm) arc (18:30:2.4mm) -- (36:2mm)
};
\node[circle,draw,inner sep=2pt,fill] (merkez) at (0,0) {};
\end{scope}
%\begin{scope}[shift={(6.8cm,0.5cm)},scale=0.8]
%\draw[fill] (0.5mm,0.5cm) rectangle (0.49cm,0.6cm);
%\draw[fill,postaction={pattern=north west lines,pattern color=white}] (0.2cm,0.5cm) -- ++ (0,-0.4cm) -- ++ (-45:1mm) --++(45:1mm) --++ (0,
%0.4cm) --cycle;
%\end{scope}
\end{tikzpicture}
\caption{General Teleoperation System}%
\label{fig:teleop}%
\end{figure}
\section{Modeling of Bilateral Teleoperation Systems}\label{sec:litchapmodeling}
The dominating modeling paradigm of bilateral teleoperation systems is the two-port network approach. Consider the following quote from
1989:
\begin{displayquote}[{\cite{hannaford89}}][.]
%In this paper, development of the analytical framework is complemented by modeling of an actual teleoperator system.
The modeling approach is to transform the teleoperation system model into an electrical circuit and simulate it using
SPICE, the electronic circuit simulation program developed at UC Berkeley
\end{displayquote}
%\end{mdframed}
As seen from Hannaford's motivation, the computer-based simulation tools are used extensively since then. Arguably,
this is one of the main reasons why network and electrical circuit based modeling dominated the teleoperation literature.
Reinforced with the circuit simulation tools, experts of the field started to construct analogies that go beyond
a mere mechanical-electrical system analogy. Consequently, the most prominent concept borrowed from these analogies is the two-port
network view of bilateral teleoperation systems. The reader is referred to \Cref{chap:apdxnetwork} for a short recap of network theory.
Today, the quoted convenience also applies to almost all physical systems, i.e., one can simulate arbitrary models via many
computational packages. Yet, it's a de facto standard to use the circuit modeling while the teleoperation devices are mostly
mechanical. Hence, it's not clear whether the benefit of such an artificial step still exists. Once the system is represented
by a mathematical model, as it is demonstrated in the later sections, the mechanical/electrical analogy is, roughly, an equivalence
based on the resulting model and works in the \enquote{from electrical to mechanical} direction too. Therefore, the
circuit based modeling approach is merely a convention rather than a requirement.
\subsection{Two-port Modeling of Teleoperation Systems}
In the teleoperation context, if one uses the \enquote{load-source} analogy for the manipulated environment
and the human, then the system models all the bilateral interaction between the load and the source
ports (as in \Cref{fig:portrep}). This modeling view is quite powerful since the components are described via
their input/output (or external) properties, i.e., across variable/through variable relations (e.g. force/velocity,
voltage/current etc.). Also, the non/linearity properties of the components are not relevant at the outset if we
are only interested in energy exchange, which is the basis of the so-called Time-Domain Passivity Methods
\cite{hannafordryu} which we also mention later in this chapter. Thus, the user, the control system, the
environment, the remote and local devices and the communication delays are seen as $1$- and $2$-ports exchanging
energy in time. Since the external behavior of the ports can be characterized completely by the power variables
associated with the terminals, e.g., voltage drop across the terminals and current flowing through them, it is
indeed very convenient to model these components with electrical ports as interacting \enquote{black boxes}
(See \Cref{fig:portrep}).
\begin{rem} We should emphasize here that, in this context, the energy exchange is used as a gauge of a
potentially unstable behavior. The motivation relies on the fact that in order to classify a system as an
unstable one, the system should exhibit unstable behavior at its port(s) and to exhibit such behavior
additional energy is needed. Hence if all the components are incapable of contributing energy in the loop,
finite energy excitations will eventually decay and the system would reach its steady state after a transient
response\footnote{For the sake of brevity, marginal stability or limit cycles also require extra energy for
the sustain.}. As one can directly identify, this is simply the rough sketch of the celebrated passivity theorem.
Based on this argument, there is a recurring theme in the literature that a teleoperation system should be
passive in order to have a stable interconnection. This is due to the hypothesis that end terminations are
also passive. However, in a few studies, this is mistakenly taken as a sufficient and necessary condition
for stability and hence creates quite some confusion for the non-experts of the field. Passivity is not an
essential feature of the teleoperation systems but only a convenient shortcut for deriving interconnection
stability conditions. We should iterate that stability is the top-priority objective and does not require
passivity by any means even if we do guarantee stability by rendering the sub-components passive.
\end{rem}
\begin{figure}
\begin{subfigure}[b]{0.5\textwidth}
\centering
\begin{tikzpicture}[>=stealth,baseline=15mm,
every node/.style={draw,minimum size=1cm}]
\node (deltas) at (0,0) {$\Delta_l$};
\node[left=of deltas] (g) {$G$};
\node[left=of g] (deltal) {$\Delta_s$};
\tikzset{every node/.style={draw,circle,inner sep=1pt,fill=white}}
\foreach \y in {s,l}{
\node at ({$(g)!0.5!(delta\y)$} |- {g.40}) (circ1\y){};
\node at ({$(g)!0.5!(delta\y)$} |- {g.-40})(circ2\y){};
};
\draw (circ1l) -- (deltal.40) (circ2l) -- (deltal.-40)
(circ1s) -- (deltas.140) (circ2s) -- (deltas.-140);
\draw[->] (circ2s) -- (g.-40) (circ1s) -- (g.40);
\draw[->] (circ2l) -- (g.-140) (circ1l) -- (g.140);
\end{tikzpicture}
\caption{}
\label{fig:portrep}
\end{subfigure}%
\begin{subfigure}[b]{0.5\textwidth}
\centering
\begin{tikzpicture}[>=stealth,scale=0.5, transform shape]
\matrix (G) [draw,matrix of math nodes,inner sep=1mm,row sep=1mm,ampersand replacement=\&]{ G_{11} \& G_{12}\\ G_{21} \& G_{22}\\};
\matrix (delta) at (0,3) [outer sep=0,draw,matrix of math nodes,inner sep=0.5mm,ampersand replacement=\&]{\Delta_s \& \\ \& \Delta_l\\};
\draw[->] (G.160) -| ++ (-0.4,0.8) node[draw,circle,fill=white,inner sep=2,label={[inner sep=0]0:$\scriptstyle -$}] {} |- (delta.160);
\draw[->] (G.200) -| ++ (-0.7,0.6) node[draw,circle,fill=white,inner sep=2,label={[inner sep=0]0:$\scriptstyle -$}] {} |- (delta.200);
\draw[<-] (G.20) -| ++ (0.4,0.5) |- (delta.20);
\draw[<-] (G.-20) -| ++ (0.7,0.5) |- (delta.-20);
\end{tikzpicture}
\caption{}
\label{fig:nom_net}
\end{subfigure}
\caption[Two representations of a 2-port network.]{Two representations of a 2-port network. Here $\Delta_s$ and $\Delta_l$ represents
the source and the load immitances respectively. In the teleoperation context, they are the human operator and the explored environment.}
\end{figure}
Clearly, thanks to this modeling method, we don't even need to know exactly what $\Delta_l,\Delta_s$ blocks are, except
their class (e.g. linear/nonlinear, time invariant/time varying etc.) to analyze the interconnection via $G$ and its port
behavior. Thus, the problem of modeling of the human arm or of the uncertain environment is circumvented. However, with
the same reasoning, the passivity property does not distinguish particular systems as long as they are passive. That is to
say, some of the crucial information is lost about these specific ports; we discard any impedance or admittance relations
shared by the port variables.
Energy based modeling is also the natural basis of bond-graphs. Bond-graphs, much like port representations, are graphical tools
to model the dynamical systems via energy balancing between subcomponents (See \cite{gawthrop} for an introduction). In other words,
the bond-graphs are built on top of the notion of bonds representing the instant energy or power exchange between nodes via edges
drawn between them. Therefore, bond-graphs already present a powerful framework for the abstraction of the bilateral interaction between
the local and the remote site. For a classical use of bond-graphs in impedance control, the reader is referred to Hogan's trilogy
(\cite{hogan:1,hogan:2,hogan:3}). There are also many studies with application focus, e.g., \cite{krishnaswamy} using hydraulic
systems for bilateral teleoperation, among many others.
\subsection{Assumptions on the Local and Remote \enquote{Ports}}
As mentioned above, network theory offers a great opportunity for modeling teleoperation systems, or better, avoiding
a refined modeling. Still, to invoke the stability analysis and synthesis results of network theory, there is a need
to distinguish $\Delta_s,\Delta_l$ further in the universum of $1$-ports. Otherwise, with no additional assumptions
there is not much we can conclude from such an interconnection since they can be any arbitrary model with arbitrary
behavior set as long as they respect the port condition. This is obviously a crude approximation of the real physical
interaction that teleoperation systems exhibit.
In the teleoperation and haptics literature, it is customary to assume the load and the source terminations as \enquote{passive}
mathematical operators (see \Cref{chap:apdxnetwork}). Starting with this hypothesis, the stability problem can be converted to
a typical energy dissipation problem. Hence the view of the designer is tuned to look out for the energy sources and interaction
between two distant media. This approach treats the human and the environment as passive $1$-port circuit elements together with
additional voltage and/or current sources modeling the intentional force input to the system. The controller(s) act as the energy regulator
preventing excess energy generation to avoid a possible instability even in the cases where extra energy does not endanger
stability or in fact needed by the user to accomplish certain task or stabilize the system.
Additionally, as summarized in \Cref{chap:apdxnetwork} and in \Cref{chap:analysis}, one can use the
network theory based conditions to assess the stability and performance conditions thanks to this hypothesis.
This brings us to the discussion of the justification of the assumption as it is generally not given in full generality in the
literature. Is it indeed valid to assume that the human can be modeled as a passive system? If one scans through the literature
about the passivity of human operators, it is the Hogan's paper \cite{hogan89}
that is almost universally cited. The striking detail is, however, that Hogan never claims that the human hand/arm is a passive
system. Instead he clearly shows that under very specific conditions, human behavior is indistinguishable from that of a
passive system:
\begin{displayquote}[{\cite{hogan89}}][.]
Thus, despite the fact that the
limb is actively controlled by neuro-muscular feedback, its apparent stiffness is equivalent to that of a completely passive
system. In the light of Colgate's recent proof [3]\footnote{Reference \cite{colgatehogan88} of this thesis. However exactness
of stability characterization for two LTI passive complex uncertainty blocks was already well-known in SSV theory (e.g., \cite{packdoyle})
and also in the classical network theory works at the time of writing. Therefore it is a misattribution.} that an apparently
passive impedance is the necessary and sufficient condition for a stable actively-controlled system to remain stable on contact
with an arbitrary passive environment, this experimental result strongly suggests that neural feedback in the human arm is carefully
tuned to preserve stability under the widest possible set of conditions
\end{displayquote}
Moreover, the task reported in the paper that is given to human operators and analyzed afterwards, can be considered as a biased one because the success
of the test is related to the passive behavior of the human. The task is, roughly speaking, holding a handle which is perturbed by
random disturbances and trying to keep the handle still at a predefined position on the 2D plane. Hence, the task is simply to mimic
a passive system. Had it been the case that measurements on the human arm would exhibit a non-symmetric stiffness matrix in the arm
model, it would simply be a failure of the test subject (regardless of the physical limitations of the human arm in general). Note
that this is a plausible situation for rehabilitation tasks. The other possibility would then be that the test subject was unable
to keep up with the changes, or using the control theory jargon, the bandwidth of the subject was lower than the required agility to
perform the test adequately. The well-known phenomenon due to such human input is the \enquote{pilot induced oscillations} in which the
pilot of an aircraft, while trying to stabilize the aircraft, via overcorrecting inputs, destabilizes the system due to many distinct
reasons (the phase lag of the pilot, response time of the aircraft etc.). We refer to the interesting report \cite{mcruer} for a more
detailed exposition. Also, if for some reason, the task at hand is to prevent the system to reach a steady state at a certain position
and the perturbations are applied accordingly i.e., to create a virtual negative potential, the results obtained from the experiments
would most probably differ from that of \cite{mussa85}. Thus, it's emphasized here that the passivity of the human is closely linked
to the requirements of the tasks.
\begin{rem}\label{rem:filament}
A particular detail should be clarified about the measurements taken in \cite{hogan89}. It is stated that:
\begin{displayquote}[{\cite{hogan89}}][.]
While normal human subjects held the handle of the manipulandum at a stable position
in the workspace, small perturbations were applied. Measurements
of the human's restoring force were made after the system
had returned to steady state following the perturbation but before
the onset of voluntary intervention by the subjects.
\end{displayquote}
Therefore, it is emphasized that only the involuntary response is taken into account during the measurements in order to
capture the natural properties of the human arm before the human correction intervenes. In fact, due to this crucial distinction
the results such as \cite{dyck2013} do not disprove Hogan's results since the voluntary input is included in the model hence the passive
dynamics, if any, is mixed with the independent force input. It is the voluntary input of the human that should not be included in the
analysis, since the human necessarily puts in energy, at the very least, to move the local device violating passivity condition trivially.
\end{rem}
We believe that it is unavoidable to introduce some concepts from the muscle physiology in order to put Hogan's
argument into some mechanical engineering perspective. It is simply impossible for us to give a detailed analysis,
however, mentioning the involved process via some mechanical analogies in order to relate the results of
Hogan and Mussa-Ivaldi (\cite{mussa85}) seems feasible. This would emphasize the reason why we think that the common inference
from their experiments in the literature is not inline with the conclusions of these studies. We refer the reader
to the physiology literature e.g., \cite{spudich,millman,offer,horowits,yildiz,geeves} and references therein for a full
treatment. Hence, we will only give a rough picture about the apparent behavior. Nevertheless the point that we want
to emphasize is, fortunately, not related to the inner workings of the human muscles.
%
%
%\makeatletter
%\pgfdeclareplotmark{o*}
%{%
%\pgfcoordinate{tpt-\the\pgf@plot@mark@count}{\pgfpointorigin}
%\pgfpathcircle{\pgfpointorigin}{\pgfplotmarksize}
%\pgfusepathqfillstroke
%}
%\makeatother
\begin{figure}%
\centering%
\begin{tikzpicture}[transform shape, scale=0.5,decoration={zigzag,amplitude=1pt,segment length=1mm}]
\begin{scope}
\begin{scope}
\begin{scope}[rotate=90]
\path[domain=0.6*pi/3:6.5*pi/3,
samples=50,
mark=o*,
mark size=2mm,
mark options={fill=white},
mark indices={1,...,4,13,14,...,30,36,37,...,40,42,43,...,49}] plot[parametric] function{3*sin(t),0.1*sin(2*t)};
\filldraw[fill=white] (tpt-49) ++(10:2mm) arc (110:40:2mm) arc (-40:156:2mm) arc (65:5:2mm) ;
\fill (tpt-42) ++(150:2mm) arc (150:210:2mm) arc (330:30:2mm);
\end{scope}
\node[draw,anchor=west,shape=ellipse,ultra thick,minimum width=1cm] (a) at ([shift={(2mm,-3.5mm)}]tpt-42) (myo) {};
\node[draw,minimum height=5cm,minimum width=7mm] (actin) at ([shift={(2.5cm,-2mm)}]tpt-46) {};
\draw (myo.east) -- ($(myo.east)!0.5!(actin.north west)$) coordinate (a);
\draw[decorate] (a) -- ($(myo.east)!0.80!(actin.north west)$)coordinate (b);
\draw (b) -- (actin.north west);
\end{scope}
\begin{scope}[yshift=-7cm]
\begin{scope}[rotate=90]
\draw[domain=0.6*pi/3:6.5*pi/3,
samples=50,draw=none,
mark=o*,
mark size=2mm,
mark options={fill=white},
mark indices={1,...,4,13,14,...,30,36,37,...,40,42,43,...,49}] plot[parametric] function{3*sin(t),0.1*sin(2*t)};
\filldraw[fill=white] (tpt-49) ++(10:2mm) arc (110:40:2mm) arc (-40:156:2mm) arc (65:5:2mm) ;
\fill (tpt-42) ++(150:2mm) arc (150:210:2mm) arc (330:30:2mm);
\end{scope}
\node[draw,rotate=-20,anchor=west,shape=ellipse,ultra thick,minimum width=1cm] (a) at ([shift={(2mm,-3.5mm)}]tpt-42) (myo) {};
\node[draw,minimum height=5cm,minimum width=7mm] (actin) at ([shift={(2.5cm,-2mm)}]tpt-46) {};
\draw (myo.east) -- ($(myo.east)!0.5!(actin.north west)$) coordinate (a);
\draw[decorate] (a) -- ($(myo.east)!0.80!(actin.north west)$)coordinate (b);
\draw (b) -- (actin.north west);
\end{scope}
\end{scope}
\begin{scope}[xshift=7cm]
\begin{scope}
\begin{scope}[rotate=90]
\draw[domain=0.6*pi/3:6.5*pi/3,
samples=50,draw=none,
mark=o*,
mark size=2mm,
mark options={fill=white},
mark indices={1,...,4,13,14,...,30,36,37,...,40,42,43,...,49}] plot[parametric] function{3*sin(t),0.1*sin(2*t)};
\filldraw[fill=white] (tpt-49) ++(10:2mm) arc (110:40:2mm) arc (-40:156:2mm) arc (65:5:2mm) ;
\filldraw[fill=white] (tpt-42) ++(150:2mm) arc (150:210:2mm) arc (330:30:2mm);
\end{scope}
\node[draw,anchor=west,shape=ellipse,ultra thick,minimum width=1cm] (a) at ([shift={(7mm,-3.5mm)}]tpt-42) (myo) {};
\node[draw,minimum height=5cm,minimum width=7mm] (actin) at ([shift={(2.5cm,-2mm)}]tpt-46) {};
\draw (myo.east) -- ($(myo.east)!0.5!(actin.north west)$) coordinate (a);
\draw[decorate] (a) -- ($(myo.east)!0.80!(actin.north west)$)coordinate (b);
\draw (b) -- (actin.north west);
\end{scope}
\begin{scope}[yshift=-7cm]
\begin{scope}[rotate=90]
\draw[domain=0.6*pi/3:6.5*pi/3,
samples=50,draw=none,
mark=o*,
mark size=2mm,
mark options={fill=white},
mark indices={1,...,4,13,14,...,30,36,37,...,40,42,43,...,49}] plot[parametric] function{3*sin(t),0.1*sin(2*t)};
\filldraw[fill=white] (tpt-49) ++(10:2mm) arc (110:40:2mm) arc (-40:156:2mm) arc (65:5:2mm) ;
\filldraw[fill=white] (tpt-42) ++(150:2mm) arc (150:210:2mm) arc (330:30:2mm);
\end{scope}
\node[draw,rotate=-20,anchor=west,shape=ellipse,ultra thick,minimum width=1cm] (a) at ([shift={(7mm,-3.5mm)}]tpt-42) (myo) {};
\node[draw,minimum height=5cm,minimum width=7mm] (actin) at ([shift={(2.5cm,-2mm)}]tpt-46) {};
\draw (myo.east) -- ($(myo.east)!0.5!(actin.north west)$) coordinate (a);
\draw[decorate] (a) -- ($(myo.east)!0.80!(actin.north west)$)coordinate (b);
\draw (b) -- (actin.north west);
\end{scope}
\end{scope}
\end{tikzpicture}
\caption[Muscle activation mechanism]{Muscle activation mechanism. The sequence is given in a clockwise (or anti-clockwise) fashion (adapted from \cite{geeves}).}%
\label{fig:lit:muscle}%
\end{figure}
The skeletal muscle activation takes place via a process described by the \emph{sliding filaments model} (\Cref{fig:lit:filament}).
The muscles consist of muscle fibers and muscle fibers are made up of \emph{myofibrils}. The myofibrils involve different types of thin
and thick filaments, mainly of the type \emph{actin}, \emph{myosin} and \emph{titin} filaments. The myosin filaments involve
extensions that can bind to the thin actin filaments. The relative motion of these filaments are produced due to these
extensions via ATP hydrolysis. Moreover, these extensions stay connected or disconnected to actin filament unless more
ATP is utilized. Hence, the muscle needs extra energy to relax which is the reason behind \enquote{\emph{rigor mortis}}
and some types of the muscle spasms when this additional energy source can not be provided or ATP can not bind to the
actin filament for some reason. Moreover, binding or detachment of these extensions is also regulated by the
$\mathrm{Ca}^{++},\mathrm{Mg}^{++}$ concentration in the muscle cell and controlled, eventually, by motor neurons using yet another
chemical trigger.
\begin{figure}%
\centering
\includegraphics[height=7cm]{\impath/literature/filament}%
\caption[The simplified illustration of the sliding filaments model for a single sarcomere.]%
{The simplified illustration of the sliding filaments model for a single sarcomere. (Source: \cite{ottenheijm})}%
\label{fig:lit:filament}%
\end{figure}
Therefore, the individual muscle activation at the basic level is analogous to a clock escapement mechanism \cite{headrick}
i.e. each time the lever arm pulls back a thread of the gear and the cycle repeats. The resulting relative motion is very much
like a graphite coming out of a mechanical pencil when pressed from the eraser cap. Obviously, muscle behavior is not rachet-like
but smooth. This is because each of these mechanisms operates independently. Thus, at each time instant, different myosin extensions
can be found at different phases of the cycle very much like a helical gear pair that are always in contact at one point.
Using this analogy, ATP molecules are used to open and to close the pencil clutch made up of myosin extensions and actin filament
is pushed forwards. Moreover, titin filaments can be thought of as the connecting rod of the pencil from the cap to the clutch which
is mostly responsible for the passive elasticity of the muscles.
In summary, the muscles have varying non/backdrivable configurations. Moreover, we can lock our muscles in place
e.g., we can try to keep our arm at one position during drilling etc. to increase the precision. Then, the arm becomes
a stiff object with inherent stiffness of the connection rod (titin), pencil clutch (myosin actin), muscle tendons and
various other involved processes resisting to the applied strain.
Coming back to our original discussion, the stiffness of the arm that has been the subject of the experiments mentioned
above is, again invoking the analogy, based on the closed clutch response of the arm. In other words, what is measured
is a cumulative spring-like behavior of the arm that is actively controlled to be kept at a certain configuration. This
is related to the \enquote{human's restoring force} given in \Cref{rem:filament}.
Having a spring-like property naturally implies that the arm is passive. But the problem with this argument, as far as
we understand, is that there is no reason to assume that the human arm model involves a symmetric positive definite
stiffness matrix at each time instant for an arbitrary trajectory. In fact, as shown clearly in \cite{mussa85}, the major eigenvector
of the stiffness matrix varies both in terms of direction and magnitude. Therefore, it might be possible to extract energy
from the human arm with some particular pathological trajectory. Just in the case of a frozen time analysis of a time-varying
operator does not imply stability, the conclusion is only valid for postural analysis at a fixed configuration of the human arm
but not for an arm trajectory.
\begin{quote}
Therefore, it is our belief that the assumption of human arm being a passive system is incorrect for most teleoperation tasks.
\end{quote}
Since it is customary in the literature to include the passivity hypothesis, let us invoke it here too for the sake of the argument.
The question of how, then, a human can possibly move anything while remaining passive is one that makes the whole story even more
complicated. The voluntary input of the human is taken as an exogenous and state independent input to the system. Hence, the human
cognitive input is an additional but independent force source acting on the handle together with the passive human arm immitance
as depicted in \Cref{fig:lit:stateindep}. In other words, the contribution of the exogenous force and the arm dynamics are assumed
to remain mutually exclusive.
\begin{figure}%
\centering
\begin{tikzpicture}
\tikzstyle{spring}=[thick,decorate,decoration={coil,pre length=0.3cm,post length=0.3cm,segment length=1mm,amplitude=2mm}]
\tikzstyle{damper}=[thick,decoration={markings,
mark connection node=dmp,
mark=at position 0.5 with
{
\node (dmp) [thick,inner sep=0pt,transform shape,rotate=-90,minimum width=15pt,minimum height=3pt,draw=none] {};
\draw [thick] ($(dmp.north east)+(2pt,0)$) -- (dmp.south east) -- (dmp.south west) -- ($(dmp.north west)+(2pt,0)$);
\draw [thick] ($(dmp.north)+(0,-5pt)$) -- ($(dmp.north)+(0,5pt)$);
}
}, decorate]
\draw (-5mm,10mm) arc (120:240:1.7mm)
arc (140:240:2mm)
arc(140:230:2mm)
arc(150:230:1.7mm)
to[out=-20,in=190] ++(1.3cm,1mm)
to[out=-5,in=200] ++(3cm,-1mm)
coordinate (kose);
\coordinate (ustkose) at ([yshift=10mm]kose);
\draw (kose) .. controls +(60:2mm) and +(-60:2mm) .. ++(0,5mm) coordinate (orta);
\filldraw[pattern=north east lines]
(orta) .. controls +(120:3mm) and +(190:2mm) .. (ustkose)
.. controls +(-10:2mm) and +(60:3mm) .. (orta);
\draw (ustkose) to[in=-15,out=160] ++(-3cm,1mm)
to[in=20,out=170]
coordinate [pos=0.6](b)
coordinate [pos=0.8](a) (-5mm,10mm);
\fill[gray] (a) |- ++(2mm,15mm) -|(b);
\draw[-DIN,ultra thick] ([shift={(2mm,10mm)}]a) -- ++(-2cm,0) node[above,midway] {$F_{exo}$};
\coordinate (yay) at ([shift={(5mm,5mm)}]orta);
\coordinate (damp) at ([shift={(5mm,-5mm)}]orta);
\coordinate (duvar) at ([shift={(2cm,0)}]orta);
\draw[thick] (orta) -| (yay) (orta) -| (damp) ([yshift=7mm]duvar) |-(-1cm,-5mm);
\fill[pattern=north east lines] ([yshift=7mm]duvar) |-(-1cm,-5mm) --++(0,-2mm) -| ([shift={(2mm,7mm)}]duvar);
\draw[spring] (yay) -- ++(15mm,0);
\draw[damper] (damp) -- ++(15mm,0);
\end{tikzpicture}
\caption[A pictorial representation of the state independent human force input]%
{A pictorial representation of the state independent human force input. By state dependence, we mean that the
vector $F_{exo}$ is completely decoupled from the arm characteristics and can be taken as a genuine exogenous input.}%
\label{fig:lit:stateindep}%
\end{figure}
The reader would spot that this is not in line with Hogan's remark since the activation state in the muscles should be altered in
order to apply some force. In other words, we need to alter the stiffness matrix of the arm to be able to move but, to the best of our
knowledge, this is not given in any teleoperation study. In other words, the
answer to the question of whether the human's restoring force would exhibit the same properties during the force exertion phase is
not known to us. It has been also shown that the applied force is also related to the muscle tension-length properties \cite{eltit}
which we have not considered in our analogy.
Therefore, we have to further separate the human force into two parts, namely, the active neuromuscular feedback force that keeps the
human arm passive and the voluntary and cognitive force input applied to the system. How this is usually performed is not clear in the
literature. To the best of our knowledge, this issue is considered (but still briefly) only in \cite[Sec. II.B]{kazeroonitsay} and
references therein.
This ambiguity becomes much more important since the control oriented focus of this thesis necessitates that we concentrate on
worst cases rather than the experiments performed within the cognitive range of human operators. In other words, we are interested in
the cases where things go wrong due to various reasons, such as sampling disturbances, measurement noises, directionality
etc. Therefore it cannot be a satisfactory argument if stability depends on the user's neuromuscular feedback or, simply, the user's
stabilization capabilities. In order to use the bilateral teleoperation devices in real-life, stability should be addressed
regardless of the set of prescribed human actions. Hogan's findings are not sufficient for supporting the passivity assumption
often found in the literature.
In summary, the passivity of the human and the environment (or the virtual environment in haptics/virtual reality applications), is
only plausible in certain cases; then it should be verified regardless, in order to assume that the corresponding mathematical models
are passive. Still, analysis and synthesis methods that invoke this assumption lead to many real-world implementations
with varying degrees of realism. An obvious follow-up question is \enquote{how can then the reported analysis and synthesis results
based on this passivity assumption lead to successful implementations which would imply the validity of the assumption?}. We need to
focus on the \enquote{successful} part in order to answer this question. If we again scan through the literature these successful
implementations are only successful under strict assumptions about the user behavior together with any combination of the following
device properties
\begin{itemize}
\item The controller stabilizes the loop via excessive apparent damping hence either or both free air or hard contact realism is lost.
\item The bandwidth of the force and position tracking are exceedingly small.
\item The remote and local devices exhibit position drifts over time.
\item The performance of the device is time-varying and not uniform.
\end{itemize}
We argue that the success of these methods is due to the conservatism of the analysis/synthesis tools and does not validate the
passivity hypotheses on the respective models. Many frequency-domain methods, which assume LTI terminations at both ends of the
teleoperation network, are reported to achieve stable bilateral interaction in real setups. While impressive, this shouldn't have
been possible had the tests not been exceedingly conservative since the real setups involve sudden contacts i.e. they are essentially
time-varying systems. We recall the well-known fact that the behavior set of the time-varying systems are significantly richer than
LTI systems. Once again, we remind the reader that the proper metrics to gauge the success of teleoperation systems are not known and
we use the typical error-norm based control design rationale.
A compact version of the argument above is given by Yokokohji and Yoshikawa in \cite{yokokohjiyoshikawa}:
\begin{quote}
Passivity of the system can be a
sufficient condition of stability only when the system interacts
passive environments. In the case of master-slave systems, if
we could assume that the operator and the environment are
passive systems, then the sufficient condition of stability is
that the master-slave system itself must be passive. Strictly
speaking, however, the operator is not passive because he/she
has muscles as the power source. Colgate et al. [21]\footnote{Reference
\cite{colgatehogan88} of this thesis.} mentioned
that even if the system has an active term, the system stability
is guaranteed unless the active term is in some way state dependent.
Obviously, the operator is passive when $\tau_{op}= 0$.
Therefore, we will give the following assumption about $\tau_{op}$:
\enquote{\emph{The operators input $\tau_{op}$ independent to the state of the
master-slave system. In other words, the operator does not
generate $\tau_{op}$ that will cause the system to be unstable.}}
Dudragne et al. [3]\footnote{Reference \cite{dudragne} of this thesis.} gave a similar assumption in order to use
the concept of passivity for stability distinction. The above
assumption seems tricky in a sense, but it is necessary to ensure
the system stability by the passivity.
\end{quote}
\noindent Finally, a supplementary remark is also given by Buerger and Hogan in \cite{buergerhogan1}:
\begin{quote}
When passivity is used as a stability objective, the only assumption
made about the environment is that it, too, is passive.
This is likely sufficient to guarantee coupled stability with humans
(though, to date, it has not been conclusively proven that
human limbs are passive; see [29]\footnote{Reference \cite{hogan89}
of this thesis.} for an argument for treating them as such). However,
given the properties of human arms
described above, passivity is unnecessarily restrictive. Our experience
has shown that some controllers that are known to be
nonpassive are adequately stable in clinical rehabilitation tasks [26]%
\footnote{Reference \cite{buergerconf} of this thesis.}.
\end{quote}
We should mention here that Hogan's paper together with other identification experiments are extremely important for many
fields and needs no motivation. The discussion above only points out that the frequently reported inference that follows from
his results is not in line with the results.
The idea of modeling the teleoperation as a two-port network seems to have multiple origins and we have no reference to
point out a common source. However, in general, the popularity of two-ports can be attributed to \cite{andersonspong,nieslotine,
rajuphd,hannaford89,yokokohjiyoshikawa}.
\subsection{Uncertain Models of Bilateral Teleoperation for Robustness Tests}\label{sec:lit:uncmodel}
Another possibility of modeling the human arm and its cognitive input is to define a reference position signal \enquote{filtered} by the
human arm impedance\footnote{The term \emph{impedance} is used in a more general sense than its common usage to denote
LTI transfer functions such that no distinction is made between linear and nonlinear or time-invariant and time-varying
operators.} e.g., \cite{leelee,kazeroonitsay}. Various studies pointed out that the identification experiments suggest a
mass-spring-damper system pattern is evident in the frequency response data of the human arm recorded under various task
performance similar to the one given in \cite{hogan89}. The general method is to instruct the human to perform a specific
task and then perturb the hardware with certain predesigned disturbance signals such that the output can be evaluated to obtain
a mathematical model. In the literature, the model structure is often set a priori to be a second order transfer function and
the parameters are optimized to minimize the mismatch between the experimental and predicted response. It is also well-known
that the human can change the inherent impedance of the arm during the task execution (see, e.g., \cite{tsujimorasso}).
Therefore, the studies are performed in the ranges where it is safe to assume that the human arm characteristics are constant
or constant up to negligible changes.
The modeling is straightforward via uncertain mass-spring-damper system differential equation manipulations. Suppose the human arm
admits the second order model
\[
M(\Delta_1)\ddot{x} + B(\Delta_2)\dot{x} + K(\Delta_3)x = F_h -F_m
\]
where $F_h,F_m$ denote the human force and the force feedback inputs, respectively. Then choosing a multiplicative or additive uncertainty
structure and via basic linear fractional transformations, the signal relations are converted to the interconnection shown in
\Cref{fig:lit:figunc}. The arrows are deliberately left out as it is up to the designer to get different immitance models.
\begin{figure}%
\centering
\begin{tikzpicture}
\matrix[draw,matrix of math nodes] (m) {\Delta_1 &&\\&\Delta_2 &\\&&\Delta_3\\};
\node[below=1cm of m,draw,minimum size=2.0cm] (g) {$G_{nom}$};
\draw (g.155) -| ++(-1cm,1cm) |- (m.west);
\draw (g.25) -| ++(1cm,1cm) |- (m.east);
\draw (g.-155) -- ++(-1cm,0cm) (g.-25) -- ++ (1cm,0cm);
\end{tikzpicture}
\caption{Uncertain model representation by taking out the uncertainty blocks}%
\label{fig:lit:figunc}%
\end{figure}
Many studies have appeared in the literature regarding such modeling and the majority of these assume a
mechanical model of order from two to five. Note that this is an assumption made a priori and only applies
to the specific task performed by the human in the experiment from which the frequency response data is
collected. The commonly utilized models, unfortunately most given for the fixed postural arm
configuration can be found in \cite{kazeroonitsay,tsujigoto,kosuge,colgate1,speich,fucavus,buergerhogan1,leungfa,
husalculoewen,andrifour,laroche}. Obtaining these measurements are time-consuming and difficult to parameterize. For this reason, although
the results along this direction are scarce, they are, as in the passivity case, very valuable.
The disadvantage of such parametrization of the human arm is contrasted with the passivity approach methods
invoking the argument of the time-varying nature of the arm parameters. It is often rightfully argued that the
uncertainty ranges in which the stiffness and damping (and partially inertial) coefficients change, are too
large to be considered in the structured singular value based robustness tools. Moreover, many auxiliary
effects such as the visual feedback, cognitive lag of the brain etc. are not considered in the identification
experiments; in the passivity approach all of these are lumped into a single port condition. Obviously,
the main difficulty is to get a model (out of hypothetical insights, identification experiments etc.) which is
not required in the passivity approach.
The papers \cite{leelee,prekopiou} offer interesting alternatives for human modeling as they attempt to incorporate
many of the aforementioned effects, but the results are prospective and yet to be utilized.
\section{Analysis}\label{sec:litchapanalysis}
The stability analysis is one of the major problems in designing stable yet high-performance teleoperation
systems. It's often not feasible to manually tune some local controllers and make test subjects use it in order to verify the
design specifications. Moreover, by relying only on the experiments, one can miss an important destabilizing scenario if the field
experiments do not cover that particular case. Hence, an a priori certificate of stability is much sought after. The stability
analysis can also give some guidelines about the parameter selection in the hardware design phase and can lead to minimized
design iterations. Therefore, having a realistic stability test is essential in building these systems.
Similar to the modeling section, the analysis in the literature extensively relies on network theory based results. In fact
this is where the network theory stands out as a complete tool for analysis and synthesis of bilateral teleoperation systems
via the hypothesis that human and the environment models are passive.
The common terminology for stability is somewhat different than that of the contemporary control theory as \emph{nominal stability}
is used for the stability properties of isolated two disjoint media; when the interaction is set up between these two media the closed-
loop stability problem is called \emph{coupled stability}. To the best of our knowledge, this terminology is introduced in
\cite{colgatehogan88} hence we refer to this paper (or Colgate's thesis \cite{colgatephd}) for more details.
It's also worth mentioning that the passivity and stability is used often interchangeably and also usually referred
to the classical texts \cite{haykin,mitra69,chen91} for the precise definitions. Hence, there is a little guesswork
required to classify the stability definitions given in the literature in order to locate which version is meant.
The important distinguishing point is that marginal stability is often accepted in the definition of stability results
since it arises frequently in lossless (hence passive) models where energy conservation is assumed. However, the analysis
results that rely on such assumptions do not guarantee asymptotic interconnection stability, but only lead to certain
passivity properties of the interconnection (see \cite[Thm. 6.1]{khalil} and \cite[Sec. V]{lawrence} for the discussion
on strict passivity).
As given in \Cref{sec:litchapmodeling}, the passivity property is crucial to many studies in the literature. The direct
physical interpretation of the abstract concepts gives even more appeal to such energy book-keeping methods. Another
advantage of passivity methods is that the nonlinear counterparts of the results are also available in the literature
and relatively easy to utilize. However, this convenience misses out many relevant details that are specific to
teleoperation systems and result with too general conditions. Let us recall a general version of the passivity theorem:
\begin{thm} The negative feedback connection of two passive systems is passive. The negative feedback connection of a passive system with
a strictly passive system is asymptotically stable.\end{thm}
Note that, this result is valid for both nonlinear and linear systems. Invoking the theorem twice on the teleoperation
system allows us to conclude that, under the passivity assumption of the human and the environment, if the two-port is
passive then the interconnection is passive. Moreover, if any of the involved operators is strictly passive the teleoperation
system is asymptotically stable.
\begin{figure}%
\centering
\includegraphics[width=0.6\columnwidth]{\impath/literature/train}%
\caption[A transparent two-port network with passive terminations]{A transparent two-port network with passive
terminations. The actuation of the railcar is taken as a state-independent input to the system. (Source: \cite{train})}%
\label{fig:lit:train}%
\end{figure}
The passivity theorem is in general not necessary for stability but only sufficient. Because there do exist stable
interconnections that involves nonpassive subsystems. In particular, the conservatism brought in by passivity
assumption is arbitrarily high (especially in the nonlinear case). Facetious as it may seem, the test also takes
into account the port terminations shown in \Cref{fig:lit:train} for a table-top joystick. We have to emphasize
that the three-carriage railcar with two cabs, is a valid, almost perfectly transparent two-port network with
passive end terminations. Hence, a regular passivity-based stability test includes those end terminations for
a simple hand-held device. In other words, if we only consider the passivity property of the involved subcomponents and do not
distinguish further, there is no possible way to distinguish a table-top teleoperation system from a train since
both are passive. It's that conservative.
The major disadvantage of the passivity methods is that the procedure is focused almost only on the
energy exchange. The performance specifications are very difficult to formulate and also difficult
to integrate into the analysis and synthesis steps using only the inner product structure. As an
example, the signals that are not port variables such as position errors, nonlinear effects etc.,
that are functions of these signals, can't be utilized easily in the performance specifications.
Same difficulty arises in the normed space structures though much more can be achieved. Similarly,
peak-to-peak gain minimization methods are not mature enough to handle any practical system without
excessive conservatism.
Another disadvantage is that the power- or the energy-based analysis, due to the inner-product structure,
can not distinguish the individual signals. Consider the ideal case where the human and the local
device is pushing each other and cancelling each other's contribution. In this case the external or observable
energy exchange based on the port variables is zero (negligible) which can not be distinguished from the case of
not touching at all (a small motion on the device).
When the network, human, and the environment models are assumed to be Linear Time Invariant (LTI), the frequency domain methods
allow us the analyze the teleoperation systems for stability and performance. The most common stabiliy analysis tool for such
models is the Llewellyn's stability criteria (also often called absolute stability theorem or unconditional stability theorem).
For linear networks, the following definitions seem to be used quite widely (modified from \cite{chen91}):
\begin{define}[Potential Instability] A two-port network is said to be potentially unstable if there exist two passive one-port
immitances that, when terminated at the ports, produce a persisting natural frequency.
\end{define}
Notice that this definition is equivalent to the common \emph{robust stability} definition which states that if a two port
network is robustly stable there exists no pair of one-port immitances that makes the network exhibit a persisting natural
frequency.
\begin{define}[Absolute Stability at $\iw_0$] A two-port network is said to be absolutely stable if it is not potentially unstable.
\end{define}
\subsection{Llewellyn Stability Criteria}\label{sec:llewellyn}
The well known conditions for stability of a two-port network, formulated in \cite{llewellyn,bolinder,rollett},
are recalled in \Cref{chap:apdxnetwork}. As shown in \cite{rollett}, the conditions stated in \Cref{thm:apdx:llw}
are invariant under immitance substitution. This result forms the basis for almost all passivity-based frequency
domain bilateral teleoperation stability analysis approaches in the literature. We also derive this theorem
from an IQC perspective and show that it is actually the passivity counterpart of the of the $D$-scalings in the
$\mu$-tools. This has also been derived in a scaled teleoperation context in \cite{poortenyokokohji} using only
Structured Singular Value (SSV) arguments.
Thanks to the frequency domain formulation, it is possible to rewrite the condition \eqref{eq:lit:llew3} as a fraction
and see the problematic regions in which the fraction gets close to or crosses to the instability, together with one of the
conditions given in \eqref{eq:lit:llew2}.
\begin{figure}%
\centering
\begin{tikzpicture}[>=stealth]
\node[draw,minimum size=1cm] (h2) {$H_2$};
\node[draw,minimum size=1cm, above=1cm of h2] (h1) {$H_1$};
\node[draw,circle,inner sep=1mm,right= 1cm of h2] (a1) {};
\node[draw,circle,inner sep=1mm,left= 1cm of h1,label={[inner sep=0]-60:$-$}] (a2) {};
\draw[->] (h1) -| (a1) node[right,midway] {$y_1$};
\draw[->] (h2) -| (a2) node[left,midway] {$y_2$};
\draw[->] (a1) -- (h2) node[above,midway] {$e_2$};
\draw[->] (a2) -- (h1) node[above,midway] {$e_1$};
\draw[<-] (a1) -- +(1cm,0) node[right] {$u_2$};
\draw[<-] (a2) -- +(-1cm,0) node[left] {$u_1$};
\end{tikzpicture}
\caption{Negative feedback interconnection}%
\label{fig:lit:passint}%
\end{figure}
\subsection{\texorpdfstring{$\mu$}{mu}-analysis}
As given in \Cref{sec:lit:uncmodel}, stability in the face of uncertainties can also be analyzed in the generalized plant framework
of robust control. After rewriting the signal relations, the teleoperation system can be written as an uncertain interconnection as
shown in \Cref{fig:lit:uncincgeneral}. In this setting, $G$ is the model of the nominal bilateral teleoperation system and $\Delta$
is a block diagonal collection of uncertainties, such as the human, the environment, delays, etc. Stability tests are based on structural
hypotheses on the diagonal blocks of the operator $\Delta$ such as gain bounds or passivity. These properties should allow us to develop
numerically verifiable conditions for the system $G$ that guarantee interconnection stability. This is intuitive because we have no access
to the actual $\Delta$ and we can only describe its components by means of indirect properties.
If the interconnection subsystems are represented in the scattering parameters, the $\mu$ test is precisely equivalent to the test of
Llewellyn's theorem and often called as Rollett's stability parameter. This is due to the well-known equivalence between the small-gain
and passivity theorem \cite{desvid}. We have to note that the equivalence is stated in terms of the stability characterization. Otherwise,
the small-gain theorem requires a normed space structure whereas passivity theorem requires an inner product space structure, hence the
applicability is relatively limited. This is also related to the fact that we need to work with power variables exclusively in the
passivity framework and this is not always convenient if the performance specifications are related to other variables.
In the literature, this analysis method often follows the scattering transformations such that, the passivity assumption avoids the
explicit modeling and then small-gain theorem is utilized mainly to handle the delay problem via with the involved norm bounded operators.
A refinement can be found in \cite{poortenyokokohji} where the authors utilize a direct $\mu$-analysis to reduce the conservatism,
however rather remarkably, it's not picked up by other studies and the analysis is mainly limited to small-gain conditions even in
the linear systems.
We could have also directly choosen to utilize the uncertain modeling of the human and the environment and utilize $\mu$-analysis for the
teleoperation system had the specific models been available.
\begin{figure}%
\centering%
\begin{tikzpicture}[>=stealth]
\node[draw,minimum size=0.75cm] (plant) at (0,0) {$G$};
\node[draw,minimum size=0.75cm] (unc) at (0,1.5) {$\Delta$};
\draw[->] (plant.west) -| ++(-5mm,10mm) node[left] {$v$} |- (unc.west);
\draw[->] (unc.east) -| ++(5mm,-5mm) node[right] {$u$} |- (plant.east);
\end{tikzpicture}
\caption{Uncertain Interconnection}%
\label{fig:lit:uncincgeneral}%
\end{figure}
\subsection{Modeling the Communication Delay}
Over the past two decades, it has been confirmed in various studies that, if present, communication delays are a major
source of instability (reports date back to 60's, e.g., \cite{sheridanferrell} and the references in \cite{andersonspong}).
Even when the delay duration $t$ is known and constant, the delay operator can be shown to be nonpassive since $e^{-st}$ is
not positive real. Hence, when combined with the passivity framework, it violates the assumptions on the uncertain operators.
At end of the 80's and early 90's, two prominent studies (\cite{andersonspong,nieslotine}) proposed to handle the delay
robustness problem using scattering transformations. This notion is best explained, in our humble opinion, by loop transformations
since the original articles refer to microwave and transmission line theories which use quite specialized terminology. One can also
find a slightly different system theoretical view of these transformations in \cite{colgate3}. If we restrict the discussion to LTI
operators\footnote{In the nonlinear case, it's a \emph{completion of square} argument to switch from the inner product structure to a
norm structure provided that the signal space is suitable for such operation.}, the key concept of the scattering transformation
or the wave variables methods is to map the closed right half plane to the closed unit disk via a special case of bijective M\"{o}bius
(or linear fractional or bilinear) transformation:
\begin{equation}
W: \Complex_+\cup \Complex_0 \mapsto \left\{ z\in\Complex \mid \abs{z} \leq 1 \right\}\, ,\, W(z) = \frac{z-1}{z+1}
\label{eq:lit:smith}
\end{equation}
One can directly verify that $1\mapsto 0,\infty\mapsto 1$ and $0\mapsto -1$ under $W$. Pictorially, the mapping is given in
\Cref{fig:lit:smith} using a Smith chart which is located at the origin. Hence, positive real transfer matrices
become norm bounded by $1$ such that we can analyze the interconnection using the small-gain theorem.
Let us demonstrate a few properties of this transformation. First, this mapping can be shown with a block diagram.
Assume that $G$ is a proper positive real LTI SISO system and let the input/output relation be given by $y=Gu$. Then,
with a standard manipulation, we obtain a feedback interconnection that leads to the mapping
\begin{equation}
W(G(s)) = \frac{G(s)-1}{G(s)+1} = -1 + \frac{2G(s)}{G(s)+1} \Longrightarrow
\begin{tikzpicture}[baseline=(g.center),scale=0.5,transform shape,every node/.style={draw,minimum size=1cm},every path/.style={->},>=stealth]
\node[draw] (g) at (0,0) {$G$};
\node[draw,left=1cm of g] (s1) {$\sqrt{2}$};
\node[draw,right=1cm of g] (s2) {$\sqrt{2}$};
\node[draw,circle, right=1cm of s2,minimum size=0,inner sep=3pt] (j) {};
\draw (s1) -- (g) node[midway,circle,fill=white,minimum size=2mm,inner sep=0] (f) {};
\draw ($(g)!0.5!(s2)$) |- ++(-1cm,-1cm) -| (f);
\draw (g) -- (s2);\draw (s2)--(j);\draw (j)--+(1cm,0);
\draw (s1) ++(-25mm,0) -- (s1);
\draw (s1) ++(-15mm,0) |- ([shift={(-1cm,1cm)}]j.center) -| (j);
\path (j)++(50:3mm) node[draw=none,minimum size=0,inner sep=3pt] {$-$};
\node[draw=none] at ([shift={(2mm,2mm)}]f.north) {$u$};
\node[draw=none] at ([shift={(-23mm,3.5mm)}]s1) {$\hat{u}$};
\node[draw=none] at ($(g)!0.5!(s2)+(0,2.7mm)$) {$y$};
\node[draw=none] at ($(j)+(0.8cm,3.2mm)$) {$\hat{y}$};
\end{tikzpicture}
\label{eq:wavetrafo}
\end{equation}
Simply following the signal paths, we also see that the input/output relation becomes
\[
\tilde{y}\coloneqq W(G(s))\tilde{u}.
\]
where
\begin{equation}
\pmatr{\tilde{y}\\\tilde{u}} = \sqrt{2}\pmatr{-1 &1\\1 &1}\pmatr{u\\y}
\label{eq:scatrafo}
\end{equation}
Note that, physical realization of this transformation requires a feedforward and
a feedback control action to which typically referred with \emph{\enquote{Wave encoding}}.
Often these variables are normalized with $\sqrt{2}$ at the outset. The seperation of $\sqrt{2}$
blocks is a matter of convention and provides symmetry in the block diagrams. Also we have,
\begin{align*}
(W\circ W)(G) &= \frac{-1}{G},\\
(W\circ W \circ W) (G) &= \frac{-1}{W(G)},\\
(W\circ W \circ W \circ W) (G) &= G\\
\end{align*}
which shows the effect of the $90^\circ$ clock-wise rotations of the Riemann sphere about the axis
parallel to the imaginary axis ($W$ is an element of M\"{o}bius group with $\circ$ operation). This
stereographic projection idea is also the main idea behind the derivation of the stability parameter
of Edwards and Sinsky (\cite{edsin}). Moreover, from \Cref{fig:lit:smith}, we can
see a visual proof of Theorem 3.1 in \cite{andersonspong} which states that;
\[
\Re{G(\iw)}\succeq 0 \iff \norm{W(G(s))}_\infty\leq 1
\]
\begin{figure}%
\centering
\begin{tikzpicture}[scale=0.6]
\begin{smithchart}[show origin,
axis background/.style={shading=mysphere},
no marks,
grid=major,
axis equal,samples=150,
]
\foreach \t in {0,0.2,0.5,1}{
\addplot+[ultra thick,domain=-10:10] (\t,\x);
}
\end{smithchart}
\begin{scope}[shift={(8cm,0)}]
\begin{axis}[axis lines=middle,
grid=major,
no marks,
axis background/.style={right color=black!50,left color=white},
xmax=1.5,
%ytick=\empty,
every axis/.append style={
extra description/.code={
\node[xshift=-5mm] at (0,1) {Im$(z)$};
\node[xshift=-5mm] at (1,0.55) {Re$(z)$};
}}]
\foreach \x in {0,0.2,0.5,1}{
\addplot+[ultra thick] coordinates {(\x,-1.5) (\x,1.5)};
}
\end{axis}
\end{scope}
\end{tikzpicture}
\caption[Mapping the closed right half plane onto the unit disc.]%
{Mapping the closed right half plane onto the unit disc.}%
\label{fig:lit:smith}%
\end{figure}
%
%\begin{figure}[b]%
%\centering
%\begin{tikzpicture}
%\begin{axis}[unit vector ratio=1 1 1,view={65}{20},hide axis,colormap/invgray,ymax=4]
%\begin{scope}[/pgfplots/restrict z to domain=-1:0];
%\addplot3[surf,shader=interp,z buffer=sort,samples=19,variable=\u, variable y=\v,domain=0:180, y domain=0:180]({cos(u)*sin(v)}, {sin(u)*sin(v)}, {cos(v)});
%\addplot3[opacity=0.5,mesh,draw=black,ultra thin,samples=19,variable=\u, variable y=\v,domain=0:180, y domain=180:360] ({cos(u)*sin(v)}, {sin(u)*sin(v)}, {cos(v)});
%\end{scope}
%\shade[right color=black!50,left color=white](axis cs:-2,0,0) -- (axis cs:2,0,0) -- (axis cs:2,2,0) -- (axis cs:-2,2,0)-- cycle;
%\draw (axis cs:0,-1.2,0) -- (axis cs:0,2,0);
%\draw[white,thick] (axis cs:-2,0.3,0) -- (axis cs:2,0.3,0);
%\draw[white,thick] (axis cs:-2,0.70,0) -- (axis cs:2,0.70,0);
%\begin{scope}[/pgfplots/restrict z to domain=-0.01:1]
%\addplot3[surf,shader=interp,z buffer=sort,samples=19,variable=\u, variable y=\v,domain=0:180, y domain=0:180]({cos(u)*sin(v)}, {sin(u)*sin(v)}, {cos(v)});
%\addplot3[opacity=0.5,mesh,draw=black,ultra thin,samples=19,variable=\u, variable y=\v,domain=0:180, y domain=180:360] ({cos(u)*sin(v)}, {sin(u)*sin(v)}, {cos(v)});
%\draw (axis cs:-2,0,0) -- (axis cs:2,0,0);
%\end{scope}
%\addplot3+[domain=0:18,samples=100,no marks,white,samples y=0] ({cos(x)*sin(5*x)},{sin(x)*sin(5*x)},{cos(5*x)});
%\addplot3+[solid,domain=0:45,samples=100,no marks,white,samples y=0] ({cos(x)*sin(2*x)},{sin(x)*sin(2*x)},{cos(2*x)});
%\end{axis}