Heuristic Considerations On Fast Axopodial

Transcription

Heuristic Considerations On Fast Axopodial
Heuristic Considerations On Fast Axopodial Contraction
In The Heliozoa
by
Bruce Eric Davis
April 1994
Institute of Statistics Mirneo Series
#2261
BMA Series #38
1
1. INTRODUCTION
The Heliozoa are a polyphyletic assemblage of predaceous freshwater and marine protozoa whose
globose cell bodies (diameter"" 8-2600 ]lm) bear many (20-several hundred) radially disposed slender
protrusions called axopodia (diameter"" 0.1-10 ]lm;
length", 30-500 ]lm).
Each axopodium is
stiffened by an axoneme consisting of an array of microtubules (numbering 3- '" 500; each
~
25 nm
in outside diameter) cross-linked in a taxonomically diagnostic pattern (Febvre-Chevalier, 1985;
Siemensma, 1991).
Axopodia function in locomotion, cell fusion, cell division, egestion, attachment to substrata, and
prey capture.
Ciliates, flagellates, and small metazoa that happen accidentally to collide with an
axopodium may stick to it owing to a superficial mucous coat discharged by extrusomes. Prey thus
ensnared may then be pulled to the cell body by fast contraction of the axopodium (duration of
contraction
< 100 ms; Suzaki et al., 1980a). Ingestion of prey through phagocytosis ensues and takes
place on a time scale of tens of minutes.
Fast axopodial contraction proceeds by the breakage of the axonemal microtubules along their
lengths, but the speed of contraction is much greater than rates of microtubule breakdown measured
in vitro and in other organisms. Hence, either some component of the mechanism of fast axopodial
contraction is faster than a similar component of these processes, or a different mechanism is
involved. The most tenable hypothesis supported by experiments to date appears to be that an action
potential causes an influx of Ca2+ across the axopodial membrane, which releases some unknown
second messenger(s) (e.g., inositol l,4,5-trisphosphate or Ca2+ itself) to cleave the axonemal
microtubules by means of a bound second messenger-activated severing factor.
A conformational
change in a centrin- or spasmin-like calcium-binding protein then drives contraction.
(A flowchart
summarizing this scheme appears on the next page.) The present paper explores some speculations as
to how fast axopodial contraction occurs; a summary of the remaining sections of the report follows
below.
2
HYPOTHETICAL CONTRACTION SCHEME
ACTION POTENTIAL
CALCIUM INFLUX
ECOND MESSENGE
_
MICROTUBULES
ONTRACTILE ORGANELL
3
SUMMARY OF REMAINING SECTIONS
2. Estimation of length constant
Standard calculations reveal that the length constant is about twice the axopodial length. This
fact, together with recordings made by Febvre-Chevalier et al. (1989) of conduction speeds up to 3
mis, imply that the
action potential propagation time is a small fraction of the total contraction time.
3. Surface electrostatic charge
Surface electrostatic charge is not expected to impede axopodial motion because the Debye
length is very small (of the order of nm).
4. Second messenger diffusion
Calcium diffusion times, bas~d on the classical value of the diffusion coefficient (Hodgkin and
Keynes, 1957), are in the 7-50 ms range for large axopodia, and in the 9-120 J-ls range for small ones.
If a recently revised
value of the diffusion coeficient is used (Allbritton et al., 1992), the diffusion
times increase to as much as 660 ms for large axopodia, suggesting that calcium diffusion might be
rate-limiting.
However, internal calcium stores, possibly activated by the action potential, may
appreciably shorten diffusion distances. Diffusion times of 0.5-4.5 ms are expected for large axopodia
in this case.
For
inositol 1,4,5-trisphosphate,
diffusion times of 1-88 ms are expected for large
axopodia, and 16 J-ls for small ones.
5. Calcium concentration to trigger contraction
For a number of protozoa under in vivo and in vitro conditions, the minimum calcium
concentration required for fast contraction is reported. To induce contraction only in the fast mode in a
certain marine heliozoan, the threshold calcium concentration in the external medium is 10- 6 M.
6. Estimate of force exerted by contractile organelle
Using published values for stresses developed by actomyosin and by spasmonemes, together
with cross-sectional areas estimated from electron micrographs, a value for the force produced by the
putative contractile organelle in large heliozoa is obtained that is close to that measured
experimentally.
4
7. Strain rate
Published data for stalk contraction (isochronous with axopodial contraction) indicate that the
time course is exponential (Febvre-Chevalier et al., 1992). A calculation shows that this observation
implies that the strain rate is constant.
8. Reynolds number
Reynolds numbers calculated on the basis of the length of the axopodium are certainly < 7 and
probably at most of the order of 10- 1. This implies that fast axopodial contraction occurs in a viscous
flow regime wherein inertial effects can be neglected.
9. Slender body theory
Treating the axopodium as a slender right circular cylinder translating axially normal to an
infinite plane wall yields estimates of the force and power of contraction, which are summarized in
section 11.
10. Sliding plane theory
Treating the axopodium as an infinite plane sliding at a specified angle into a stationary,
infinite plane wall yields estimates of the force and power of contraction, which are also summarized in
section 11.
11. Table of estimates of force and power
The caricatures of the two previous sections are applied to a large and a small heliozoan to
yield estimates of force and power during contraction that are similar to or below those determined by
experiments.
12. Estimates of energy uptake from food
A final consideration is how much energy, in terms of the amount obtained from prey capture,
would be spent in the contraction that captures the prey. The slender body theory of section 9 is used
in the calculation.
13. Detailed research plan outline
The preceding sections constitute but a preliminary survey of the problem of fast axopodial
5
contraction. This final section sets forth a detailed outline of a theoretical and experimental research
plan to elucidate the mechanism. A team of investigators would require several years to carry out this
ambitious program.
6
2. ESTIMATION OF THE LENGTH CONSTANT FOR PROPAGATION
OF AN ACTION POTENTIAL IN FAST AXOPODIAL CONTRACTION
According to the idealized passive one-dimensional cable model, the potential across the membrane
decays exponentially:
= V(O) e -Ixll "0 ,
V(x)
(1)
(2)
where V = potential across the membrane (mV), x = position along the length of the cable (cm),
"0 =length
constant (cm), d
= cable
diameter (cm),
Rm
= membrane
resistance (0 cm 2), and
Ri = resistivity of cytoplasm (0 cm) (Nossal and Lecar, 1991, p. 249). Note that
over which the membrane potential falls to
i of its initial value.
That is, V("o)
"0 is the distance
= V~O)
. Shepherd
"0 vs. d for three values of ~~ appropriate for unmyelinated
In the sequel, I read "0 from the middle curve of Fig. 13.
(1974, Fig. 13, p. 64) gives a graph of
1
mammalian nerve processes.
From Fig. 13 one has that for d = 0.1 JJm, corresponding to axopodia of Ci/iophrys marina and
Heterophrys marina,
"0
~
100 JJm, which is twice the typical axopodial length (50 JJm) of these
heliozoa (Davidson, 1975).
For Echinosphaerium nucleoli/um, d = 5-10 JJm (TUney and Porter, 1965, p. 341), for which
Shepherd's graph indicates
"0
~
700-1000 JJm. Tilney and Porter (1965, pp. 328, 341) state that the
axopodia were as much as 400 JJm long.
The foregoing results are borne out by direct calculation of the length constant from equation (2).
Measurements of
am for Heliozoa have not been made, but for ciliates, values corrected to allow for
the surface area of the ciliary membranes are 14.3 kO cm 2 for Paramecium, and 73 kO cm 2 for Stentor
(summarized by Wood, 1982, p. 538). [By comparison, for smooth muscle, Rm
al., 1983, p. 23).] Typically, Ri = 100 0 cm
(Jack et
kO cm (Jack et al., 1983, p.23), and for axopodia
Substituting these values in equation (2), we find that for
am = 14.3 kO
"0 = 2 x 102-2 x 103 JJm, whereas for Rm =73 kO cm2, ~o = 4 x 102_4 x 103 JJm.
Observe that
d = 1 x 10-5_1 x 10-3 cm.
cm 2,
= 0.1
= 40 kO cm 2
7
these length constants are comparable to those read from Shepherd's graph for unmyelinated
mammalian nerve processes.
GONGL USIONS
If fast axopodial contraction is actually triggered by an action potential in the heliozoa named
above, and if values of
~
I
are close to those for mammalian neural processes, then Ao appears to be
slightly greater than the typical lengths ofaxopodia.
Hence, as Dr. Tom Kepler points out
(conversation, 6/23/93), we may be confronted with an ambiguous intermediate case where Ao "" Lo '
rather than AO
Note that if AO
For Ao
~
~
~
Lo or AO
~
Lo ' where Lo is the initial (uncontracted) length of the axopodium.
Lo ' regeneration would be necessary if an action potential were to serve as a trigger.
Lo ' attenuation of an action potential would be negligible;
thus an action potential could
easily serve as a trigger in this case.
The values of AO found above are greater than the lengths ofaxopodia. Thus, we may conclude
that the action potential undergoes negligible attenuation and the axopodial cytoplasm must be nearly
isopotential. As Wood (1989, pp. 361-362) points out, action potentials function chiefly in protozoa to
amplify transmembrane potentials and Ca2+ fluxes, a role which has been well-documented in Stentor
and Paramecium and which would clearly be of importance in triggering fast axopodial contraction.
In Actinocoryne contractilis, action potentials propagate from head to stalk base at 3 m s-l
(Febvre-Chevalier et al., 1989, p. 241), which implies spread over the maximum axopodiallength (500
Jlm) in 2 x 10- 1 ms, much less than the observed latency (7
±3
ms;
Febvre-Chevalier and Febvre,
1992, p. 587).
Wood (1989, p. 368) has suggested that the propagation of electrical impulses and cell contraction
observed in certain ciliates (e.g., Stentor, Vorticella) can be accounted for by the spread of action
potentials along the membranes enclosing intracellular Ca2+ stores. Wood notes that these vesicles are
commonly in apposition with the plasma membrane, and that their small diameters and the low ionic
strength of their contents imply a length constant smaller than that for the surrounding cytoplasm.
(Wood's assumption that the ionic strength of the vesicle contents is low seems doubtful to me.)
8
Hence, the vesicle stores would not be isopotential, and action potentials could serve as signals to bring
about the sequential release of Ca2+. It is conceivable that such a scheme might operate in the larger
Heliozoa to couple the passage of an action potential along the axopodial membrane to the discharge of
intraaxopodial Ca2+ required to power the contractile organelle.
(See, e.g., Matsuoka et al., 1984,
Figs. 3-4 for electron micrographs showing apposition of Ca2+ vesicle and plasma membranes in
Echinosphaerium akamae.)
9
3. ON TIlE ELECTROSTATIC CHARGE AT THE CELL SURFACE
The aim of the present section is to show that electrostatic forces due to surface electrostatic
charge can have only a negligible effect on the motion of a contracting axopodium. This conclusion is
supported by a simple calculation.
Owing to lengthy extracts from publications, I enclose my own
parenthetical comments in brackets.
De Loof (1986, p.254) says that "From biochemical data, it is well established that the cell
surface carries a net negative charge (Dolowy, 1984). [Matsuoka et al. (1984a, 1986) and Shigenaka et
al. (1989a) confirm this generalization for Echinosphaerium akamae; see Vommaro et al. (1993) for a
recent determination of negative surface charge in a f1agellat.e.]
Fixed surface charges result in the
formation of a diffuse electrical double layer in which the charge at. t.he surface is balanced by charges
of opposite sign in the medium and by the dipole moment of water molecules immediately adjacent to
the surface."
Probstein (1989) discusses diffuse electrical double layers at several points
III
his text.
The
following are selected excerpts.
"Electroneutrality or the absence of charge separation holds closely in aqueous electrochemical
solutions, although not necessarily organic solutions, everywhere except in thin regions near charged
boundaries. These regions are termed double layers or Dcbye sheaths and have thicknesses on the order
of 1-10 nm. The double layer is important when we consider very small charged particle interactions
and
charged
surface
phenomena,
but
is generally
unimportant
with
respect
to
bulk
flow
characteristics." (Probstein, 1989, p. 43)
Probstein (1989, p. 187, eqn. 6.4.5) gives the Debye shielding dist.ance or Dchye lcngth for an
infinite plane double layer as
1
)2
2F2z2c •
A - ( £RT
D -
(1)
where AD is the Debye length (m), R is the gas constant (8.314 .J K-Imol- l ), T is the absolute
temperature (K), F is the Faraday constant (9.648 x 10 4 C mor l ), z is the valence of the ion in
question, c is its concentration (mol m- 3 ), and
£
is the dielectri<; constant (permittivity) of water. Note
10
that (=
(O(r'
where
(0
is the permittivity of a vacuum ( = 8.851
relative permittivity of water. At 25° C (
= 298 K), = 78.3.
(r
X
10- 12 C 2
J'r1 m 2),
and
(r
is the
It seems reasonable to use this value in
d(
the sequel because dT is small (Lide. 1993, 6-148) and 298 K is near t.he t.emperatures at which cert.ain
heliozoa have been cultured. In particular, Ando and Shigenaka (1989) grew Echinosphaerium akamae
at 20° C (= 293 K), and Davidson (1975) noted generation t.imes for Heterophrys marina and
Ciliophrys marina at
~
21° C (
= 297
K). For a symmetrical electrolyte in aqueous solution at 298 K,
the Debye length is given by
nm
(Probstein, 1989, p. 187, eqn. 6.4.6).
(2)
Probstein observes that "for a univalent electrolyte the Debye
length is thus about 1 nm for a concentration of 10 2 mol m- 3 and 10 nm for 1 mol m- 3 ."
Now I compute approximately t.he Debye length of a double layer due to Ca 2 + on the surface of
H. marina and C. marma.
mM
= 10
mol m- 3 .
Taking z
Davidson (1975) states that for the seawater used, c = [Ca 2 +J = 10
=2
(divalent electrolyte) and substituting the foregoing value of c in
equation (2) above yields '\0 = 1.5 nm. Since'\o is much less t.han the diameter (l00 nm) and length
of an axopodium (50 pm), and the distance between evaginations of the contracted axopodial
membrane (Davidson, 1975, Figs. 86-91), it must be that electrost,atic forces offer negligible resistance
to contraction.
SO}vIE FURTHER CONSlDERA TJONS
1.
Is equation (1) for the Debye length of an infinite plane double layer the same for interface
shapes appropriate for heliozoa?
Equation (1) also holds for spheres (Russel, 1987, p. 23) and is
accurate within a factor of 2 for cylinders (Katchalsky et aI., 1966, pp. 306-307). Thus the conclusion
drawn in the present section is unchanged for these more realistic shapes.
2.
c
How does the Debye length calculation turn out for Na +? For Na +, appropriate values are
= 460 mM = 460 mol m- 3 and z = 1.
3.
ions?
Inserting these data in equation (2) affords '\0
= 0.45 nm.
How can one calculate the Debye length for a double layer involving a realistic mixture of
For seawater, one may estimate a lower bound on th~ Debye length by taking z = 2 as the
11
maximum valence and letting c = 533 mol m- 3 be the total concentration of the principal cations. The
latter figure is arrived at by considering K+ (10 mM), Na+ ('160 mM), Ca 2 + (10 mM), and Mg 2 + (53
mM). We assume that the number of cations and anions in solution is equal. Then equation (2) yields
AD
= 0.21 nm as a
4.
lower bound.
How does contact with anion-exchange resrn beads (Matsuoka et al., 1986;
Ando and
Shigenaka, 1989) provoke fast contraction? One may conjecture that anion-exchange resin beads make
the potential at the outer face of the axopodial membrrlne more negative without altering the potential
at the cytoplasmic face, thereby decreasing the electric field across the membrane and opening voltagegated Ca 2+ channels (see Hille, 1992, pp. 460-461; Jack et al., 1983, pp. 3'18-3'19).
5. Probstein (1989, pp. 187-188) includes useful additional commentary on the Debye length:
"From equation [1] or [2] it can be seen that AI) decreases inversely as the square root of the
concentration. Physically this is a result of the fact that. t.here are more counterions per unit of depth.
The Debye length also decreases with increasing valency because fewer ions are required to equilibrate
the surface charge. More importantly, AD increases as the square root of RT. That is, without thermal
agitation the double layer would collapse to an indefinitely thin layer."
"From the above considerations we can now define
solutions.
what is meant by electrically neutral
If the dimensions of the system L are much larger than AI), then whenever local charge
concentrations arise or external potentials are introduced into t.he solution they are shielded out in a
distance short compared with L, leaving the bulk of the solution free of large electric potentials or
fields. Based on a Debye length of 1-10 nm, the assumption of electrical neutrality is generally justified
for the problems so far considered.
However, . . . in the case of very small charged microscopic
capillaries, such as are characteristic of membranes and finely porous media, the double layer is central
to the calculation of the solut.e and ion fluxes."
On p. 188, Probstein derives the Debye-Hiickel approximation and obtains therefrom an
alternative expression for the Debye length:
¢J = ¢J w exp ( -
~~
),
(3)
where
ePw
is the wall potential.
"The Debye length is thus seen to be the ~ decay distance for the
potential and electric field at low potentials.
Close to the charged surface where the potential is
relatively high and Debye-H iickel approximation inapplicable, the potential decreases faster than the
exponential fall-off indicated."
"The potential
ePw
can be related to the charge density at the surface by equating the surface
charge with the net space charge in the diffuse part of the double layer."
13
4. CALCULATION OF THE MEAN FIRST PASSAGE TI.\IES FOR Ca~H AND IP
3
DIFFUSION FROJ\I THE AXOPODIAL MEMBRANE TO THE AXONEME
The present section estimates the mean first passage times for diffusion of Ca 2+ and IP3 (inositol
1,4,5-trisphosphate).
These chemical species are typical second messengers in cells, and may be
responsible for triggering fragmentation of the axonemal microtubules by binding to a
microtubule-
severing factor (perhaps akin to the microtubule-cleaving factor isolated by Vale (1991) from Xenopus
oocytes or the F-actin-severing protein, gelsolin) postulated here.
I make the following simplifying and idealizing assumptions:
1. There are no interactions among messenger molecules or ions.
2. There is no axial gradient in messenger concentration. That is, the messenger concentration is
space-clamped, so
3.
~~ = 0, where z is the axial coordinate.
Diffusion occurs in a two-dimensional annulus such that the inner aspect of the axopodial
membrane is the source of Ca 2+, whereas the axoneme, viewed as a disk, is a sink that binds Ca2+
irreversibly. I assume that there are no other sources or sinks of Ca2+. (See the annexed diagram.)
1----
4.
A
----I
Geometrical hindrance due to the filamentous structure of cytoplasm is negligible.
This
assumption seems reasonable because mesh spacings in the cytoskeleton are ...... 100 nm (Gershon et al.,
1985, p. 5033), but calculations show that diffusion of molecules of radius 0.3 nm would begin to be
appreciably impeded at spacings from 10-15 nm (Blum et al., 1989, p. 1000).
14
If a denotes the diameter of the axoneme and A represents the diameter of the axopodium, then
the mean first passage time
T
for diffusion to the axonemal disk is
T
where D is the diffusion coefficient
A2 (A)
a ,
(1)
= 8D In
For Ca2+,
(Hardt, 1981).
typical published values of D in
cytoplasm are 6 x 10-10 m 2 s-1 (Hodgkin and Keynes, 1957), 4 x 10- 10 m 2 s-1 (Kargacin and Fay, 1991),
and 3 x 10-10 m 2 s-1
DCa = 3.8 X 10- 11 m 2 s-1
(Roberts, 1993, p. 74).
However, a recent determination reports
(Albritton et al., 1992).
Arbitrarily, I shall take 5 x 10-10 m 2 s-1 as a
middling compromise "classical" value of DCa' and I shall call 3.8 x 10- 11 m 2 s-1 the "revised" value.
For IP 3, Albritton et al. (1992) found D1P3 = 2.83 X 10- 10 m 2 s-l. Measurements of a and A are seldom
reported in the literature.
Echinosphaerium nucleolilum
Hoffman et al. (1983, p. 375) found that isolated axonemes measured
which they maintain is consistent with electron micrographs by Tilney.
~
1.5 pm in diameter,
According to Tilney and
Porter (1965), the axopodia are 5-10 pm in diameter at levels remote from the tip. Using the classical
value of DCa' and putting a = 1.5 pm and A
T
=10 pm in equation (1), we obtain
= 5 X 10 1 ms.
Similarly, for a = 1.5 pm and A = 5 pm,
T
Hence
= 7 ms.
15
7 ms ~
T
~ 5 X 10 1 ms (classical).
Repeating the foregoing calculations with DCa = 3.8 X 10- 11 m 2 s-l yields
9 X 101 ms ~
T
~ 6 x 10 2 ms (revised).
For IP3' equation (1) affords
Echinosphaerium aka mae
From an electron micrograph of a cross-section of an axopodium in Matsuoka and Shigenaka
(1984b, Fig. 3), one can readily estimate the diameters of the axoneme and axopodium.
magnification factor is x 33,000, 1 cm
<-+
Since the
0.30 pm on the published micrograph. The diameters of the
images of the axoneme and axopodium are 3 cm and 12 cm, respectively, so a= 0.9 pm and A = 3.6
pm. Thus in the classical case, we obtain for the Ca2+ mean first passage time
(3.6 X 10<6 m)2 In (12)
- 8(5 x 10- 10 m2 s<l)
3
T _
T
=4.5 ms.
We may repeat the calculation for another electron micrograph of a cross-section probably taken
at a different level (Shigenaka et al., 1989b, Fig. 2). Here the a:<oneme image diameter is 6 em, the
axopodium image diameter is 14 cm, and the magnification is x 88,000.
a
= 0.7 pm, A = 1.6 pm, and
Hence 1 cm
<-+
O.lIpm,
16
T
= 0.54 ms.
It follows that
~ T ~
0.54 ms
4.5 ms (classical).
Similar calculations using the revised value of DCa yield the range of times
7 ms ~
T
~ 6 X 10 1 ms (revised).
For IP3' we obtain the time interval
Ciliophrys marina
Davidson (1975) determined that the axoneme is a triad of microtubules in which the links are 15
nm long. Assuming that the microtubules are 25 nm in diameter, it seems reasonable to approximate
the axoneme as a disk of diameter 50 nm. Therefore, considering Ca2+ diffusion and taking a
and A = 0.1 J.lm,
7
(1 X 10- m)2 In (100)
- 8(5 x 10- 10 m 2 s·l)
50
T _
T
= 9 J.lS (classical).
The same calculation with DCa = 3.8 X 10- 11 m 2 s-1 gives
T
= 1.2 x 102
J.ls (revised).
= 50 nm
17
For IP 3,
TIP
3
= 16 J.ls.
GaNGL USIONS
To assess the mean first passage times computed above, we consider the observed latencies for
fast axopodial contraction. Davidson (1975, p. 42) remarked that in high-speed films of Heterophrys
marina "there is apparently a latent period of 10-20 ms before contraction actually begins, but this is
difficult to observe in many cases because the time period between frames is greater than or equal to
the latent period". Febvre-Chevalier et al. (1989, p. 239) reported more reliably that the latency for
axopodial and stalk contraction in Adinocoryne contractilis following external mechanical and
electrical stimulation and intracellular current injection is 7 ± 3 ms (n = 80). Since the axopodia of A.
contractilis are from 50-500 J.lm long (Febvre-Chevalier and Febvre, 1992, p. 587), a range which
includes the axopodia treated here, we are probably justified in using the latency of this heliozoan as a
benchmark. It should be pointed out, however, that no data appear to be available on the diameters of
A. contractilis axopodia.
In the large axopodia of heliozoa of the genus Echinosphaerium, Ca2+ diffusion times based on
the classical value of the diffusion coefficient range between 0.54 ms and 50 ms, whereas those founded
on the revised value lie between 7 ms and 6 x 10 2 ms. That the larger estimates exceed the measured
latencies noted above would suggest that Ca2+ diffusion might be rate-limiting in large axopodia.
However, internal Ca2+ stores (Matsuoka and Shigenaka, 1984, 1985; Matsuoka et al., 1985), possibly
discharged by the action potential, may appreciably shorten diffusion distances.
For IP3' time
18
discrepancies are harder to reconcile than for Ca2+ because IP 3 derives from the hydrolysis of a plasma
membrane-bound precursor, phosphatidylinositol 4,5-bisphosphate (PIP 2), and is not released from
cytoplasmic stores.
In Echinosphaerium, 0.9 ms ~ TI?3 ~ 8 X 10 1 ms.
Observe that the even smaller
diffusion coefficient of ATP in cytoplasm (D AT ? = 1.50 x 10- 10 m 2 s-1;
Kushmerick and Podolsky,
1969) implies that the use of cyclic nucleotides (e.g., cAMP) as second messengers probably would not
expedite signaling in large axopodia.
In the smallest axopodia, exemplified by those of Ci/iophrys marina, the very brief mean first
passage times show that second messenger diffusion occurs amply fast. This conclusion comports with
observations of the time course of ciliary beat reversal (cilium diameter ~ 0.2 J.lm): Ca2+ influx from
the medium is both necessary and sufficient to induce beat reversal in Paramecium (summarized by
Harold, 1986, pp. 416-419, 501). There is no ultrastructural evidence for intraaxopodial Ca2+ stores
in C. marina or H. marina; indeed, apart from the putative contractile element, the axopodia of the
latter heliozoan lack mitochondria and other recognizable organelles (Davidson, 1975).
Although the role of Ca2+ as the source of chemical potential for the spasmoneme of vorticellids
is well-established, few studies have been made of second messenger signaling in the contractile
appendages of protozoa. Evans et al. (1988) used drug treatments to implicate IP 3 , cyclic nucleotides,
and Ca2+ in the control of tentacle contraction in the suctorian He/iophrya erhardi.
Suctorian
tentacles are comparable in size to axopodia, but their latency (3.7 ± 0.1 s) and contraction time
(18.2 ± 0.2 s) are three orders of magnitude greater (Evans et al., 1988, p. 384; mean ± SE, n
= 161).
Moreover, the tentacular membrane is inexcitable and contraction results from the relative sliding of
intact microtubules, possibly driven by actomyosin.
It is of interest that in connection with his studies on muscle, A.V. Hill (1948) derived an
analytical solution for the problem of inward diffusion of a substance liberated from the wall of a
cylinder of radius a.
To obtain this result, Hill first solved the initial-boundary value problem for
diffusion from an annular source between coaxial cylinders of radii a and b (a>b). The concentration
c(r,t) is governed by
19
c(r,t) = cO' b ~ r ~ a
c(r,O) = 0, 0 ~ r ~ b
dc
- 0, r -- a
dr c(O,t) <
00
'v't > o.
The solution is
where ai > 0, J 1(ai)
== 0, and J k( ) is the kth order Bessel function of the first kind.
Hill then let the thickness of the annular source shrink to zero in the limit as b ... a in equation
(2). Accordingly, conservation of mass implies
(3)
By the mean-value theorem with c E (aibja , ai)'
(4)
But J 1(ai)
== O. Thus,
(5)
From Gradshteyn and Ryzhik (1965, Section 8.472, Formula 1, p. 967), Bessel functions of order v
obey the recurrence relation
20
z Jz Zv (z) + II Zv (z) = z Zv-l (z).
Putting Zv ( )
= Jv (
) with v
= 1, and evaluating at z = c, we may write
(6)
Since Otib/a
< c < Oti '
C
-+
Oti as b
-+
a, and
Therefore, from (6),
and by substitution in (5),
(7)
as b
-+
a.
Insertion of expressions (3) and (7) into equation (2) gives
(8)
Hill (1948, Fig. 2, p. 450) plots
c~r,t) vs.
00
D; for values pertinent to diffusion in muscle.
a
21
5. [Ca2 +J IN RELATION TO CONTRACTILE, OTHER MOTILITY PROCESSES IN PROTOZOA,
INCLUDING THE HELIOZOA
Most of the calcium concentrations reported below pertain to the external medium and are denoted
by rCa 2+ Je'
Intracellular free calcium concentrations are denoted by [Ca-~+ Jj .
Sources of data are
given parenthetically after the scientific names of the organisms.
Heliozoa
A ctinocoryne contraclilis (Febvre-Chevalicr and Febvre, 19S9)
Total inhibition of contraction.
64% slow contraction (6-15 s).
13% fast contraction.
23% no contraction.
27% slow contraction.
67% fast contraction.
6% no contraction.
Only fast contraction.
The authors consider (p. 587) that the threshold for contraction is [Ca 2+Je = 10- 7 M.
Heterophrys marina and Ciliophrys marina (Davidson, 1975, p. 49)
22
No contraction; axopodia "become very brittle and are
easily broken".
Transformation to fast-swimming flagellate.
Axopodia contract to ::::: one-half their extended length.
Normal contraction.
Echinosphaerium akamae
(Matsuoka and Shigenaka. 1984, pp. 426-427;
MatslJoka et al., 1985,
pp. 69-70)
"Hardly any axopodial contraction."
(I have not
found a statement of the threshold [Ca 2+l e .)
Axopodia shorten slowlv in the presence of the divalent cation ionophore A23187 (1 J1.g mr 1)
and Ca 2+. (Time to shorten to 49% of extended length = 10 min.) A much smaller extent of
shortening was seen at lower [Ca 2+l e.
Echinosphaerium eichhorni (Schliwa, 1976, 1977)
Slow contraction was observed for [Ca2 +l e as low as 10- 5 M in the presence of 0.5 J1.g mr 1
A23187. The speed of contraction increases with [Ca 2 +l e : half-maximal contraction was observed
after 35 min in 1 x 10- 4 M Ca 2 + and 12 min in 5 x 10- 4 1\1 Ca2 +. In tile most concentrated
solution tried ([Ca 2+]e = 1 x 10- 3 M), axopodia shorten to 20% of their extended length within
4 min of exposure.
Is A23187 a carrier or a channel? See Harold (1986, pr. 360-362) for a comparison of
23
transport rates for these classes of ionophores. The roint is that the Ca 2+ flux due to A23187
may be much less than that for natural memhrane channels. See also Schliwa's papers for refs. on
A23187.
Vorticellid Spasmonemes
Vorticella (Amos, 1971; Weis-Fogh, 1975, pp. 8"1-85)
Glycerinated stalks remain extended in 10- 8 f"Y Ca 2+. The threshold concentration for
contraction is 5 x 10- 7 M. Stalks remain contrilctcd ill. [Ca 2+]e = 10- 6 I\L The external work
done in contraction in vivo can be accounted for hy a change of [Ca 2 +]i from 10- 8 to 10- 6 M.
See Amos on the rapid cycling of glycerinated spasornoneme on a time scale comparable
to the process in vivo.
Zoothamnium genicu/atllm (Weis-Fogh and Amos, 19(2)
Glycerinated colonies were fully extended in [Ca 2 +]e = 10- 8 1\1, and contracted with stalk
flexure in 10- 5 M Ca 2+
"a,<;
in life but taking several seconds and without suhsequent relaxation"
(p. 303). The isolated spasmoneme shows arpropriate hehavior at the same values of [Ca 2+]e'
Ciliate Cell Body Contraction
B/epharisma japonicllm (Matsuoka et aI., 1991)
Results for Triton-extracted models:
The threshold concentration for model elongation is [Ca 2+]e = 3 x 10- 7 M. The authors
estimate that [Ca2+]. = 10- 7 M. Drug treatment implies that elongation is ATP-dependent,
1
•
24
whereas contraction is ATP-independcnt. Elollgated models contract when transferred from
[Ca2 +]e
= 10- 10 M (with ATP) to [Ca 2+]e = 10- 0 1\1 (without ATP).
Stentor (Huang and Mazia, 1975, pp. 393-391)
Results for EGT A-treated cells:
Cells remain extended in 10- 8 1\1 Ca 2 +.
Cells contract in 10- 6 M Ca 2+.
Spirostomum (Hawkes and Holberton, 1975; Ishida and Shigenaka, 1988)
Hawkes and Holbert,on found that rnyonemcs do not. evince velocit.y-Ioad relations
appropriate for muscle, but behave instead like an elastomer. (That is, the force developed
is a constant independent of load.) These authors reckoned (p.601) that enough energy for
the work of contraction could be obtained by varying [Ca2 +]i from 10- 7 M to 10- 5 M, thereby
binding 1.93 x 10- 14 mol Ca2 +.
Ishida and Shigenaka's more recent researches on Triton-
extracted cells are consistent with the foregoing range of concentrations.
Ciliary/Flagellar Beat Reversal
Blepharisma japonicum (Matsuoka et aI., 1991)
Results for Triton-extracted models:
Models swim forward for [Ca 2+]e
< 10- 6 ~'l. but swim backward or tumble for [Ca 2+]e > 10-6 M.
Chlamydomonas reinhardtii (Melkonian and Rohenek, 1984, p. 244)
Results for isolated flagellar apparatus:
2S
Same as for B/epharis71la, viz., forward swimming for [Ca 2 +]e < 10- 6 ~I, backward swimming for
[Ca 2 +]e> 10- 6 M. [Melkonian et al. (1002, p. 187) assert that "as in other eukaryotic cells, the
the cytosolic free Ca2+ concentration in the nagellate green algae is presumably also kept at a low
and constant level of about 5 x 10- 8 l'vl Ca~H".]
Paramecium (literature epitome in Gregson et aI., ] 99:lb, p. (98)
Results for det.ergent-extracted cell models:
Forward swimming for [Ca 2 +]e:::; 10- 7 M, backward swimming for [Ca 2 +]e ~ 10- 6 M.
Contractile Dinophycean Flagella
Peridinium inconspicullm (Hohfeld et aI., ] 088, p. 21)
Results on the isolated transverse nagellum:
"Flagella supercoil within a few seconds t.o a highly cont.racted st.at.e" in 5 x 10- 5 M Ca 2+.
Ceratium tripos (Maruyama, 1985; H6hfeld et. aI., 1988)
Results on reactivated longitudinal nagella:
Contraction requires [Ca 2+]e ~ ] x 10- 5 l\L
Oxyrrhis marina (Godart and Huitorel. ]902)
Results on longitudinal flagella still attached to permeabilized cells:
[Ca 2 +]e = 1 x 10- 7 M
50% contraction
[Ca 2+]e = 3 x 10- 7 M
50% autotomy
[9a2+]e = 9 x 10- 7 M
00% autot.omy
26
Prymnesioph vcean Haptonemata
Ghrysochromu/ina acantha (l\lelkonian et aI., 1992, p. 207; Gregson et aI., 1993b)
Gregson et a!. (pp. 686, 695) show that haptonemal coiling is Ca 2+-dependent.
They establish that "an external concentration of hetween 10- 7 1'v1 and 10- 6 ]',,1 Ca2+ is the
threshold below which the frequency of coiling on cell death is reduced" in cells treated with
Ca2+ jEGTA buffers before fixation. Melkoniall et al. state that it is not known whether centrin
causes coiling of the haptonema.
GONGL USJONS
The axopodia of A. contracli/is and E. n!.:amne can undergo fast contraction in solutions spanning
the same range of [Ca 2 +]e values (typically ...... 10- 7_10- 5 :\1) t.hat. permit.s Ca 2 +-dependent motility in
most other intact and permeabilized protozoa invest.igated to date.
By contrast, the axopodia of H.
marina and G. marina contract normally only at much higher external Ca 2+ concentrations (10- 2-10- 1
M). It may be that the absence of vesicular Ca 2 + stores in these very thin motile appendages means
that an especially steep Ca 2 + concentration gradient must he set up to supply enough chemical
potential energy to drive contraction.
In this connection, it is of interest that in Paramecium, in
which an influx of Ca 2 + across the ciliary m~mbrane serves merely as a signal for heat reversal, not as
a source of chemical potential energy for beating, the threshold rCa2+]e for beat reversal is only 10- 6
M.
27
6. ESTIMATION OF THE FORCE EXERTED BY THE CONTRACTILE ORGANELLE
According to Bereiter-Hahn (1987, Table 2, p.10), Portzehl (1951) found that the stress exerted by
an actomyosin thread prepared from skeletal muscle is 2.5 x 10 4 N m- 2.
Interestingly, the tensile
stresses exerted by vorticellid spasmonemes in vivo and in vitro are of the same order of magnitude
(Weis-Fogh, 1975, Table I, p.94):
In vivo:
10 4 N m- 2
Vorticella
Garchesium
4 x 10 4 - 8 x 10 4 N m- 2
In vitro:
Zoothamnium
Now suppose that the contractile organelle of the Heliozoa can develop stresses of the same order of
magnitude as the stresses produced by actomyosin and spasmonemes. For simplicity, I shall take this
stress to be 5 x 10 4 N m- 2.
Suppose further that in Echinosphaerium akamae the X-body is the
contractile organelle. Then one JVay approximate the cross-sectional area of the image of the X-body
In
a certain particularly clear electron micrograph (Shigenaka et al., 1989b, Fig. 2) as the area
enclosed by a rectangle measuring 2 cm x 4 cm. Taking 1 cm
1 x 10- 13 m 2.
+-+
1.1 x 10- 7m, it follows that 8 cm 2
The force exerted is thus (5 x 10 4 N m- 2)(1 x 10- 13 m 2)
=5 x 10-9
N
+-+
= 5 x 103 pN.
Note that this value is very close to the force of fast axopodial contraction, (4.1 ± 0.9) x 103 pN,
measured by Suzaki et al. (1992a) in E.
akamae. It is comparable to the estimated drag (8.6 x 10 3
pN) on the cell body of a contracting Vorticella (Amos, 1975b), but greater than the forces exerted by
swimming sperm of various phyla (5-350 pN; Green, 1988).
GONGL USIONS
The foregoing argument demonstrates that the force ofaxopodial contraction could be produced
chemomechanically by an actomyosin- or spasmin-like Ca2-f:-binding protein.
Although there is
28
ultrastructural and experimental evidence for proteinaceous intraaxopodial contractile organelles (see,
e.g., Davidson, 1975;
Shigenaka et al., 1982;
Matsuoka and Shigenaka, 1985),
their chemical
composition has not been investigated. Centrin or a centrin-like protein (possibly a spasmin) has been
detected by immunofluorescence in the base of the microtubule-containing contractile stalk and near
the microtubule organizing centers for the axopodial axonemes of Pseudopedine//a elastica (Koutoulis et
al., 1988; Melkonian et aI., 1992, p. 207). It is not known whether centrin or a homologue of centrin
also occurs within the axopodia of this heliozoan.
29
7. AN OBSERVATION ON THE TIME COURSE OF FAST AXOPODIAL/STALK CONTRACTION
The exponential shortening curves reported by Febvre-Chevalier (1981) and Febvre-Chevalier and
Febvre (1992) for the stalk and axopodia of Actinocoryne contracti/is imply that
t ~t
= -k, L(O) = Lo' k>O
(1)
or
f
= -k, f(O) = 0, k>O.
(2)
That is, the strain rate, f, is constant. Note that this result is perfectly consistent with the definition
of natural or true strain, fT' appropriate for large deformations (Wainwright et al., 1976, eqn. 2.4, p.9):
(3)
I show this by writing
fT
= In (t~ = - kt
(4)
and differentiating throughout with respect to t:
(5)
whence
(6)
which is identical with equation (1) above. 0
Now I determine the value of k and the functional form of L = L(t).
The solution of equation (1) is
L(t)
= Lo exp( -
kt)
(7)
whence
k
o)
=r1ln (LL.
(8)
From Febvre-Chevalier and Febvre (1992, p. 588) the mean half-time for stalk contraction in
Actinocoryne contracti/is is 4 ms (n = 20). Hence equation (8) yields
k
=! In 2 =0.17 IDs-I.
(9)
30
The mean initial stalk length is 185 J.Lm (n = 20). Thus, the time course of shortening, equation (7),
becomes
L(t) = 185 exp( - 0.17t).
(10)
Now I estimate L = L(t) for the individual time course of stalk contraction shown in Fig. 3, p. 588.
The half-time is 2.2 ms. Hence
k
= 2~2 In 2 =0.31 ms- 1
The curve in Fig. 3 begins somewhat above 150 J.Lm.
(11)
Let x denote the length in excess of 150 J.Lm.
Then taking the correspondence 1.1 em...... 50 J.Lm, we have
14
0.2 -- 50
x => x -- 9. 1 ,.."m .
(12)
Since this is but an estimate, I round to 10 J.Lm, whence
Lo = 160 J.Lm.
(13)
L(t) = 160 exp( - 0.3lt).
(14)
Therefore
I checked this expression for t
= 6.5 ms, which gave L == 21 J.Lm, which comports with Fig. 3.
GONGL USIONS
The constancy of strain rate during fast axopodial contraction in A. contraetilis has not been
pointed out previously.
Time courses of contraction are unknown for other heliozoa, but have been
measured for ciliates, including peritrichs (Vorticella: Jones et al., 1970bj Katoh and Naitoh, 1992),
heterotrichs (Spirostomum: Hawkes and Holberton, 1974; Stentor: Newman, 1972), and a suctorian
(Discophrya:
Hackney and Butler, 1981). The authors cited did not fit curves analytically to these
data, and as the published graphs are variously irregular and in some cases both concave and convex, it
is doubtful wheter exponentials would be satisfactory.
The observed exponential time course of fast axopodial contraction in A. contraetilis lacks an
explanation. An increase in drag as the shortening axopodium moves toward the cell body, length- or
speed-dependent chemomechanical properties of the contra.ctile organelle, and an increase in
31
intraaxopodial viscosity as contraction proceeds might singly or jointly account for the exponential
time course, but it has not been possible to quantify these effects. In Spirostomum, the force of cell
body contraction is independent of speed in solutions of various viscosities (Hawkes and Holberton,
1975, p. 598;
Amos et al., 1976, p. 293).
It should be kept in mind that further experiments are
needed to determine if an exponential time course holds generally for fast axopodial contraction in the
Heliozoa.
32
8. REYNOLDS NUMBERS FOR FAST AXOPODJAL CONTRACTION IN THE HELIOZOA
I compute Re = ~l, where U is the speed of contraction (m s-l) and v is the kinematic viscosity of
water (v ~ 10- 6 m 2s- 1). I assume that the charactcristic Icngth I (m) is the distance through which the
axopodium contracts.
Note that owing to crumpling (huckling) of the axopodial membrane during
contraction, there may be more than one vaillc of Re t.o consider.
Echinosphaerium akamae:
From Suzaki et al. (1992a, p. 432),
_ (2.6x 10- 3 m s-I)(J.72x lO- tJ rn) _
0- 1
R.e -6 ? -1
- 4 xI.
10 m- s
A ctinocoryne contractilis:
Febvre-Chevalier (1981, p. 340) says that whcn the hase of the head or stalk is stimulated, the
axopodia contract to two-thirds of their initial extended length in 2 ms.
However, she does not say
what the extended length is. Febvre-Chevalicr and Fcbvre (1992, Fig. 1, (d )-(f)) indicate a contraction
time == 2.4 ms, and note (p. 587) that the axopodia are 50-500 pm long, Thus, roughly, the speeds of
contraction range between
~ (5 x 1O-~3 m) = 2 x 10- 2 m s-1 and 2 x 10 m s-I,
2 x 10
s
between
and
(2 x 10- 1 m s-I)(~ 5 x 10- 4 m)
10-6 m2s- 1
_
=
I.
Accordingly, Re must lie
33
That is,
0.7 < Re < 7.
Heterophrys marina and Ciliophrys marina:
From Davidson (1975), the typical length of an axopodium is 50llm, and an axopodium contracts
by ::::: 90% ( ::::: 45 /lm) in 20 ms. Hence,
u
= 4.5 x 1O-~) m = 2.2 x 10- 3
Tn
s-l.
2.0 x 10-- s
Thus,
Actinophrys sol:
A typical contraction entails a displacement of ::::: 100 11m in < 1 s (Ockleford and Tucker, 1973,
pp. 375-376). Hence U> 100 /lm s-l, and
R e>
(10-4 m s-I)(10-4 m) _ 10- 2
-G ') -1
.
10 m- s
The following chart summarizes values of the Reynolds number for fast axopodial contraction in
some heliozoa.
34
REYNOLDS NUMBERS FOR FAST AXOPODJAL CONTRACTION IN SOME HELIOZOA
HELIOZOAN
U
(m s-l)
Echinosphaerium
akamae
2.6 x 10- 3
Actinocoryne
contractilis
'"" (2 x 10- 2_
2 x 10- 1)
Heterophrys marina
and
Ciliophrys marina
Actinophrys sol
I
(m)
Re
1.72 x 10- 4
0.4
'" (3 x 10- 5_
3 x 10-4 )
0.7-7
REFERENCE
Suzaki et al.
(1992, p. 432)
Febvre-Chevalier
(1981); FebvreChevalier and Febvre
(1992)
'"" 2.2 x 10- 3
4.5 x 10- 5
0.1
Davidson (1975)
> 1 x 10-4
1 x 10- 4
> 10- 2
Ockleford and
Tucker (1973)
By way of comparison, Childress (1977, p. 4) gives for Paramecium Re'"" 10- 2 for a beating cilium,
and Re '"" 0.2 for swimming of the whole organism, where the Re values are based on the length of the
cilium and cell body, respectively.
Childress (1981, Table 1.1, p. 4) classifies Paramecium as a
"Stokesian" swimmer.
e
Happel and Brenner (1983, p. 3) seem to suggest that the creeping motion (Stokes) equations can
safely be applied for Re < 5.
Some published compilations of Re values for various organisms follow.
35
From Vogel (1981, p. 67):
Re
A luge wh:z.le swimming
A
run~
swimming
A duck flying
A luge
:l.t
~t
~
10 m sOl
s~e
the
300.000,000.
speed
30,000,000
20 m s"t
dr~gonfly going
A copepod in
~t
300,000
7 m SOl
30,000
pulse of 20 em SOl
300
Flight of the sm:z.llest flying insectS
30
An
A.
invertebr~te 12.CY~,
0.3 mm long,
moving at 1 mm sOl
0.3
se~
urchin sperm advancing
the species at 0.2 mm sOl
0.03
From Sleigh and Blake (1977, p. 249):
TABLE
1. Dimensions of features of the body and cilia for representative examples of the range of ciliated organisms.
Body
length
Name
Group
(L)~
Swimmingt
velocity
body
~S-I lengths
S-I
Uronema
Tetrahymena
Didinium
Paramecium
Veliger (Aplysia)
Blepharisma
Opalina
Spirostomum
Convoluta
Pleurobrachia
Protozoa
Protozoa
Protozoa
Protozoa
Mollusc larva
Protozoa
Protozoa
Protozoa
Flatworm
Ctenophore
2S
70
130
210
11 SO
480
12S0
1000
ISO
?
350
600
50
1000
600
400
1000
2000
15000
7SOOO
46
7
10
5
?
2
0'12
1
0'3
5
Rb
Number
of body
cilia
(n)
Cilia
length
(l)1Jl1I
10- 2
10- 2
10- 1
10- 1
200·
S
Soo·
17S0
SOOO
7
10- 1
10- 2
1
1
10 3
200
7000·
lOS
lOS.
7 x 106
200(108)
12·S
12
56t
7·5
IS
12
7·S
SOOt
-
Rc
nl
Metachronism
10- 3
10- 2
10- 2
10-%
10- 1
10-%
10- 2
10- 2
10-%
10
10 3
?
10 3
?
2·2 x 10· Dexioplectic
6 x 10· Dexioplectic
1-1 X 10· Laeoplectic
S x 10·
?
1·5 X 10 6 Symplectic
1·2 x 106
?
5 x 10 7 Laeoplectic
lOS (l010) Antiplectic
3'S
X
·Excludes compound cilia; tindicates compound cilia; tfigures for swimming velocity are "maximal" values, particularly for Pleura-
e
brachia.
R and R c are Reynolds numbers based on cell body length 'and cilium length, respectively.
b
36
9. A CRUDE ESTIMATE, ACCORDING TO SLENDER BODY THEORY, OF THE DRAG
AND POWER DISSIPATION IN FAST AXOPODIAL CONTRACTION
I use an asymptotic formula derived by de Mestre and Russel (1975) and corrected by Brennen and
Winet (1977, pp. 351-351) for the tangential force coefficient, Cs ' of a slender right circular cylinder
translating along its axis normal to an infinite plane wall:
2a
II
I
I
h
The formula in question is appropriate for a
«: hand
C _
S -
L
2~ ~1:
2irf.l
In (Lao)-1.75
(Brennen and Winet, 1977, Table 2, p. 350), where f.l
(1)
= 1 x 10- 3 kg m- 1 s-l is the dynamic viscosity of
water. The drag per unit length is given by
(2)
where U is the speed of contraction. The total drag on an axopodium is thus
F tot = Cs UL o ,
(3)
37
where Lo is the initial (uncontracted) length of the axopodiuTll. The power expended per unit length is
(4)
and the total power dissipated is
(5)
The foregoing formulae are used in the sequel.
Ciliophrys marina and Heterophrys marina:
Here Lo = 50 /-lm, a = 0.05 Jim, and U = 2.5 x 10- 3 m s-I (Davidson, 1975).
560h, which likely satisfies the condition a ~ h demanded by equation (I ).)
=3xlO- 6 N m- I
= 3 pN /-lm- 1
F tot
= (3 pN /-lm- I )(50 pm) = 150 pN
P s =(3xlO- 6 N m- 1)(2.5xIO- 3 m $-1)
= 7.5 x 10- 9 W m- 1
= 7.5 fW
J.lm- I
P tot = (7.5 f\N J.lm- 1)(50 J.lm)
= 3.8 x 10 2 fW
(Observe that a =
Hence
38
Echinosphaerium akamae:
From Suzaki et al. (1992a, pp. 431-432), I take Lo = 200 11m, U = 2.6 x 10- 3 m s-l, and I estimate
that a = 2 pm, noting that the radius of the glass needle is 4 lun. (Observe that a = 510h, which may
satisfy the relation a
«:
h required by equation (1).) Therefore
F _ 271"(1 x 10- 3 kg m- I s-I)(2.fi x 10- 3 m s-l)
s In 100 - 1.7.5
=6xlO- 6 Nm- l
= 6 pN pm-I.
The total force exerted in a maximal contraction of 172 IHn (the greatest displacement of the
needle observed - see p. 431 and Table 1, p. 432) is thus
which is of the same order of magnitude as the experimental value, (4.1
P s = (6 x 10- 6 N m- l )(2.fi x 10- 3 m
5-
1)
= 2 x 10- 8 W m- I
P tot
= (1 x 10- 9 N)(2.fi x 10- 3 m 5- 1)
= 1 x 10- 12 W
= 1 x 10 3 fW
± 0.9) x
10 3 pN.
39
CONCLUSTONS
Although it is encouraging that Slender Body Theory yields a value of drag comparable to the
measured force of contraction
III
E. akamae, it should he emphasized that the assumptions of the
theory are not closely obeyed. The surface of the axopodillm has a complicated time-dependent form
and its slenderness diminishes as contraction proceecls.
Moreover, the axopodium always maintains its
continuity with the cell body so that there is no fluid gap, and the actual motion is unsteady. It has
not been possible to estimate the importance of these discrepancies.
40
10. SLIDING PLANE THEORY
{)=a.
STATIONARY
PUl.NE
u
{)=o
SLIDING PUl.NE
The problem concerns two-dimensional Stokes flow.
Introducing the stream function 1/J for
mathematical convenience, we solve the biharmonic equation
(1)
We use polar coordinates and let the plane given by 0 = 0 move at constant speed, while the plane at
0=
Q
remains stationary (see diagram above). Hence the boundary conditions are
(2)
ur
=Uo =0
at 0
=
0'.
The stream function 1/J satisfies the relations
(3)
whence the boundary condtions (1) become
r1 81/J
80
= - U,
~1/Jr
v = 0
at 0 = 0
(4)
- 0
0
r1 81/J
80 = 0, 81/J
8r at = Q.
41
Taylor (1960) and Moffatt (1964) found that the problem has a similarity solution of the form
1/J(r,B) =
Y
2.1. _
'f/ -
1/Jrr
l'
f(B).
1
(5)
1
+ r 1/J r +"2
1/JBB'
l'
(6)
Hence, inserting (5) into (6),
y27j) =
Vf + <10d
1'\
2
;)
whence
which reduces to
(7)
(Note the singularity at
l'
= 0.)
Consider
(8)
The characteristic eq uatioll of (8) is
or
Hence m
= ± i are roots of multiplicity two,
whence
42
Taylor (1960) and Moffat (1964) found that the problem has
ljJ(r,B)
= r frO).
it
similarity solution of the form
(5)
(6)
Hence, inserting (5) into (6),
whence
which reduces to
(7)
(Note the singularity at r
= 0.)
Consider
(8)
The characteristic equation of (8) is
or
?
?
(m- + 1)- = O.
Hence m
= ± i are roots of multiplicity two, whence
43
Taylor (1960) and Moffatt (1964) found that the problem has a similarity solution of the form
1/;(r,B)
=
l'
f(B).
(.5)
(6)
Hence, inserting (5) into (6),
whence
which reduces to
(7)
(Note the singularity at
l'
= 0.)
Consider
(8)
'1'1\1: c11;lrrt.ct,eristic eClilrltioll of (8) is
4
oJ
m +2mor
Hence m
= ± i are
roots of multiplicity two, whence
+1=
0
44
[(0) = AcosO + BsinO + COcosO
+ DOsinO,
['(0) = - AsinO + BcosO + C(cosO - OsinO)
where' denotes
(9)
+ D(sinO + OcosO),
(10)
l(r
Now (4) implies that
EN
[(0) = &'
(11 )
,
f(O)
.
= r1 EN.
80°
Thus the boundary conditions (3) may be recast as
f(O) = 0, ['(0) = - U
(12)
[(a)
= ['(0) = o.
From (9) and (10) we have
•
['(a)
f(O) = 0 :::} A = 0
(13)
['(0) = - U :::} B + C = - U
(14)
[(0) = 0 :::} Bsina + Cocoso + Dosino = 0
(15)
= 0 :::} Bcosa + C(cosa -
osino)
+ D(sina + ocoso)
= O.
(16)
Eliminating B successively between (14) and equations (15) and (16):
- Bcosa - Ccoso = Ucoso
Bcoso + C(cosa - asino)
+ D(sino + acoso) = 0
- Cosina + D(sinn+ocoso) = Ucosa
(17)
- Bsina - Csino = Usina
Bsina + Cacoso
+ Dosino =
0
C(acosa - sina)
+ Dosino =
Usino
(18)
45
Solving (17) and (18) simultaneously by Cramer's rule:
U cosa
sinO'
+ acosO'
U sina
asina
- aSIllO'
si nO' + O'coSO'
acosa - sina
as III a
C
•
whence
(19)
From (14),
B = - (U + C).
(20)
Therefore, inserting (19) into (20) gives
(21)
From (15), we may write
D =
- (Bsina + Cacosa)
asina
whence by substitution for Band C,
D=
U(a-sina cosO')
.'
0'2 _ sin 2 O'
(22)
46
Recalling that t/J = rf(O), and inserting the expressions (13), (19), (21), (22) for A, B, C, D into (9)
yields
t/J=
U~
[-a 2 sinO+sin 2 a(OcosO)+(a-sinacosa)OsinO].
2
2
0' - SIn 0'
In the case of normal incidence,
= f.
0'
t/J =
(23)
Thus, the streamlines are given by
--JlL (- ~2 sinO + 0 cosO + f 0 sinO).
~-1
4
(24)
The following figure (slightly modified to conform with my notation) from Batchelor (1967, p. 225)
depicts some typical streamlines.
•
"'---
~-::....-_--
0'
-,
~
l'.
,
Figure +8.%. Two-<iimensional Row in a cotner due to ODe rigid plme
sli~ aD anomer (arbitrary =its for !f).
The tangential stress on the sliding plane is the shear stress TrO evaluated at 0 = 0 (Leal, 1992; p.
143, above eqn. (4-70); p. 154):
TrO 1 0
= 0 = JJ {r %r
C/)+¥ ~~r}o = 0
=
o.
From (3) and (23),
Uo
= - ~t/J =
vr
~
2 2 [ - 0'2
0' - SIn 0'
sinO + sin 2 0' (0 cosO) +
(0' -
sinO' cosa)O sinO]
(0' -
sinO' cosa)O sinO].
whence
r
g(:0) = r(a
ar\
2
-
U.
2
SIn 0')
[ - 0'2
sinO + sin2 0' (0 cosO) +
(25)
47
Therefore,
(26)
Also from (3) and (23},
ur
w
=}88 O· =
a
U.
2
-Sill
2
a
[-a 2 cosO+sin 2 a(cosO-OsinO)+(a-sinacosa)(sinO+OcosO)]
whence
u
} 88 Or =
2 U. 2
[a 2 sinO - sin 2 a (2sinO + 0 cosO)
r(a -sm a)
+ (a -
sina cosa)(2cosO - 0 sinO)].
Accordingly,
8pr I
u
-
_
~O 0 - 0 -
2U
(
.
2
• 2
a-smacosa).
r(a -sm a)
(27)
Thus, by substitution of (26) and (27) in (25),
T rO
. 2
I 0 -- 0 = r(a 2 2~U
-sm a)
(.
)
a - sma cosa .
(28)
To apply sliding plane theory to axopodial contraction, I consider the following arrangement:
•
U=a
u
STATIONARY
PLANES
8=0
o
SLIDING PLANE
Hence, from (28), the drag stress on both sides of the moving plane is, by symmetry,
liD
=
4~U. 2 (
2
a -.
sma cosa ) .
r(a -sm a)
The criterion for the self-consistency of the assumption of negligible inertia is
(29)
48
- rU
Re-zr
where Re denotes the Reynolds number, and
1/
~
1,
(30)
is the kinematic viscosity of water (Batchelor, 1967, pp.
225-226). Hence the solution is valid within a neighborhood of the intersection 0 of the stationary and
moving planes given by
1/
(31)
r~U·
Now we determine the range of validity of sliding plane theory for parameter values characteristic
of fast axopodial contraction in some heliozoa. We shall take
1/
= 1 x 10- 6 m 2 s-l.
Heterophrys marina and Gi/iophrys marina
Since U = 2.5 x 10- 3 m s-1 (Davidson, 1975), r ~ 4 x 10-4 m = 400 j1.m.
•
length of the axopodium, 50 j1.m, is
k(400
Note that the initial
j1.m) •
Echinosphaeri1l.m akamae
Here U = 2.6 x 10- 3 m s-1 (Suzaki et al., 1992a). Thus, again, roughly, r ~ 400 j1.m. Observe that
the initial length of the axopodium, :::::: 200 j1.m, is! (400 j1.m).
GONGL USIONS
I suggest that it may be useful to apply Sliding Plane Theory to H. marina and G. manna. E.
akamae seems not to satisfy the criterion for self-consistency.
Values of force and power for H. marina appear in the table comprising Section 11. It can be seen
that Sliding Plane Theory yields values about two orders of magnitude smaller than estimates founded
on Slender Body Theory and Stokes's law. Given that the finite dimensions, time-varying shape, and
unsteady motion ofaxopodia grossly violate the hypotheses of Sliding Plane Theory, it is not surprising
that the results depart drastically from those predicted by Slender Body Theory and Stokes's law.
49
However, I do not know whether these values of force and power are accurate, for they lie in the same
range as has been reported for swimming spermatozoa of various phyla (5-350 pN; Green, 1988; 2-200
fW; Pedley and Kessler, 1992), and single-celled algae (diameter 10 /lm, speed 100 /lm s-l) have been
estimated to expend 2 fW in locomotion (Pedley and Kessler, 1992, p. 318).
•
50
11. ESTIMATES OF FORCE AND POWER IN FAST AXOPODIAL CONTRACTION
HELIOZOAN
Echino&phaerium
akamae
FORCE
POWER
(pN)
(fW)
4 X 103
103
._-----
103
SPECIFIC POWER
(W kg-I)
--------
HOW VALUES
REMARKS AND REFERENCES
OBTAINED
Glass needle deflection
10°
SBT
Suzaki et al. (1992)
Specific power based on est. mass of
entire axopodium; de Mestre and
Russel (1975); Brennen and Winet
(1977)
103
Heterophrys
102 •
103
5 X 101_1 X 10 2 t
5 X 10°_2 X 10 2
(1.1-7.5) X 102 •
SBT
Specific power based on est. mass
of X-body; idem
.Cak./measurement by
Davidson; tCak. by
B.E.D. from Davidson
cine data, Stokes's law
applied to transport
marina
e
Specific power based on est. mass
of entire axopodium; Davidson
(1975)
of attached carbon
sphere
1.5 X 10 2
3.8 X 10 2
1.8 X 104
SBT
Specific power based on est. mass
( "'" 10- 14 g) of central filament
de Mestre and Russel (1975);
Brennen and Winet (1977)
~
8 X 10°
2 X 101
-_.._----
SPT,
a
=f
-
90% contraction; Taylor (1962),
Moffat (1964), Batchelor (1967),
Sherman (1990), Leal (1992)
5 X 10°
1 X 101
4 X 10°
1 X 10 1
_.. _-----_ ...
---- ..... ----
SPT,
a
SPT, a
= 23
1r
idem
=
3:
idem
Slender Body Theory (SBT) and Sliding Plane Theory (SPT) presuppose Stokes flow and constant speed. See the text
and references cited for details of additional assumptions. Only external hydrodynamic resistance is considered in the
calculations of power and specific power. Sliding Plane Theory is of doubtful applicability at the size scale of E.
akamac.
51
12. SOME ROUGH ESTIMATES OF ENERGY UPTAKE FROM FOOD IN HELIOZOA
Laybourn-Parry (1992, pp. 83-84) remarks that measured protozoan digest.ive (= assimilation)
efficiencies range between 22 and 76%. As a conservative estimate, I shall take the efficiency to be 20%
for the protozoa considered below.
Actinophrys sol (Heliozoa)
Patterson and Hausmann (1981, Fig. 12, p. 44) observe that in A cfin ophrys sol the resorption of
the fluid contents of the food vacuole normally is completed t1 h after capture of one Co/pidium
colpoda.
[From Patterson and Hausmann (1981, p. 40), I gather that the prey is destroyed by lysis
typically within 25-35 min after capture.] For simplicity in the ensuing calculation, I shall assume that
energy uptake occurs uniformly throughout this 4 h period.
colpoda, but I have a value for the smaller C. campy/um.
generally somewhat larger than C. campy/urn:
length
20
J lack the heat of combustion of C.
[According to Blick (1972), C. co/poda is
C. colpoda length = 100-150 JIm;
C. campy/um
= 50-120 JIm.] Laybourn-Parry (1984, p. 140) gives for C. campy/um a heat of combustion of
± 1.5 J mg- 1 ash-free dry mass. Note that (Ash-free dry mass) ~ (Dry mass).
For the sake of
convenience in a rough calculation, and realizing that Colpidium lacks a mineral skeleton, I take the
foregoing masses to be equal.
My reading of Fig. 3.7, p. 76 in Laybourn-Parry (1984) indicates an approximate cell volume range
of (1-12) x 10 4 Jlm 3 for C. campy/urn. I shall assume a cell volume of 6 x 10 4 l.Lm 3. If, following Caron
and Goldman (1990, legend of Table 7.4, p. 291), we assume a typical protozoan density of 1.0 g cm- 3
and use the empirical relation
Dry protozoan mass = 0.2(Wet protozoan mass),
then for one C. campylum cell
Dry mass
~ 0.2(6 x 104 Jlm 3 )(1.0 gjl0 12 J.Lm 3 )
= 1 x 10- 8 g
(1)
52
= 10 ng.
[By comparison, the dry mass of a well-fed Podophrya fixa rangcs from ~ 4-10 ng (Laybourn, 1976,
Table 1, p. 206), and a large Didinium or small Stentor havc a dry mass of
~
200 ng (Laybourn-Parry,
1984, p. 99). Thus, I think my estimate is reasonable.]
Therefore, the heat of combustion of one C. campylum is
(20.15 J mg- 1)(1 x 10- 8 g)(10 3 mg g-l)
= 2 x 10- 4 J.
The energy taken up by A. sol is thus
(A. sol assimilation efficiency) ( C. campyillm hCi'lt of combustion)
(2)
= 0.20(2 x 10- 4 J)
= 4 x 10- 5 J.
The average power uptake by A. sol is
Energy uptakc
Time for assimilation
(3)
4 x 10- 5 J
= 3 X 10 6 fW.
Owing to the generally larger size of the actual prey, C. colpoda, the probably small difference
between dry mass and ash-free dry mass, and the low assimilation efficiency assumed, I consider this
figure to be a conservative estimate of power uptake.
According to Patterson and Hausmann (1981, p. 46), an A. sol which has not fed for at least 24 h
has a body diameter dB
~
45 pm. Hence the volume of the cell body is
(4)
= t1r(45 t1m )3
= 4.8 x 10 4 pm 3 .
53
The same typical A. sol has
~
42 axopodia, each of length Lo
~
100 11m and basal diameter d A = 1.25
pm. Thus the aggregate volume of the axopodia is
VA = 42 X /2 il"Lod A
2
(5)
= ~~( 100 Ilm)(1.25)2
= 1.8 x 103 pm 3.
The total volume of one A. sol
IS
Vtot=VB+VA
(6)
= (4.8 x 10 4 + 1.8 x 10 3 )
/lm 3
=5.0 x 104 11m3.
N6te that this volume is of the same order of magnitude as that of Colpidillm.. [A. sofs gluttony is not
unprecedented among heliozoa.
Dragesco (1964) descrihes the voracit.y of Echinosphaerillm. eichhorni.
And Capriulo (1990, p. 221) points out that aloricate ciliates (e.g., Didinium.) can ingest prey larger
than themselves.] From the cell volumes (pm 3 ) and respiration rates (nl 02 /lm- 3 h- 1) given by Caron
et al. (1990, Table 8.1, p. 309), I calculate the power dissipated in respiration by Colpidium to be (4.76.2) x 10 5 fW. Hence, my presumably conservative estimat.e of power uptake by A. sol is 5-6 times the
respiratory power expenditure of Colpidium.
I conclude t.hat A. sol
likely obtains energy from
Colpidium at a rate sufficient to sustain itself, if not to grow and reproduce.
Spongodrymus (Radiolaria) in relation to Echinosphaerillm (Heliozoa)
Anderson et al. (1984) measured the uptake of 14C-labeled algal and crustacean prey by a solitary
radiolarian of the genus Spongodrymus. According to Anderson (l983, p. 181), Spongodrymus catches
its prey by means of swiftly contracting axopodia.
Actinopoda.)
diameter of
(Like heliozoa, radiolaria are members of the
Anderson et al. (1984, p. 206) note that the species they studied had a cell body
~
800 pm, comparable to that of the freshwater heliozoan Echinosphaerium eichhorni,
which may attain a cell body diameter of over 1000 pm ( = 1 mm;
Barrett, 1958, p. 206;
Febvre-
Chevalier, 1985, p. 310). Thus, there are grounds for supposing that power uptake rates (derived from
54
measured predation rates) for Spongodrymus may be of the same order of magnitude as those for
Echinosphaerium.
Anderson et al. (1984, p. 205) found that when the dinonagcllate A mphidinium carterae was
supplied at a density of 4.2 x 10 3 cells (ml culture mediurnf 1, one Spongodrymus ingested 983 cells in 2
h.
[Anderson et al. (1984, p. 206) caution that the rates ohserved are "ad libitum feeding rates and
probably do not represent algal [sic] predation rates in the natural environment where the algal prey
density may be lower by several orders of magnitude".] The caloric content of one dinoflagellate cell
can be estimated by means of the linear regression formula (Hitchcock, 1982, Eqn. 4, p. 371)
log(ncal ceW 1) = 0.80 log(cell volume in 11m3)
+ 0.90; r=0.93, n=8.
(7)
This formula is founded on data for eight species of dinonagellates, whence n = 8.
From Hitchcock (1982, p. 3(5), the mean volume of A. carterae (based on measurements on a
minimum of 25 cells) is 818
j.lm
3. Hence, regression formula (7) above yields
log (neal cen- 1) = 0.80 log (818)
+ 0.90
= 3.2
Therefore, one cell has a caloric content of
10 3.2 neal
= 1.7 x 10 3 neal
= 1.7 x 10- 6 cal
= i.l x 1O- 6 .J (1 cal = 4.184.1).
Thus the caloric content of 983 cells is
(7.1 x 10- 6 J ceW 1)(983 cells)
= 7.0 x 10- 3 .J.
Assuming conservatively an assimilation efficiency of 20%, the energy uptake of Spongodrymus is
1.4 x 10- 3 .J.
If we suppose that this energy is transduced over a period of 4 h (the same length of time as for A.
s00,
then the power uptake is
55
= 9.7 x 10- 8 J s-l
= 9.7 x 10 7 fW.
Since the cell body of Spongodrymus is :::: 800 jJm in diameter, its volume is approximately
VB = i7l"(800 jJm)3
- ?_. 7 x
10 8 jJm.
3
I have no information on the number or linear dimensions of the axopodia of Spongodrymus. If we
assume that their collective volume is roughly 10% of the cell body volume, then we may take the total
volume of one Spongodrymus to be
From Caron et al. (1990, Table 8.1, pp. 309-310), I extract the following data on a heterotrich and
a foraminiferan for comparison with my estimates for Spongodrymus. Values for Spongodrymus are in
italics.
ORGANISM
Spiro.ftomum
VOLUME
RESPIRATORY POWER/POWER UPTAKE FROM PREY
(jJm3 )
(fW)
1.2 x 10 7
6.7 X 10 7
1.25 X 108
3.3 X 10
3x 108
9.7X 10
amhiguum
(heterotrich)
Roulina
leei
(foraminiferan)
Spongodrymu.f
(radiolarian)
8
7
56
According to Anderson (1983, pp. 121, 123), Spongodrymus contains symbiotic algae;
perhaps
these make a significant contribution to its energy budget.
Now I compare an estimate of the energy dissipated against hydrodynamic drag in axopodial
contraction with the energy acquired in feeding by Spongodrymus and A. sol.
From Suzaki et al. (1992, Table 1, p. 432), the contraction time for E. akamae is
172 p.m
2600 p.m s-l
= 6.5 x 10- 2 s.
The total power dissipated is, by Slender Body Theory (cf. Section 9),
3 x 10 3 fW
= 3 x 10- 12 J s-l.
Hence, the total energy expended is
(3 x 10- 12 J s-1)(6.5 x 1O- 2s)
=2xlO- 13 J.
13
Observe that this is only
X 10100 = 1 x 10-8 % of the energy taken in by Spongodrymus and
3
-13
1.4 x 10- J
x 10 5
100 = 5 x 10- 7 % of the energy obtained by A. sol in the feeding bouts recounted above.
4 x 10- J
(2
(2
J)
J)
The actual percentages are likely larger because I take no account of drag on attached prey, power
expended against internal viscous resistance and in breaking down microtubules, etc.
General Comparative Considerations !ill Power Dissipation in Microorganisms
Earlier in this section I have shown that for A. sol, roughly,
For Tetrahymena pyriformis, whose cell volume is of the same order of magnitude, measured
respiration rates are generally ,.., 10- 5 nl 02 p.m- 3 h- 1 (Caron et al., 1990, Table 8.1, p. 310). Hence,
assuming equality of respiration rates, I estimate that the metabolic power of one A. sol is ,.., 10-6 W.
If we suppose that the power dissipated by A. sol in axopodial contraction is the same as in E. akamae
57
(likely an overestimate because axopodia of A. sol are smaller), then the fraction of metabolic power
( = fraction of metabolic energy) spent to overcome hydrodynamic drag is approximately
12
( 10- 6
10- W
W) 100 = 10-4%.
Note that even if all ::::: 42 axopodia contract at once, the fraction is only about 10- 3 %.
compare the foregoing figure with the following estimate due to Bray (1992, p. 6):
Swimming consumes only a
small fraction of the cell's energy
The power required to propel a cd! in W'ateris quite small. Itmai-b~ calculated
::s the viscous drag multiplied by the velocity. For a spherical cd!
(61ta11v) X v
power
where
11
viscosity of water (lO-2 g/ cm sec)
a
radius of cell
v
= velocity of cell
If the cd! h::s a r.o.dius of 1 J.UI1 and is trd.vding at 10 J.UIl!sec, che power consumed is 2 X lOo11 e.rgsIsec (2 X 1O-l8 J/sec).
What is this power requirement in tc:III1S of ATP molecules, che principal currency of energy in the cell? Hydrolysis of one gram mole of ATP releases
about 40KJ of useful. energy; hydrolysis of a single ATP molecule, abOUt 10'1' J.
The cd! in the above calculation chert:.:are requires che"hydrolysis of at le::sc 20
molecules of ATP every second in order to maintain its speed of 10 ~sec. Ifwe
::ssume a low efficiency of a few' percenc.ge points., chen che ATP hydrolysis due
to swimming might be 1000 molecules per second. Since che mecboiic me of a.
typica.l cd! is around 10' ATP molecules per second, ho~, ~g is noe
a ma.jor energy case.
'.,
Hence, Bray's fictive coccus devotes only 10- 2% of its metabolic power to locomotion.
It is useful to
58
An analogous estimate was made by Pedley and Kessler (1902, p. 318):
The energy stored within microorganisms is acquired by various
means such as photosynthesis. When an algal cell swims it generally uses
a very small amount of its stored energy, so that for any given experiment
of several hours duration, or for an overnight interval of active swimming
by the cell. energy consumption and supply need not be considered.
The supply and consumption time scales are not as well separated for
bacteria.
From the Stokes fonnula one may calculate, for a "typical" algal cell
(radius 10 ,Urn. speed 100 ,Urn) a swimming power of2 x 10- 15 J S-I cell-I.
The light input can be taken as l W m - 2; using aphotosynthetic conversion
efficiency of 3%,and the cell's area, the power input is estimated at 10- 11
J s - 1 cell- I. Thus, for algae, swimming requires only a small fraction of
the input energy. Estimates for spermatozoan swimming power range from
2 x 10- 15 to 2 X \0-13 J S-I ceil-I (Bishop 1962); in that case the fluid
- medium in which they swim can also supply energy. 1be temperature
increase or the fluid due to the swimming power dissipated by a typical
cell population is quite negligible. For \07 cells em - J, the temper:uure rises
by 10-' °c s - I. If 1 W m-7! light intensity is absorbed and converted to
heat by the cells, the temperature rise from that source can be as high
-:: -as-iO:~'oc S-I. Illumination can be a source of convection currents in
experiments that investigate the ir.teractions of organisms with light.
ENERGY
Hence, Pedley and Kessler's fictive alga devotes only 2 x 10- 2% of its power input to propulsion.
Fenchel (1986, p. 83) remarks (referring to Fenchel and Finlay, 1983) that "motility in growing
protozoa can at most account for 0.1-0.7% of the energy budget".
Now I adduce data for certain nonheliozoan contractile protozoa.
From Hawkes and Holberton (1975) I have for Spirostomum ambiguum the following facts:
Typical extended length
= 1.5 x 10- 1 cm = 1.5 mm (p. 599)
Fast contraction occurs to
Cell radius
~
43% of extended length (p. 599)
=75 pm for extended cell=> cell diameter = 150 pm
(p.597)
Speed of contraction = 20 cm s-1 (p. 595)
Hawkes and Holberton (1975, p. 601) estimate that the energy dissipated in doing work against the
viscous external medium in one body contraction (complete in 10 ms) is 2.13 x 10- 3 erg = 2.13 x 10- 10
J.
Amos (1971, p.127j 1975b, p. 413) used Stokes's law to estimate the drag on the cell body of
59
Vorticella. Amos treats the cell body as a sphere of radius 20 J.lm, pulled 80 J.lm at an average speed of
23 mm s-l. He finds that the viscous drag is 8.6 x 10- 9 N. Hence, the work done to overcome drag in
one contraction is
(8.6 x 10- 9 N)(8.0 x 10- 5 m)
= 6.9 x 10- 13 J.
It seems likely that this figure is lower than the true value owing to the steep increase in drag as a
sphere closely approaches a plane wall at normal incidence [See Power and Power (1993, Fig. 2, p. 61).]
Also, the calculation omits consideration of drag on the stalk.
The subjoined table epitomizes the foregoing discussion.
ESTIMATED ENERGIES OF FAST CONTRACTION FOR SOME PROTOZOA
ORGANISM
TYPICAL DISPLACEMENT
(Jlrn)
TYPICAL SPEED
(J.lrn s·l)
TYPICAL TIME
(rns)
ENERGY
(J)
Spiro8tomum
ambiguum
855
2 X 10 5
-4
2.13 X 10. 10
Vorticella
conl1allaria
80
2.3 X 10 4
-3
6.9 X 10- 13
172
2.6 X 103
65
2 x 10- 13
:
Echino8phaerium
a.camae
60
13. OUTLINE OF A RESEARCH PLAN TO ELUCIDATE
THE MECHANISM OF FAST AXOPODIAL CONTRACTION
I. INTRODUCTION
A. Synopsis of the Heliozoa
A polyphyletic assemblage of predaceous freshwater and marine protozoa, here considered to
embrace pedinellids (Siemensma, 1991).
Globose cell bodies (diameter'" 8-2600 pm) bear many ( '" 20-several hundred; I do not know
the max. number yet found) radially disposed slender protrusions called axopodia (diameter
'" 0.1-10 pm; length '" 30-500 pm). Each axopodium is stiffened by an axoneme consisting of
an array of microtubules (3- '" 500) cross-linked in a taxonomically diagnostic pattern (FebvreChevalier, 1985; Siemensma, 1991).
Axopodia serve several functions, e.g., locomotion, cell fusion, cell division, egestion,
attachment to substrate, and prey capture.
B. Fast axopodial contraction
Axopodia can contract rapidly (duration of contraction '" 4-60 ms; absolute speed of
contraction '" 2.6-60 mm/s; relative speed of contraction '" 13-300 axopodial lengths/s;
extent of contraction = shortening per unit initial axopodial length = 66- '" 100%) in response
to various artificial stimuli and contact with prey (Febvre-Chevalier, 1981; Febvre-Chevalier
and Febvre, 1992; Suzaki et al., 1992a).
The mechanism of contraction is unknown. Contraction proceeds by breakage of the axonemal
microtubules (but see abstract by Suzaki et al., 1992b), requires calcium ions in the external
medium, is preceded by an action potential in at least one species, and may be driven by
putative contractile elements visible in some electron micrographs.
C. Speed disparity conundrum
I
[I.e., fast axopodial contraction entails microtubule (MT) breakdown, yet contraction speeds
~ MT disassembly speeds typically measured in vivo and in vitro.]
Although the speed of fast axopodial contraction is not extraordinary in comparison with that
of other forms of protozoan motility (e.g., max. swimming speed of Uronema
1.15 mm/s 46
cell lengths/s [Sleigh and Blake, 1977]; avg. cell body speed owing to spasmoneme contraction
in Vorticella = 23 mm/s [Amos, 1975b, p. 413]; cell body contraction speed in Spirostomum,
Stentor '" 200 mm/s '" 100 cell lengths/s [Huang and Mazia, 1975, p. 389]), it is exceptional
for a process depending on the disintegration of MTs (Amos and Amos, 1991; Febvre-Chevalier
and Febvre, 1992).
=
=
Cf. observations by differential interference contrast (DIC) video microscopy of MT shortening
in various cell types and in vitro:
In reticulopodia of Reticulomyxa, max. endwise disassembly speed '" 20 pm/s (Chen and
61
Schliwa, 1990). Caplow (1992) remarks that this is the greatest rate of MT depolymerization
yet measured in vivo.
By contrast, the mean endwise disassembly rate of MTs showing dynamic instability in CRO
fibroblasts is only 32.2 ± 17.7 pm/min (duration of shortening = 4.3 ± 3.3 Sj n=18j Shelden
and Wadsworth, 1993).
[See Keith and Farmer (1993) for warnings on the use of
photobleaching to monitor MT dynamics.]
In vitro endwise disassembly speed for phosphocellulose-purified (hence presumably MAP-free)
beef brain MTs in high-calcium ([calcium ion]= 5 mM) solution ..... 1.8 J.lm/s (Gal et al., 1988).
[Aside: Note that Zigmond (1993) reports a milder speed discrepancy for F-actin.]
Cf., as other instances of fast MT breakdown, axopodial contraction in the spumellarian
radiolarian Spongodrymus; contraction ofaxopodia and axopodium-like stalk in pedinellids
(allied with and possibly ancestral to actinophryid heliozoans); autotomy of flagella In
Chlamydomonas, diverse taxa. Find estimates/measurements of the speeds of these processes.
D. Insufficiency of MT breakdown to produce contraction
Cf. experiments of Tilney, Davidson, Schliwa, Salmon, others.
Cf. particle and chromosome movement in vitro by MT depolymerization
Remarks
by
Oster,
Stossel,
polymerization/ depolymerization.
others
on
cell
movement
and
filament
Cf. neurite retraction.
E. Buckling, bending conundrum
Why don't axopodia buckle locally or globally or propagate bending waves during normal
contraction? [Cf. cilia, non-retractile and retractile flagella, haptonemata. Note that Lee (1989)
claims that some haptonemata do not coil.]
II. GOALS
A.
Characterize the main physical and biological principles governing fast axopodial
contraction.
In particular, I propose scaling relations and compare other protozoan groups, both closely and
distantly related. (Comparisons with technological artifacts may be helpful.)
B. Make experimentally testable predictions, suggest further experiments and theory (point out
simplifying assumptions) to settle outstanding questions.
[C. Moot a theory/model of how contraction occurs.]
III. WARNINGS
A. Accepting that the Heliozoa are polyphyletic, there may be
contraction.
> 1 basic mechanism of
62
B. The pitfall of adaptationism
Perhaps the diversity of axonemal MT patterns is not a response to various selection pressures.
[Recall Gould and Lewontin (1979).]
Axopodia serve functions other than prey capture by fast contraction. (E.g., slow surface
transport, locomotion, cell fusion, egestion.) These functions likely influence mechanical design.
IV. SUMMARY OFAXOPODIAL MORPHOLOGY IN THE EXTENDED AND CONTRACTED
STATES
A. Early observations of axoneme birefringence, etc. by Engelmann, MacKinnon, Roskin,
Brandt, others.
B. A possible dichotomy: Is it true that centrohelidians have only non-tapering axopodia
whereas other heliozoan orders have tapering axopodia? Settle the issue by consulting Smith
and Patterson (1986) and marking the table, synopsis in Margulis et al. (1990). Also see
Patterson and Hedley (1992, p. 21) for a summary of spp.
Relation between tapering and size?
Do (non-) tapering axopodia always have (non-) tapering axonemes?
•
How does tapering scale?
Cf. dendrites, axons, reticulopodia, etc.
Electrical implications of tapering?
Hydrodynamic implications of tapering?
C.
Basal origin ofaxonemes on nuclear plaques or centroplasts
These microtubule organizing centers (MTOCs) would seem to forbid appreciable axial sliding
of intact MTs [but see abstract by Suzaki et al. (1992b)]. Also note constraint on axial sliding
imposed by cell body diameter.
A telescoping mechanism involving many relatively short segments possibly could provide high
speed without axopodial buckling. Do Davidson's (1975) observations of haptocyst motion tell
against telescoping in the heliozoans he studied?
D. Miscellaneous Axopodial Organelles
Putative contractile elements seen by electron microscopy (EM) in some genera, not seen in
others. Note various conditions of fixation, etc. List all species in which putative contractile
elements have been found.
Mitochondria in the axopodia of larger spp. (only?) need not imply an ATP requirement for
contraction, may serve other motility, general metabolic processes. Functional implications of
dilated, tubular cristae in mitochondria of the cell bodies or axopodia of most heliozoan taxa?
See Patterson (1986, Fig. 4.24 , p. 60; p.62).
Extrusomes
63
See Bardele, Grain, Davidson, Hausmann and Patterson, others.
E. The condition of the axopodial membrane, axopodial contents following contraction.
Cf. suctorian tentacles, single myocytes (Krueger et al., 1992).
Does the axopodium thicken appreciably upon contraction?
v.
MT PROPERTIES
A. Supramolecular structure
2-D surface crystals and the passage of dislocations, other defects.
B. Polarity [possible implications of Mitchison (1993)?]
C. Flexural rigidity
D. Persistence length
E. Euler buckling, bending as cantilever
F. Low Euler buckling, bending loads imply utility of battens for structural stability
•
Measurements, observations, theory on other built-up cell structures containing MTs, F-actin.
G. Waugh (1989) on mechanics of the marginal band of newt RBCs
H. Is an axoneme necessary for mechanical support, other functions?
[Other functions might include service as track for molecular motors, but recall the
experiments of Edds (1975a,b).]
Cf. pseudopodia of filose amoebae, axons, reticulopodia, other cell protrusions containing
loosely organized bundles of microtubules. N.B. Do pseudopodia of filose amoebae contain MTs
or do they derive support from F-actin, other filaments? How large are prey taken by filose
amoebae?
Measurements, observations, and theory on built-up cell structures containing MTs, F-actin.
I. Dynamic instability [see data summarized by Erickson and O'Brien (1992)]
J. Disassembly speed increased by calcium ions
K. What is the minimum radius of curvature of a single MT below which breakage occurs?
1. What is the minimum radius of curvature of a cilium, flagellum, or axopodium below which
breakage occurs?
M. See Hill and Kirschner's theory (Hill, 1987, pp. 51-56) for the steady-state (Le., constant
shortening speed) loss of subunits from linear supramolecular aggregates under compressive
force. Also see Hill (1987, pp. 24-28) and Gordon and Bro~land (1987).
64
N. Protofilament numbers in heliozoans
See Jones, Tilney, Cachon et al.
O. Biochemical peculiarities of heliozoan tubulins
See Luduena.
P. Stable and unstable populations of MTs: Categorize heliozoan axonemal MTs, recall
possible function of histones. See Tilney, Schliwa, Matsuoka et al. (1984b), others on colchicinesensitivity, etc.
Q. See Pryer et al. (1992) on influence of MAPs on brain MT dynamic instability in vitro. Also
see Gelfand and Bershadsky (1991).
VI. THE AXONEME
A. Function as armature (beam-column) for the axopodium
B. Hypothetical design(s) for this purpose
Why aren't axonemes simple circular palisades of MTs?
C.
Diversity of actual axoneme patterns, classification of symmetries, comparison with
hypothetical design(s) noted in B.
•
Summary of axoneme structure after Anderson (1988, pp. 242-243).
Actinophryid axoneme structure.
Compare areal density of MTs with that for neurites and Allogromia reticulopodia (Travis and
Bowser, 1991).
D. General principles/concepts of structure, stability, and strength.
Consider tapering, entasis (cf. Math. Intelligencer article) in relation to beam-column theory.
See esp. Ashby (1991), Parkhouse (1987), Wainwright et al. (1976).
Structure of axoneme such as to promote spread of damage?
E. How does the axoneme fail in contraction?
Does the structure of the axoneme promote spread of damage?
Does "intercalary destabilization" in the stalk and axopodial axonemes of Actinocoryne = shear
band formation? (Febvre-Chevalier and Febvre, 1992)
Does evidence for stalk oscillation in Actinocoryne indicate possible modes of failure?
Can I make suggestions as to the statistical distribution of lengths of MT fragments
immediately after contraction?
What predictions could be derived from a scaling arguIl1;ent comparing the stalk and axopodia
of Actinocoryne? What, e.g., would an argument from geometric similarity suggest? Also,
65
consider the possible effects of parallel axonemes on the speed of contraction, etc. Recall that
the stalk contains parallel axonemes.
Apart from the simple kinetic arguments adduced earlier, can the endwise disassembly model
be debunked by treating it as a Stefan problem?
Vale's (1991) MT-severing factor or a hypothetical gelsolin-like protein (cutting MTs rather
than F-actin) would seem to act too slowly. But would such a severing factor have to cut all of
the protofilaments in order to be effective?
Is there evidence for or against the existence of periodic/intermittent breakpoints in the
axoneme?
Do recent studies of AI/ogromia, axon MT assembly bear on the foregoing question?
Presumably the axoneme must fail reliably when stimulated appropriately. The axoneme may
not be perfectly homogeneous; possibly there exists a distribution of breaking strengths along the
axoneme or within a given section, whence global failure of the axoneme is governed by both
deterministic and stochastic effects. See Gaines and Denny (1993).
Upper bound on speed of breakdown/contraction given by elastic wave speed in the axoneme?
VII. SPEED LIMITATIONS
A. Triggering of contraction, spread of excitation
Role of action potential [see Major's (1993) papers and refs. therein]' implications of space
constant, comparison with other protozoans [see Febvre-Chevalier et al. (1989) review].
Electrical implications ofaxopodial tapering?
Electrical phenomena associated with extrusome discharge and contractile element actuation?
Diffusion of calcium ions from calcium channel pores (Barritt, A.V. Hill, many others).
Calcium waves probably are too slow.
B. Scaling relations for contraction
Clues from the stalk of Actinocoryne?
C. Speed limitations inherent in hypothetical breakdown schemes (e.g., series, parallel, etc.)
Upper bound on speed of breakdown/contraction given by elastic wave speed in the axoneme?
VIII. SQME CONTRACTION MECHANISMS THAT QAli BE RULED Q!IT. QR SHOWN I.Q BE
IMPROBABLE
A. Actomyosin
66
Are Davidson's (1975) arguments against this valid? (But see Edds, Febvre-Chevalier.)
Cf. speeds of cytoplasmic streaming in characean algae.
B. Dynein, other motor molecules
Cf. speed of MT extrusion from ciliary and flagellar axonemes, etc. Also see Amos and Amos
(1991), Peskin et al. (1993) for calculation of max. Brownian ratchet speeds, other data?
Cf. Sheetz et al. (1992) on motor-induced retraction of actin-supported filopodia.
C. Marangoni effect
See Probstein (1986) and Edwards et al. (1992).
D. Membrane tension
See Manton (1964, 1968), Davidson (1975), Peskin et al. (1993), Deiner et al. (1993), others.
[Note that the foregoing schemes in C and D presuppose prior or concomitant disassembly of
MTs.]
E. Myonemes
There is no ultrastructural evidence for myonemes of acantharian, peritrich form; however, I
should inquire into conditions for satisfactory preservation. Note that the rowing movements of
the axopodia of Sticholonche are produced by contraction/relaxation of proximal nonactin
filaments.
IX. THE CHEMOMECHANICS OF THE CONTRACTILE ELEMENT
A. Evidence for a centrin/spasmin-like contractile element
Fluorescent anti-centrin antibody binding in the pedinellid Pteridomonas.
B. Evidence against a centrin/spasmin-like contractile element
C. Phase transition in a gel
Cf. papers by Tanaka, Katchalsky, others on gels.
Cf. work of Nanavati (Duke postdoc.) on secretory granule matrix. Note voltage-dependence of
swelling.
Cf. research on gels as active materials.
X. EXTERNAL HYDRODYNAMICS
A. Impulsively started flow at low Reynolds number?
achieve steady speed?
XI. INTERNAL HYDRODYNAMICS
What is the characteristic time to
67
A. Possible relevance of lubrication theory
B. Pressure-driven flow at low Reynolds number
Is the pressure gradient required unreasonably large? See ref. cited in Spero (1982).
Cf. possible piston-like action of euglenid, other protozoan ingestion devices.
C. Transient discharge at low Reynolds number in tubes of small bore
D. Steady low Reynolds number flow in tubes of small bore
XII.
COMPARISONS WITH AND CLUES FROM OTHER PROTOZOA, MT-SUPPORTED
STRUCTURES, CELL PROTRUSIONS
A. Cf. heliozoan axoneme patterns with putative diffusion-limited aggregation (DLA), other
patterns in axonemes of radiolaria and acantharia
Mechanical and morphogenetic implications of tessellations vs. DLA?
See Ashby, Dyson, Lakes, Mandelbrot, Schroeder, others on scaling. Connection between
Dyson's scaling relation and scaling relations for strength of honeycombs, other hierarchical
structures?
Cf. handedness of actinophryid heliozoan axonemes, radiolarian axonemes
B. Suctorian feeding and prehensile tentacles
C. Piston of Erythropsis
D. Haptonemata
E. Euglenid nemadesms and subpellicular MTs
F. Axostyle MTs
G. Importance of "extracellular matrix" in reinforcing reticulopodia of Astrammina
XIII. MISCELLANEOUS
A. Axopodia as "nanosystems" (sensu Drexler, 1992)
B. Scaling of prey capture in heliozoan "diffusion" feeding
See Fenchel's (1984) analysis of the scaling of prey interception in bacterivorous marine
heliozoans.
C. Stirring of medium by axopodia
Verify Suzaki et al.
Echinosphaerium akamae.
(1980)
observations
of spontaneous
axopodial
contraction
lD
See Anderson's (1983) observations of spontaneous axopodial contraction in the spumellarian
68
radiolarian Spongodrymus.
See Purcell (1978) for theory on the efficacy of stirring at low Reynolds number.
D. Time course of contraction in Heterophrys marina cannot be deduced from Davidson's (1975)
observations of haptocyst motion.
E. A physicist's view of general aspects of fast axopodial contraction
XIV. DISCUSSION
A. Probably only a small fraction of the organism's energy budget is spent on contraction
Give rough calclulations and recall that microorganisms spend relatively little energy on
motility.
Energy budgets of protozoa are ably treated by Fenchel (1987, pp. 55-56).
B. Slow reextension compared to other protozoa
Cf. longitudinal flagellum of Ceratium tripos: 1-5 s [Hohfeld et at. (1988) citing Maruyama
(1981)]; Stentor cell body reextension: 10-60 s [Huang and Mazia (1975, p. 403)].
Cf. vorticellids, suctorians, other ciliates.
C. Size-dependence, scaling of reextension times
Note that the low speed of reextension probably entails negligible drag.
Consider work to extend the axopodial membrane, diffusive/convective transport of materials,
theory of length control devised for rodlike viruses [see Peskin et at. (1993); Nossal and Lecar
(1991)].
D. Outstanding questions and decisive experiments for their resolution
E. Concluding summary, prospectus
69
REFERENCES
Adam, G. and Delbruck, M. 1968. Reduction in dimensionality in biological diffusion
processes. In A. Rich and N. Davidson (eds.) Structural Chemistry and
Molecular Biology. W.H. Freeman, San Francisco, pp. 198-215. (QD461 S917)
Adams, D.S. 1992. Mechanisms of cell shape change: The cytomechanics of cellular
response to chemical environment and mechanical loading. J. Cell Biol. 117(1):
82-93. [Experiments on Physarum cytoplasm.]
Aidley, D.J. 1989. The Physiology of Excitable Cells. 3rd ed. Cambridge University
Press, Cambridge. (QH 631 A36 1989)
Akkas, N. (ed.) 1987. Biomechanics of Cell Division. Plenum, New York. (QH605
.N27 1986)
Akkas, N. (ed.) 1990. Biomechanics of Active Movement and Deformation of Cells.
Springer-Verlag, Berlin. (QH611 N36)
Alberts, B., Bray, D., Lewis, J., Raff, M., Roberts, K., and Watson, J.D. 1989.
Molecular Biology of the Cell. 2nd ed. Garland, New York. (Vet. Med. QH
581.2 M64)
Albrecht-Buehler, G. 1990. In defense of "nonmolecular" cell biology. International
Review of Cytology 120: 191-241.
Albritton, N.L., Meyer, T., and Stryer, L. 1992. Range of messenger action of calcium
ion and inositoll,4,5-trisphosphate. Science 258: 1812-1815.
Alexander, R. MeN.
1979.
The Invertebrates.
Cambridge University Press,
Cambridge. (QL 362 A43) [See esp. pp. 64-65 for calculation of calcium ion
influx into Paramecium during ciliary beat reversal.]
Alexander, R. MeN. 1982. Locomotion of Animals. Blackie, Glasgow and London.
(Vet. Med. QP 301 A296 1982)
Alexander, R. MeN. 1985. The ideal and the feasible:
evolution. Biol. J. Linn. Soc. 26: 345-358.
Physical constraints on
Alexander, S.P. and Rieder, C.L. 1991. Chromosome motion during attachment to
the vertebrate spindle:
Initial saltatory-like behavior of chromosomes and
quantitative analysis of force production by nascent kinetochore fibers. J. Cell
Biol. 113(4): 805-815. [Includes important discussion of cytoplasmic viscosity.]
AI-Khazzar, A.R., Butler, R.D., and Earnshaw, M.J. 1983. The effect of ionophore
A23187 and ruthenium red on tentacle contraction and ultrastructure in
Discophrya collini: Evidence of calcium fluxing from intracellular reservoirs.
Protoplasma 117: 158-166.
Allison, A.C., Hulands, G.H., Nunn, J.F., Kitching, J.A., and MacDonald, A.C. 1970.
The effect of inhalational anaesthetics on the microtubular system in
Aetinosphaerium nucleofilum. J. Cell Sci. 7: 483~499.
70
Alt, W. 1990. Mathematical models and analysing methods for the lamellipodial
activity of leukocytes. In N. Akkas (ed.) Biomechanics of Active Movement and
Deformation of Cells. Springer-Verlag, Berlin, pp. 403-42. (QH 611 N36)
Amos, L.A. and Amos, W.E. 1991. Molecules of the Cytoskeleton. Macmillan
Education Ltd., Houndsmill and London. [Reviewed by K. Hausmann in Europ.
J. Protistol. 28: 364. (August 21, 1992)]
Amos, W.E. 1971. Reversible mechanochemical cycle in the contraction of Vorticella.
Nature 229: 127-128.
Amos, W.B. 1972. Structure and coiling of the stalk in the peritrich ciliates Vorticella
and Carchesium. J. Cell Sci. 10: 95-122.
Amos, W.E. 1975a. Structure and protein composition of the spasmoneme. In L.
Bolis, H.P. Maddrell, and K. Schmidt-Nielsen (eds.) Comparative Physiology:
Functional Aspects of Structural Materials. North-Holland, Amsterdam, pp. 99104. (QH510 157 1974) [See p. 103 for notion that contraction results from
"summation in series of many microscopic shortening events."]
Amos, W.B. 1975b. Contraction and calcium binding in the vorticellid ciliates. In S.
Inoue and R.E. Stephens (eds.) Molecules and Cell Movement. Raven Press, New
York, pp. 411-436. (QH 647 M64) [IMPORTANT REVIEW ARTICLE. See p.
411 for remark on bending (as opposed to collapse) of stalk. Also, see A.
Durham's comment in the ensuing discussion of the paper (p. 435). Durham says
that calcium-induced configurational changes occur in apochlorotic beet spot
virus.]
Amos, W.B., Routledge, L.M., and Yew, F.F. 1975c. Calcium-binding proteins in a
vorticellid contractile organelle. J. Cell Sci. 19: 203-213.
Amos, W.B., Routledge, L.M., Weis-Fogh, T., and Yew, F.F. 1976. The spasmoneme
and calcium-dependent contraction in connection with specific calcium binding
proteins. In C.J. Duncan (ed.) Calcium in Biological Systems. Symposia of the
Society of the Society for Experimental Biology Number XXX. Cambridge
University Press, Cambridge, pp. 273-301. (QH302 S622 no. 30) [MAJOR
REVIEW]
Andersen, R.A. and Wetherbee, R. 1992. Microtubules of the flagellar apparatus are
active during prey capture in the chrysophycean alga Epipyxis pulchra.
Protoplasma 166: 8-20. [See Wetherbee and Andersen (1992).]
Anderson, E. and Beams, H.W.
1960.
The fine structure of the heliozoan,
Echinosphaerium nucleofilum. J. Protozool. 7(2): 190-199.
Anderson, O.R. 1983. Radiolaria. Springer-Verlag, New York. (NRL QL368 R2 A5
1983)
Anderson, O.R., Swanberg, N.R., and Bennett, P. 1984. An estimate of predation
rate and relative preference for algal versus crustacean prey by a spongiose
skeletal radiolarian. Marine Biology 78: 205-207. (Nat. Resources Library)
Anderson, O.R., Moss, M.L., and Skalak, R.
1987.
The cytoskeletal and
71
biomineralized supportive structures in radiolaria. In J. Bereiter-Hahn, O.R.
Anderson, and W.-E. Reif (eds.) Cytomechanics: The Mechanical Basis of Cell
Form and Structure. Springer-Verlag, New York.
Anderson, O.R. 1988. Comparative Protozoology: Ecology, Physiology, Life History.
Springer-Verlag, New York. (Vet. Med. QL 366 A53)
Ando, M. and Shigenaka, Y. 1989. Structure and function of the cytoskeleton in
Heliozoa: 1. Mechanism of rapid axopodial contraction in Echinosphaerium. Cell
Motility and the Cytoskeleton 14: 288-301.
Andre, J.C., Bouchy, M., and Donner, M. 1987. Diffusion in organized assemblies and
in biological membranes. Biorheology 24: 237-272. (QC 189 B5) [Roughly
written, with many typographical mistakes.]
Ashby, M.F. 1991. Materials and shape. (Overview No. 92) Acta Metallurgica et
Materialia 39: 1025-1039. (TN1 .A28)
Atkins, A.G. and Mai, Y.-W. 1988. Elastic and Plastic Fracture: Metals, Polymers,
Ceramics, Composites, Biological Materials. Ellis Horwood, Chichester. (TA409
A76 1985)
Atluri, S.N. and Amos, A.K. 1988. Large Space Structures: Dynamics and Control.
Springer-Verlag, Berlin. (TL910 L37 1988)
Baba, S.A. 1972. Flexural rigidity and elastic constant of cilia.
459-467.
J. expo BioI. 56:
Bardele, C.F.
1972.
Cell cycle, morphogenesis and ultrastructure in the
pseudoheliozoan Clathrulina elegans. Zeitschrift fur Zellforschung 130: 219-242.
(QH301 .Z47)
Bardele, C.F. 1974. Transport of materials in the suctorian tentacle. In M.A. Sleigh
and D.H. Jennings (eds.) Transport at the Cellular Level. Symposia of the
Society for Experimental Biology Number XXVIII. Cambridge University Press,
Cambridge.
Bardele, C.F. 1975. The fine structure of the centrohelidan heliozoan, Heterophrys
marina. Cell Tissue Res. 161: 85-102. (QH 301 .Z47) [Not cited by Davidson
(1975).]
Bardele, C.F. 1976. Particle movement in heliozoan axopods associated with lateral
displacement of highly ordered membrane domains. Z. Naturforsch. [C] 31: 190194. (Q3 Z392 -6th floor)
Bardele, C.F.
1977a.
Organization and control of microtubule pattern in
centrohelidan heliozoa. J. Protozool. 24: 9-14.
Bardele, C.F. 1977b. Comparative study ofaxopodial microtubule patterns and
possible mechanisms of pattern control in the centrohelidan Heliozoa
Acanthocystis, Raphidiophrys, and Heterophrys. J. Cell Sci. 25: 205-232.
Barrett, J.M. 1958. Some observations on Actinosphaerium nucleofilum n. sp., a new
fresh water actinophryid. J. Protozool. 5: 205-209.
72
Barritt, G.J. 1992. Communication within Animal Cells. Oxford University Press,
Oxford. (QH604.2 B37 1992)
Barta, E. and Liron, N. 1988. Slender body interactions for low Reynolds numbers.
Part I: Body-wall interactions. SIAM J. Appl. Math. 48: 993-1008.
Barta, E. and Liron, N. 1988. Slender body interactions for low Reynolds numbers.
Part II: Body-body interactions. SIAM J. Appl. Math. 48(6): 1262-1280.
Batchelor, G.K. 1967. An Introduction to Fluid Dynamics.
Press, Cambridge. (QA911 B33)
Cambridge University
Bayley, P.M., Schlistra, M.J., and Martin, S.R.
1990.
Microtubule dynamic
instability: Numerical simulation of microtubule transition properties using a
lateral cap model. J. Cell Sci. 95: 33-48.
Belar, K. 1922. Untersuchungen an Actinophrys sol Ehrenberg. I. Die Morphologie
des Formwechsels. Archiv fur Protistenkunde 46: 1-96.
Belar, K. 1924. Untersuchungen an Actinophrys sol Ehrenberg. II. Beitrage sur
Physiologie des Formwechsels. Archiv fUr Protistenkunde 48: 371-434.
Belhadri, A., Bayle, D., and Brugerolle, G. 1992. Biochemical and immunological
characterization of intermicrotubular cement in the feeding apparatus of
phagotrophic euglenoids: Entosiphon, Peranema, and Ploeotia. Protoplasma
168: 113-124.
Bell, G. 1988. Sex and Death in Protozoa: The History of an Obsession. Cambridge
University Press, Cambridge. (QL366 B45 1988) [The theme of the book is
announced in the title of the first chapter:
"The question of protozoan
immortality". Bell discusses and presents data culled from the researches of
Belar on Actinophrys sol. Belar maintained this heliozoan in culture for more
than 1000 generations.]
Bell, G.1. 1978. Models for the specific adhesion of cells to cells. Science 200: 618627.
Bereiter-Hahn, J. 1987. Mechanical principles of architecture of eukaryotic cells. In
J. Bereiter-Hahn, O.R. Anderson, and W.-E. Reif (eds.) Cytomechanics: The
Mechanical Basis of Cell Form and Structure. Springer-Verlag, New York, pp. 330.
Bereiter-Hahn, J., Anderson, O.R., and Reif, W.-E. (eds.) 1987. Cytomechanics:
The Mechanical Basis of Cell Form and Structure. Springer-Verlag, New York.
(Chapel Hill Zoology Library: QH 611 .C97)
Bereiter-Hahn, J. 1978. A model for microtubular rigidity. Cytobiologie 17: 298300. (QH 573 C89)
e
Berk, D.A., Clark, A., and Hochmuth, R.M. 1992. Analysis of lateral diffusion from a
spherical cell surface to a tubular projection. Biophys. J. 61: 1-8.
Bershadsky, A.D. and Gelfand, V.1. 1981. ATP-dependent regulation of cytoplasmic
73
microtubule disassembly. Proc. Nat. Acad. Sci. USA 78: 3610-3613.
Bird, R.B., Armstrong, R.C., and Hassager, O. 1987. Dynamics of Polymeric Liquids:
Volume 1. Fluid Mechanics. (2nd ed.) John Wiley, New York. (TA357 D95
1987 v.1)
Birkhoff, G. 1983. Numerical fluid dynamics. SIAM Review 25(1): 1-34. [John von
Neumann Lecture. Includes a masterly summary of analytical fluid dynamics.]
Bjerknes, M. 1986. Physical theory of the orientation of astral mitotic spindles.
Science 234: 1413-1416.
Black, M.M. and Smith, W.
1988.
Regional differentiation of the neuronal
cytoskeleton with an appendix: Diffusion of proteins in the neuron cell body-mathematical approximations and computer simulations. In R.J. Lasek and M.M.
Black (eds.) Intrinsic Determinants of Neuronal Form and Function. Alan R.
Liss, New York, pp. 463-486. (UNC-Greensboro QP 361 M4 1988)
Blake, J.R. and Sleigh, M.A. 1974. Mechanics of ciliary locomotion. BioI. Rev. 49:
85-125. (QH 301 C26) [IMPORTANT REVIEW.]
Blake, J.R. 1971. A note on the image system for a Stokeslet in a no-slip boundary.
Proc. Camb. Philos. Soc. 70: 303-310.
Blake, J.R. 1974. Singularities of viscous flow. Part II. Applications to slender-body
theory. J. Eng. Math. 8: 113-124.
Blick, H. 1972. Ciliated Protozoa. World Health Organization, Geneva.
Bloodgood, R.A. 1978. Unidirectional motility occurring in association with the
axopodial membrane of Echinosphaerium nucleofilum. Cell BioI. Int. Rep. 2:
171-176.
Bloodgood, R.A. 1987. Glycoprotein dynamics in the Chlamydomonas flagellar
membrane. Adv. Cell BioI. 1: 97-130.
Bloom, M., Evans, E., and Mouritsen, O.G. 1991. Physical properties of the fluid
lipid-bilayer component of cell membranes: A perspective. Quarterly Reviews of
Biophysics 24(3): 293-397. [Note that this article does not treat properties of the
cell cortex.]
Blum, J.J. and Hines, M. 1979. Biophysics of flagellar motility. Quarterly Review of
Biophysics 12: 103-180. [Probably dated.]
Blum, J.J., Lawler, G., Reed, M., and Shin, 1. 1989. Effect of cytoskeletal geometry
on intracellular diffusion. Biophys. J. 56: 995-1005.
Bovee, E.C. and Sawyer, T.K. 1971. Feeding on gastrotrichs by the heliozoan
Actinophrys sol. Trans. Am. Microsc. Soc. 90: 365-369. (QH 201 A3) [No
observations of fast axopodial contraction.]
Bovee, E.C. 1991. Sarcodina. In F.W. Harrison and J.O. Corliss (eds.) Microscopic
Anatomy of Invertebrates. Volume 1. Protozoa. Wiley-Liss, New York, pp. 161259. (Vet. Med. R QL 363 M53 1991 v.1) [Nice EM images and convenient
74
organization, but poorly written.
arrays (p. 38), etc.]
See matter
In
this volume on axostyle MT
Bowser, S.S., DeLaca, T.E., and Rieder, C.L. 1986. Novel extracellular matrix and
microtubule cables associated with pseudopodia of Astrammina rara, a
carnivorous Antarctic foraminifer. J. Ultrastruct. Mol. Struct. Res. 94: 149-160.
Bowser, S.S., Alexander, S.P., Stockton, W.L., and DeLaca, T.E. 1992. Extracellular
matrix augments mechanical properties of pseudopodia in the carnivorous
foraminiferan Astrammina rara: Role in prey capture. J. Protozool. 39(6): 724732. (November-December) [VALUABLE LIST OF REFS.]
Brandt, K. 1878. Uber die Axenfaden der Heliozoen und die Bewegungen von
Actinosphaerium. Sber. Ges. naturf. Freunde Bed., pp. 171-178.
Bray, D. 1992. Cell Movements. Garland, New York. (QH 647 B73)
Brennen, C. and Winet, H. 1977. Fluid mechanics of propulsion by cilia and flagella.
Annual Review of Fluid Mechanics 9: 339-398. (QC145 A57 v. 9 1977) [VERY
IMPORTANT. See esp. Table 2, p. 350 ("Wall effects on resistive coefficients for
the translation of straight slender cylinders") and the accompanying discussion on
pp. 349-351.]
Briere, C. and Goodwin, B.C. 1990. Effects of calcium input/output on the stability
of a system for calcium-regulated viscoelastic strain fields. J. Math. BioI. 28:
585-593.
Brodland, G.W. and Gordon, R. 1990. Intermediate filaments may prevent buckling
of compressively loaded microtubules. Trans. ASME J. Biomech. Eng. 112: 319321.
Brokaw, C.J. 1988. Bending wave propagation by microtubules and flagella. Math.
Biosci. 90: 247-263.
Brokaw, C.J.
1989.
Direct measurements of sliding between outer doublet
microtubules in swimming sperm flagella. Science 243: 1593-1596.
Brown, P.A. and Berlin, R.D. 1988. Factors that determine the spacing between
microtubules: Their possible role in cytoarchitecture. In R.J. Lasek and M.M.
Black (eds.) Intrinsic Determinants of Neuronal Form and Function. Alan R.
Liss, New York, pp. 307-321. (UNC-Greensboro QP 361 M4 1988)
Brown, R.C. 1994. The pore size distribution of model filters produced by random
fragmentation described in terms of the Weibull distribution. Chern. Eng. Sci.
49(1): 145-146.
Brugerolle, G. and Mignot, J.-P.
1983.
Caracteristiques ultrastructurales de
l'helioflagelle' Tetradimorpha (Hsiung) et leur interet pour l'etude phyletique des
heliozoaires. J. Protozool. 30: 473-480. [Tetradimorpha has a somewhat
irregular axonemal MT pattern containing triads. Fig. 4, pA75 is a longitudinal
section of the cell body which shows (spectacularly) the axoneme of a posterior
axopodium coursing through the cytoplasm. This axoneme actually arises from a
centroplast situated at the anterior end.]
75
Brugerolle, G. and Mignot, J .-P. 1984a. The cell characteristics of two helioflagellates
related to the centrohelidian lineage: Dimorpha and Tetradimorpha. Origins of
Life 13: 305-314.
Brugerolle, G. and Mignot, J.-P. 1984b. Les caracteristiques ultrastructurales de
l'helioflagelle' Dimorpha mutans Gruber (Sarcodina, Actinopoda) et leur interet
phyletique. Protistologica 20: 97-112. [Important paper; good EM images of
quincuncial arrangement of axonemal MTs.]
1985.
Ultrastructure d' He driocystis pellucida
(Heliozoa:
Brugerolle, G.
Desmothoracida) et de sa forme migratrice flagellee. Protistologica 21: 259-265.
Burns, R.G. 1981. Microtubules and mitosis. In P.C.L. John (ed.) The Cell Cycle.
Cambridge University Press, New York.
1987.
Stimulus-response coupling in the
Butler, R.D. and McCrohan, C.R.
contraction of tentacles of the suctorian protozoan Heliophrya erhardi. J. Cell Sci.
88: 121-127.
Buxbaum, R.E., Dennerll, T., Weiss, S., and Heidemann, S.R. 1987. F-actin and
microtubule suspensions as indeterminate fluids. Science 235: 1511-1514.
Buxbaum, R.E. and Heidemann, S.R. 1988. A thermodynamic model for force
integration and microtubule assembly during axonal elongation. J. theor. BioI.
134: 379-390.
Buxbaum, R.E. and Heidemann, S.R. 1992. An absolute rate theory model for
tension control of axonal elongation.
[Preprint dated 7/91, accepted for
publication in J. theor. Biol.] J. theor. BioI. 155: 409-426.
Cachon, J., Cachon, M., Febvre-Chevalier, C., and Febvre, J. 1973. Determinisme de
l'edification des systemes microtubulaires stereoplasmiques d'Actinopodes. Arch.
Protistenk. 115: 137-153.
Cachon, J. and Cachon, M. 1974. Les systemes axopodiaux. Annee Biologie 13:
523-560. (QH301 A486) [Includes comparative treatment of linkage schemes for
Echinosphaerium and critique of Harris (1970b).]
Cachon, J. and Cachon, M.
1975.
Role des microtubules dans les courants
cytoplasmiques des axopodes. C.R. Acad. Sci. Paris 280: 2341-2343.
Cachon, J., Cachon, M., Tilney, L.G., and Tilney, M.S. 1977. Movement generated
by interactions between the dense material at the ends of microtubules and nonactin-containing microfilaments in Sticholonche zanclea. J. Cell BioI. 72: 314338.
Cachon, J. and Cachon, M. 1978. Sticholonche zanclea Hertwig. A reinterpretation of
its phylogenetic position based upon new observations on its ultrastructure.
Archiv fur Protistenkunde 120: 148-168.
Cachon, J. and Cachon, M. 1981a.
BioSystems 14: 313-326.
Cachon, J. and Cachon, M.
1981b.
Movement by non-actin filament mechanisms.
Polymorphism of tubulin reassembly in the
76
microtubular system of a heliozoan. II. Cytochemical effects of mechanical
shocks and ultrasounds. BioI. Cell. 40: 23-32. (QH 201 J55)
Cachon, J., Cachon, M., and Bruneton, J.-N. 1981c. An ultrastructural study of the
effects of very high frequency ultrasounds on a microtubular system (axopods of a
heliozoan). BioI. Cell. 40: 69-72. (QH 201 J55)
Caldwell, D.G. and Taylor, P.M. 1990. Chemically stimulated pseudo-muscular
actuation. Int. J. Engng. Sci. 28(8): 797-808. (TAl 173) [Protozoa mentioned
in the Introduction.]
Caplow, M. 1992. Microtubule dynamics. Current Opinion in Cell Biology 4: 58-65.
Caplow, M., Shanks, J., Breidenbach, S., and Ruhlen, RL. 1988. Kinetics and
mechanisms of microtubule length changes by dynamic instability. J. BioI. Chem.
263: 10943-10951.
Capriulo, G.M. 1990. Feeding-related ecology of marine protozoa. In G.M. Capriulo
(ed.) Ecology of Marine Protozoa. Oxford University Press, New York, pp. 186259. [See esp. pp. 202-205 on heliozoa and pp. 233-235 on ciliates, incI. suetoria.]
(QL362 E29 1990)
Caron, D.A. and Goldman, J.C. 1990. Protozoan nutrient regeneration. In G.M.
Capriulo (ed.) Ecology of Marine Protozoa. Oxford University Press, New York,
pp. 283-306. (QL362 E29 1990)
Caron, D.A., Goldman, J.C., and Fenchel, T. 1990. Protozoan respiration and
metabolism. In G.M. Capriulo (ed.) Ecology of Marine Protozoa.' Oxford
University Press, pp. 307-322. (QL362 E29 1990)
Caspar, D.L.D. 1980. Movement and self-control in protein assemblies. Biophys. J.
32: 103-135. [Excellent but chiefly treats viruses. See description of T4 tail
sheath and the discussion following the paper.]
Cassimeris, L., Pryer, N.K., and Salmon, E.D. 1988. Real-time observations of
microtubule dynamic instability in living cells. J. Cell BioI. 107: 2223-2231.
Cheer, A., Vincent, J.-P., Nuticelli, R., and Oster, G. 1987. Cortical activity in
vertebrate eggs I: The activation waves. J. theor. BioI. 124: 377-404.
Cheer, A. and Koehl, M.A.R. 1987. Paddles and rakes: Fluid flow through bristled
appendages of small organisms. J. theor. BioI. 129: 17-39.
Cheer, A. and Koehl, M.A.R. 1987. Fluid flow through filtering appendages of
insects. IMA Journal of Mathematics Applied in Medicine and Biology 4: 185199.
Chen, J., Kanai, Y., Cowan, N.J., and Hirokawa, N. 1992. Projection domains of
MAP2 and tau determine spacings between microtubules in dendrites and axons.
Nature 360: 674-677. (17 December 1992)
Chen, Y.-T. and Schliwa, M. 1990. Direct observation of microtubule dynamics in
Reticulomyxa: Unusually rapid length changes. and microtubule sliding. Cell
Motility and the Cytoskeleton 17: 214-226. [Important discussions of tubulin
77
cone. dependence and rates of
discussion/speculation, pp. 224-225.]
MT
extension
on
p.
221;
useful
Cheng, L.Y. 1987a. Deformation analysis in cell and developmental biology. Part I:
Formal methodology. Trans. ASME J. Biomech. Eng. 109: 10-17.
Cheng, L.Y. 1987b. Deformation analysis in cell and developmental biology. Part II:
Mechanical experiments on cells. Trans. ASME J. Biomech. Eng. 109: 18-24.
[Table 1, p.19 sets forth typical values of mech. props. of the unfertilized sea
urchin egg.]
Childress, S. 1977. Mechanics of Swimming and Flying.
Mathematical Sciences, New York University.
Courant Institute of
Childress, S. 1981. Mechanics of Swimming and Flying. Cambridge University Press,
Cambridge. (Natural Resources Library QP 310 .F5 C48 1981) [A slightly
revised and augmented version of Childress (1977).]
Chretien, D., Metoz, F., Verde, F., Karsenti, E., and Wade, R.H. 1992. Lattice
defects in microtubules:
Protofilament numbers vary within individual
microtubules. J. Cell BioI. 117(5): 1031-1040.
Clough, R.W. and Penzien, J.
York. (TA654 C6 cop.3)
1975.
Dynamics of Structures.
Condeelis, J. 1992. Are all pseudopods created equal?
Cytoskeleton 22: 1-6.
McGraw-Hill, New
Cell Motility and the
Cordova, N.J., Vale, R.D., Oster, G.F. 1991. Dynein-microtubule interactions. In L.
Peliti (ed.) Biologically Inspired Physics. Plenum Press, NY, pp. 207-215.
Cordova, N.J., Ermentrout, B., and Oster, G.F. 1992. Dynamics of single-motor
molecules: The thermal ratchet model. Proc. Natl. Acad. Sci. USA 89: 339-343.
Correia, J.J. and Williams, RC. 1983. Mechanisms of assembly and disassembly of
microtubules. Ann. Rev. Biophys. Bioeng. 12: 211-235.
Cortese, J.D., Schwab, B., Frieden, C., and Elson, E.L. 1989. Actin polymerization
induces a shape change in actin-containing vesicles. Proc. Natl. Acad. Sci. USA
86: 5773-5777.
Coue, M., Lombillo, V.A., and McIntosh, J.R 1991. Microtubule depolymerization
promotes particle and chromosome movement in vitro. J. Cell BioI. 112: 11651175.
Cox, RG. 1970. The motion of lone; slender bodies in a viscous fluid.
General theory. J. Fluid Mech. 44(4): 791-810.
Part 1.
Crank, J. 1975. The Mathematics of Diffusion. (2nd ed.) Clarendon Press, Oxford.
Crank, J. 1976. Diffusion mathematics in medicine and biology.
Institute of Mathematics and its Applications 12: 106-112.
Cummings, F.W.
1985.
Bulletin of the
A pattern-surface interactive model of morphogenesis.
J.
78
theor. BioI. 116: 243-273. [Speculation (p. 266) on the prevalence of cylinders in
biology.]
Curry, A. and Butler, R.D. 1976. The ultrastructure, function, and morphogenesis of
the tentacle in Discophrya sp. (Suctorida) Cileatea. J. Ultrastruct. Res. 56: 164176.
Daily, B., Elson, E., and Zahalak, G.!. 1984. Cell poking. Determination of the elastic
area compressibility modulus of the erythrocyte membrane. Biophys. J. 45: 671682.
Dall, D.J. 1983. A theory for the mechanisms of polar filament extrusion in the
microspora. J. theor. BioI. 105: 647-659. [Nonmathematical.]
Davidson, L.A. 1972. Ultrastructure of the heliozoan Heterophrys marina and of the
helioflagellate Ciliophrys marina. J. Ultrastruct. Res. 38: 219. [Abstract]
Davidson, L.A. 1973. Contractile axopodia of the centrohelidan heliozoan Heterophrys
marina. J. Cell BioI. 59: 71a (Abstract 141). [Abstract of talk given at the 13th
annual meeting of the American Society for Cell Biology.]
Davidson, L.A. 1975. Studies of the actinopods Heterophrys marina and Ciliophrys
marina: Energetic and structural analysis of their contractile axopodia. Ph.D.
thesis, University of California, Berkeley. [Summary in Dissertation Abstracts
International, Vol. 37, July 1976, (114-B)-(115-B).]
Davidson, L.A. 1976. Ultrastructure of the membrane attachment sites of the
extrusomes of Ciliophrys marina and Heterophrys marina (Actinopoda). Cell
Tissue Res. 170: 353-365.
Davidson, L.A. 1982. Ultrastructure, behavior, and algal flagellate affinities of the
helioflagellate Ciliophrys marina, and the classification of the helioflagellates
(Protista, Actinopoda, Heliozoea). J. Protozool. 29: 19-29.
Dawson, R.M.C., Elliott, D.C., Elliott, W.H., and Jones, K.M. (eds.) 1969. Data for
Biochemical Research. Clarendon, Oxford. [See Vet. School Library for 1986 ed.]
de Duve, C. 1984. A Guided Tour of the Living Cell. Volume One. Scientific
American Books, New York. (QH581.2 D43 1984 v.1) [pp. 210-211 give de
Duve's enthusiastic, illustrated appreciation of the Heliozoa]
Deiner, M., Tamm, S.L., and Tamm, S. 1993. Mechanical properties of ciliary
axonemes and membranes as shown by paddle cilia. J. Cell Sci. 104: 1251-1262.
De Loof, A. 1986. The electrical dimension of cells: The cell as a miniature
electrophoresis chamber. International Review of Cytology. 104: 251-352. [An
excellent general review which devotes relatively little space to the subtitle.]
Dembo, M., Harlow, F., and Alt, W. 1984. The biophysics of cell surface motility. In
A.S. Perelson, C. DeLisi, and F.W. Wiegel (eds.)
Cell Surface Dynamics:
Concepts and Models. Marcel Dekker, New York, pp. 495-541.
Dembo, M. 1986. The mechanics of motility in
50: 1165-1183.
dis~ociated
cytoplasm. Biophys. J.
79
Dembo, M. and Harlow, F. 1986. Cell motion, contractile networks, and the physics
of interpenetrating reactive flow. Biophys. J. 50: 109-121.
Dembo, M., Maltrud, M., and Harlow, F. 1986.
contractile networks. Biophys. J. 50: 123-137.
Numerical studies of unreactive
Dembo, M., Torney, D.C., Saxman, K., and Hammer, D. 1988. The reaction-limited
kinetics of membrane-to-surface adhesion and detachment. Proc. Roy. Soc. Lond.
B 234: 55-83.
Dembo, M. 1989. Mechanics and control of the cytoskeleton in Amoeba proteus.
Biophys. J. 55: 1053-1080.
Dembo, M. 1989. Field theories of the cytoplasm. Comments on Theoretical Biology
1(3): 159-177.
de Mestre, N.J. 1973. Low Reynolds number fall of slender cylinders near boundaries.
J. Fluid Mech. 58: 641-656.
de Mestre, N.J. and Russel, W.B. 1975. Low Reynolds number translation of a
slender cylinder near a plane wall. J. Eng. Math. 9: 81-91. [Very important
paper--includes graphs quantifying wall effect. Brennen and Winet (1977, p. 351)
remark that there are typographical errors in some of the formulae in this paper.]
Dentler, W.L.
1981.
Microtubule-membrane interactions in cilia and flagella.
International Review of Cytology 72: 1-47.
Dimitrov, D.S. and Jain, R.K. 1984. Membrane stability. Biochim. et Biophys. Acta
779: 437-468. [Roughly written review but contains many data; Marangoni
instability mentioned on pp. 451-452.]
Dobell, C. 1917. On Oxnerella maritima, nov. gen., nov. spec., a new heliozoan and
its method of division; with some remarks on the centroplast of the Heliozoa.
Quart. J. Micr. Sci. Ser. 2 62: 515-538.
Dobell, C. 1932. Antony van Leeuwenhoek and his "Little Animals". Harcourt, Brace
and Co., New York. (QH31 L55D6) [On p.381, Dobell refers to Johann Conrad
Eichhorn (1718-1790), pastor of the Church of St. Catherine at Danzig. Dobell
also gives two references to Eichhorn's' protozoological researches in the
bibliography.]
Dolowy, K. 1984. Bioelectrochemistry of cell surfaces. Prog. Surf. Sci. 15: 245-368.
(QD506 .P76)
Dong, C. and Skalak, R. 1992. Leukocyte deformability: Finite element modeling of
large viscoelastic deformation. J. theor. BioI. 158: 173-193. (21 September)
Doyle, J.F. 1989. Wave Propagation in Structures. An FFT-Based Spectral Analysis
Methodology. Springer-Verlag, New York. (QA935 D693 1989)
Dragesco, J. 1964. Capture et ingestion des proies chez Actinosphaerium eichhorni
(Rhizopoda, Heliozoa). Arch. Zool. Exp. Gen. 10:1: 163-175. (QL1 A72)
80
Drexler, K.E.
1992.
Nanosystems:
Molecular Machinery, Manufacturing, and
Computation. John Wiley, New York. (T174.7 D74 1992)
Dullien, F.A.L. 1992. Porous Media: Fluid Transport and Pore Structure. 2nd ed.
Academic Press. (QC 175.2 D84)
Durrschmidt, M. and Patterson, D.J. 1987. On the organization of the Heliozoa
Raphidiophrys ambigua Penard and R. pallida Schulze. Annales des Sciences
Naturelles Zoologie et Biologie Animale 8: 135-155. (QH3 A62) [Includes data
on axonemal MTs, cell size, interpretive drawing.]
Dustin, P. 1978. Microtubules. 1st ed. Springer-Verlag, Berlin. (QH 603 M44 D87
1978)
Dustin, P. 1984. Microtubules.
D87 1984)
2nd ed.
Springer-Verlag, Berlin. (QH 603 M44
Dyson, F.J. 1966. The search for extraterrestrial technology. In R.E. Marshak (ed.)
Perspectives in Modern Physics: Essays in Honor of Hans A. Bethe. John Wiley,
New York, pp. 641-655. (QC774 B4M3) [See esp. pp. 646-647 on the construction
of light rigid structures.]
Easwaran, C. V. and Kokal, S.L. 1992. A theory for transient discharge in tubes with
applications to a new type of viscometer. SIAM J. AppI. Math. 52(6): 15011508. (December 1992)
Eckert, R. and Naitoh, Y. 1970. Passive electrical properties of Paramecium and
problems of ciliary coordination. J. Gen. PhysioI. 55: 467-483.
Edds, K.T. 1975a. Motility in Echinosphaerium nucleofilum. 1. An analysis of
particle motions in the axopodia and a direct test of the involvement of the
66:
145-155.
[Fig. 1, p.147, is an excellent
axoneme.
J. Cell BioI.
photomicrograph showing the irregular tapering profile of a single axopodium and
the curvature of the cell body surface in vivo. The photomicrograph includes a 10
pm scale bar. According to the "Materials and Methods", the optical setup
produced "an optical section thickness considerably below 1 pm". The article
concentrates on intraaxopodial motions.]
Edds, K.T. 1975b. Motility in Echinosphaerium nucleofilum
II. Cytoplasmic
contractility and its molecular basis. J. Cell BioI. 66: 156-164. [Reports the
discovery of actin (identified by HMM decoration, double-helicity, and diameter)
in the cytoplasm. Edds does not say that he was able to find actin in the
axopodiaj
his experimental technique would seem to preclude accurate
localization.]
Edds, K.T. 1981. Cytoplasmic streaming in a heliozoan. BioSystems 14: 371-376.
Edgar, L.A. 1982. Diatom locomotion: A consideration of movement in a highly
viscous situation. Br. PhycoI. J. 17: 243-251.
Edwards, D.A., Brenner, H., and Wasan, D.T. 1991. Interfacial Transport Processes
and Rheology. Butterworth-Heinemann, Boston.. (QD 506 E38)
81
Elson, E.L. 1988. Cellular mechanics as an indicator of cytoskeletal structure and
function. Annual Review of Biophysics and Biophysical Chemistry 17: 397-430.
Engelhardt, H., Schuster, S.C., and Baeuerlein, E. 1993. An Archimedian spiral: The
basal disk of the Wolinella flagellar motor. Science 262: 1046-1048. (November
12) [Heliozoa axonemes mentioned in transitu on p.1047.]
Engelmann, T.W. 1875. Contractilitat und Doppelbrechung. Pfliigers Archiv fur die
gesammte Physiologie des Menschen und der Thiere 11: 432-464.
Erickson, H.P. and O'Brien, E.T. 1992. Microtubule dynamic instability and GTP
hydrolysis. Annu. Rev. Biophys. Biomoi. Struct. 21: 145-166. (QH505 A1A55
v.21 1992)
Etienne, E.M. 1970. Control of contractility of Spirostomum by dissociated calcium
ions. J. Gen. Physioi. 56: 168-179. [Cited by Huang and Pitelka (1973), p. 707.
Supposedly Etienne used aequorin (microinjected) to demonstrate the role of Ca2+
in the onset of contraction.]
1981.
Polarity of some motility-related
Euteneuer, U. and McIntosh, J.R.
microtubules. Proc. Nati. Acad. Sci. USA 78(1): 372-376.
Evans, E.A. 1986. Membrane mechanics and cell adhesion. In G.W. SchmidSchonbein, S.L-Y. Woo, and B.W. Zweifach (eds.) Frontiers in Biomechanics.
Springer-Verlag, New York, pp. 3-17. (QP303 F69 1986)
Evans, E.A. 1993. New physical concepts for cell amoeboid motion. Biophys. J. 64:
1306-1322.
Evans, R.L., McCrohan, C.R., and Butler, R.D. 1988. Tentacle contraction in
Heliophrya erhardi (Suctoria): The role of inositol phospholipid metabolites and
cyclic nucleotides in stimulus-response coupling. Exp. Cell Res. 177: 382-390.
Febvre, J. 1981. The myoneme of the Acantharia (Protozoa): A new model of
cellular motility. BioSystems 14: 327-336. (QH 301 C87) [See p. 330 for graph
of time course of contraction.]
Febvre-Chevalier, C.
1973a.
Un nouveau type d'association des microtubules
axopodiaux chez les heliozoaires. Protistologica 9: 35-43.
[A study of an
undescribed marine centrohelidian (presumably Aetinocoryne contractilis) with
nice EM pictures, axoneme diagrams with angles clearly labeled, and a helpful
chart. By Markham rotation (Fig. 10, p. 39), the author purports to show that
the axonemal MTs have only 10 protofilaments.]
Febvre-Chevalier, C.
1973b.
Hedraiophrys hovassei, morphologie, biologie et
cytologie. Protistologica 9: 503-520.
Febvre-Chevalier, C. 1975. Etude cytologique de Gymnosphaera albida Sassaki, 1894
(Heliozoaire Centrohelidie'). Protistologica 11: 331-344. [See Jones (1976).]
Febvre-Chevalier, C. 1980. Behaviour and cytology of Aetinocoryne contractilis, nov.
gen., nov. sp., a new stalked heliozoan (Centrohelida): Comparison with the other
related genera. J. mar. BioI. Assn. UK 60: 909-9.28.
82
Febvre-Chevalier, C. 1981. Preliminary study of the motility processes in the stalked
heliozoan Actinocoryne contractilis. BioSystems 14: 337-343.
Febvre-Chevalier, C. and Febvre, J. 1982. Locomotion processes in some marine
pela~ic and benthic protozoa.
Annales de l'Institut Oceanographique, Paris
58(S): 137-142.
Febvre-Chevalier, C. and Febvre, J.
1984. Axonemal microtubule pattern of
Cienkowskya mereschkovskyi and a revision of heliozoan taxonomy. Origins of Life
13: 315-338.
Febvre-Chevalier, C. 1985. IV. Class Heliozoea Haeckel 1866. In J.J. Lee, S.H.
Hutner, and E.C. Bovee (eds.) An fllustrated Guide to the Protozoa. Society of
Protozoologists. Lawrence, Kansas, pp. 302-317. (QL366 14 1985)
Febvre-Chevalier, C. and Febvre, J. 1986. Motility mechanisms in the actinopods
(Protozoa):
A
review
with
particular
attention
to
axopodial
contraction/extension, and movement of nonactin filament systems. Cell Motility
and the Cytoskeleton 6: 198-208.
Febvre-Chevalier, C., Bilbaut, A., Bone, Q., Febvre, J. 1986. Sodium-calcium action
potential associated with contraction in the heliozoan Actinocoryne contractilis.
J. expo Bio!. 122: 177-192.
Febvre-Chevalier, C., Bilbaut, A., Febvre, J., and Bone, Q. 1989. Membrane
excitability and motile responses in the protozoa, with particular attention to the
heliozoan Actinocoryne contractilis. In P.A.V. Anderson (ed.) Evolution of the
First Nervous Systems. Plenum Press, New York, pp. 237-253. (QP356.15 N38
1989)
1992. Microtubule disassembly in vivo:
Febvre-Chevalier, C. and Febvre, J.
Intercalary destabilization and breakdown of microtubules in the heliozoan
Actinocoryne contractilis. J. Cell Bio!. 118(3): 585-594.
Fenchel, T. 1984. Suspended marine bacteria as a food source. In M.J.R. Fasham
(ed.) Flows of Energy and Materials in Marine Ecosystems: Theory and
Practice. Plenum Press, New York and London, pp. 301-315. (Nat. Resources
Library QH541.5 S3 F57 1984) [See esp. pp. 305, 307, 308-309 for quantitative
discussion of bacterivory in heliozoa.]
Fenchel, T. 1987. Ecology of Protozoa: The Biology of Free-Living Phagotrophic
Protists. Science Tech Publishers, Madison, Wisconsin. (NRL QL366 F46 1987)
Fitzharris, T;, Bloodgood, R.A., and McIntosh, J.R. 1972. Particle movement in the
axopodia of Echinosphaerium: Evidence concerning the role of the axoneme. J.
Mechanochem. Cell Motility 1: 117-124.
Fowler, P.W. 1992.
December 1992)
Vocabulary for fuzzy symmetry.
Nature
360:
626.
(17
Freifelder, D. 1982. Physical Biochemistry. (2nd ed.) W.H. Freeman, San Francisco.
[See esp. pp. 57-62 on polarization microscopy and form
(QH 345 F72)
birefringence. The Vet. School Library has a more recent ed.]
83
Freund, L.B. 1990. Dynamic Fracture Mechanics. Cambridge University Press,
Cambridge. (TA409 F77 1990--8th floor, D.H. Hill)
Friedel, J. 1964. Dislocations. Pergamon Press, Oxford. (QD945 F713 1964)
Fujiwara, K. and Tilney, L.G. 1975. Substructural analysis of the microtubule and its
p'olymorphic forms. Annals of the New York Academy of Sciences 253: 27-50.
lSee esp. p. 35 and pp. 38-39 for matter on Echinosphaerium, but heliozoans are
treated throughout this article. Tilney et al. (1973) deals with some of the same
topics.]
Fukui, Y. 1993. Toward a new concept of cell motility: Cytoskeletal dynamics in
amoeboid movement and cell division. International Review of Cytology 144:
85-127.
Fung, Y.C. 1990. Biomechanics: Motion, Flow, Stress, and Growth.
Verlag, New York. [See esp. pp. 132-140] (QP 303 F86)
Springer-
Gaines, S.D. and Denny, M.W. 1993. The largest, smallest, highest, lowest, longest,
and shortest: Extremes in ecology. Ecology 74(6): 1677-1692.
Gal, V., Martin, S., and Bayley, P. 1988. Fast disassembly of microtubules induced
by M~2+ or Ca2+. Biochem. Biophys. Res. Comm. 155(3): 1464-1470. (QH 301
.B355) [IMPORTANT]
Gallez, D. and Coakely, W.T. 1986. Interfacial instability at cell membranes. Prog.
Biophys. Molec. BioI. 48: 155-199. (QH 505 Al P76) [See p. 161 for matter on
breakdown ofaxopodia under high pressure. Article contains many data, matter
on cell adhesion, phagocytosis, etc.]
Galloway, J.W. 1991. Macromolecular asymmetry. In G.R. Bock and J. Marsh
(eds.) Biological Asymmetry and Handedness. (Ciba Foundation Symposium
162.) John Wiley and Sons, Chichester, pp. 16-29 [discussion on pp. 29-35]. (QP
517 S83 B55 1991)
Gelfand, V.I. and Bershadsky, A.D. 1991. Microtubule dynamics: Mechanism,
regulation, and function. Annual Review of Cell Biology 7: 93-116. (QH 573
A54 v.7 1991)
Gershon, N.D., Porter, K.R., and Trus, B.L. 1985. The cytoplasmic matrix: Its
volume and surface area and the diffusion of molecules through it. Proc. Nati.
Acad. Sci. USA 82: 5030-5034.
Gilman, J.J. 1969.
(QC191 G5)
Micromechanics of Flow in Solids.
McGraw-Hill, New York.
Gittes, F., Mickey, B., Nettleton, J., and Howard, J. 1993. Flexural rigidity of
microtubules and actin filaments measured from thermal fluctuations in shape. J.
Cell BioI. 120(4): 923-934.
Godart, H. and Huitorel, P. 1992. Effects of calcium on the longitudinal flagellum of
Oxyrrhis marina. BioI. Cell. 76: 365-372.
Goodwin, B.C.
1989.
Unicellular morphogenesis.
in
w.n.
Stein and F. Bronner
84
(eds.) Cell Shape: Determinants, Regulation, and Regulatory Role. Academic
Press, San Diego, pp. 365-391. (QH611 C425) [Mostly deals with ciliates and
Acetabularia. The Heliozoa are mentioned and figured on p. 367, reference being
made to the spherical cell body symmetry, not the axonemal MTs.]
Goodwin, B.C. and Briere, C. 1991. Generic dynamics of morphogenesis. In E.
Mosekilde and L. Mosekilde (eds.) Complexity, Chaos, and Biological Evolution.
Plenum Press, New York, pp. 281-301.
(QH505
N333
1991)
[Treats
Acetabularia. The appendix is an exposition of the Goodwin-Trainor model of
Ca2+ - cytogel interaction.]
Gordon, R. and Brodland, G.W.
1987. The cytoskeletal mechanics of brain
Cell
morphogenesis:
Cell state splitters cause primary neural induction.
Biophysics 11: 177-238.
Gould, S.J. and Lewontin, R.C.
1979. The spandrels of San Marco and the
Panglossian paradigm: A critique of the adaptationist programme. Proc. R. Soc.
Lond. B 205: 581-598.
Gradshteyn, 1.S. and Ryzhik, 1.M.
Academic Press, New York.
1965.
Table of Integrals, Series, and Products.
Grain, J. 1986. The cytoskeleton of protists: Nature, structure, and functions.
International Review of Cytology 104: 153-250. [AN IMPORTANT REVIEW]
Granick, S. 1991. Motions and relaxations of confined liquids. Science 253: 13741379.
Grebecki, A. and Hausmann, K.
1992. Surface coat shedding and axopodial
movements in Actinophrys sol. Europ. J. ProtistoI. 28: 390-397. (November 20,
1992)
Grebecki, A. and Hausmann, K. 1993. Motor behaviour of prey during first steps of
food capture by Actinophrys sol. Europ. J. Protista!. 32: 157-164.
Green, D.P.L. and Purves, R.D. 1984. Mechanical hypothesis of sperm penetration.
Biophys. J. 45: 659-662.
Green, D.P.L. 1988. Sperm thrusts and the problem of penetration. BioI. Rev. 63:
79-105. (QH 301 C26)
Gregson, A.J., Green, J.C., and Leadbeater, B.S.C. 1993a. Structure and physiology
of the haptonema in Chrysochromulina (Prymnesiophyceae). 1. Fine structure of
the flagellar jhaptonematal root system in C. acantha and C. simplex. J. Phycol.
29: 674-686.
Gregson, A.J., Green, J.C., and Leadbeater, B.S.C. 1993b. Structure and physiology
of the haptonema in Chrysochromulina (Prymnesiophyceae). II. Mechanisms of
haptonematal coiling and the regeneration process. J. PhycoI. 29: 686-700.
Grell, K.G. and Meister, A. 1984a. Beitrage zur Ultrastruktur der Konjugation von
Ephelota gemmipara R. Hertwig (Suctoria) Protistologica 20(1): 65-86.
Grell, K.G. and Benwitz, G.
1984b.
Die Ultrastruktur von Ephelota gemmipara
85
Hertwig und E. plana Wailes (Suctoria): Ein Vergleich. I. Die Adulte Form.
Protistologica 20(2): 205-233. [THIS PAPER IS OF CHIEF INTEREST]
Grell, K.G. and Benwitz, G. 1984c. Die Ultrastruktur von Ephelota gemmipara
Hertwig und E. plana Wailes (Suctoria): Ein Vergleich. II. Der Schwarmer.
Protistologica 20(3): 437-461.
Gronenberg, W., Tautz, J., and Holldobler, B. 1993. Fast trap jaws and giant
neurons in the ant Odontomachus. Science 262: 561-563.
Grunbaum, B. and Shephard, G.C. 1987. Tilings and Patterns. W.H. Freeman, New
York. (QA 166.8 G78 1987)
Gueron, S. and Liron, N. 1992. Ciliary motion modeling and dynamic multicilia
interactions. Biophys. J. 63: 1045-1058. (October) [VERY IMPORTANT. See
esp. the remarks on pp. 1054-1055 concerning the range of influence. ALSO
NOTE THE VALUABLE LIST OF REFERENCES.]
Gunther, B. and Morgado, E. 1984. Transport similarity: Dimensional analysis of
diffusion at the cellular level. J. theor. BioI. 108: 437-449.
Hackney, C.M. and Butler, R.D. 1981. Electrically induced tentacle retraction in the
suctorian protozoan Discophrya collini (Root). J. Protozool. 28(2): 151-157.
[IMPORTANT PAPER]
Hackney, C.M. and Butler, R.D.
1981.
Tentacle contraction in glycerinated
Discophrya collini and the localization of HMM-binding filaments. J. Cell Sci 47:
65-75. ]
Hackney, C.M. and Butler, R.D. 1981. Distribution of calcium in the suctorian
Discorhrya collini: An X-ray microanalytical study. Tissue and Cell 13: 453459.
Hameroff, S.R.
1987.
Ultimate Computing:
Biomolecular Consciousness and
Elsevier.
[Contains discussion of axoneme formation in
Nanotechnology.
heliozoans.]
Hameroff, S.R., Rasmussen, S., and Mansson, B. 1988. Molecular automata in
microtubules: Basic computational logic of the living state? In C. Langton (ed.)
Artificial Life.
Santa Fe Institute Studies in the Sciences of Complexity.
Addison-Wesley, pp. 521-553.
Hameroff, S.R., Lahoz-Beltra, R., Samsonovich, A., Scott, A., and Rasmussen, S.
1991. Learning in the cytoskeleton? Models of back-propagation and synaptic
plasticity via microtubule/MAP2 phonon automata. [Preprint sent by Stuart
Hameroff, August 1991]
Hammer, D.A. and Lauffenburger, D.A. 1987. A dynamical model for receptormediated adhesion to surfaces. Biophys. J. 52: 475-487.
Happel, J. and Brenner, H. 1983. Low Reynolds Number Hydrodynamics With Special
Applications to Particulate Media. Martinus Nijhoff, The Hague. (Textiles
Library QA 911 .H35 1983) [A paperback reprin,t of the 1973 edition.]
8tl
Hara, R., Asai, H., and Naitoh, Y. 1985. Electrical responses of the carnivorous
ciliate Didinium nasutum in relation to discharge of the extrusive organelles. J.
expo BioI. 119: 211-224.
Hardt, S. 1978. Aspects of diffusional transport in microorganisms. In S.R. Caplan
and M. Ginzburg (eds.) Energetics and Structure of Halophilic Microorganisms.
Elsevier/North-Holland Biomedical Press, Amsterdam, pp. 591-597.
(UNCChapel Hill QR 97 .S3 E45 1978)
Hardt, S. 1981. The diffusion transit time: A simple derivation. Bull. Math. BioI.
43: 89-99.
Hargittai, 1. and Pickover. C.A. (eds.)
Singapore. (QA483 S68 1992)
1992.
Spiral Symmetry.
World Scientific,
Harold, F.M. 1986. The Vital Force: A Study of Bioenergetics. W.H. Freeman, New
York. (QH 510 H37 1986) [Possibly the best general introduction I have seen.]
Harris, J.E. 1961. The mechanics of ciliary movement. In J.A. Ramsay and V.B.
Wigglesworth (eds.) The Cell and the Organism. Cambridge University Press,
Cambridge, pp. 22-36. (QH581 R36) [Harris mentions haptonemata on p.33 and
p.35.]
Harris, W.F. 1970. Defects in surface crystals. Ph.D. thesis, University of Minnesota.
[Summary published in Dissertation Abstracts International, Vol. 31, Jan. 1971,
3872-B (Reference Collection R LB7 M52).]
Harris, W.F. 1970. The arrangement of the axonemal microtubules and links of
Echinosphaerium nucleofilum. J. Cell BioI. 46: 183-187.
Harris, W.F. and Scriven, L.E. 1970. Cylindrical crystals, contractile mechanism of
bacteriophages and the possible role of dislocations in contraction. J. theor. BioI.
27: 233-257.
Harris, W.F. and Scriven, L.E. 1971. Moving dislocations in cylindrical crystals cause
waves of bending. I. Kink angle and wave types. J. Mechanochem. Cell Motility
1: 33-40.
Harris, W.F.
1972a. Twisting and mechanochemical conversion of energy at
dislocations in cylindrical crystals. J. Mechanochem. Cell Motility 1: 147-149.
Harris, W.F. 1972b. Dislocations in tobacco mosaic virus.
[See esp. refs. to other papers by Harris.]
Nature 240: 294-297.
Harris, W.F. and Robison, W.G. Jr. 1973. Dislocations in microtubular bundles
within spermatozoa of the coccid insect Neosteingelia texana and evidence for slip.
Nature 246: 513-515.
Hausmann, K. 1978. Extrusive organelles in protists. International Review of
Cytology 52: 197-268. [See pp. 238-239 on heliozoans.] (QH 573 15 v.52 1978)
Hausmann, K. and Patterson, D.J. 1982. Pseudopod formation and membrane
production during prey capture by a heliozoan (fe~ding by Actinophrys, II.). Cell
Motility 2: 9-24.
87
Hausmann, K., Linnenbach, M., and Patterson, D.J. 1983. The effects of taxol on
microtubular arrays: In vivo effects on heliozoan axonemes. J. Ultrastructure
Res. 82: 212-220.
Hausmann, K. and Mulisch, M. 1987. Determination of body shape in protists by
cortical structures. In J. Bereiter-Hahn, O.R. Anderson, and W.-E. Reif (eds.)
Cytomechanics: The Mechanical Basis of Cell Form and Structure. SpringerVerlag, New York, pp. ?-?
Hawkes, RB. and Holberton, n.v. 1974. Myonemal contraction of Spirostomum. 1.
Kinetics of contraction and relaxation. J. Cell PhysioI. 84: 225-236. (QP1 J6)
Hawkes, R.B. and Holberton, n.v. 1975. Myonemal contraction of Spirostomum. II.
Some mechanical properties of the contractile apparatus. J. Cell PhysioI. 85(3):
595-602. (QP1 J6)
Hegstrom, RA. and Kondepudi, D.K.
Am. 262(1): 108-115.
1990. The handedness of the universe. Sci.
Heidemann, S.R. and Buxbaum, RE. 1990. Tension as a regulator and integrator of
axonal growth. Cell Motility and the Cytoskeleton 17: 6-10.
Herrmann, H. 1986. Asymmetry and symmetry in cellular organization. Computers
and Mathematics with Applications 12B(1/2): 155-167. (Q172.5 S95 S94 1986)
[An undistinguished article.]
Heuser, J., Reese, T.S., Dennis, M.J., Jan, Y., Jan, L., and Evans, L. 1979. Synaptic
vesicle exocytosis captured by quick freezing and correlated with quantal
transmitter release. J. Cell BioI. 81: 275-300.
Heuser, J.E. 1986. Different structural states of a microtubule cross-linking molecule,
captured by quick-freezing motile axostyles in protozoa. J. Cell BioI. 103: 22092227. (December) [Important paper]
Heuser, J.E., Zhu, Q., and Clarke, M. 1993. Proton pumps populate the contractile
vacuoles of Dictyostelium amoebae. J. Cell BioI. 121(6): 1311-1327. [See p. 1325
for remarks on the spasmoneme; also see pp. 1323-1324 for obs. that isolated
contractile vacuoles behave like simple osmometers.]
Higdon, J.J.L. 1979. The generation of feeding currents by flagellar motions. J. Fluid
Mech. 94: 305-330. [IMPORTANT PAPER--TREATS FEEDING BY SESSILE
STALKED FLAGELLATE]
Hill, A.V. 1948. On the time required for diffusion and its relation to processes in
muscle. Proc. Roy. Soc. Lond. B 135: 446-453.
Hill, T.L. 1987.
York.
Linear Aggregation Theory in Cell Biology. Springer-Verlag, New
Hille, B. 1992. Ionic Channels of Excitable Membranes.
Sunderland, Massachusetts.
(2nd ed.)
Sinauer,
Hillman, D.E. 1988. Parameters of dendritic shape and substructure: Intrinsic and
88
extrinsic determination?
In R.J. Lasek and M.M. Black (eds.)
Intrinsic
Determinants of Neuronal Form and Function. Alan R. Liss, New York, pp. 83113. (UNC-Greensboro QP 361 M4 1988)
Hines, M. and Blum, J.J. 1978. Bend propagation in flagella. I. Derivation of
equations of motion and their simulation. Biophys. J. 23: 41-57. [See esp. pp.
47-49 (note square root dropped on RHS of eqn. (25)), pp. 50-51, and refs. to
papers by Lubliner. Article treats shear resistance to sliding.]
Hiram, Y. and Avinoam, N. 1984. On mechanical aspects of vesicular transport. J.
theor. BioI. 111: 91-113.
Hiramoto, Y. 1974. Mechanics of ciliary movement. In M.A. Sleigh (ed.) Cilia and
Flagella. Academic Press, New York, pp. 177-196. (QP 310 C5S53) [Many
suggestive back-of-the-envelope-style calculations.]
Hirokawa, N. and Yorifugi, H. 1986. Cytoskeletal architecture of reactivated crayfish
axons, with special reference to cross bridges among microtubules and between
microtubules and membrane organelles. Cell Motil. Cytoskel. 6: 458-468.
Hitchcock, G.L. 1982. A comparative study of the size-dependent organic composition
of marine diatoms and dinoflagellates. J. Plankton Res. 4(2): 363-377.
Hochmuth, R.M. 1990. Cell biomechanics:
Biomech. Eng. 112: 233-234.
A brief overview.
Trans. ASME J.
Hodgkin, A.L. and Keynes, R.D. 1957. Movements of labelled calcium in squid giant
axons. J. Physiol. 138: 253-281.
Hoffman, E.J. 1977. Detachment ofaxopodia from the heliozoan, Echinosphaerium
nucleofilum. Exp. Cell Res. 110: 75-78.
Hoffman, E.J., Price, D.A., and Krumpelman, J.A. 1983. Isolation ofaxopodial
axonemes from mass-cultured Echinosphaerium nucleofilum. J. Protozool. 30(2):
371-375.
.
Hohfeld, I., Otten, J., and Melkonian, M. 1988. Contractile eukaryotic flagella:
Centrin is involved. Protoplasma 147: 16-24. [AN IMPORTANT REVIEW
AND SOURCE OF REFERENCES.]
Holwill, M.E.J. 1974. Hydrodynamic aspects of ciliary and flagellar movement. In
M.A. Sleigh (ed.) Cilia and Flagella. Academic Press, New York, pp. 143-175.
HolwilL M.E.J. 1977. Low Reynolds number undulatory propulsion in organisms of
different sizes. In T.J. Pedley (ed.) Scale Effects in Animal Locomotion.
Academic Press, New York, pp. 233-242.
Hoogenraad, H.R. 1927. Beobachtungen iiber den Bau, die Lebensweise und die
Entwicklung von Hedriocystis pellucida Hertw. et Less. Arch. Protistenk. 58:
321-342.
Hotani, H. and Miyamoto, H. 1990. Dynamic features of microtubules as visualized
by dark-field microscopy. Adv. Biophys. 26: 135,-156.
89
Hotani, H., Lahoz-Beltra, R., Combs, B., Hameroff, S., and Rasmussen, S. 1991.
Microtubule dynamics, liposomes and artificial cells: In vitro observation and
cellular automata simulation of microtubule assembly/disassembly and membrane
morphogenesis. [Preprint sent by Stuart Hameroff, August 1991]
Hou, L., Lanni, F., and Luby-Phelps, K. 1990. Tracer diffusion in F-actin and Ficoll
mixtures: Toward a model for cytoplasm. Biophys. J. 58: 31-43.
Hovasse, R. 1965a. Ultrastructure des axopodes de I'Heliozoaire.
Paris 260: 61-99.
C.R. Acad. Sci.,
Hovasse, R. 1965b. Ultrastructure comparee' des axopodes chez deux heliozoaires des
genres Actinosphaerium et Raphidiophrys. Protistologica 1: 81-88.
Huang, B. and Pitelka, D.R. 1973. The contractile process in the ciliate Stentor
coeruleus. 1. The role of microtubules and filaments. J. Cell BioI. 57: 704-728.
[The authors remark (p.708) that "...the contractile capabilities of Stentor are not
altered when cells are bathed in Ca-free medium." The concluding section is a
worthwhile summary.]
Huang, B. and Mazia, D. 1975. Microtubules and filaments in ciliate contractility. In
S. Inoue and R.E. Stephens (eds.) Molecules and Cell Movement. Raven Press,
New York, pp. 389-409. (QH 647 M64) [IMPORTANT REVIEW ARTICLE]
Huang, B., Mengerson, A., and Lee, V.D. 1988. Molecular cloning of cDNA for
caltractin, a basal body-associated Ca- binding protein: Homology in its protein
sequence with calmodulin and the yeast CDC31 gene product. J. Cell BioI. 107:
133-140.
Hutchinson, G.E. 1967. A Treatise on Limnology. Volume II. Introduction to Lake
Biology and the Limnoplankton. John Wiley, New York. (QH98 H82 v.2) [See
esp. Chapter 20: "The Hydromechanics of the Plankton". On pp. 275-276, H.
epitomizes the views of R. Margalef (1957) on the effect of superficial electrostatic
charge on the sinking of plankton.]
Hutchinson, G.E. 1982. Life in air and water.
Discovery 16(1): 3-9. (QH1 D5)
(D.C. Marsh Lecture for 1981)
Hyams, J.S. and Stebbings, H. 1979. Microtubule associated cytoplasmic transport.
In K. Roberts and J.S. Hyams (eds.) Microtubules. Academic Press, London, pp.
487-530. (QH 603 M44 M52)
Ichikawa, M. and Chou, T. 1979. Analysis of deformation and contraction in
cylindrical crystals due to dispirations or dislocations. J. theor. BioI. 78: 129142.
Ikegawa, S., Ishida, H., Shigenaka, Y., and Masuyama, E. 1987. Cytoskeleton of
heliozoan axopodia in Echinosphaerium akamae. ZooI. Sci. 4: 993. [Abstract]
Ingber, D.E. and Folkman, J. 1989. Tension and compression as basic determinants
of cell form and function: Utilization of a cellular tensegrity mechanism. In
W.D. Stein and F. Bronner (eds.) Cell Shape: Determinants, Regulation, and
Regulatory Role. Academic Press, San Diego, pp..3-31.
90
Ishida, H. and Shigenaka, Y. 1988. Cell model contraction in the ciliate Spirostomum.
Cell Motility and the Cytoskeleton 9: 278-282.
Jack, J.J.B., Noble, D., and Tsien, R.W.
Cells. Clarendon Press, Oxford.
1983.
Electric Current Flow in Excitable
Janmey, P.A., Hvidt, S., Oster, G.F., Lamb, J., Stossel, T.P., and Hartwig, J.H. 1990.
Effect of ATP on actin filament stiffness. Nature 347: 95-99. [The conclusions
of this paper are gainsaid by Pollard et al. (1992).]
Janmey, P.A. 1991. Mechanical properties of cytoskeletal polymers. Current Opinion
in Cell Biology 3: 4-11.
Johnson, K.A. and Borisy, G. 1979. Thermodynamic analysis of microtubule selfassembly in vitro. J. Mol. BioI. 133: 199-216. [Many data and graphs; a
difficult paper in respect of sheer detail. The 6 S subunit to which the authors
refer is the tubulin dimer (see p. 216).]
Johnson, K.A. 1980. Thermodynamics of microtubule assembly.
443-445. [A poster summary.]
Biophys. J.
32:
Johnson, S.W. and Wetzel, J.P. 1990. Engineering, Construction, and Operations in
Space II: Volume 2. American Society of Civil Engineers, New York. (TL797 S6
1990 v.2)
Jones, A.R., Jahn, T.L., and Fonesca, J.R. 1970. Contraction of protoplasm. III.
Cinematographic analysis of the contraction of some heterotrichs. J. Cell.
Physiol. 75: 1-8. [Graphs of time course of contraction.]
Jones, A.R., Jahn, T.L., and Fonesca, J.R. 1970. Contraction of protoplasm. IV.
Cinematographic analysis of the contraction of some peritrichs. J. Cell. Physiol.
75: 9-20. [Data and views on the initiation and propagation of contraction in the
stalks of vorticellids. Also, matter on latency and refs. to Japanese work in
English. See esp. the authors' warning in the final paragraph.]
Jones, J.C.R. and Tucker, J.B. 1981. Microtubule-organizing centres and assembly of
the double-spiral microtubule pattern in certain heliozoan axonemes. J. Cell Sci.
50: 259-280. [Authors remark (p. 263) that axoneme spirals ccw when viewed
from tip-to-base. Contrast with Kitching and Craggs (1965) and Tilney et al.
(1973) on Echinosphaerium nucleofiIum.]
Jones, W.C. 1976. The ultrastructure of Gymnosphaera aibida Sassaki, a marine
axopodiate protozoon. Phil. Trans. Roy. Soc. Lond. B 275: 349-384.
Kajiwara, K. and Ross-Murphy, S.B. 1992. Synthetic gels on the move. Nature 355:
208-209. (January 16)
Kamimura, S. and Takahashi, K.
1981. Direct measurement of the force of
microtubule sliding in flagella. Nature 293: 566-568.
Kamimura, S. and Mandelkow, E. 1992. Tubulin protofilaments and kinesindependent motility. J. Cell BioI. 118(4): 865-875.
91
Kao, H.P., Abney, J.R., and Verkman, A.S. 1993. Determinants of the translational
mobility of a small solute in cell cytoplasm. J. Cell BioI. 120: 175-184. (Jan.)
Kargacin, G. and Fay, F.S. 1991. Ca2+ movement in smooth muscle cells studied
with one- and two-dimensional diffusion models. Biophys. J. 60: 1088-1100.
Karr, T.L., Kristofferson, D., and Purich, D.L. 1980. Mechanism of microtubule
depolymerization. Correlation of rapid induced disassembly experiments with a
kinetic model for endwise depolymerization. J. BioI. Chern. 255: 8560-8566.
[September 1980] (QP501 J7)
Katchalsky, A., Lifson, S., Michaeli, I., and Zwick, M.
1960.
Elementary
Size and Shape Changes
mechanochemical processes. In A. Wassermann (ed.)
of Contractile Polymers:
Conversion of Chemical into Mechanical Energy.
Proceedings of seminars held at University College, London) Pergamon Press,
xford, pp. 1-40. (Textiles Library QD471 S52 1957)
g
Katchalsky, A., Alexandrowicz, Z., and Kedem, O. 1966. Polyelectrolyte solutions.
In B.E. Conway and R.G. Barrads (eds.) Chemical Physics of Ionic Solutions.
John Wiley, New York, pp. 295-346.
Katoh, K. and Naitoh, Y. 1992. A mechanosensory mechanism for evoking cellular
contraction in Vorticella. J. expo BioI. 168: 253-267.
Katz, D.F., Blake, J.R., and Paveri-Fontana, S.L. 1975. On the movement of slender
bodies near plane boundaries at low Reynolds number. J. Fluid Mech. 72: 529540.
Katzir-Katchalsky, A. 1976. Biophysics and Other Topics: Selected Papers by
Aharon Katzir-Katchalsky. Academic Press, New York. (QH505 K35 1976)
Kawachi, M., Inouye, I., Maeda, 0., and Chihara, M. 1991. The haptonema as a
food-capturing
device:
Observations
on
Chrysochromulina
hirta
(Prymnesiophyceae). Phycologia 30(6): 563-573. [IMPORTANT P APER-MANY DATA; ALSO VALUABLE FOR REFS.]
Keith, C.H. and Farmer, M.A. 1993. Microtubule behavior in PC12 neurites:
Variable results obtained with photobleach technology. Cell Motility and the
Cytoskeleton 25: 345-357.
Kerst, A., Chmielewski, C., Livesay, C., Buxbaum, R.E., and Heidemann, S.R. 1990.
Liquid crystal domains and thixotropy of filamentous actin suspensions. Proc.
Nat. Acad. Sci. USA 87: 4241-4245. (June)
Kim, S. and Karrila, S.J. 1991. Microhydrodynamics: Principles and Selected
Applications. Butterworth-Heinemann, Boston. (QA 911 K49)
In E.E. Bittar (ed.)
Cell motility.
King, C.A. and Preston, T.M.
1992.
Fundamentals of Medical Cell Biology. Volume 5B. Jai Press, Greenwich, CT,
pp. 147-184. (QH581.2 F86 1991 v.5B)
Kitching, J.A. 1957. Effects of high hydrostatic pressures on Actinophrys sol. J. expo
BioI. 34: 511-525.
92
Kitching, J.A.
1960. Responses of the heliozoan Actinophrys sol to prey, to
mechanical stimulation, and to solutions of proteins and certain other substances.
J. expo BioI. 37: 407-416.
Kitching, J.A. 1964. The axopods of the sun animalcule Actinophrys sol (Heliozoa).
In R.D. Allen and N. Kamiya (eds.) Primitive Motile Systems in Cell Biology.
Academic Press, New York.
Kitching, J.A. and Craggs, S. 1965. The axopodial filaments of the heliozoon
Actinosphaerium nucleofilum. Exp. Cell Res. 40: 658-660. [Authors note (p.
658) that "to an observer situated outside the organism, all the spirally wound
sheets of fibres trace a cw course from the central axis of the rod (=axoneme)
outwards". This agrees with Tilney et al. (1973).]
Koloclney, M.S. and Wysolmerski, R.B. 1992. Isometric contraction by fibroblasts
and endothelial cells in tissue culture: A quantitative study. J. Cell BioI. 117(1):
73-82.
Koshland, D.E., Mitchison, T.J., and Kirschner, M.W. 1988. Polewards chromosome
movement driven by microtubule depolymerization in vitro. Nature 331: 499504.
Koutoulis, A., McFadden, G.l., and Wetherbee, R. 1988. Spine-scale reorientation in
Apedinella radians (Pedinellales, Chrysophyceae): The microarchitecture and
immunocytochemistry of the associated cytoskeleton. Protoplasma 147: 25-41.
Kristensen, B.T., Engdahl Nielsen, L., and Rostgaard, J. 1974. Variations in
myoneme birefringence in relation to length changes in Stentor coeruleus. Exp.
Cell Res. 85: 127-135.
Kristofferson, D., Karr, T.L., and Purich, D.L. 1980. Dynamics of linear polymer
disassembly. J. BioI. Chem. 255: 8567-8572. [September] (QP501 J7)
Krueger, J.W., Denton, A., and Siciliano, G. 1992. Nature of motions between
sarcomeres in asynchronously contracting cardiac muscle cells. Biophys. J. 61:
145-160. [Treats single, isolated myocytes. See the Discussion, esp. pp. 157-159;
stress data, nice diagram on p. 158. Also see the refs. following this and the
preceding paper by Krueger et al.]
Kuhn, W., Ramel, A., and Walters, D.H. 1960. Conversion of chemical into
mechanical energy by homogeneous and cross-striated polymeric systems. In A.
Wassermann (ed.) Size and Shape Changes of Contractile Polymers: Conversion
of Chemical into Mechanical Energy. (Proceedings of seminars held at University
College, London) Pergamon Press, Oxford, pp. 41-77. (Textiles Library QD471
S52 1957)
Kushmerick, M.J. and Podolsky, R.J. 1969. Ionic mobility in muscle cells. Science
166: 1297-1298.
LaBarbera, M. 1984. Feeding currents and particle capture mechanisms in suspension
feeding animals. Am ZooI. 24: 71-84.
Lakes, R. 1993. Materials with structural hierarchy. Nature 361: 511-515. (11 Feb.)
93
Larsen, J. 1985. Ultrastructure and taxonomy of Actinomonas pusilla, a heterotrophic
member of the Pedinellales (Chrysophyceae) Br. Phyc. J. 20: 341-355.
Lasek, R.J. 1988. Studying the intrinsic determinants of neuronal form and function.
In R.J. Lasek and M.M. Black (eds.) Intrinsic Determinants of Neuronal Form
and Function. Alan R. Liss, New York, pp. 3-58. (UNC-Greensboro QP 361 M4
1988)
Lauffenburger, D.A. 1991. Models for receptor-mediated cell phenomena: Adhesion
and migration. Annual Review of Biophysics and Biophysical Chemistry 20:
387-414.
Laybourn, J. 1976. Respiratory energy losses in Podophrya fixa Muller in relation to
temperature and nutritional status. J. gen. Micr. 96: 203-208.
Laybourn-Parry, J. 1984. A Functional Biology of Free-Living Protozoa. University
of California Press, Berkeley and Los Angeles. (QL 366 L25 1984)
Laybourn-Parry, J. 1992. Protozoan Plankton Ecology. Chapman and Hall, London.
(QL366 L26 1992)
Leadbeater, B.S.C.
1983.
Distribution and chemistry of microfilaments in
choanoflagellates, with special reference to the collar and other tentacle systems.
Protistologica 19: 157-166. [Not useful. Actin "demonstrated" by HMM
decoration. No indication of the speed of these apparently actin-based systems.]
Leal, L.G. 1980. Particle motions in a viscous fluid. Annu. Rev. Fluid Mech. 12:
435-476.
Leal, L.G. 1992. Laminar Flow and Convective Transport Processes: Scaling
Principles and Asymptotic Analysis. Butterworth-Heinemann, Boston. (QA 929
L44 1992) [VERY IMPORTANT. Note correction to eqns. (2-4) and (2-5) on
•
page 18: p should be replaced by p.]
Lee, R.E. 1989. Phycology.
(QK566144 1989)
2nd ed.
Cambridge University Press, Cambridge.
Lewin, R.A. 1992. What the haptonema is for. Nature 356: 195-196.
Lewis, M.A. and Murray, J.D. 1991. Analysis of stable two-dimensional patterns in
contractile cytogel. J. Nonlinear Sci. 1: 289-311.
Li, Y. and Tanaka, T. 1992. Phase transitions of gels. Annu. Rev. Mater. Sci. 22:
243-277. (TA401 A825 v.22 1992)
Lide, D.R. (ed.) 1993. CRC Handbook of Chemistry and Physics. (74th ed., 19931994) CRe Press, Boca Raton.
Lighthill, M.J. 1952. On the squirming motion of nearly spherical deformable bodies
through liquids at very small Reynolds numbers. Communications of Pure and
Applied Mathematics 5: 109-118. (QAl C718)
Lighthill, M.J. 1975. Mathematical Bio/luiddynamics. Society for Industrial and
Applied Mathematics, Philadelphia. (QP310 S95 L50)
94
Lighthill, M.J. 1976. Flagellar hydrodynamics. SIAM Review 18(2): 161-230.
[Lighthill mentions axopodia in a nonmathematical paragraph on p. 170.]
Little, M., Quinlan, R.A., Hoffman, E.J., and Luduena, R.F. 1983. Identification and
characterization ofaxopodial tubulins from Echinosphaerium nucleofilum. Eur. J.
Cell BioI. 31: 55-61.
Luby-Phelps, K., Taylor, D.L., and Lanni, F.
cytoplasm. J. Cell BioI. 102: 2015-2022.
1986.
Probing the structure of
Luby-Phelps, K., Castle, P.E., Taylor, D.L., and Lanni, F. 1987. Hindered diffusion
of inert tracer particles in the cytoplasm of mouse 3T3 cells. Proc. NatI. Acad.
Sci. USA 84: 4910-4913.
Luby-Phelps, K., Lanni, F. and Taylor, D.L. 1988. The submicroscopic properties of
cytoplasm as a determinant of cellular function. Annual Review of Biophysics
and Biophysical Chemistry 17: 369-396.
Luft, J.H. 1971. Ruthenium red and violet. 1. Chemistry, purification, methods of
use for electron microscopy and mechanism of action. Anat. Rec. 171: 347-368.
Lugt, H.J. 1983. Vortex Flow in Nature and Technology. John Wiley, New York.
(QC159 L8413 1983) [See esp. pp. 48-52 on low Re flows; also, Re table on p. 79.]
MacDonald, A.C. and Kitching, J.A. 1967. Axopodial filaments of Heliozoa. Nature
215: 99-100. [Mostly not helpful; speculation on intraaxopodial transport.]
Machemer, H. 1984. Mechanoresponses in protozoa. In G. Colombetti, F. Lenci, and
P.-C. Song (eds.) Sensory Perception and Transduction in Aneural Organisms.
Plenum Press, New York, pp. 179-209. (QR97 L5 N37 1984) [VALUABLE
REVIEW. On p.189, Machemer describes the use of piezoelectric probes to
deliver mechanical stimuli. Esp. attention is paid to the ca. 12 J-Lm long sensory
tail cilium.]
MacKinnon, D.L. 1909. Optical properties of contractile organs in Heliozoa.
PhysioI. 38: 254-258.
J.
Maddox, J. 1992. Is molecular biology yet a science? Nature 355: 201.
Major, G., Evans, J.D., and Jack, J.J.B. 1993. Solutions for transients in arbitrarily
branching cables: 1. Voltage recording with a somatic shunt. Biophys. J. 65:
423-449.
Malev, V.V., Gromov, D.B., Komissarchik, Ya. Yu., and Brudnaya, M.S.
Interfacial models of nerve fiber cytoskeleton. Biophys. J. 63: 1101-1117.
1992.
Mandelbrot, B.B. 1983. The Fractal Geometry of Nature. W.H. Freeman, New York.
Mandelkow, E.-M., Mandelkow, E., and Milligan, R.A. 1991. Microtubule dynamics
and microtubule caps: A time-resolved cryo-electron microscopy study. J. Cell
BioI. 114(5): 977-991.
Mandelkow, E.-M. and Mandelkow, E. 1992. Microtubule oscillations. Cell Motility
95
and the Cytoskeleton 22: 235-244.
Manton, I. 1964. Further observations on the fine structure of the haptonema in
Prymnesium parvum. Archiv fur Mikrobiologie 49: 315-330. (QK504.A7)
Manton, I. 1968. Further observations on the microanatomy of the haptonema in
Chrysochromulina chiton and Prymnesium parvum. Protoplasma 66: 35-53.
Margulis, 1. and Sagan, D. 1986. Origins of Sex. Yale University Press, New Haven
and London. (QH 481 M27 1986)
Margulis, L., Corliss, J.O., Melkonian, M., and Chapman, D.J.
Protoctista. Jones and Bartlett, Boston. (QR 74.5 H36)
1990.
Handbook of
Mark, J.E. and Erman, B. 1988. Rubberlike Elasticity: A Molecular Primer. John
Wiley, New York. (QD381 M379 1988)
Marsland, D., Tilney, L.G., and Hirshfield, M. 1971. Stabilizing effects of D20 on the
microtubular components and needle-like form of heliozoan axopods: A pressuretemperature analysis. J. Cell Physiol. 77: 187-194.
Maruyama, T. 1981. Motion of the longitudinal flagellum in Ceratium tripos
Dinoflagellida): A retractile flagellar motion. J. Protozool. 28(3): 328-336.
Beautiful
micrographs.]
f
Maruyama, T. 1982. Fine structure of the longitudinal flagellum in Ceratium tripos,
a marine dinoflagellate. J. Cell Sci. 58: 109-123. [See Fig. 15, p. 121 for
schematic illustration of relaxed and retracted flagellum.]
Maruyama, T. 1985. Ionic control of the longitudinal flagellum in Ceratium tripos
(Dinoflagellida). J. Protozool. 32: 106-110.
Matsuno, K. 1989a. Cell motility: An interplay between local and nonlocal
measurement. BioSystems 22: 117-126. (QH 301 C87)
Matsuno, K. 1989b. Protobiology:
Raton. (QH 505 M27 1989)
Physical Basis of Biology.
CRC Press, Boca
Matsuo, E. and Tanaka, T. 1992. Patterns in shrinking gels. Nature 358: 482-485.
(6 August) [Color photographs of sundry morphologies.]
,
Matsuoka, T., Sugiyama, M., and Shigenaka, Y. 1984. Electrophoretic migration of
the heliozoan, Echinosphaerium akamae. Cytobios 40: 107-114.
Matsuoka, T. and Shigenaka, Y. 1984. Localization of calcium during axopodial
contraction in the heliozoan, Echinosphaerium akamae. Biomed. Res. 5: 425432.
Matsuoka, T., Shigenaka, Y., and Naitoh, Y. 1985. A model of contractile tubules
showing how they contract in the heliozoan Echinosphaerium. Cell Structure and
Function 10: 63-70.
Matsuoka, T. and Shigenaka, Y. 1985. Calcium localization during microtubular
disassembly in heliozoan cells. J. Electron Microsc. 34(1): 33-37.
96
Matsuoka, T., Ishigame, K., Muneoka, Y. and Shigenaka, Y. 1986. External factors
which induce the rapid contraction of heliozoan axopodia. Zool. Sci. 3: 437-443.
Matsuoka, T., Watanabe, Y., Kuriu, T., Arita, T., Taneda, K., Ishida, M., Suzaki, T.,
and Shigenaka, Y.
1991.
Cell models of Blepharisma:
Ca2+-dependent
modification of ciliary movement and cell elongation. Europ. J. Protistol. 27:
371-374.
McMahon, T.A. and Bonner, J.T. 1983. On Size and Life. Scientific American
Books, New York. [See graph on p. 149 of stalk height vs. stalk diameter in the
fruiting bodies of Dictyostelium.]
McMahon, T.A. 1984. Muscles, Reflexes, and Locomotion.
Press, Princeton. (QP321 M338 1984)
Princeton University
Melkonian, M. 1990. Centrin-mediated cell motility in eukaryotic cells. In W. Alt
Biological Motion.
(Proceedings of a Workshop,
and G. Hoffmann (eds.)
Konigswinter, Germany, March 16-19, 1989) Lecture Notes in Biomathematics,
Volume 89. Springer-Verlag, Berlin, pp. 117-118. (QP 301 B56 1990)
Melkonian, M. and Robenek, H. 1984. The eyespot apparatus of flagellated green
algae: A critical review. Prog. Phycol. Res. 3: 193-268.
Melkonian, M., Beech, P.L., Katsaros, C., and Schulze, D. 1992. Centrin-mediated
cell motility in algae. In M. Melkonian (ed.) Algal Cell Motility. Chapman and
Hall, New York, pp. 179-221. (QK565 A375 1991)
Meyer, T. 1991. Cell signalling by second messenger waves. Cell 64: 675-678. [Brief
review with data summaries in tables and elem. math. treatment.]
Mignot, J.-P. and Brugerolle, G. 1992. Cell division in the heliozoan Dimorpha
mutans and evolution of centrosomal organization during the cell cycle. Biology
of the Cell 72: 51-60. (QH 201 V55)
Miki-Noumura, T. 1990. Dark-field microscopic study of microtubules in solution.
International Review of Cytology 122: 65-104.
Minton, A.P. 1992. Confinement as a determinant of macromolecular structure and
reactivity. Biophys. J. 63: 1090-1100.
Mitchison, T.J. 1993. Localization of an exchangeable GTP binding site at the plus
end of microtubules. Science 261: 1044-1047.
Mizushima-Sugano, J., Maeda, T., and Miki-Noumura, T. 1983. Flexural rigidity of
singlet microtubules estimated from statistical analysis of their contour lengths
and end-to-end distances. Biochim. Biophys. Acta 755: 257-262. (QP 501 B55)
Moffatt, H.K. 1964. Viscous and resistive eddies near a sharp corner. J. Fluid Mech.
18: 1-18.
Morel, J.E. and Bachouchi, N. 1988. Muscle contraction and movement of cellular
organelles: Are there two different types of mechanisms for their generation?
J. theor. BioI. 132: 83-96.
97
Moreton, R.B. and Amos, W.B. 1979. Electrical recording from the contractile ciliate
Zoothamnium geniculatum Ayrton. J. expo BioI. 83: 159-167.
Murota, K. and Ikeda, K. 1992. On random imperfections for structures of regularpolygonal symmetry. SIAM J. AppI. Math. 52(6): 1780-1803. (December 1992)
Nanavati, C. and Fernandez, J.M. 1993. The secretory granule matrix: A fast-acting
smart polymer. Science 259: 963-965. [See discussion by Travis (1993).]
Narayanan, R. (ed.) 1982. Axially Compressed Structures:
Applied Science Publishers, London. (TA646 A94 1982)
Stability and Strength.
Narayanan, R. and Roberts, T.M. (eds.) 1991. Structures Subjected to Dynamic
Loading: Stability and Strength. Elsevier Applied Science, London. (TA654 S79
1991)
Needham, D. and Hochmuth, R.M. 1990. Rapid flow of passive neutrophils into a 4
p'm pipet and measurement of cytoplasmic viscosity. Trans. ASME J. Biomech.
Eng. 112: 269-276.
Newman, E. 1972. Contraction in Stentor coeruleus: A cinematic analysis. Science
177: 447-449.
Nicklas, R.B. 1984.
Motility 4: 1-5.
A quantitative comparison of cellular motile systems.
Nisbet, B. 1984. Nutrition and Feeding Strategies in Protozoa.
London. (Vet. Med. QL364 N57 1984)
Cell
Croom Helm,
Nishi, T. Kobayashi, M., and Shigenaka, Y. 1986. Membrane activity and its
correlation with vacuolar contraction in the heliozoan Echinosphaerium. J. Exp.
ZooI. 239: 175-182.
Nishi, T., Gotow, T., and Kobayashi, M. 1988. Changes in electrical connection
during cell fusion in the heliozoan, Echinosphaerium akamae. J. expo BioI. 135:
183-191.
Nitsche, L. 'and Brenner, H. 1989. Eulerian kinematics of flow through spatially
periodic models of porous media. Archive for Rational Mechanics and Analysis
107(3): 225-292.
Nossal, R. 1987. In vitro polymerization of complex cytoplasmic gels. In P.S. Russo
(ed.) Reversible Polymeric Gels and Related Systems. American Chemical
Society, Washington, DC, pp. 224-235.
Nossal, R. 1988. On the elasticity of cytoskeletal networks. Biophys. J. 53: 349-359.
Nossal, R. and Lecar, H.
Redwood City.
1991.
Molecular and Cell Biophysics.
Addison-Wesley,
Ochiai, T., Hara, R., and Asai, H. 1983. Contraction of the spasmoneme and coiling
of the sheath in the glycerinated stalk of Vorticella. Cytobios 36: 95-105.
98
o ckleford, C.D. and Tucker, J.B.
1973. Growth, breakdown, repair, and rapid
contraction of microtubular axopodia in the heliozoan Actinophrys sol.
J.
Ultrastructure Res. 44: 369-387.
Ockleford, C.D. 1974. Cytokinesis in the heliozoan Actinophrys sol. J. Cell Sci. 16:
499-517.
o ckleford, C.D. 1975. Ultraviolet light microbeam irradiation of the microtubules in
single heliozoan axopodia. Exp. Cell Res. 93: 127-135.
Odell, G.M. 1977. Amoeboid motions. In R.C. DiPrima (ed.) Modern Modeling of
Continuum Phenomena. (Lectures in Applied Mathematics, Vol. 16) American
Mathematical Society, Providence, pp. 191-220. (QA808.2 S85 1975 cop. 2)
Odell, G.M. and Bonner, J.T. 1986. How the Dictyostelium discoideum grex crawls.
Phil. Trans. R. Soc. Lond. B 312: 487-525.
Okuno, M. and Hiramoto, Y. 1979. Direct measurements of the stiffness of
echinoderm sperm flagella. J. expo BioI. 79: 235-243.
Osada, Y., Okuzaki, H., and Hori, H. 1992. A polymer gel with electrically driven
motility. Nature 355: 242-244. (16 January)
Oster, a.F., Perelson, A.S., and Tilney, L.G. 1982. A mechanical model for
elongation of the acrosomal process in Thyone sperm. J. Math. BioI. 15: 259265.
Oster, a.F. and Odell, G.M. 1984. The mechanochemistry of cytogels.
12D(1984): 333-350. (QC 20.7 N6 153)
Oster, a.F. and Perelson, A.S. 1985. Cell spreading and motility:
lamellipod. J. Math. BioI. 21: 383-388.
Physica
A model
Oster, a.F. and Perelson, A.S. 1987. The physics of cell motility. J. Cell Sci. SuppI.
8: 35-54. (QH 623. B7)
Oster, a.F. 1988. Biophysics of the leading lamella.
Cytoskeleton 10: 164-171.
Cell Motility and the
Oster, a.F. and Moore, H.-P.H. 1989a. The budding of membranes. In A. Goldbeter
Cell to Cell Signalling:
From Experiments to Theoretical Models.
(ed.)
Academic Press, London, pp. 171-187. (QH 604.2 C44448 1989)
Oster, a.F., Cheng, L.Y., Moore, H.-P.H., and Perelson, A.S.
formation in the Golgi apparatus. J. theor. BioI. 141: 463-504.
1989b.
Vesicle
Oster, a.F. 1989c. Cell motility and tissue morphogenesis. In W.D. Stein and F.
Bronner (eds.) Cell Shape: Determinants, Regulation, and Regulatory Role.
Academic Press, San Diego, pp. 33-61. (QH611 C425 1989)
Otter, T., Bourns, B., Franklin, S., Rieder, C., and Salmon, E.D. 1987. Hydrostatic
pressure effects on cytoskeletal organization an.d ciliary motility: A calcium
hypothesis. In H.W. Jannasch, R.E. Marquis, and A.M. Zimmerman (eds.)
99
Current Perspectives in High Pressure Biology. Academic Press, London, pp. 7593. (QH 313 C8 1987) [The authors suggest (p. 76) that "pressure may induce its
disruptive effects by elevation of the concentration of free calcium ion in the
cytosol". See esp. the section "Mechanisms of pressure-induced disruption of MT
and actin microfilament assembly" and the refs.]
Parkes, E.W. 1974.
(TA645 P3 1974)
Braced Frameworks.
(2nd ed.)
Pergamon Press, Oxford.
Parkhouse, J.G. 1984. Structuring: A process of material dilution. In H. Nooshin
(ed.) Third International Conference on Space Structures. Elsevier, London, pp.
367-374.
Parkhouse, J.G. 1987. Damage accumulation in structures. Reliability Engineering
17: 97-109. (TS173 R44)
Patterson, D.J. 1979. On the organization and classification of the protozoan
Actinophrys sol Ehrenberg, 1830. Microbios 26: 165-208. (QRl M7) [Mostly
treats encystment, excystment; disappointing dearth of morphometry.]
Patterson, D.J. and Hausmann, K. 1981. Feeding by Actinophrys sol
Heliozoa): 1. Light microscopy. Microbios 31: 39-55.
(Protista:
Patterson, D.J. and Hausmann, K. 1982. Morulate bodies in actinophryid Heliozoa:
A fixation artefact derived from microtubules? Cell Struct. Func. 7: 341-348.
Patterson, D.J. and Fenchel, T. 1985. Insights into the evolution of Heliozoa
(Protozoa, Sarcodina) as provided by ultrastructural studies on a new species of
flagellate from the genus Pteridomonas. BioI. J. Linn. Soc. 24: 381-403. (QH301
B42)
Patterson, D.J. and Larsen, J. (eds.) 1991. The Biology of Free-Living Heterotrophic
Flagellates. Clarendon Press, Oxford. (QL368 F5 B46 1991)
Patterson, D.J. and Hedley, S. 1992. Free-Living Freshwater Protozoa:
Guide. CRC Press, Boca Raton. (QL 366.5 P37)
A Color
Pedley,. T.J: and. Kessler, .J.O. 1992. Hydrodynamic phenomena in suspensions of
sWImmmg mIcroorganIsms. Annual Review of Fluid Mechanics 24: 313-358.
(QC 145 A57 v. 24 1992)
Perelson, A.S. and Coutsias, E.A. 1986. A moving boundary model of acrosomal
elongation. J. Math. BioI. 23: 361-379.
Peskin, C.S., Odell, G.M., and Oster, G.F. 1993. Cellular motions and thermal
fluctuations: The Brownian ratchet. Biophys. J. 65: 316-324.
Phillips, D.M. 1974. Structural variants in invertebrate sperm flagella and their
relationship to motility. In M.A. Sleigh (ed.) Cilia and Flagella. Academic
Press, New York, pp. 379-402.
Pollack, G.H. 1990. Muscles and Molecules: Uncovering the Principles of Biological
Motion. Ebner and Sons, Seattle. (QP 321 P57 1990) [See esp. Chapters 6
("Helix-Coil Transition in the Myosin Rod") 7 ("Helix-Coil Transition as the
100
Prime Mover") and 15 ("Machina Carnis:
spasmoneme is discussed on p.254 and p. 256.]
The Final Product").
The
Pollard, T.D., Goldberg, I., and Schwarz, W.H. 1992. Nucleotide exchange, structure,
and mechanical properties of filaments assembled from ATP-actin and ADP-actin.
J. BioI. Chern. 267: 20339-20345. (October 5)
Portzehl, H. 1951. Muskelkontraktion und Modellkontraktion.
Zeitschrift fUr Naturforschung 6B: 355-361. (Q3 Z391)
II.
Mitteilung.
Power, H. and de Power, B.F. Second-kind integral equation formulation for the slow
motion of a particle of arbitrary shape near a plane wall in a viscous fluid. SIAM
J. Appl. Math. 53(1): 60-70.
Pozrikidis, C. 1992. Boundary Integral and Singularity Methods for Linearized
Viscous Flow. Cambridge University Press, Cambridge. (TA 357.5 V56 P69
1992) [VERY IMPORTANT]
Probstein, R.F.
1989.
Physicochemical Hydrodynamics:
Butterworth-Heinemann, Boston. (QD 453.2 P75)
An
Introduction.
Pryer, N.K., Walker, R.A., Skeen, V.P., Bourns, B.D., Soboeiro, M.F., and Salmon,
E.D. 1992. Brain microtubule-associated proteins modulate microtubule dynamic
instability in vitro: Real-time observations using video microscopy. J. Cell Sci.
103: 965-976.
Purcell, E.M. 1977. Life at low Reynolds number. American Journal of Physics 45:
3-11. (QC1 A5)
Rahat, M., Parnas, I., and Nevo, A. 1969. Extensibility and tensile strength of the
stalk "muscle" of Carchesium. Exp. Cell Res. 54: 69-76.
Rahat, M., Pri-Paz, Y., and Parnas, I. 1973. Properties of stalk "muscle" contractions
of Carchesium sp. J. expo BioI. 58: 463-471.
Rasmussen, S., Karampurwala, H., Vaidyanath, R., Jensen, K.S., and Hameroff, S.
1990. Computational connectionism within neurons: A model of cytoskeletal
automata subserving neural networks. Physica D 42: 428-449.
Rieder, C.L. 1979. High-voltage electron microscopy reveals new components in the
axonemes and centroplast of the centrohelidian Raphidiophrys ambigua. J. Cell
Sci. 40: 215-234. [Indicates presence of central filament like that reported by
Davidson (1975) in Heterophrys, but suggests function in pattern formation.]
Rikmenspoel, R. 1982. Ciliary contractile model applied to sperm flagellar motion. J.
theor. BioI. 96: 617-645.
Ripoll, D., Marchi, E., and Benegas, J. 1981.
depolymerization. J. theor. BioI. 93: 53-62.
A new approach to microtubule
Ritter, E. 1954. Cytoplasmic structures and cytochemistry of Actinosphaerium
eichhorni. J. Protozool. 1: Suppl., p. 9.
Rizzuto, R., Brini, M., Murgia, M., and Pozzan, T. 1993. Microdomains with high
101
Ca2+ close to IP 3 -sensitive channels that are sensed by neighboring mitochondria.
Science 262: 744-747.
Roberts, A.M. 1981. Hydrodynamics of protozoan swimming. In M. Levandowsky
and S.H. Hutner (eds.) Biochemistry and Physiology of Protozoa. Volume 4.
(2nd ed.) Academic Press, New York, pp. 5-66. (NRL QL369.2 L87 1979 vA)
Roberts, W.M.
74-76.
1993. Spatial calcium buffering in saccular hair cells. Nature 363:
Roskin, G. 1925. Uber die Axopodien der Heliozoa und die Greiftentakeln der
Ephelotidae. Archiv fur Protistenkunde 52: 207-216.
Roth, L.E., Pihlaja, D.J., and Shigenaka, Y. 1970a. Microtubules in the heliozoan
axopodium. 1. The gradion hypothesis of allosterism in structural proteins. J.
Ultrastructure Res. 30: 7-37.
Roth, L.E. and Shigenaka, Y. 1970b. Microtubules in the heliozoan axopodium. II.
Rapid degradation by cupric and nickelous ions. J. Ultrastructure Res. 31: 356374.
Roth, L.E. and Shigenaka, Y. 1975. Microtubules in the heliozoan axopodium. IV.
Cell-surface regulation by using hyaluronate. Cytobiologie 12: 79-92.
Roth, L.E. and Pihlaja, D.J. 1977. Gradionation:
patterning. J. Protozool. 24(1): 2-9.
Hypothesis for positioning and
Routledge, L.M., Amos, W.B., Gupta, B.L., Hall, T.A., and Weis-Fogh, T. 1975.
Microprobe measurements of calcium binding in the contractile spasmoneme of a
vorticellid. J. Cell Sci. 19: 195-201.
Routledge, L.M. 1978. Calcium-binding proteins in the vorticellid spasmoneme.
Extraction and characterization by gel electrophoresis. J. Cell BioI. 77: 358-370.
Rubenstein, D.l. and Koehl, M.A.R. 1977. The mechanisms of filter feeding: Some
theoretical considerations. Am. Natur. 111: 891-894.
Rudzinska, M.A. 1973. Do Suetoria really feed by suction? BioScience 23: 87-94.
Russel, W.B. 1987. The Dynamics of Colloidal Systems.
Press, Madison. (QD549 R79 1987)
University of Wisconsin
Salisbury, J.L. 1982. Calcium-sequestering vesicles and contractile flagellar roots. J.
Cell Sci. 58: 433-443.
Salisbury, J.L., Baron, A., Surek, B., and Melkonian, M. 1984. Striated flagellar
roots: Isolation and partial characterization of a calcium-modulated contractile
organelle. J. Cell BioI. 99: 962-970.
Salisbury, J.L., Sanders, M.A., and Harpst, 1. 1987. Flagellar root contraction and
nuclear movement during flagellar regeneration in Chlamydomonas reinhardtii. J.
Cell BioI 105: 1799-1850.
Salisbury, J.L., Baron, A.T., and Sanders, M.A. 1988. The centrin-based cytoskeleton
102
of Chlamydomonas reinhardtii. J. Cell BioI. 107: 635-642.
Salisbury, J.1. 1989. Centrin and the algal flagellar apparatus. J. Phycol. 25: 201206.
Salmon, E.D. 1975a. Spindle microtubules: Thermodynamics of in vivo assembly and
role in chromosome movement. Annals of the New York Academy of Sciences
253: 383-406. [van't Hoff plots based on Inoue's simple equilibrium model]
Salmon, E.D. 1975b. Pressure-induced depolymerization of spindle microtubules. II.
Thermodynamics of in vivo spindle assembly. J. Cell BioI. 66: 114-127.
Salmon, E.D., Saxton, W.M., Leslie, R.J., Karow, M.L., and McIntosh, J.R. 1984a.
Diffusion coefficient of fluorescein-labeled tubulin in the cytoplasm of embryonic
cells of a sea urchin: Video image analysis of fluorescence redistribution after
photobleaching. J. Cell BioI. 99: 2157-2164.
Salmon, E.D., Leslie, R.J., Saxton, W.M., Karow, M.L., and McIntosh, J.R. 1984b.
Spindle microtubule dynamics in sea urchin embryos:
Analysis using a
fluorescein-labeled tubulin and measurements of fluorescence redistribution after
laser photobleaching. J. Cell BioI. 99: 2165-2174.
Salmon, E.D. 1989. Microtubule dynamics and chromosome movement. In J.S.
Hyams and B.R. Brinkley (eds.) Mitosis: Molecules and Mechanisms. Academic
Press, New York.
Sato, M. and Rodan, G.A. 1989. Bone cell shape and function. In W.D. Stein and F.
Bronner (eds.) Cell Shape: Determinants, Regulation, and Regulatory Role.
Academic Press, San Diego, pp. 329-362. [Data on mechanical properties of
cytoplasm, etc.] (QH 611 C425 1989)
Scheibel, L.W. 1992. Role of calcium/calmodulin-mediated processes in protozoa.
International Review of Cytology 134: 165-242. [Nothing on heliozoans, but may
be very useful.]
Schliwa, M. 1976. The role of divalent cations in the regulation of microtubule
assembly. In vivo studies on microtubules of heliozoan axopodium using the
ionophore A23187. J. Cell BioI. 70: 527-540. [The extension rate for axopodia is
ca. 6 J.Lm/min in Echinosphaerium.]
Schliwa, M. 1977. Influence of calcium on intermicrotubule bridges within heliozoan
axonemes. J. Submicr. CytoI. 9: 221-228.
Schliwa, M., Euteneuer, U., Bulinski, J.L., and Izant, J.G. 1981. Calcium lability of
cytoplasmic microtubules and its modulation by microtubule-associated proteins.
Proc. NatI. Acad. Sci. USA 78: 1037-1041.
Schmid-Schonbein, G.W., Skalak, R., and Chien, S. 1986. Leukocyte protopodia. In
G.W. Schmid-Schonbein, S.L-Y. Woo, and B.W. Zweifach (eds.) Frontiers in
Biomechanics. Springer-Verlag, New York, pp. 33-47. (QP303 F69 1986)
Schmid-Schonbein, G.W. 1990.
Biophysics 17: 107-135.
Leukocyte biophysics:
An invited review.
Cell
103
Schmidt, W.J. 1944. Polarisationsmikroscopische Beobachtungen an Actinosphaerium
eichornii. Biologisches Zentralblatt 64: 5-14. (QH 301 B 528)
Schmidt-Nielsen, K. 1977. Problems of scaling: Locomotion and physiological
correlates. In T.J. Pedley (ed.) Scale Effects in Animal Locomotion. Academic
Press, New York, pp. 1-21. (QP 301 S2)
Schroeder, M. 1991. Fractals, Chaos, Power Laws. W.H. Freeman, New York.
Schulgasser, K. and Witztum, A. 1992. On the strength, stiffness and stability of
tubular plant stems and leaves. J. theor. BioI. 155: 497-515.
Segalman, D., Witkowski, W., Adolf, D., and Shahinpoor, M. 1992. Electricallycontrolled polymeric gels as active materials in adaptive structures. In G.J.
Knowles (ed.) Active Materials and Adaptive Structures. Institute of Physics
Publishing, Bristol, pp. 335-345. (TA418.9 S62 A36 1991)
Segel, L.A. 1991. The infinite and the infinitesimal in models for natural phenomena.
Reviews of Modern Physics 63(2): 225-238. [April 1991]
Shahinpoor, M. 1992.
Conceptual design, kinematics and dynamics of swimming
robotic structures using active polymer gels. In G.J. Knowles (ed.) Active
Materials and Adaptive Structures. Institute of Physics Publishing, Bristol, pp.
91-95. (TA418.9 S62 A36 1991)
Shaw, C.T. 1992. Using Computational Fluid Dynamics. Prentice Hall, New York.
(QA911 S453 1992)
Sheetz, M.P., Wayne, D.B., and Pearlman, A.L. 1992. Extension of filopodia by
motor-dependent actin assembly. Cell Motility and the Cytoskeleton 22(3): 160169.
Shelden, E. and Wadsworth, P. 1993. Observation and quantification of individual
microtubule behavior in vivo: Microtubule dynamics are cell-type specific. J. Cell
BioI. 120(4): 935-945.
Shepherd, G.M. 1974. The Synaptic Organization of the Brain:
Oxford University Press, New York.
An Introduction.
Shigenaka, Y., Roth, L.E., and Pihlaja, D.J. 1971. Microtubules in the heliozoan
axopodium. III. Degradation and reformation after dilute urea treatment. J.
Cell Sci 8: 127-152.
Shigenaka, Y., Watanabe, K., and Kaneda, M. 1974. Degrading and stabilizing effects
of Mg2+ ions on microtubule-containing axopodia. Exp. Cell Res. 85: 391-398.
Shigenaka, Y., Tadokoro, Y., and Kaneda, M. 1975. Microtubules in protozoan cells.
I. Effects of light metal ions on the heliozoan microtubules and their kinetic
analysis. Annot. Zool. Japon. 48: 227-241. (QL1 A56) [Very poor English;
some passages almost unreadable. The ions in question are Sr, Ca, and Mg.
Lineweaver-Burk plots.]
104
Shigenaka, Y. 1976. Microtubules in protozoan cells. II. Heavy metal ion effects on
degradation and stabilization of the heliozoan microtubules. Annot. Zoo1. Japon.
49: 164-176. [Hg2+, Cd 2 +, and Zn2+]
Shigenaka, Y. and Kaneda, M. 1979a. Studies on the cell fusion of heliozoans. IV.
An electron microscopical study on the fusion process accompanied with axopodial
degradation. Annot. Zool. Japon. 52: 28-39. (QL1. A56)
Shigenaka, Y., Maruoka, T., Toyohara, A., and Suzaki, T. 1979b. Studies on the cell
fusion of heliozoans. V. Cell membrane fluidity and surface coat in close relation
to the fusion reaction. Annot. Zoo1. Japon. 52: 163-178. (QL1 A56)
Shigenaka, Y., Toyohara, A., Suzaki, T. and Watanabe, S. 1979c. Studies on the cell
fusion of heliozoans. VI. Enhancement of the fusion reaction induced by cold and
colchicine treatment. Cell Structure and Function 4: 199-207. (QH 573.C397)
Shigenaka, Y., Maruoka, T., Toyohara, A., and Suzaki, T. 1979d. Microtubules in
protozoan cells.
IV.
Effects of sulfhydryl-blocking reagents on axopodial
degradation and reformation in Heliozoa. Cell Struct. Funct. 4: 23-34.
Shigenaka, Y., Watanabe, K., and Suzaki, T. 1980. Taxonomic studies on two
heliozoans, Echinosphaerium akamae sp. nov. and Echinosphaerium ikachiensis sp.
nov. Annot. Zoo1. Jap. 53: 103-119. (QL1.A56)
Shigenaka, Y., Yano, K., Yogosawa, R., and Suzaki, T. 1982. Rapid contraction of
the microtubule-containing axopodia in a large heliozoan, Echinosphaerium. In H.
Sakai, H. Mohri, and G. Borisy (eds.) Biological Functions of Microtubules and
Related Structures. Academic Press, New York, pp. 105-114.
Shigenaka, Y. and Iwata, S. 1984. Resorption of microtubules and nucleus during the
encystation process in Echinosphaerium nucleofilum. Cytobios 41: 113-126.
(QH 573.C9)
Shigenaka, Y., Iwata, S., and Toyohara, A. 1985. Ultrastructural changes during the
cyst formation process in a heliozoan, Echinosphaerium akamae.
Microbios 44:
141-155. (QR l.M7)
Shigenaka, Y., Yano, K., and Imada, M. 1989a. Axopodial contraction evoked by
electrical stimulation and its ultrastructural analysis in a heliozoan
Echinosphaerium. Zoo1. Sci. 6(1): 45-53. (QL 1.Z738)
Shigenaka, Y., Maesako-Tabata, J., and Imada, M. 1989b. Electron microscopy of the
continous membrane system in a large heliozoon Echinosphaerium. Eur. J.
Protisto1. 24: 174-180. (QH 274 P7)
Siemensma, F.J. 1991. Heliozoea. In D. Matthes (ed.) Protozoenfauna Band 2:
Nackte Rhizopoda und Heliozoea. Gustav Fischer Verlag, Stuttgart, pp. 171-290.
[The whole volume is reviewed by D.J. Patterson in Europ. J. Protist. 27: 202203 (1991). Borrowed from the M.C.Z. Library at Harvard, Feb. 1994.]
Silvester, N.R. and Sleigh, M.A. 1985. The forces on microorganisms at surfaces in
flowing water. Freshwater Biology 15: 433-448. (QH96 .A1 F73)
105
Skalak, R. and Zhu, C. 1990. Thermodynamics and mechanics of active cell motions.
In N. Akkas (ed.) Biomechanics of Active Movement and Deformation of Cells.
Springer-Verlag, Berlin, pp. 155-183. (QH611 N36)
Skalak, R., Dong, C., and Zhu, C. 1990. Passive deformations and active motions of
leukocytes. Trans. ASME J. Biomech. Eng. 112: 295-302.
Sleigh, M.A. and Blake, J.R. 1977. Methods of ciliary propulsion and their size
limitations. In T.J. Pedley (ed.) Scale Effects in Animal Locomotion. Academic
Press, New York, pp. 243-256. (QP301 S2)
Sleigh, M.A. 1984. Motile systems of ciliates. Protistologica 20: 299-305. (QH274
P7) [IMPORTANT REVIEW WITH REFS. TO SUCTORIAN TENTACLES]
Sleigh, M.A. 1989. Protozoa and Other Protists. Edward Arnold, London.
Smith, R.McK. and Patterson, D.J. 1986. Analyses of heliozoan interrelationships:
An example of the potentials and limitations of ultrastructural approaches to the
study of protistan phylogeny. Proc. Roy. Soc. Lond. B 227: 325-366.
Smith, S.B., Finzi, L., and Bustamante, C. 1992. Direct mechanical measurements of
the elasticity of single DNA molecules by using magnetic beads. Science 258:
1122-1126. (11 November)
Spero, H.J. 1982. Phagotrophy in Gymnodinium fungiforme (Pyrrhophyta):
peduncle as an organelle of ingestion. J. Phycol. 18: 356-360.
The
Spielman, L.A. 1977. Particle capture from low-speed laminar flows. Annual Review
of Fluid Mechanics 9: 297-319. (QC145 A57 v. 9 1977)
Steuer, E.R., Wordeman, 1., Schroer, T.A., and Sheetz, M.P. 1990. Localization of
cytoplasmic dynein to mitotic spindles and kinetochores. Nature 345: 266-268.
Stossel, T.P., Janmey, P.A., and Zaner, K.S. 1987. The cortical cytoplasmic actin
gel. In J. Bereiter-Hahn, O.R. Anderson, and W.E. Reif (eds.) Cytomechanics.
Springer-Verlag, Berlin, pp. 131-153.
Stossel, T.P. 1990. How cells crawl. Am. Sci. 78: 408-423.
Sugiyama, M., Ikegawa, S., Masuyama, E., Suzaki, T., Ishida, M., and Shigenaka, Y.
1992. Isolation and properties of the axopodial cytoskeleton of a heliozoan,
Echinosphaerium akamae. Eur. J. Protistol. 28: 214-219. (QH274 P7)
Suzaki, T., Shigenaka, Y., and Takeda, Y. 1978. Studies on the cell division of
heliozoans. II. Active role ofaxopodia on division process and transformation of
remnants of cytoplasmic connecting bridge to new axopodia. Cell Struct. Fund.
3: 209-218.
Suzaki, T., Shigenaka, Y., Watanabe, S., and Toyohara, A. 1980a. Food capture and
ingestion in the large heliozoan, Echinosphaerium nucleofilum. J. Cell Sci. 42:
61-79.
Suzaki, T. and Shigenaka, Y. 1982. Intra-axopodial particle movement and axopodial
surface motility in Echinosphaerium akamae. In H. Sakai, H. Mohri, and G.G.
106
Borisy (eds.)
Biological Functions of Microtubules and Related Structures.
Academic Press, New York, pp. 91-103.
Suzaki, T., Ando, M., Ishigame, K., Shigenak, Y., and Sugiyama, M.
1992a.
Structure and function of the cytoskeleton in Heliozoa. 2. Measurement of the
force of rapid axopodial contraction in Echinosphaerium. Eur. J. Protistol. 28:
430-433. (November 20, 1992)
Suzaki, T., Ando, M., and Shigenaka, Y.
1992b.
Quantitative polarization
microscopical analysis of microtubule dynamics in a heliozoan Echinosphaerium
during axopodial contraction. Cell Motility and the Cytoskeleton 23(4): 308.
(abstract)
Tamm, S.L. and Tamm, S. 1976. Rotary movements and fluid membranes in termite
flagellates. J. Cell Sci. 20: 619-639. [EM pictures of single spiral axostyle MT
arrays with chirality given in each case: the "row of MTs spirals inward in a cw
direction as viewed from the anterior end." Authors note that the spacing
between axostylar MTs is much less regular than in the heliozoan axoneme.]
Tanaka, T. 1987. Gels. In C. Nicolini (ed.) Structure and Dynamics of Biopolymers.
Martinus Nijhoff, Dordrecht, pp. 237-257. (QP624 N368 1986)
Taylor, E.W. 1963. Brownian and saltatory movements of cytoplasmic granules and
the movement of anaphase chromosomes. In A.L. Copley (ed.) Proceedings of
the Fourth International Congress on Rheology.
Part 4.
Symposium on
Biorheology.
(Brown University, Providence, RI
August 26-30,
1963)
Interscience, New York, pp. 175-191. (TP156.R45)
Taylor, G.I. 1960. Similarity solutions of hydrodynamic problems. In N.J. Hoff and
W.G. Vincenti (eds.) Aeronautics and Astronautics. Pergamon Press, Oxford,
pp. 21-28. (TL505 .D87)
Taylor, G.I. 1966. The force exerted by an electric field on a long cylindrical
conductor. Proc. Roy. Soc. A 291: 145-158. [Reprinted as paper 42 in vol. 4 of
the Collected Works, pp. 486-500. (QC3 T37 vA)]
Taylor, G.I. 1972. Low Reynolds number flows. In A. Shapiro (ed.) fllustrated
Experiments in Fluid Mechanics. (Based on material in films produced under
direction of the National Committee for Fluid Mechanics Films.) MIT Press,
Cambridge, pp. 47-54. (QC145.2 N37)
Tenchov, B.G. and Yanev, T.K. 1986. Weibull distribution of particle sizes obtained
by uniform random fragmentation. J. Colloid and Interface Sci. 111(1): 1-7.
Thompson, N.L. and Burghardt, T.P. 1985. The influence of diffusion on the rate of a
reversible quasi-unimolecular reaction in one, two or three dimensions.
Biophysical Chemistry 21: 173-183.
Tilney, L.G. and Porter, K.R. 1965. Studies on microtubules in Heliozoa. I. The fine
structure of A ctinosphaerium nucleofilum (Barrett) with particular reference to
the axial rod structure. Protoplasma 60: 317-344.
Tilney, L.G., Hiramoto, Y., and Marsland, D. 1966.. Studies on the microtubules in
the Heliozoa. III. A pressure analysis of the role of these structures in the
107
formation and the maintenance of the axopodia of Aetinosphaerium nucleofilum.
J. Cell BioI. 29: 77-95.
Tilney, L.G. and Porter, K.R. 1967. Studies on the microtubules in Heliozoa. II.
The effect of low temperature on these structures in the formation and
maintenance of the axopodia. J. Cell BioI. 34: 327-343.
Tilney, L.G. 1968. The assembly of microtubules and their role in the development of
cell form. In M. Locke (ed.) The Emergence of Order in Developing Systems.
Developmental Biology Supplement, Vol. 2. Academic Press, New York, pp. 63102.
Tilney, L.G. 1968. Studies on the microtubules in Heliozoa. IV. The effect of
colchicine on the formation and maintenance of the axopodia and the
redevelopment of pattern in Aetinosphaerium nucleofilum. J. Cell Sci. 3: 549562.
Tilney, L.G. and Byers, B. 1969. Studies on the microtubules in Heliozoa. V.
Factors controlling the organization of microtubules in the axonemal pattern in
Echinosphaerium (Aetinosphaerium) nucleofilum. J. Cell BioI. 43: 148-165.
Tilney, L.G. 1971. How microtubule patterns are generated. The relative importance
of nucleation and bridging of microtubules in the formation of the axoneme of
Raphidiophrys. J. Cell BioI. 51: 837-854.
Tilney, L.G., Bryan, J., Bush, D.J., Fujiwara, K., Mooseker, M.S., Murphy, D.B., and
Snyder, D.H. 1973. Microtubules: Evidence for 13 protofilaments. J. Cell BioI.
59: 267-275. [Axopodial MTs of E. nucleofilum fixed in situ with glutaraldehyde
/tannic acid have 13 protofilaments. Authors remark (p. 275) that the axoneme
spirals ccw when viewed from base to tip.]
Tissot, 0., Pierres, A., Foa, C., Delaage, M., and Bongrand, P. 1992. Motion of cells
sedimenting on a solid surface in a laminar shear flow. Biophys. J. 61: 204-215.
[See esp. pp. 213-214 on the mechanical behavior of cell protrusions.]
Toyohara, A., Shigenaka, Y., and Utsumi, K. 1977. Disturbance in arrangement of
heliozoan axopodia and their degradation with concanavalin A treatment. Cell
Struct. Func. 2: 307-320.
Toyohara, A., Shigenaka, Y., and Mohri, H. 1978. Microtubules in protozoan cells.
III. Ultrastructural changes during disintegration and reformation of heliozoan
microtubules. J. Cell Sci. 32: 87-98.
Toyohara, A., Suzaki, T., Watanabe, S., and Shigenaka, Y. 1979. Ultrastructure of
heliozoan Echinosphaerium nucleofilum axopodia revealed by whole mounted
negative staining technique. J. Electron Microsc. 28: 185-188. (QC373 .E4J6)
Travis, J. 1993. Unexpected intelligence turns up in a cellular gel. Science 259: 893894.
Travis, J.L. and Bowser, S.S. 1991. The motility of foraminifera. In J.J. Lee and
O.R. Anderson (eds.) Biology of Foraminifera. Academic Press.
Troyer, D. 1975. Possible involvement of the plasma membrane in saltatory particle
108
movement in heliozoan axopods. Nature 254: 696-698.
Tucker, J.B. 1977. Shape and pattern specification during microtubule bundle
assembly. Nature 266: 22-26.
Tucker, J.B. 1979. Spatial organization of microtubules. In K. Roberts and J.S.
Hyams (eds.) Microtubules. Academic Press, London, pp. 315-357.
Tucker, J.B. 1984. Spatial organization of microtubule organizing centers and
microtubules. J. Cell BioI. 99: 55s-62s.
Turner, J.G.R. and Webster, J. 1991. Mass and momentum transfer on the small
scale: How do mushrooms shed their spores? Chemical Engineering Science
46(4): 1145-1149. (TP1 C67) [IMPORTANT PAPER. 3 mm/s descent speed in
air of free-falling spores.]
Vainshtein, B.K. 1986. Symmetry of biological macromolecules and their associations.
Computers and Mathematics with Applications 12B(1/2): 237-269. (Q172.5 S95
S94 1986)
Valberg, P.A. and Feldman, H.A. 1987. Magnetic particle motions within living cells.
Measurement of cytoplasmic viscosity and motile activity. Biophys. J. 52: 551561.
Vale, R.D. and Oosawa, F. 1990. Protein motors and Maxwell's demons: Does
mechanochemical transduction involve a thermal ratchet? Adv. Biophys. 26: 97134.
Vale, R.D. 1991. Severing of stable microtubules by a mitotically-activated protein in
Xenopus egg extracts. Cell 64: 8,27-839. [See esp. pp. 830-831, p. 834 on the
effect of MAPs.]
Vallee, R.B., Shpetner, H.S., and Paschal, B.M. 1990. Potential roles of microtubuleassociated motor molecules in cell division. In G.W. Conrad and T.E. Schroeder
(eds.) Cytokinesis: Mechanisms of Furrow Formation During Cell Division.
Annals of the New York Academy of Sciences, Vol. 582. New York.
Vogel, S. 1981. Life in Moving Fluids: The Physical Biology of Flow. Willard Grant,
Boston. [Added mass discussed on pp. 275-276.] (QH505 V63)
Vommaro, R.C., Attias, M., and De Souza, W. 1993. The cell surface charge of Bodo
sp. (Kinetoplastida: Bodonidae). Europ. J. Protistol. 29: 32-37.
Wagenknecht, T. and Bloomfield, V.A. 1975. Equilibrium mechanisms of length
regulation in linear protein aggregates. Biopolymers 14: 2297-2309. (QD471
.B6)
Wainwright, S.A., Biggs, W.D., Currey, J.D., and Gosline, J.M. 1976.
Design in Organisms. John Wiley, New York. (QP303 M4 1976)
e
Mechanical
Watters, C. 1968. Studies on the motility of the Heliozoa. I. The locomotion of
Actinosphaerium eichhorni and Actinophrys sp. J. Cell Sci. 3: 231-244.
Waugh, R.E. and Hochmuth, R.M.
1987.
Mechani~al equilibrium of thick, hollow,
109
liquid membrane cylinders. Biophys. J. 52: 391-400. [Study prompted by
observation of "tethers" extending from RBCs in shear flow.]
Waugh, R.E. and Erwin, G. 1989. Flexural rigidity of marginal bands isolated from
erythrocytes of the newt. J. Cell BioI. 108: 1711-1716. (May)
Waugh, R.E., Song, J., Svetina, S., and Zeks, B. 1992. Local and nonlocal curvature
elasticity in bilayer membranes by tether formation from lecithin vesicles.
Biophys. J. 61: 974-982.
Webster, J., Proctor, M.C.F., Davey, R.A., and Duller, G.A. 1988. Measurement of
the electrical charge on some basidiospores and an assessment of two possible
mechanisms of ballistospore propulsion. Trans. Br. mycoi. Soc. 91: 193-203.
Weis-Fogh, T. and Amos, W.B. 1972. Evidence for a new mechanism of cell motility.
Nature 236: 301-304.
Weis-Fogh, T. 1975. Principles of contraction in the spasmoneme of vorticellids. A
new contractile system. In L. Bolis, H.P. Maddrell, and K. Schmidt-Nielsen
(eds.) Comparative Physiology: Functional Aspects of Structural Materials.
North-Holland, Amsterdam, pp. 83-98. (QH510 157)
Weiss, D.G., Langford, G.M., and Allen, R.D. 1987. Implications of microtubules in
cytomechanics: Static and motile aspects. In J. Bereiter-Hahn, O.R. Anderson,
and W.-E. Reif (eds.) Cytomechanics: The Mechanical Basis of Cell Form and
Structure. Springer-Verlag, New York, pp. 100-?
Welch, G.R. 1992. An analogical "field" construct in cellular biophysics: History and
present status. Prog. Biophys. molec. BioI. 57: 71-128.
Wilson, D.E. 1966. The physiology, locomotion and taxonomy of some heliozoa and
helioflagellats. Ph.D. thesis, University of California, Los Angeles. [Not cited by
Davidson (1975). Summary published in Dissertation Abstracts International,
Vol. 27, Number 6, December 1966, 2202-B. I read this dissertation 1/93. A
shoddy piece of work befitting supervision by the unreliable T.L. Jahn (Wilson
suggests that the axonemal MTs are made of F-actin!), but contains many refs. to
old studies.]
Winet, H. 1973. Wall drag on free-moving ciliated microorganisms. J. expo BioI. 59:
753-766.
Wood, D.C.
1982.
Membrane permeabilities determining resting, action and
mechanoreceptor potentials in Stentor coeruleus. J. Compo Physioi. A 146: 537550. (QH301 Z46)
Wood, D.C. 1989. The functional significance of evolutionary modifications found in
the ciliate, Stentor. In P.A.V. Anderson (ed.) Evolution of the First Nervous
Systems. Plenum Press, New York, pp. 357-371. (QP356.15 N38 1989)
Wu, T.Y. 1977. Introduction to the scaling of aquatic animal locomotion. In T.J.
Pedley (ed.) Scale Effects in Animal Locomotion. Academic Press, New York,
.
pp.203-232. [IMPORTANT PAPER]
Yamada, K. and Asai, H. 1982. Extraction and some properties of the protein spastin
110
B from the spasmoneme of Carchesium polypinum. J. Biochem. 91: 1187-1195.
(QP501 .J6*)
Yanagida, T., Arata, T., and Oosawa, F. 1985. Sliding distance of actin filament
induced by a myosin crossbridge during one ATP hydrolysis cycle. Nature 316:
366-369.
Yates, G.T. 1986. How microorganisms move through water. Am. Sci. 74: 358-365.
(Q1 .A55)
Yoneda, M. 1973. Tension at the surface of sea urchin e~gs on the basis of "liquid
drop" concept. Advances in Biophysics 4: 153-190. [pp. 181-182: "Most of the
values for the tension at the surface of the unfertilized (sea urchin) egg lie
between 0.1 and 0.2 dyn/cm . . . . For an egg of about 100 jJm in diameter, a
tension at the surface of 0.1-0.2 dyn/cm corresponds to an internal pressure of 40
to 80 dyn cmo2 ." N.B. I do not know how these pressures were calculated.]
Yoneda, M. and Nagai, R. 1988. Structural basis of cytoplasmic streaming in
characean internodal cells. A hydrodynamic analysis. Protoplasma 147: 64-76.
Zhu, C. and Skalak, R. 1988. A continuum model of protrusion of pseudopod in
leukocytes. Biophys. J. 54: 1115-1137.
Zhu, C., Skalak, R., and Schmid-Schonbein, G.W. 1988. One-dimensional steady
continuum model of retraction of pseudopod in leukocytes. J. Biomech. Eng.
111: 69-77.
Zigmond, S.H. 1993. Recent quantitative studies of actin filament turnover during
cell locomotion. Cell Motility and the Cytoskeleton 25: 309-316.
Zimmerman, B., Moestrup, 0., and Hallfors, G. 1984. Chrysophyte or heliozoan:
Ultrastructural studies on a cultured species of Pseudopedinella (Pedinellales ord.
nov.), with comments on species taxonomy. Protistologica 20(4): 591-612.
[Discussion and EM images of MT triads supporting the tentacles.]