Sei sulla pagina 1di 34

DE PARIS SUD

UNIVERSITE
U.F.R SCIENTIFIQUE DORSAY
Licence et Magist`
ere de Physique
2014-2015

Classical Electromagnetism
-

Lecture Notes

26

CHAPTER 1. Introduction and overview - Electromagnetism

1.8

Additional reading

The following article 16 , describes one application of radiation pressure.

16. Dachwald et al., Int. AAAF Symposium on Space Propulsion, 2002.

More than 55% of the NEO population has inclinations


larger than 10 , more than 30% has inclinations larger than
20 . Reaching such inclinations with spacecraft requires a very
large v.

Utilizing solely the freely available solar radiation


pressure for propulsion, solar sailcraft provide a wide
range of opportunities for low-cost interplanetary missions, many of which are difficult or impossible for any
other type of conventional spacecraft due to their large
v-requirement. Many of those high-energy missions
are of great scientific relevance, such as missions to
Mercury and to Near Earth Objects (asteroids and
short period comets) with highly inclined or retrograde orbits1 . Within the inner solar system (including the main asteroid belt) solar sailcraft are especially
suited for multiple rendezvous and sample return missions due to their (at least in principle) unlimited vcapability. Even missions to the outer solar system
may be enhanced by using solar sailcraft, albeit the
solar radiation pressure decreases with the square of
the sunsail distance. For such missions solar sailcraft may gain a large amount of energy when first
approaching the sun, thereby performing a so-called
solar photonic assist maneuver that turns the trajectory into a hyperbolic one [4][5][13]. Such trajectories
allow reasonable transfer times to the outer planets
(and to near interstellar space) without the need to
perform any gravity assist maneuver. However, without the use of additional propulsive devices and/or an
aerocapture maneuver at the target body, only fast
fly-bys can be achieved due to the associated large
hyperbolic excess velocities.

Introduction

The magnitude and direction of the solar radiation


pressure (SRP) force acting on a flat solar sail due
to the momentum transfer from solar photons is completely characterized by the sun-sail distance and the
sail attitude. The latter is generally expressed by the
sail normal vector n, whose direction is usually described by the sail clock angle and the sail cone
angle (Figure 1). Figure 2 gives a picture of the
forces exerted on a flat and perfectly reflecting solar
sail (ideal sail) of area A by the solar radiation pres-

Solar Sailcraft Orbital Mechanics

Several mission studies for high-energy interplanetary solar sailcraft missions have been carried out at
DLR [4][5][8][9] and elsewhere [13][15]. Most of them
require a rather demanding sailcraft performance to
keep mission durations short (see Table 1). However,
taking the current state-of-the-art in engineering of
ultra-lightweight structures into account, solar sailcraft of the first generation will be of relatively moderate performance. For such near-term solar sailcraft
few mission examples can be found in the literature.
The aim of this paper is to narrow down this gap
and to get a lower bound on solar sailcraft performance for interplanetary missions that are under consideration. It will be shown, that challenging scientific missions are feasible at relatively low cost, even
with moderate performance sailcraft of the first generation. This will be demonstrated below by the trajectory analysis of a proposed sample return mission
to Near Earth Asteroid 1996FG3 (mission duration
approx. 9.4 years).

Solar sailcraft provide a wide range of opportunities for high-energy low-cost


missions. To date, most mission studies require a rather demanding performance that will not be realized by solar sailcraft of the first generation.
However, even with solar sailcraft of moderate performance, scientifically
relevant missions are feasible. This is demonstrated with a Near Earth Asteroid
sample return mission and various planetary rendezvous missions.

Universitat der Bundeswehr M


unchen, Neubiberg
Institut f
ur Raumfahrttechnik
Phone: +49-89-6004 2138 Fax: +49-89-6004 2138
E-Mail: bernd.haeusler@unibw-muenchen.de

1
German Aerospace Center (DLR), Cologne
Institute of Space Sensor Technology and Planetary Exploration
Phone: +49-2203-601 {3001|3028} Fax: +49-2203-601 4655
E-Mail: {bernd.dachwald|wolfgang.seboldt}@dlr.de

ausler2
Bernd Dachwald1 , Wolfgang Seboldt1 and Bernd H

PERFORMANCE REQUIREMENTS FOR NEAR-TERM


INTERPLANETARY SOLAR SAILCRAFT MISSIONS

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

Sailcraft performance
ac [ mm/s2 ] [ g/m2 ]
0.5
16.0
0.7
11.4
0.75
10.7
0.85
9.4
1.0
8.0
1.0
8.0
1.0
8.0
1.0
8.0
1.25
6.4

Transfer time
[ yr]
1.4
10.4
3.3
3.0
6.8
0.6
1.0
1.2
1.2
[13]
[4][5]
[4][8]
[4]
[15]
[13]
[13]
[13]
[15]

References

P A( er n) er

= 2P A cos2 n

= 2P A( er n)2 n =

FSRP = Fr + Fr =

and making use of er er = 2( er n) n:

Fr = P A( er n) er

Fr =

sure P acting on the sails center of surface. From the


geometry of Figure 2 the total SRP force FSRP can be
easily calculated:

Figure 2: Perfect reflection

Figure 1: Definition of the sail clock angle and the


sail cone angle

for the respective SRP force in a distance r from the


sun. Thus, to experience a reasonable acceleration,
solar sailcraft must be large and very lightweight.
The orbital dynamics of solar sailcraft is in many respects similar to the orbital dynamics of other spacecraft, where a small continuous thrust is applied to
modify the spacecrafts orbit over an extended period
of time. However, other continuous thrust spacecraft
may orient its thrust vector in any desired direction

for the effective pressure acting at 1 AU on a solar sail


that is oriented normal to the sun-line and


1 AU 2
FSRP = (Peff )1AU
A cos2 n
r

.
(Peff )1AU = 2(P0 )1AU = 7.757 106 N/m2

Thus, in the case of perfect reflection, the thrust


force is always along the direction of the sail normal
vector n. At 1 AU, the solar radiation pressure is
.
(P0 )1AU = 4.563 106 N/m2 . Therefore, the effective pressure (force per unit area) acting on an ideal
sail normal to the sun-line is twice the solar radiation
.
pressure, 2(P0 )1AU = 9.126 106 N/m2 . However, a
real solar sail is not a perfect reflector and a thorough trajectory analysis must take into account the
optical properties of the real sail. Since in this case a
small but significant fraction of the incoming sunlight
is absorbed or reflected non-specularly, a tangential
force component is acting on a real solar sail, so that
FSRP is no longer along the direction of n. However, for preliminary mission analysis this tangential
force component may be neglected, since the resulting
small angular deviation of FSRP from the sail normal
can be compensated by the sail steering strategy for
interplanetary transfer trajectories (where > 55.5
is not required [10]). Nonetheless, an overall sail efficiency parameter should be used, which takes into
account the reduced magnitude of FSRP due to the
non-perfect reflectivity of the sail including its deflection/warping under load. Assuming a conservative
sail efficiency of 0.85 (aluminum coated plastic
film), we get

Table 1: Fast solar sailcraft missions using advanced sailcraft (rendezvous, if not stated otherwise)

Mercury
Pluto (fly-by)
(4) Vesta
2P/Encke
21P/Giacobini-Zinner
Venus
Mars
(433) Eros
(1566) Icarus

Target body

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

ms
A

mp
ms + mp
m
=
= s +
A
A
A

the characteristic acceleration ac is defined as the


maximum acceleration at 1 AU solar distance. It

is defined accordingly as the specific mass of sailcraft including the payload (index p). It should
be noted, that the term payload stands for the
total sailcraft except the solar sail assembly (i.e.
except the propulsion system).

the sailcraft loading

is defined as the mass of the sail assembly (the


sail film and the required structure for storing,
deploying and tensioning the sail, index s) per
unit area. Thus, the sail assembly loading is the
key parameter for the performance of a solar sail
and the efficiency of its structural design.

s =

the sail assembly loading

Before talking about performance of near-term solar


sailcraft, the most common performance definitions
should be given. The performance of solar sailcraft
may be expressed by the following parameters:

Solar Sailcraft Performance


Parameters

Figure 3: Spiralling inwards and outwards

and vary its thrust level within a wide range, whereas


the thrust vector of solar sailcraft is constrained to lie
on the surface of a bubble directed away from the
sun (see Figure 3). Nevertheless, by controlling the
sail orientation relative to the sun, solar sailcraft can
gain orbital angular momentum (if FSRP et > 0) and
spiral outwards away from the sun or lose orbital
angular momentum (if FSRP et < 0) and spiral inwards towards the sun.

ac
5.93 mm/s2

The square solar sail consisted of four CFRP (Carbon Fiber Reinforced Plastics) booms with a specific
mass of 101 g/m and of four triangular sail segments
made of aluminum-coated (0.1 m) plastic films with

Figure 4: Fully deployed 20 m20 m solar sail at DLR

In December 1999 a ground-based demonstration of


solar sailcraft technology was performed at the German Aerospace Center (DLR) at Cologne, where a
20 m 20 m solar sail was successfully deployed in a
simulated zero-g environment and ambient environmental conditions (Figure 4) [6][14].

DLR Ground-based Demonstration of


Solar Sail Technology

where = GMsun .

Using the lightness number, the SRP force acting


on the sail can be written as
m
FSRP = 2 cos2 n
r

the lightness number , which is independent


from solar distance, is defined as the ratio of the
SRP acceleration experienced by a solar sail normal to the sun line and the solar gravitational
acceleration (5.93 mm/s2 at 1 AU)

Using the characteristic acceleration, the SRP


force acting on the sail can be written as


1 AU 2
FSRP = mac
cos2 n
r

(Peff )1AU A = mac = Aac =


mp
)Aac
= (s +
A
(Peff )1AU
ac =
m
s + Ap

can be calculated via

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

Near Earth Asteroids (NEAs) are a promising category of target bodies for a first solar sailcraft mission,
since they can be accessed relatively easily and since
they are of great scientific interest. Therefore, in August 2000, a dedicated mission for the exploration of
NEAs with solar sailcraft (ENEAS) was proposed by
DLR in cooperation with the Westf
alische WilhelmsUniversit
at at M
unster (Germany) as a candidate

DLR Mission Proposal for ENEAS

The booms were rolled up in a 60 cm 60 cm


65 cm-sized deployment module, from where they
unfolded automatically. After deployment they returned to their tubular shape with high bending and
buckling strength. Subsequently, the four sail segments were deployed by ropes. To assess the handling behavior of different sail materials, the sail segments were made of three different aluminum-coated
plastic films, 12 m polyethylene terephtalate (PET,
Mylarr ), 7.5 m polyimide (PI, Kaptonr ) and 4 m
polyethylene naphthalate (PEN). All segments were
reinforced along the three edges of the triangle to
prevent rips. The specific mass of the sail film
was 18.9 g/m2 for the Mylarr -segment, 12.4 g/m2 for
the Kaptonr -segment and 10.5 g/m2 for the PENsegment. The deployment module and the cross section of the booms for this ground-based demonstration
were dimensioned for a 40 m 40 m solar sail, which
was too large for an in-door demonstration. For the
structural sizing of the booms two load cases were considered, bending due to the SRP force and buckling
due to sail deployment and sail tensioning forces.
According to FEM (Finite Element Method) calculations, similar booms could be used also for larger sails
[7].

Figure 5: DLR deployable CFRP boom

a thickness between 4 and 12 m. The booms consisted of two CFRP shells that were bonded at the
edges to form a tubular shape, so that they can be
pressed flat and rolled up (Figure 5).

FSRP,c

A
s
ms
mp
m

ac

(50 m)2
29.2 g/m2
73 kg
65.5 kg
138.5 kg
55.4 g/m2
1/42.4
0.14 mm/s2
19.5 mN

For propulsionless attitude control, the solar sail


and the micro-satellite would be separated by a
commercially available 10 m collapsible control mast,
which is housed inside the deployment module in its
stowed configuration. This control mast is attached to
the deployment module via a two degree of freedom
actuator gimbal, which allows to rotate the mast including the attached micro-satellite with respect to
the sail (Figure 6). Thus, by rotating the control
mast, the center of mass (CM) can be offset from
the center of pressure (CP). The resulting external
torque may be used to rotate the sail about any CMintersecting axis parallel to the sail plane (this attitude control concept was originally proposed by [1]).
1996FG3 was chosen as the target body for the
ENEAS mission, since 1996FG3 has orbital elements
not too different from that of Earth and since it is an
object of exceptional scientific interest. Observations
indicate that 1996FG3 is a binary asteroid, consisting of a central body with a rotation period of about
3.60 hours and a satellite with an orbital period of
about 16.15 hours. The determined average bulk density is 1.4 0.3 g/cm3 which is highly suggestive of
a rubble pile structure [12]. ENEAS is intended to

Table 2: Parameters for the ENEAS solar sailcraft

Sail area
Sail assembly loading
Sail assembly mass
Payload mass
Total sailcraft mass
Sailcraft loading
Lightness number
Characteristic acceleration
Characteristic SRP force

Figure 6: DLR ENEAS solar sailcraft with deployed


control mast (artists view)

within the German small satellite program for extraterrestric sciences [3][14]. Based on the successful deployment experiment described above, ENEAS
(Figure 6) was intended to feature a deployable 50 m
50 m solar sail that would be capable to transport a
micro-satellite with a mass of 65.5 kg to a NEA within
less than five years. Table 2 summarizes the ENEAS
parameters.

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

(Peff )1AU
m ,
s + s2p

one can see that the performance depends on three


design parameters, the sail assembly loading s , the
payload mass mp and the side length s (or area s2 ) of
the solar sail, defining a three-dimensional solar sail
design space. Figures 8 and 9 show parametric sections of this design space for a fixed s = 29.2 g/m2

ac =

Looking at the equation for the characteristic acceleration of solar sailcraft with a square sail,

Near-Term Solar Sailcraft


Performance

The flight time could be reduced by 45 days (3%),


reducing at the same time the C3 requirement from
4 km2 /s2 to 0 km2 /s2 , thus permitting a reduction of
launch cost. The accuracy of the trajectory generated
by the artificial neural network is r < 11 000 km for
the relative distance to the target body at rendezvous
and v < 43 m/s for the relative velocity (even without performing a local fine tuning of the trajectory)
[2].

Figure 7: 1996FG3 rendezvous trajectory for


ac = 0.14 mm/s2

determine the physical properties and the evolution of


the 1996FG3 system.
Trajectory optimization using the calculus of variations revealed, that the ENEAS sailcraft can reach
1996FG3 in 4.5 years (1640 days), if it is inserted directly into an interplanetary trajectory with a hyperbolic excess energy of C3 = 4 km2 /s2 . However, more
recently performed trajectory optimization, based on
artificial neural networks and evolutionary algorithms
produced a better trajectory for the same launch date,
which is closer to the (unknown) global optimum (Figure 7) [2].

It should be noted that mp and s can be chosen independently, whereas s (s) is a function of s with s /s < 0, since
the mass of the booms and the deployment module scale less
than linearly with the sail area. However, we are on the safe
side, when we assume s /s = 0 to keep calculations simple.

By different combinations of the three design parameters any desired characteristic acceleration can
be achieved2 . An increase in payload mass can, for
example, be offset with a proportional increase of s2
or with a (not inversely proportional) decrease of s .

Figure 9: The characteristic acceleration ac as a


function of s and mp for s = 50 m

Figure 8: The characteristic acceleration ac as a


function of s and mp for s = 29.2 g/m2

and a fixed s = 50 m respectively (as for the ENEAS


sailcraft). As can be seen from the diagram in Figure
8, a characteristic acceleration of up to 0.265 mm/s2
can be achieved without any payload. For a smaller
ac a positive payload mass can be accommodated, depending on the sail size. To achieve a characteristic
acceleration beyond 0.265 mm/s2 , the sail assembly
loading has to be further reduced (Figure 9).

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

1
1 + mp /s s2
1
m p =
1 + s s2 /mp
2
s =
1 + s s2 /mp
s =

3
similar to the Rosetta mission to comet 46P/Wirtanen,
which will have three intermediate gravity assist maneuvers
(Mars-Earth-Earth) and a trip time of approximately nine years

Q4: What is the maximum payload mass to get ac,min


for the specified s and s?

Q2: What is the minimum characteristic acceleration


ac,min to perform the mission in Tmax ?

Q1: What is the maximum acceptable mission duration Tmax ?

Q3: What is the expected sail assembly loading s and


sail dimension s for near-term solar sailcraft?

Figure 10: 1996FG3 rendezvous trajectory for


ac = 0.10 mm/s2

Answer to Q1: At present, the maximum acceptable mission duration seems to be determined by the
trip time required with chemical propulsion, including (eventually multiple) gravity assist maneuvers.
Due to the relatively large v-requirement of about
6 10 km/s for a mission comparable to ENEAS-SR,
but with chemical propulsion, such a mission would
require either an expensive launch vehicle and heavy
spacecraft, resulting in a short trip time of a few years,
or several gravity assists, resulting in a long trip time3 .
Since our approach aims at low-cost missions, only the
gravity assist option seems to be a reasonable conventional alternative. Thus, for the ENEAS-SR mission
we assume a total mission duration of more than ten
years as not acceptable.
Answer to Q2: Trajectory calculations show, that
an ENEAS-SR mission to 1996FG3 can be achieved
even with a characteristic acceleration of 0.10 mm/s2
in 9.40 years, including a rendezvous trajectory of 6.27
years (2290 days, Figure 10), 340 days of operations
at the asteroid and an Earth return trajectory of 2.20
years (805 days, Figure 11).

Answer to Q3: The diagram in Figure 12 shows the


required sail size for different sail assembly loadings
and payload masses, to obtain a characteristic acceleration of 0.10 mm/s2 . Based on the experiences with
the ground-based solar sail technology demonstration
described above, we consider a maximum sail size of
70 m 70 m with a sail assembly loading of 29.2 g/m2
(sail film + booms + deployment module) as a realistic however still challenging baseline for the
ENEAS-SR mission.
Answer to Q4: The specified s and s yield a payload mass of 237 kg to get a characteristic acceleration

The ENEAS sailcraft was intended to rendezvous


1996FG3 for remote sensing with a minimum scientific
payload mass of 5 kg (CCD camera + IR spectrometer
+ magnetometer). To study the 1996FG3 system in
more detail, it would be necessary to place a lander on
the surface of the asteroid (e.g. for mass spectrometry
and/or alpha-proton spectrometry). Some investigations (e.g. micro-structure and isotope analysis) to determine the age and the evolution of 1996FG3 could
be achieved only by taking samples of the asteroid
back to Earth. Due to their unlimited v-capability,
solar sailcraft are especially capable to perform such
sample return missions. However, compared to the
ENEAS rendezvous mission, the payload mass has to
be increased considerably. The key questions for the
ENEAS-SR (sample return) mission design are:

ENEAS with Sample Return

For the ENEAS sailcraft, we have s = 0.537,


mp = 0.473 and s = +0.946. As can be seen,
the side length of the sail is the most critical parameter with respect to the ENEAS sailcraft performance.
Thus, an increase in performance is best done by increasing the size of the solar sail.
If costs can be described by (known or estimated)
functions of the three design parameters, then the optimum (cost minimal) sailcraft design for a given performance can be determined.

with

ac

=
ac

Those design sensitivities can be determined quantitatively using sensitivity functions, which provide an
indication of the relative importance of each design
parameter for a given point in the solar sail design
space [10]. The sensitivity function for any design parameter {s , mp , s} may be written as

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

Since for solar sailcraft of moderate performance gaining orbital energy in the Earths gravitational field is
difficult and time consuming, the launcher will insert
the ENEAS-SR solar sailcraft directly into an interplanetary trajectory with a hyperbolic excess energy
of C3 = 0 km2 /s2 . After the injection, the sail and

ENEAS-SR Mission Scenario

of 0.10 mm/s2 . Current research at our department


indicates that it should be possible to realize such a
mission within the specified mass budget, including a
lander of about 140 kg and a sample return capsule
of about 40 kg. Table 3 summarizes the ENEAS-SR
parameters.

Figure 12: The side length s of the solar sail that is


required to achieve a characteristic acceleration of 0.10 mm/s2 as a function of s
and mp

Figure 11: 1996FG3 sample return trajectory for


ac = 0.10 mm/s2

FSRP,c

s
ms
mp
m

ac

29.2 g/m2
143 kg
237 kg
380 kg
77.6 g/m2
1/59.3
0.10 mm/s2
38.0 mN

Those quasi-stationary hovering positions are unstable but can be stabilized using a feedback control
loop to sail attitude alone [11]. Hovering near the asteroid, the (likely complex) gravitational field of the
target body is studied, so that a coarse gravitational
field model can be determined. Thereafter, the lander with the Earth return capsule is separated from

Figure 13: Hovering at the asteroid

the attitude control mast are deployed in a 3-axis stabilized mode. Then the sail is oriented to follow a
pre-calculated attitude profile, leading to an optimal
interplanetary transfer trajectory. During the transfer, the ENEAS-SR solar sailcraft will run almost autonomously, so that ground monitoring will be carried
out on a weekly basis only. At the end of the transfer
trajectory the solar sailcraft will be making a rendezvous with 1996FG3 within its gravitational sphere
of influence (Hill-sphere) of between 70 km radius (at
perihelion) and 150 km radius (at aphelion). Even in
the near-field of the asteroid, the SRP acceleration
of between 0.05 mm/s2 (at aphelion) and 0.21 mm/s2
(at perihelion) is larger than the asteroids gravitational acceleration (0.01 to 0.00005 mm/s2 in a distance ranging from 5 to 50 km), so that the sailcraft
is able to hover on an artificial equilibrium surface in
the hemisphere that is opposite to the sun (Figure 13).

Table 3: Parameters for the ENEAS-SR solar sailcraft

Sail assembly loading


Sail assembly mass
Payload mass
Total sailcraft mass
Sailcraft loading
Lightness number
Characteristic acceleration
Characteristic SRP force

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR
Sail area
A
(70 m)2

We have investigated the minimum solar sailcraft performance requirements for various interplanetary missions. We were able to show, that the characteristic

Summary

We have investigated the performance of near-term solar sailcraft also for rendezvous missions with celestial
bodies other than Near Earth Objects. Table 4 gives
the minimum rendezvous times for solar sailcraft with
a characteristic acceleration of 0.10 to 0.20 mm/s2 for
several target bodies. It shows that even with nearterm solar sailcraft planetary rendezvous mission are
feasible within the inner solar system, if relatively long
trip times can be tolerated. However, a characteristic
acceleration of 0.10 mm/s2 seems to be a lower bound
for rendezvous missions within the inner solar system.

Other Promising Missions for


Near-Term Solar Sailcraft

the solar sail to go into closer orbit about the asteroid. While measuring the asteroids gravitational field
with increasing accuracy, the orbit of the lander is continuously lowered until a safe landing trajectory can
be computed (some or all of those extensive computations may be performed on Earth). Once landed,
the sample is fed directly into the Earth return capsule and brought back to the hovering sailcraft. In
this mission phase, the sailcraft is waiting edge-on
(so that no SRP force is acting on the sail) at the
L2 Lagrange point for the lander in order to assist
the rendezvous. The lander design, the sample extraction mechanisms and the subsystems required to
rendezvous the waiting sailcraft require further studies and are beyond the scope of this paper. Since
1996FG3 is a binary system, it would be interesting
to land and extract samples from both bodies to investigate the origin and the collisional evolution of the
1996FG3 system. Since the gravitational acceleration
is very low near the asteroid and the required v for
the lander less than 10 m/s, a cold gas system with a
propellant mass of less than 4 kg will suffice to perform all operations. After rendezvous with the hovering sailcraft, the re-docked ENEAS-SR solar sailcraft
returns the sample to Earth. The return trajectory is
much faster than the transfer trajectory to 1996FG3
since no rendezvous is required at Earth. Thus, the
sailcraft may arrive with a relatively large hyperbolic excess velocity of about 8.4 km/s. The gravitational acceleration of Earth adds another 11.2 km/s,
so
that the Earth reentry velocity may reach about
8.42 + 11.22 km/s = 14.0 km/s. Finally, just before
the arrival of the ENEAS-SR solar sailcraft at Earth,
the return capsule is separated from the lander and
injected into an Earth reentry trajectory, where it
is decelerated by atmospheric friction and breaking
parachutes.
20.1
6.8
6.2

Mercury
Venus
Mars

Transfer time [ yr]


for ac [ mm/s2 ]
0.10 0.15 0.20
8.3
5.9
4.2
4.6
2.9
2.0
9.2
7.5
5.1

[4] M. Leipold. Solar Sail Mission Design. Doctoral thesis, Lehrstuhl f


ur Flugmechanik und
Flugregelung; Technische Universitat M
unchen,
1999. DLR-FB-2000-22.

[3] E. K. Jessberger and W. Seboldt et al. ENEAS


exploration of near-earth asteroids with a sailcraft. Proposal for a Small Satellite Mission
within the Extraterrestric Sciences Program of
DLR, August 2000.

[2] B. Dachwald and W. Seboldt. Optimization of


interplanetary rendezvous trajectories for solar
sailcraft using a neurocontroller. Monterey, August 2002. AIAA/AAS Astrodynamics Specialist
Conference. AIAA-2002-4989, in preparation.

[1] F. Angrilli and S. Bortolami. Attitude and orbital


modelling of solar-sail spacecraft. ESA Journal,
14:431446, 1990.

References

acceleration must be at least 0.10 mm/s2 in order to


avoid unacceptable long mission durations even for
relatively easily accessible inner solar system bodies.
A 70 m 70 m solar sail with a sail assembly loading
of 29.2 g/m2 (sail film + booms + deployment module) was considered to be a realistic however still
challenging near-term baseline. With this solar sail,
a characteristic thrust of 38 mN can be achieved. The
characteristic acceleration defining the mission duration depends on the actual payload mass mp and
ranges from 0.265 mm/s2 (mp = 0 kg) to 0.10 mm/s2
(mp = 237 kg). We have also demonstrated, that a
sample return mission to a Near Earth Asteroid with
such a solar sail is feasible within a mission duration of
approx. 9.4 years. In addition to the scientific value
of such a mission, the demonstration of the technical capabilities of solar sail propulsion in deep space
would be a central objective.

Table 4: Minimum transfer times to the inner planets


for solar sailcraft with a characteristic acceleration of 0.10, 0.15 and 0.20 mm/s2 (vmin
denotes the minimum impulsive v-values
for elliptical non-coplanar Hohmann-like orbit transfers with zero hyperbolic excess velocities at both ends of the trajectory)

vmin
[ km/s]

Target
body

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

[15] J. L. Wright. Space Sailing. Gordon and Breach


Science Publishers, Philadelphia, 1992.

[14] W. Seboldt, M. Leipold, M. Rezazad, L. Herbeck, W. Unkenbold, D. Kassing, and M. Eiden.


Ground-based demonstration of solar sail technology. Rio de Janeiro, 2000. 51st International
Astronautical Congress. IAF-00-S.6.11.

[13] C. G. Sauer. Optimum solar-sail interplanetary trajectories.


San Diego, August 2000.
AIAA/AAS Astrodynamics Conference. 76-792.

[12] S. Mottola and F. Lahulla. Mutual eclipse events


in asteroidal binary system 1996FG3 : Observations and a numerical model. Icarus, 146:556
567, 2000.

[11] E. Morrow, D. J. Scheeres, and D. Lubin. Solar sail orbit operations at asteroids. Journal of
Spacecraft and Rockets, 38:279286, 2001.

[10] C. R. McInnes. Solar Sailing. Technology, Dynamics and Mission Applications. Springer
Praxis Series in Space Science and Technology.
SpringerPraxis, Berlin, Heidelberg, New York,
Chicester, 1999.

[9] M. Leipold, W. Seboldt, S. Lingner, E. Borg,


A. Herrmann, A. Pabsch, O. Wagner, and
J. Br
uckner. Mercury sun-synchronous polar orbiter with a solar sail. Acta Astronautica, 39(14):143151, 1996.

[8] M. Leipold, E. Pfeiffer, P. Groepper, M. Eiden,


W. Seboldt, L. Herbeck, and W. Unkenbold. Solar sail technology for advanced space science
missions. Toulouse, 2001. 52nd International Astronautical Congress. IAF-01-S.6.10.

[7] M. Leipold, C. E. Garner, R. Freeland, A. Herrmann, M. Noca, G. Pagel, W. Seboldt,


G. Sprague, and W. Unckenbold. ODISSEE
a proposal for demonstration of a solar sail in
earth orbit. Acta Astronautica, 45(4-9):557566,
1999.

[6] M. Leipold, M. Eiden, C. E. Garner, L. Herbeck, D. Kassing, T. Niederstadt, T. Kr


uger,
G. Pagel, M. Rezazad, H. Rozemeijer, W. Seboldt, C. Sch
oppinger, C. Sickinger, and W. Unkenbold. Solar sail technology development and
demonstration. Laurel, 2000. 4th IAA International Conference on Low-Cost Planetary Missions. IAA-L-0707.

[5] M. Leipold. To the sun and pluto with solar sails


and micro-sciencecraft. Acta Astronautica, 45(49):549555, 1999.

6th International AAAF Symposium on Space Propulsion: Propulsion for Space Transportation of the XXIst Century, Versailles, France, 14-16 May 2002
Copyright (c) 2002 by DLR

Chapter 2

Electrostatics
Introduction
Chapter 1 contains in principle all laws and properties necessary to solve any problem in electromagnetism. The rest of the course will therefore be applications of these laws and properties in
specific cases.

As stated in section 1.1.4.a, in the static regime (or steady state), where the sources {, J } do
not vary in time, the coupling between the electric and magnetic field vanishes. The electric field
is only due to static charges, whereas the magnetic field is only created by constant currents. One
can therefore study the electric field - Electrostatics - and the magnetic field - Magnetostatics
- separately.
In this chapter, we will concentrate on Electrostatics. In section 2.1, we will briefly go over the
relations valid in electrostatics. These are only the retranscription of the relations seen in chapter
1 in the case of time-independent sources.

2.1
2.1.1

Maxwells equations applied to Electrostatics and consequences


Coulomb force

The force exerted by a charge q on a charge q , at a distance r apart is :

F =

qq

u qq
40 r2

(2.1)

where
u qq is the unit vector going from the source charge q to the charge q and 0 the vacuum
permittivity or permittivity of free space :
0 8, 85 1012 F.m1 = 8, 85 1012 kg1 .m3 .A2 .s4
In the case of multiple charges {qi }i acting on a charge q , Coulombs law can be written as:

F =

q X qi

ui
40 i ri2

(2.2)

where ri is the distance between the source charge qi and the charge q , and
ui is the unit vector

going from the source charge qi to the charge q .


32

2.1. MAXWELLS EQUATIONS APPLIED TO ELECTROSTATICS AND CONSEQUENCES33


In the case of a distribution of charges {(P )} acting on a charge q placed at point M ,
Coulombs law can be written as:

q
(P )
F (M ) =
P M dV
(2.3)
40
P
M3
(V )
Note that these three expressions of Coulombs law are strictly identical, albeit the form they
are written in has been chosen for convenience.

2.1.2

Maxwells equations in electrostatics

The decoupling between the electric and magnetic fields reduces the number of equations to be
taken into account. The Maxwells equations of electrostatics are :
Maxwell-Gauss

Maxwell-Faraday

2.1.3

Electrostatic field

2.1.3.a

Electrostatic field

div E = . E =
0


rot E = E = 0

(2.4)
(2.5)

The field created by a charge q at point M situated at a distance r is :

E (M ) =

ur
40 r2

(2.6)

where
u r is the unit vector going from the source charge q to point M.
In the case of multiple charges {qi }, the field created at point M can be written as:

E (M ) =

1 X qi

ui
40 i ri2

(2.7)

where ri is the distance between the source charge qi and point M , and
ui is the unit vector going
from the source charge qi to point M .

as:

In the case of a distribution of charges {(P )}, the field created at point M can be written

(P )
1
P M dV
(2.8)
E (M ) =
40
P
M3
(V )

2.1.3.b

Typical values

The table below gives the orders of magnitude for a few electric fields :

FM antenna at 100 km distance


100 W bulb at 1 meter distance
Between the atmosphere and the ground
Laser pointer
Shorting in air
High intensity laser

0.5 mV.m1
50 V.m1
1 - 100V.m1
102 V.m1
106 V.m1
108 - 1012 V.m1

34

CHAPTER 2. Electrostatics

2.1.3.c

Conservation of charge

The continuity equation (or law of conservation of charge), in statics, gives:

div J = 0

(2.9)

J is then a vector with conservative flux.

Figure 2.1: Flux tube for the current density.


Proof : Figure 2.1 shows a flux tube for the current density. This means that for all points on


non-hatched surface of (), J .dS = 0. In other words, no current density goes in or out of the
tube except at the two ends S1 and S2 . Let us moreover suppose that the flux tube is sufficiently

at all points of S , with


the outgoing normal to S , and
at
small to have J = J1
n
n
J = J2
n
1
1
1
1
2

all points of S2 , with n2 the outgoing normal to S2 . Then :

0 =
div J
(V )


=
J .dS
(S )+()+(S2 )
1


=
J .dS +
J .dS +
J .dS
(S1 )

()

(S2 )

J 1 S1 J 2 S2

Hence J1 S1 = J2 S2 . The flux of current density going through S1 is conserved when going out of
S2 .

This property is found for all vectors A having div A = 0.


2.1.3.d

Gauss law

Let us consider a closed surface (S) delimiting a volume (V ) which contains a charge volume
density and a total charge Qint . Maxwell-Gausss equation gives Gauss law:

Qint
E .dS =
dV =
(2.10)
0
(S)
(V ) 0
This relation is one of the most useful relations in electrostatics. As we will see in section 2.2, along
with the use of symmetries and invariances, it allows to determine the electric field in numerous
simple situations.

2.1. MAXWELLS EQUATIONS APPLIED TO ELECTROSTATICS AND CONSEQUENCES35


2.1.3.e

Circulation of electrostatic field

Faradays law in electrostatics gives the circulation of electrostatic field :


E . dl = 0

(2.11)

(C)

There is no work of the electrostatic force around a circular contour.


2.1.3.f

Discontinuity equations at interfaces

The discontinuity equation for the electrostatic field is :

n 12
E2 E1 =
0

(2.12)

where E i is the electric field in the medium i, is the surface charge density at the interface

between the two media, and


n 12 is the unitary vector, normal to the interface, directed from
medium (1) to medium (2).

2.1.4

Electrostatic potential

2.1.4.a

Electrostatic potential

Definition : The electrostatic potential is defined by

E = grad

(2.13)

or equivalently :
(A) (B) =


E . dl

(2.14)

The potential created by a charge q at point M situated at a distance r is :


(M ) =

q
40 r

(2.15)

In the case of multiple charges {qi }i , the potential created at point M can be written as:
(M ) =

1 X qi
40 i ri

where ri is the distance between the source charge qi and point M .

(2.16)

36

CHAPTER 2. Electrostatics

In the case of a distribution of charges {(P )}, the potential created at point M can be
written as:

1
(P )
(M ) =
dV
(2.17)
40
(V ) P M
Note : The electrostatic potential (r) of a group of charges vanishes at infinity. This is the
origin of the convention we will often use :
(r) 0 when r +
2.1.4.b

(2.18)

Potential propagation

In electrostatics, the potential propagation (see section 1.3.4) is written under the form of
Poissons equation :

=
(2.19)
0
In the specific case where there is no charge ( = 0), the propagation equation reduces to
Laplaces equation :
= 0
(2.20)
The Poisson (or Laplace) equation will have an unique solution if boundary conditions are
specified on a closed surface :
Either the potential is defined on a closed surface. This is called the Dirichlet boundary
condition.
Or the electric field is defined on a closed surface. This is called the Neumann boundary
condition.
Proof : Let us suppose that 1 and 2 obey to the same Poisson equation = 0 inside a
volume (V ). Let :
U = 2 1
Let us impose Dirichlet boundary condition on the closed bounding surface (S) of the volume (V ).
Then, inside (V ) :

U = 2 1 = +
=0
0
0
and on (S), U = 0.
We now need to prove Greens first identity :



+ grad . grad dV =
(V )

(S)

dS
n

For this, let us apply the divergence theorem ( (V ) div A dV =


A .dS =
A .
n dS) to A =

grad :


 

grad .
n dS
div grad dV =
(S)
(V )


+ grad . grad dV =
dS
(V )
(S) n

2.1. MAXWELLS EQUATIONS APPLIED TO ELECTROSTATICS AND CONSEQUENCES37


Now, let us apply Greens first identity to = = U :


 2 
U U + grad U
dV =
(V )

(S)

U
dS
n

Given the boundary conditions and the fact that U = 0 in (V ) :



2
grad U dV = 0
(V )

which means that grad U = 0 inside (V ). Hence U is constant in (V ) and since U = 0 on (S),
U = 0 in all (V ). Therefore 1 = 2 and the solution is unique.

2.1.5

Field lines and equipotential surfaces

2.1.5.a

Field lines

Definition : Electric field lines are the lines that are, at all points in space, tangent to

the electric field E . They can be determined through the relation :

E (M ) dl = 0

(2.21)

where dl is an infinitesimal vector along the field line, centered around point M .
2.1.5.b

Equipotential surfaces

Definition : Equipotential surfaces are surfaces where the scalar potential (M ) is


constant.
2.1.5.c

Properties

Property #1 : The electrostatic potential and the electric field reproduce all symmetries
and invariances of the charge distribution (P ) that create this potential and this field.

Property #2 : Equipotential surfaces are normal to the electric field E at all points and
therefore are normal to the field lines.
Proof : Equation 2.14 shows that if points A and B are points sufficiently close to one another
on the same equipotential surface :
(A) (B) = 0 =


E . dl


with dl defining the equipotential surface. Therefore E dl .
This relation is only valid in electrostatics. In the general case, the term in
taken into account.

A
t

has to be

Property #3 : Field lines cannot be closed. They begin and end either on a charge or at
infinity.
Proof : Let us suppose that (C) is a closed field line. In that case, by definition of a field line

( E parallel to dl ) :


E dl 0
E . dl =
(C)

(C)

38

CHAPTER 2. Electrostatics

However, equation 2.11 gives :


E . dl = 0

(C)

These two conditions impose :

(C)

E . dl = 0 and therefore E = 0 .

Property #4 : In regions with no charge, E conserves its flux : div E = 0


Proof : The proof is the same as the one given in section 2.1.3.c.

In other words, when the field lines move apart, the norm of E diminishes.

Property #5 : The electrostatic potential cannot be maximum nor minimum in a point


where there is no charge.
Proof : Let A be a point where the potential is maximum. Then there exists a volume V
delimited by a surface such that all points M in this volume have :
M (M ) (A)
Let us chose the volume V such that is an equipotential surface. Then :


E .dS > 0
()

unless E = 0 . But Gauss law also gives :

Qint


E .dS =
0
()

So that Qint > 0.

2.1.6

Energetics in electrostatic

2.1.6.a

Electrostatic energy

Definition : The volume density of electrostatic energy is given by :


u=

0 E 2
2

(2.22)

This is a particular case of equation 1.24.


As a consequence, the total electrostatic energy U in a volume V can be written as :

U=
u dV
(V )

U=

1
2

0 E 2 dV 0

(2.23)

(V )

This energy is always positive. It implies that (+) = 0 and that the presence of a field adds
energy to the system.

39

2.2. EXAMPLES OF USUAL CHARGE DISTRIBUTIONS


2.1.6.b

Relation with the electrostatic potential

In section 1.3.2, we have seen that q corresponds to the potential energy of a charge q placed
in a potential . In other words, if the charge q is brought from infinity (where = 0) to a point
x where the electrostatic potential is , the work done on the charge is
W = q

If the potential is created by an ensemble of N charges qj (j = 1, 2, ..., N ) at positions


rj , then

the potential energy created on the charge qi placed at position ri is :


Wi =

N
qj
qi X

40 j=1 |
rj
ri |

Then, the total potential energy for all charges is :


W =

N
X
i=1

Wi =

N
1 X X qi qj
1 X qi qj
=

40 i=1 j<i | rj ri |
80
|
rj
ri |
(i,j)

where the terms i = j are not included in the sum. This can be rewritten as :
W =
where i =

2.2
2.2.1

1
40

qj

j |
rj
ri |

1X
qi i
2 i

is the potential created by all other charges on qi at


ri .

Examples of usual charge distributions


Example #1 : Uniformly charged infinite wire

Let us consider an uniformly charged infinite wire as schematized figure 2.2.1. The uniform
linear charge density is . Let us determine the electrostatic field and potential created by such a
system.

Figure 2.2: Uniformly charged infinite wire.

40

CHAPTER 2. Electrostatics

The problem is of cylindrical symmetry. We will therefore use the cylindrical coordinates. Since
the charge distribution (r, , z),
 and the electrostatic potential conserves the symmetries of the
charge distribution, (r, , z).
 Hence :

d
E = grad =
ur
dr
Another way of obtaining this result is to say that all point M of space belongs to two symmetry
,


planes : the one defined by {
u
r uz } and the one defined by {ur , u }. Since at these points M the


electric field must belong to all symmetry planes, E k
ur .
We will apply Gauss law on the following Gauss surface : a cylinder of height h and radius r :


Qint
E .dS =
0
(SGauss )


h
E .dS +
E .dS +
E .dS =
0
(S1 )
(S2 )
(Slat )
h
0 + 0 + 2rhE(r) =
0

E(r) =
20 r

E =
u
r
20 r

From the relation between and E , one derives : = 2


ln(r) + constant. If one imposes = 0
0
"

r

on the conductor, one obtains : = 20 ln a .

2.2.2

Example #2 : Uniformly charged infinite plane

Let us consider an uniformly charged infinite plane as schematized figure 2.2.2. The uniform
surface charge density is . Let us determine the electrostatic field and potential created by such
a system.

Figure 2.3: Uniformly charged infinite plane.

41

2.2. EXAMPLES OF USUAL CHARGE DISTRIBUTIONS

The problem is of cartesian symmetry. We will therefore use the cartesian coordinates. Since
(x,
 y,
 z), and the electrostatic potential conserves the symmetries of the charge distribution,
(x,
 y,
 z). Hence :

d
uz
E = grad =
dz


Moreover, the {Oxy} plane is a symmetry plane, so that - using E k
uz :



E (SP (x, y, z)) = SP E (x, y, z)





E (x, y, z) = SP E (x, y, z)

E (x, y, z) = E (x, y, z)
We will apply Gauss law on the following Gauss surface : a cylinder of height h and radius R,
centered around the {Oxy} plane :


Qint
E .dS =
0
(SGauss )


R2
E .dS +
E .dS +
E .dS =
0
(S1 )
(S2 )
(Slat )
E(z)R2 E(z)R2 + 0 =
E(z)

R2
0

20

u
z
20

From the relation between and E , one derives : = 20 z + constant. If one imposes = 0 on
the conductor, one obtains : = 20 z.

2.2.3

Example #3 : Uniformly charged sphere

Let us consider an uniformly charged sphere as schematized figure 2.2.3. The uniform volume
charge density is . Let us determine the electrostatic field and potential created by such a system.

42

CHAPTER 2. Electrostatics
Figure 2.4: Uniformly charged sphere.

The problem is of spherical symmetry. We will therefore use the spherical coordinates. Since
the charge distribution (r, , 
), and the electrostatic potential conserves the symmetries of the
charge distribution, (r, , 
). Hence :

d
E = grad =
ur
dr
are symmetry planes.
Another way of obtaining this result is to say that all planes containing
u
r

Since the electric field must belong to all symmetry planes, E k ur .


We will apply Gauss law on the following Gauss surface : a sphere of radius r :


Qint
E .dS =
0
(SGauss )

3
.dS
= 4R
E
u
u
r
r
30
(SGauss )
2
4R3
E(r)r2 sind d =
30
=0 =0
E(r)4r2

E(r)

4R3
30
R3
30 r2
R3

u
r
30 r2

R3
From the relation between and E , one derives : = 3
+ constant. If one imposes = 0 on
0r
3

R
the conductor, one obtains : = 3
+
0r

2.3
2.3.1

R2
30 .

Conductors
Electrostatic field and potential inside a conductor

An electrical conductor is a material that contains numerous free electrons. This means that
these electrons - usually from the outer shells of the atoms constituting the conductor - are able to
freely move inside the material. In particular, if a conductor is submitted to an external electric

field E ext , the electrons will be set in motion by this field and will only stop once equilibrium is

achieved, that is to say that E total = 0 .

Property #1 : The electrostatic field inside a conductor is zero :

E conductor = 0

(2.24)

It follows that the scalar potential inside a an electrical conductor is constant.


Property #2 : An electrical conductor is an equipotential region.
= constant

(2.25)

43

2.3. CONDUCTORS

2.3.2

Charges inside a conductor

Let us consider an electrical conductor. As we have seen, the electrostatic field inside the

conductor E int is zero : E int = 0 . Then Maxwell-Gauss equation gives that the volume charge

density inside the conductor is conductor = 0 div E = 0 :


Property #3 : In a conductor at equilibrium, there are no volume charges. All excess
charges are at the surface of the conductor :
conductor = 0

(2.26)

In practice, all charges accumulate on a few atomic layers at the surface of the conductor. These
are therefore surface charges.

2.3.3

Electric field at the vicinity of a conductor


Let us consider the conductor schematized on the opposite figure and
use the discontinuity equations at its interface with another medium

where reigns an electrostatic field E ext :

E ext E int = E ext =


u
z
0

Thus, under the action of an external magnetic field, charges accumulate at the surface of the
conductor. These surface charge, in turn, create an electric field such that, in the conductor the
total field is zero.

Property #4 : Outside, but at the immediate vicinity of an electrical conductor, the external
electric field is :

n
E ext =
0

(2.27)

where
n is the vector normal to the conductor surface.

2.3.4

Application #1 : radiation pressure for a conductor placed in a


zero electric field
We here consider the point Me external to the conductor, such as
indicated in the opposite figure. One can decompose the electric field
at point Me into two components :

i. The field created by the surface charges which immediately face


the point Me on the conductor, which are on the surface dS. Since
Me can be taken infinitely close to the conductor, while still being
exterior to the conductor, it sees the surface dS under a solid angle of approximately
2, that is to say that the surface charges appear, for Me to be distributed on an infinite
plane. This component is therefore equal to the field created by an uniformly charged plane

.
s
: E1 = 2
u
z
0

ii. The field E2 created by all other surface charges on the conductor.
We will call the surface of the conductor from which the surface
dS is extracted.
Since the field in Me is :

u
E = E 1 + E2 =
z
0

44

CHAPTER 2. Electrostatics

one deduce that all charges on create a field at point Me :

E2 =
u
z
20
Now, let us calculate the force exerted by the surface charges of on those on dS :

u
dF = dq E2 (Me ) = s dS
z
20
2

This force is equivalent to a pressure since proportional to dS. The pressure therefore is P = 2s0 .
This pressure is called the radiation pressure.
A few comments :

To say that the field E1 is the same than the one created by an infinite charged plane might
look like an approximation. However, it can be shown that it is the exact result. Indeed, the
electric field outside the conductor is :

u
E ext = E1 (Me ) + E2 (Me ) =
z
0
The field inside the conductor (at Mi ) can also be decomposed into a component due to the

field created by dS : E1 (Mi ) = E1 (Me ), and a component due to the field created by

which is equal to E2 (Mi ) = E2 (Me )


Eint = 0 = E1 (Me ) + E2 (Me )


Hence :

E1 (Me )

E2 (Me )

u
z
20
s

u
z
20

It is interesting to note that the field in the vicinity of a conductor, at point Me only depends
on the local surface charge density s (Me ), i.e. the one carried by the small conductor surface
dS. The charges of the rest of the conductor (on ), organize themselves so as to cancel,
inside the conductor, the field created by those carried by dS.

2.3.5

Application #2 : radiation pressure for a conductor placed in an


external electric field

In the presence of an constant external electric field E 0 , let us proceed with an analogous reasoning. The total external electric field can
be written :

E ext = E0 + E1 (Me ) + E2 (Me ) =


u
z
0

The field inside the conductor, at point Mi can also be decomposed into

a component due to the field created by the charges on dS, which value is E1 (Mi ) = E1 (Me ),

and a component due to the field created by the charges on , which value is E2 (Mi ) = E2 (Me ) :

Eint = 0 = E0 E1 (Me ) + E2 (Me )


Hence :

E1 (Me )

E0 + E2 (Me )

u
z
20
s

u
z
20

45

2.3. CONDUCTORS

This time, the charges on organize themselves so as to exactly compensate the field created
by charges on dS, taking into account the external electric field. The radiation pressure exerted on

dS corresponds to the force exerted on the surface charges on dS due to the external field E0 and
2

due to the charges on . Its expression still is P = 2s0 .

N.B. 1 : If E0 = 0 , the result is the same as in the previous paragraph.

N.B. 2 : A conducting object placed within an uniform electrostatic field E0 charges itself at
its surface, so that the electric field outside but at the immediate vicinity of this object is not
necessarily uniform !
N.B. 2 : In this particular case, one cannot apply the principle of superposition in a simple manner

(i.e. a charged conductor in E0 = a charged conductor in zero field + a static field). This is due
to the fact that the boundary conditions, at infinity in particular, are not trivial.

2.3.6

High voltage breakdown

Let us now mention the dependence of the field with respect to shape. More specifically, we
will qualitatively discuss the field near a sharp conductor.
First, let us remark that if one takes two spheres of radius R1 and R2 R1 , both charged with
Q. According to the calculation made section 2.2.3, the electric fields respectively verify :
4R12 E1

4R22 E2

Q
0
Q
0

Hence :
E1

E2

Q
4R12 0
Q
4R22 0

As can be seen, the field near the smaller sphere is larger than the one in the vicinity of the
larger sphere. In other words, any sharp point in a material has a large electric field near its surface.
This result is important for practical realizations. Indeed, above a certain electric field, the air
will break down. In this case, a loose charge - whether electron or ion - will be accelerated by the
field, will then collide with an atom, knocking out an electron of this atom. The additional ion and
electron will, in turn, be accelerated and a cascade process will begin, resulting in a great number
of charged particules in the air. Their motion creates a discharge or a spark.
Application #1 : If one wants to put a conductor to a high potential, one must avoid the
conductor surface to have any sharp surface features. Otherwise the air will discharge towards this
tip.
Application #2 : This field effect is used by field-emission microscopy to image the material of
a thin needle 1 . This technique has enabled, since 1955, to image materials with a resolution of a
few Angstr
oms, well before the advent of local probes such as the Scanning Tunneling Microscope
(STM) or the Atomic Force Microscopie (AFM).

1. For explanations on the working principle, see for example : Field Ion Microscopy, B. Gault et al., Atom
Probe Microscopy, accessible on the Internet.

46

2.4

CHAPTER 2. Electrostatics

Electrostatic dipole

The notion of electrostatic dipole is important both in physics and chemistry. Indeed, in
chemistry, it can model the behavior of polar molecules - such as HCl for instance - under a
magnetic field. In physics, we will see in the next chapter that it is at the basis of the microscopic
description of dielectric materials.

2.4.1

Definition

Definition : An electrostatic dipole is a distribution of static charges {qi }i placed at


points Pi , of finite spatial extension around point A, and for which the electric dipole moment

p (in C.m) is non zero :


X

p =
qi APi
(2.28)
i

Example : The simplest electrostatic dipole is a charge doublet {+q, q} (q > 0). If +q is placed

at point P , q at point N , and A the center point of [N P ], and N P = d


u :

p = +q AP q AN = q N P = qd
u

Note : A dipole which moment is not affected by the application of an external electric field is
called rigid.

Definition : The dipole approximation consists in considering the effect of an electrostatic


dipole at point M such that r = AM .
In other words, one neglects the spatial extension of the real charge distribution {qi }i around

point A and replaces it by a dipole


p placed in A. Therefore, a dipole can always be modeled by
two equivalent charges of opposite sign, centered at point A.

2.4.2

Electrostatic field and potential created by an electrostatic dipole

2.4.2.a

Electrostatic potential created by an electrostatic dipole

In the dipole approximation, the electrostatic potential created at point M by an electrostatic

dipole
p placed at point A such that AM = r
u , with
u an unit vector, is :
(M ) =

p .
u
40 r2

(2.29)

47

2.4. ELECTROSTATIC DIPOLE

Figure 2.5: Schematic representation of a dipole.


Proof :
(M )

X
i

1
Pi M

qi
40 Pi M

Pi M


2 1/2



2
OM OPi

1/2

with the geometry described figure 2.5 :



1/2
1
.
u
=
r2 + OPi2 2r
r OPi
Pi M
!1/2
.
1
OPi2
2r
u
r OPi
=
1+ 2
r
r
r2
so that at first order in

OPi
r

(M )

:
=

!1/2
.
2r
u
OPi2
r OPi
1+ 2
r
r2
i
!

.
X qi
r
u
r OPi
+ ...
1+
40 r
r2
i

.
X qi
X qi
u
r OPi
+
40 r
40 r2
i
i

X
u
r
.
qi OPi
0+
2
40 r
i

qi
40 r

due to the global neutrality of the dipole. In the present case :

u
r p
(M )
40 r2
p cos

40 r2

48
2.4.2.b

CHAPTER 2. Electrostatics
Electrostatic field created by an electrostatic dipole

In the dipole approximation, the electrostatic field created at point M by an electrostatic dipole

p placed at point A such that AM = r


u , with
u an unit vector, is :

3 (
p .
u ) .
u
p
E (M ) =
(2.30)
3
40 r
Proof : Taking the same definitions as above :
V (M )

Moreover :

3 (
p .
u ) .
u
p
40 r3

=
=

2.4.2.c

p cos
40 r2

grad
2p cos

3
4
0r
p sin 3
40 r
0

p cos
+ p sin

3p cos
u
u
u
r
r

40 r3
2p cos
4 r3
p sin0

40 r3
0

Field lines and equipotential surfaces for an electrostatic dipole

From equation 2.29, the equipotential surfaces are such that r2 = k cos with k a contant.
Field lines are obtained by taking, at all points, the normal to the equipotential surfaces. Field
lines and equipotential surfaces are schematized figure 2.6.

49

2.4. ELECTROSTATIC DIPOLE

Figure 2.6: Schematic representation of the equipotential surfaces and field lines of a
dipole. Taken from http://www.pstcc.edu/departments/natural_behavioral_sciences/Web%
20Physics/Experim%2001.htm. The dipole is here modeled by two opposite charges.

Note : Bear in mind that the above expressions are obtained in the dipole approximation, which
means that they are not valid within the charge distribution.

2.4.3

Mechanical action of an external electric field on a electrostatic


dipole

2.4.3.a

Electrostatic dipole in an external electric field

In the dipole approximation, an electrostatic dipole


p , of center A, placed in an external electric

field E is submitted to forces such that :




the resultant force is : F (A) =


E (A)
p . grad
A

the angular momentum is : M(A) =


p E (A)

Proof : Let us first calculate the resultant force (first order approximation) :
X

qi E (Pi )
F (A) =
i

Then the angular momentum is :

P 




q
E
(A)
+
AP
.
grad
E
(A)

i
x
i
x
P i 




qi Ey (A) + APi . grad Ey (A)

i
P 





i qi Ez (A) + APi . grad Ez (A)
P

qi APi . grad Ex (A)

Pi

q AP . grad E (A)
Pi i i y

qi APi . grad Ez (A)
 i 

p . grad A E (A)

M(A)

APi qi E (Pi )

APi qi E (A)

p E (A)

2.4.3.b

Application in the case of a rigid electrostatic dipole

If the considered dipole is rigid, the charges composing the dipole do not move with respect to

one another.
p is then constant in norm. Then the resulting force is :

F (A)

Ue

grad A (Ue )

p . E (A)

50

CHAPTER 2. Electrostatics

Proof :
Ue

qi (Pi )

X
i

X
i

qi ext (Pi ) +

qi int (Pi )




qi ext (A) + APi . grad ext (A) + constant taken to be 0




0+
p . E (A)

The force exerted on the dipole therefore derives from the potential energy Ue . Then, the dipole

tends to align with the external field E (M ).


Example #1: Solvation of ions - Some solvent molecules, such as water for instance, are polar,
which means they possess an electrostatic dipole moment (figure 2.4.3.b). When ions are dissolved
in a solvent, the solvent molecules organize themselves around the ion in order to minimize the
electrostatic energy (figure 2.4.3.b). In chemistry, the standard unit for molecular dipole moments
is the Debye (1 D = 3.331030 C.m). A water molecule has a dipole moment of 1.83 D.

Figure 2.7: Schematization of electrostatic dipole moment of a molecule of water. Taken from http:
//guweb2.gonzaga.edu/faculty/cronk/CHEM101pub/L18-index.cfm?L18resource=water.

Figure 2.8: Schematization of the hydration of Na+ by water molecules. Taken from http://www.
science.uwaterloo.ca/~cchieh/cact/applychem/hydration.html.

Example #2: Van der Waals interactions - Van der Waals forces 2 group different forces, all
proportional to 1/r7 . You will derive some of these in the Electrostatics exercise sheet. These
2. These intermolecular forces have first been predicted by the Dutch physicist J.D. van der Waals in his PhD
thesis, defended in 1873.

2.5. METHODS IN ELECTROSTATICS

51

forces originate from dipole-dipole interactions 3 and explain most of intermolecular interactions.
They have numerous applications, from scotch tape to biology 4 and including chemistry or polymer
science.

2.5

Methods in electrostatics

There are numerous solving methods in electrostatics which we will not review here 5 . For the
purpose of this course, you will mainly be using one of the following methods :
Directly compute the electric field using symmetries, invariances and :

E =
u dV
2
40
r
(V )
Directly compute the electrostatic potential using symmetries, invariances and :

1
dV
=
40
(V ) r
Determine the electric field using symmetries, invariances and Gauss theorem :

Qint


E .dS =
0
Also note that, for a problem with spherical symmetry, the method of separation of variables as seen in Quantum Mechanics for example - can be used.

2.5.1

Method of images

We will however detail one important method : the Method of images. This method can be
useful in the case of one or a few point charges in the presence of boundary surfaces that impose
discontinuities of the electric field. For instance, point charges in the presence of conductors.
The method consists in replacing the real problem with discontinuities - for eg. the conducting
planes - by an equivalent problem in an enlarged region, with no boundaries, where one or a few
virtual charges - the image charges -, are placed outside the region of interest so that :
1. these image charge simulate the boundary conditions ;
2. the electrostatics equation - the Poisson equation - remain unchanged in the region of interest.
The virtual problem with image charges is usually easier to solve. Then, thanks to the unicity
theorem (see section 1.6.2), the electric field that is found for the equivalent-charge problem is the
same, in the region of interest, than the one in the real problem.
Example : single charge in the presence of an infinite conducting plane

3. See exercise sheet on electrostatics.


4. See section 2.6
5. An introduction to those can be found in chapters 2, 3 and 4 of J.D. Jackson, Classical Electrodynamicsand
in chapters 5 to 8 of E. Weber, Electromagnetic Theory.

52

CHAPTER 2. Electrostatics

Figure 2.9: Schematization of the equivalence between a real problem and an equivalent-image
problem.

The method of images can be understood in the simple example schematized figure 2.9.
Let us consider a charge q sitting in vacuum in the presence of an infinite conducting plane,
at a distance d, occupying a semi-infinite space. As stated earlier, the potential in a conductor is
constant and since it extends to infinity, the corresponding potential is = 0. In particular, = 0
is imposed at the boundary between the vacuum and the conductor.
Moreover, in the region of interest - outside the conductor - the electric field is determined by
the Poisson equation : = 0 .
Let us now consider another problem where there is no conductor, but two point charges q and
q, separated by 2d. Then, in the middle of these two charges = 0 and in the region where
there previously was no conductor, the electric field is determined by the same Poisson equation
as before : = 0 .
The two problems therefore have the same Poisson equation in the volume of interest - the
vacuum in the initial problem -, with the same boundary conditions - = 0 where the boundary
between vacuum and conductor was before -, so that the solution, in the region of interest, is the
same for the two problems.
The corresponding field lines are schematized figure 2.10. The plain lines are the field lines for
the real problem, whereas the dashed lines correspond to the additional field lines corresponding
to the virtual problem, outside the region of interest.

2.6. ADDITIONAL READING

53

Figure 2.10: Field of a charge near a plane conducting surface, found by the method of images.
Taken from R.P. Feynman, The Feynman Lectures, Chapter 6.

2.5.2

Numerical methods

For more complex problem - with no obvious symmetries for example - different solving softwares
are available. They generally require the user to enter the geometry of the boundary conditions
and the position of charges. They compute the electric and/or magnetic fields. Some applications
directly solve Maxwells equations, but these are practical for 2D systems, or systems with high
symmetries.
For problems where the exact solution would be too costly in computer time, softwares using
finite element analysis (FEA) can be used. These sets of numerical methods are used to find
approximate solutions to partial differential equations - in electrostatics or any other field -.

2.6

Additional reading

The following article 6 , describes the mechanism by which geckos adhere to the surface they
walk on. The ultimate mechanism at play implies van der Waals forces.

6. Autumn et al., Integr. Comp. Biol., vol. 42, p.1081, 2002.

KELLAR AUTUMN2

AND

ANNE M. PEATTIE3

Mechanisms of Adhesion in Geckos1

1 From the Symposium Biomechanics of Adhesion presented at the


Annual Meeting of the Society for Integrative and Comparative Biology, 26 January 2002, at Anaheim, California.
2 E-mail: autumn@lclark.edu
3 Present address of Anne Peattie is Department of Integrative
Biology, University of California, Berkeley, California 94720.

INTRODUCTION
Geckos seem to defy gravity as they run along
smooth vertical surfaces at up to 20 body lengths per
second (Autumn et al., 1999a), and even upside down
on the ceiling. Over two millennia ago, Aristotle commented on the ability of the gecko to run up and
down a tree in any way, even with the head downwards (Aristotle/Thompson, 1918, Book IX, Part 9).
How geckos adhere has been a gripping topic of scientific research for well over a century (Cartier, 1872;
Haase, 1900; Gadow, 1901; Weitlaner, 1902; Schmidt,
1904; Hora, 1923; Dellit, 1934; Mahendra, 1941;
Maderson, 1964; Ruibal and Ernst, 1965; Hiller, 1968,
1969, 1975; Gennaro, 1969; Russell, 1975, 1986; Williams and Peterson, 1982; Stork, 1983; Schleich and
Kastle, 1986; Irschick et al., 1996; Autumn et al.,
2000; Liang et al., 2000; Autumn et al., 2002). Rapid
locomotion on a vertical surface requires the ability to
generate parallel (frictional) forces equal to or greater
than body weight (Fig. 1A). Detachment is perhaps
even more important than attachment. After all, even
the most common household adhesives are more than
sufficient to hold statically the weight of a large gecko
(50 g), but repeated and rapid detachment without significant detachment forces is beyond the capability of
any current synthetic adhesive. The secret of geckos
adhesive capabilities lies in the structure and function

1081

of their feet (Russell, 2002) and in the adhesive toe


pads borne on the underside of each digit. These pads
consist of a series of modified lamellae (scansors; Fig.
1B), each one covered with uniform arrays of similarly-oriented hair-like bristles (setae; Fig. 1C) formed
from b-keratin (Wainwright et al., 1982; Russell,
1986). A single seta of the tokay gecko (Gekko gecko;
Fig. 1D) is approximately 100 microns in length and
5 microns in diameter (Ruibal and Ernst, 1965; Russell, 1975; Williams and Peterson, 1982). The setae of
the tokay gecko (as well as most others) branch at the
tips into 1001,000 structures known as spatulae (Fig.
1E). A single spatula consists of a stalk with a flattened, roughly triangular end, where the apex of the
triangle connects the spatula to its stalk. Spatulae are
approximately 200 nm at their widest edge (Ruibal and
Ernst, 1965; Williams and Peterson, 1982). While the
tokay is currently the best studied of any adhesive
gecko species, there exist many hundreds of species
with adhesive toe pads, encompassing an impressive
range of morphological variation at the seta, scansor,
and toe levels, which has yet to be fully characterized.
The results discussed in this paper are based predominantly on observations of isolated tokay gecko setae
and whole-animal dynamics of the house gecko Hemidactylus garnoti, which shares the tokays setal morphology (but not its toe morphology).
While the structures of many gecko setae are well
documented, a full understanding of their function has
been more elusive. Haase (1900) noted that adhesion
is load-dependent and only occurs in one direction:
proximally along the axis of the toe. Haase was also
the first to suggest that geckos stick by intermolecular

SYNOPSIS.
The extraordinary adhesive capabilities of geckos have challenged explanation for millennia,
since Aristotle first recorded his observations. We have discovered many of the secrets of gecko adhesion,
yet the millions of dry, adhesive setae on the toes of geckos continue to generate puzzling new questions and
valuable answers. Each epidermally-derived, keratinous seta ends in hundreds of 200 nm spatular tips,
permitting intimate contact with rough and smooth surfaces alike. Prior studies suggested that adhesive
force in gecko setae was directly proportional to the water droplet contact angle (u) , an indicator of the
free surface energy of a substrate. In contrast, new theory suggests that adhesion energy between a gecko
seta and a surface (WGS) is in fact proportional to (1 1 cosu) , and only for u . 608. A reanalysis of prior
data, in combination with our recent study, support the van der Waals hypothesis of gecko adhesion, and
contradict surface hydrophobicity as a predictor of adhesion force. Previously, we and our collaborators
measured the force production of a single seta. Initial efforts to attach a seta failed because of improper 3D
orientation. However, by simulating the dynamics of gecko limbs during climbing (based on force plate data)
we discovered that, in single setae, a small normal preload, combined with a 5 mm displacement yielded a
very large adhesive force of 200 microNewton (mN), 10 times that predicted by whole-animal measurements.
6.5 million setae of a single tokay gecko attached maximally could generate 130 kg force. This raises the
question of how geckos manage to detach their feet in just 15 ms. We discovered that simply increasing the
angle that the setal shaft makes with the substrate to 308 causes detachment. Understanding how simultaneous attachment and release of millions of setae are controlled will require an approach that integrates
levels ranging from molecules to lizards.

Department of Biology, Lewis & Clark College, Portland, Oregon 97219

INTEGR. COMP. BIOL., 42:10811090 (2002)

1082

K. AUTUMN
AND

A. M. PEATTIE

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Mechanism of setal attachment


Using new microelectromechanical systems
(MEMS) force measurement techniques (Chui et al.,
1998), Autumn et al. (2000) measured the force production of a single gecko seta (Fig. 1F). Initial efforts
to attach a single seta failed to generate adhesive forces above that predicted by friction because we could
not achieve the proper orientation of the seta in 6 degrees of freedom (i.e., translation in, and rotation
about, all three axes). The angle of the setal shaft was
particularly important in achieving an adhesive bond
(Fig. 1G). When we simulated the dynamics of gecko
limbs during climbing (based on force plate data; Autumn et al., 1999a, b), we discovered that a small normal preload force (Fig. 2A) yielded a shear force of
;40 mN, six times the force predicted by whole-animal measurements (Irschick et al., 1996). Proper orientation, preload, and drag yielded 10 to 20 times the
frictional force measured with the seta oriented with
spatulae facing away from the surface (Fig. 2B). The
small normal preload force (Fig. 2A), combined with
a 5 mm displacement yielded a very large shear force
of 200 mN, 32 times the force predicted by wholeanimal measurements (Irschick et al., 1996; Fig. 2C).
The discovery that maximal adhesion in isolated setae
requires a small push perpendicular to the surface, followed by a small parallel drag, explained the load dependence and directionality of adhesion observed at
the whole-animal scale by Dellit (1934), and was consistent with the hypothesis that the structure of individual setae and spatulae is such that a small preload
and rearward displacement is necessary to engage adhesion (Ruibal and Ernst, 1965; Hiller, 1968). In their
resting state, setal stalks are recurved proximally.
When the toes of the gecko are planted, we believe

figure with our own observations.) The difficulty of


first isolating and manipulating a single seta, and then
finding an instrument capable of measuring micronewton forces in two dimensions, complicated our attempt
to demonstrate how the function of a single seta might
contribute to adhesion at the organismal level. We will
now describe how we confronted these difficulties in
our experimental procedure.

1083

FIG. 1. Gecko adhesive structures and methods used to measure setal adhesive function. Images (A) and (B) by Mark Moffett. Figure modified
from Autumn et al. (2000), Nature.
A) Ventral view of a tokay gecko (Gekko gecko) climbing a vertical glass surface. Arrows represent the forces acting on the geckos feet
as it climbs.
B) Ventral view of the foot of a tokay gecko, showing seta-bearing scansors.
C) Setae are arranged in a nearly grid-like pattern on the ventral surface of each scansor. In this scanning electron micrograph, each
diamond-shaped structure is the branched end of a group of four setae clustered together in a tetrad. Box shows seta enlarged in (D).
D) Single isolated gecko seta used in measurements in Autumn et al. (2000). Box shows spatulae enlarged in (E).
E) Spatular tips of a single gecko seta.
F) Isolated seta adhering to a micro-electromechanical system (MEMS) cantilever capable of measuring forces parallel and perpendicular
to the surface. Arrow shows direction of manipulation during the experiment, simulating parallel forces generated during vertical locomotion
of the gecko.
G) Single seta adhering to a 25 mm aluminum bonding wire capable of measuring detachment force perpendicular to the surface. Arrow
shows direction of manipulation during the experiment, simulating perpendicular movement during detachment of the foot. (a is the angle
between the setal stalk and the wire.

MECHANICS OF GECKO ADHESION


In a phylogenetic comparison of the forces produced
by pad-bearing lizards, Irschick et al. (1996) showed
that two front feet of a tokay gecko (Gekko gecko)
produced 20.1 N of force parallel to the surface with
227 mm2 of pad area. The foot of a tokay bears approximately 3,600 tetrads of setae per mm2, or 14,400
setae per mm2 (Schleich and Kastle, 1986; personal
observation). Consequently, a single seta should produce an average force of 6.2 mN, and an average shear
stress of 0.090 N mm22 (0.9 atm). (Note that Autumn
et al. (2000) used a value of 5,000 setae per mm2 taken
from Ruibal and Ernst (1965). In this paper, we use
the more accurate value of 14,400 setae per mm2
(Schleich and Kastle [1986]; we have confirmed this

forces (Adhasion), noting that under this hypothesis


the attractive force should increase as the space between the feet and the substrate decreases. Setae are
recurved such that their tips point proximally, leading
Dellit (1934) to hypothesize that setae act like hooks,
catching on surface irregularities (microinterlocking).
Ruibal and Ernst (1965) later postulated that, while the
seta is engaged, the spatulae lie flat against the substrate. It was clear to them that these flattened tips
increased the realized contact area, increasing frictional force. This was an important step in understanding
setal mechanics but it did not significantly change predictions associated with a microinterlocking hypothesis. This implied that adhesion should be stronger on
rougher surfaces, and that inverted locomotion should
be difficult, if not impossible, since friction (in the
classical sense) only operates parallel to the plane of
locomotion, leaving no vertical force component to oppose gravity.
The turning point in the study of gecko adhesion
came with a series of experiments by Hiller (1968),
who suggested that the material properties of the substrate, rather than its texture, determined the strength
of gecko adhesion. In demonstrating that adhesion was
a molecular phenomenon rather than a mechanical one,
his discovery effectively refuted the microinterlocking
and friction hypotheses and paved the way for the research we describe here.

GECKO ADHESION MECHANISMS

K. AUTUMN
AND

A. M. PEATTIE

FIG. 2. Single seta measurements. Figure modified from Autumn et al. (2000), Nature.
A) Submaximal force of single seta parallel to the surface with a known perpendicular preload, as a function of time. Perpendicular preload
is designated by the dashed line. ts represents the time when the seta began to slide off the sensor. The initial perpendicular force need not be
maintained during the subsequent pull. Diagrams show the stages of setal movement corresponding to the force record from the MEMS
cantilever (Fig. 1F). Arrows indicate the direction of applied force to the seta. Vertical arrow indicates a parallel force, and a horizontal arrow
indicates a perpendicular force.
B) Setal force parallel to the surface during attachment as a function of perpendicular preload force. Setal force was taken to be the adhesive
force at the time just prior to sliding (ts; Fig. 2A). The solid line represents a seta with spatulae projecting toward the surface. The dashed
line represents the setal force with spatulae projecting away from the surface (parallel force 5 0.25 perpendicular preload 20.09; r2 5 0.64;
F 5 13; df 5 1,9; P 5 0.007). The force produced by the inactive, non-spatular region increased with normal or perpendicular force, typical
of materials with a coefficient of friction equal to 0.25. The perpendicular preloading force that could be applied attained a maximum (near
15 mN), because greater forces resulted in the setal buckling.
C) Maximal force after a maximum preload (15 mN) of a single seta parallel to the surface as a function of time. Diagrams show the
stages of setal movement corresponding to the force record from the MEMS cantilever (Fig. 1F). Arrows indicate the direction of force applied
to the seta. Vertical arrow indicates a parallel force; horizontal arrow indicates a perpendicular force. The maximum force (;200 mN) following
the small rearward displacement (5 mm) was 32 times that predicted from maximal whole animal estimates (see text). The large increase in
force during the rearward displacement may be caused by an increase the number of spatulae contacting the surface.
D) Change in the orientation of the setae may facilitate detachment. Setal angle (a) with the surface at detachment as a function of
perpendicular force. Filled symbols represent seta pulled away from the surface until release. Open symbols represent seta held at a constant
force as angle is increased. Each symbol shape represents a different seta. Data collected with wire gauge (Fig. 1G). Setal angle at detachment
changed by only 15% over the entire range of perpendicular forces. This observation is consistent with an adhesive model where sliding stops
when pulling at greater than the critical setal angle and hence stress can increase at a boundary, causing fracture of the contact.

1084

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Integration of setal mechanics, functional


morphology of the foot, and dynamics of locomotion
It is important to emphasize that without integrating
dynamics at a larger scale (body and legs), the function
of the seta would likely still remain unknown. This
underscores the importance of an integrative approach

Mechanism of setal detachment


The surprisingly large forces generated by single setae raised the question of how geckos manage to detach their feet so rapidly (15 ms; Autumn et al., 1999a)
with no measurable detachment forces (Autumn et al.,
1999b). We discovered that increasing the angle between the setal shaft and the substrate beyond 308
caused detachment (Autumn et al., 2000; Fig. 1G; Fig.
2D). It is likely that as the angle of the setal shaft
increases, stress increases at the trailing edge of the
seta, causing fracture of the spatula-substrate bonds.
Similarly, at the scansor level, geckos unusual toe
peeling behavior (digital hyperextension; Russell,
1975, 1981) may aid in reducing or eliminating detachment forces by detaching only a small number of
setae at any moment. How this peeling behavior results
in reaching the critical angle of detachment is still unclear, but the two are almost certainly linked.

that the setae are bent out of this resting state, flattening the stalks between the toe and the substrate such
that their tips point distally. This small preload and a
micron-scale displacement of the toe or scansor proximally may serve to bring the spatulae (previously in
a variety of orientations) uniformly flush with the substrate, maximizing their surface area of contact. Adhesion results and the setae are ready to bear the load
of the animals body weight.
All 6.5 million (Schleich and Kastle, 1986; Irschick
et al., 1996) setae of a 50 g tokay gecko attached maximally could theoretically generate 1,300 N (133 kg
force) of adhesive forceenough to support the
weight of two humans. This suggests that a gecko need
only attach 3% of its setae to generate the greatest
forces measured in the whole animal (20N; Irschick et
al., 1996). Less than 0.04% of a geckos setae attached
maximally are needed to support its weight of 50 g on
a wall. At first glance, gecko feet seem to be enormously overbuilt. On further consideration, however,
there are some clear advantages to possessing as many
setae as possible (at which point we might ask what
factors actually limit a geckos adhesive capacity). It
is unlikely that all setae are able to achieve the same
orientation simultaneously. The proportion of spatulae
attached may be greatly reduced on rough surfaces
(particularly those with roughness on the same scale
as spatulae or setae; Autumn and Gorb in preparation).
On dusty or exfoliating surfaces, attachment to a wellanchored substrate will not be possible for every seta.
Large forces generated by perturbations during locomotion (e.g., recovering from a fall, predator avoidance, or station-keeping in high winds) may also utilize a greater proportion of geckos adhesive capacity.

1085

to answering biological questions (Lauder, 1991; Savageau, 1991; Ryan et al., 1998; Dickinson et al., 2000;
Autumn et al., 2002). How attachment and detachment
of millions of setae during locomotion is integrated
with the function of the scansor, toe, foot, leg, and
body remains a topic of great interest and ongoing
research (Sponberg et al., 2001; Russell, 2002). Russell has suggested that in the tokay (Gekko gecko), the
perpendicular preload and 5 mm drag requirements
(Autumn et al., 2000) are controlled by hydrostatic
pressure in the highly derived blood sinuses, and lateral digital tendon system, respectively (Russell,
2002).
Since gecko setae require a preload in the normal
axis for adhesion, large forces could potentially be associated with attachment of the foot. The tremendous
adhesive capacity of gecko setae suggests that large
forces could also occur during detachment. In fact, no
measurable ground reaction forces were associated
with either attachment or detachment during vertical
climbing on a force plate of the house gecko Hemidactylus garnoti (Autumn et al., 1999a, b), indicating
that these actions are either mechanically decoupled
from the center of mass in this species, or so small as
to be undetectable.
The absence of detachment forces is consistent with
(1) the mechanism of detachment and (2) the anatomy
of the foot. Geckos peel their toes up and away from
the substrate (digital hyperextension) rather than attempting to detach an entire foot at once, much like
removing a piece of tape. Peeling minimizes peak forces by spreading detachment out over time. Since the
muscles responsible for digital hyperextension (interossei dorsales; Russell, 1975) are located in the foot,
detachment does not have to be coupled mechanically
to the center of mass, as would be the case if the gecko
used its leg musculature to break the adhesive bonds
in the foot.
The absence of attachment forces is a more complicated issue, with at least three possible explanations.
As Russell (2002) has suggested, inflation of the digital blood sinuses (such as those present in individuals
of Hemidactylus and Gekko) may satisfy the preload
requirement of the setae during attachment without
generating measurable forces acting on the center of
mass. However, control of inflation and deflation of
the sinuses remains to be demonstrated. This mechanism would not be available to those species that lack
blood sinuses.
A second potential explanation is that setal preload
and drag are a consequence of force development during the stride. Climbing geckos use all four feet similarly to produce positive fore-aft forces parallel to the
surface that propel the gecko upwards (Autumn et al.,
1999b). Left legs apply a force to the right while right
legs apply a force to the left. Therefore, all four feet
pull medially, probably dragging the setae to engage
them fully, increasing the force of attachment (Autumn
et al., 1999b). However, geckos front legs pull the
center of mass into, while hind legs push the center of

GECKO ADHESION MECHANISMS

K. AUTUMN
AND

Unsupported mechanisms: glue, suction,


electrostatics, and microinterlocking
Since geckos lack glandular tissue on their toes,
sticky secretions were ruled out early in the study of
gecko adhesion (Wagler, 1830; Cartier, 1874; Simmermacher, 1884). The idea that the individual setae acted
as miniature suction cups was first debated in the insect adhesion literature (Blackwall, 1845; Hepworth,
1854), but was later proposed for gekkonid lizards by
Simmermacher (1884). However, there are no data to
support suction as an adhesive mechanism, and the
adhesion experiments carried out in a vacuum by Del-

MOLECULAR MECHANISM OF GECKO ADHESION


While the mechanism of setal attachment and detachment in geckos is now understood in mechanical
terms (Autumn et al., 2000), the molecular mechanism
underlying adhesion in setae has remained unclear.
Adhesion can be caused by at least 11 different types
of intermolecular surface forces at the interface between solids (Israelachvili, 1992; Gay, 2002), and cannot always be distinguished from friction, a complicated phenomenon in and of itself (Persson, 1999).
The study of surface forces is currently an active area
of research (Kinloch, 1987; Israelachvili, 1992, 2001;
Christenson, 1993; Noy et al., 1995; Thomas et al.,
1995; Gay and Leibler, 1999; Kunzig, 1999). The
complexity of the problem stems from the difficulty of
knowing what materials are actually interacting at the
molecular scale (Israelachvili, 1992; Persson, 1999).
Thus it is not surprising that adhesion in geckos has
remained a challenging problem.

mass away from, the vertical substrate (Autumn et al.,


1999b), generating a net moment pitching the anterior
toward the surface and counteracting the tendency of
the head to fall away from the surface (Alexander,
1992). Front legs do not push into the vertical substrate
during or after foot contact. Thus, these results do not
support the hypothesis that the setae become preloaded
as a consequence of force development during the
stride. While this is possible for the hind feet, it is
difficult to reconcile with the negative normal forces
produced by the front feet (Autumn et al., 1999b), unless the attachment force is so small as to be undetectable. The force necessary to bend even thousands
of setae into an adhesive orientation is probably quite
small (by our estimate, at most 10 mN). In this case,
we may have observed no measurable attachment forces simply because we cannot measure them.
A third possibility is that attachment is a reversal of
the peeling process of toe detachment, which we believe to be decoupled from the center of mass. The
geckos foot may approach the substrate without pressing into it, re-applying its adhesive by gradually extending (unrolling) its toes against the surface, at
which point they are ready to bear the load of the
animals weight. In this case, setal preload forces
would be spread out over time, and would likely be
far below the resolution of our force plate (61 mN).

1086

Potential intermolecular mechanisms: van der Waals


and capillary forces
Hiller (1968, 1969, 1975) showed that adhesive
force was correlated with the water droplet contact angle of the surface, and thus with the surface energy of
the substrate, providing the first direct evidence that
intermolecular forces are responsible for adhesion in
geckos. Ongoing research is now directed toward understanding the precise nature of these intermolecular
forces.
Intermolecular capillary forces are the principal
mechanism of adhesion in many insects (Gillett and
Wigglesworth, 1932; Edwards and Tarkanian, 1970;
Lee et al., 1986; Lees and Hardie, 1988; Brainerd,
1994; Dixon et al., 1990), frogs (Emerson and Diehl,
1980; Green, 1981; Hanna and Barnes, 1991) and even
mammals (Rosenberg and Rose, 1999). Unlike these
animals, geckos lack glands on the surface of their feet
(Wagler, 1830; Cartier, 1872; Dellit, 1934; Mahendra,
1941). This in itself does not preclude the role of thin
film capillary adhesion (von Wittich, 1854, quoted directly in Simmermacher, 1884; Stork, 1980; Scherge
and Schaefer, 1998) since a monolayer of water molecules (presumably present in the environment) can
cause strong attraction between surfaces (Baier et al.,
1968; Israelachvili, 1992). The apparent inverse correlation between adhesive force and hydrophobicity, as
inferred from the water droplet contact angle (Hiller,
1968) suggests that the polarity of the surface might
be an important factor in the strength of adhesion (Fig.
3). Two points are worth noting with regard to this
hypothesis: (1) Such a monolayer would have to be
ubiquitous and relatively pure in continuous patches in
order for the gecko to take advantage of it, and (2)
Hiller (1968) did not favor a capillary mechanism for
gecko adhesion since adhesive force did not decrease
completely to zero on all hydrophobic surfaces.
An alternative mechanism is that geckos adhere by

lit (1934) suggest that suction is not involved. Furthermore, our measurements of 10 atm of adhesion
pressure (Autumn et al., 2000) strongly contradict the
suction hypothesis.
Electrostatic attraction (Schmidt, 1904) is another
possible mechanism for adhesion in gecko setae. Experiments using X-ray bombardment (Dellit, 1934)
eliminated electrostatic attraction as a necessary mechanism for setal adhesion since the geckos were still
able to adhere in ionized air. However, electrostatic
effects could possibly enhance adhesion even if another mechanism is operating (Maderson, 1964).
Friction (Hora, 1923) and microinterlocking (Dellit,
1934; the climbers boot model of Mahendra, 1941)
may also play a secondary role, but the ability of geckos to adhere while inverted on polished glass, and the
presence of large adhesive forces on a molecularly
smooth SiO2 MEMS semiconductor (Autumn et al.,
2000) show that surface irregularities are not necessary
for adhesion, and may in fact be an impediment (Autumn and Gorb, in preparation)

A. M. PEATTIE

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

A
6pD 3

(1)

In this equation, A represents the Hamaker constant, a


function of the volume and polarizability of the molecules involved. For most solids and liquids, the Hamaker constant lies between 4 3 10220 and 4 3 10219
J, so it cannot affect an estimate of force by more than
a power of ten. Far more important to the estimate is
the separation distance (D) between the two surfaces
since force of adhesion scales inversely with the third
power of this distance. This implies that at small sep-

Force per area (N m22 ) 5

van der Waals interactions alone (Stork, 1980; Autumn


et al., 2000). This is a very intriguing hypothesis since
van der Waals interactions are the weakest of all intermolecular forces, but also the most universal. An
adhesive utilizing these very weak interactions would
be capable of sticking to nearly any natural surface.
However, a vast number of these interactions would
have to operate simultaneously in order to generate a
significant amount of force. That is, there must be a
large, real contact area between the organism and the
substrate. The highly branched setae on gecko toes
may be suited for maximizing contact area. If this is
the case, setal morphology would have a greater effect
on force of adhesion than surface chemistry of the seta
or substrate. In other words, the geometry of the adhesive would be more important than the chemistry.
The strength of van der Waals attractions is highly
dependent on the distance between surfaces, increases
with the polarizability of the two surfaces, and is not
directly proportional to surface polarity (Israelachvili,
1992). The observation that geckos cannot adhere to
Polytetrafluoroethylene (PTFE; Hiller, 1968) is consistent with the van der Waals hypothesis, since PTFE is
only weakly polarizable.
The following equation estimates the force (per
area) due to van der Waals interactions between two
planar surfaces (Israelachvili, 1992):

FIG. 3. Normalized force versus surface polarity (measured by water droplet contact angle, u) from Hiller (1968, 1969). Hillers (1968)
measurements of whole geckos on surfaces of varying polarity (circles) showed a correlation between polarity and force (y 5 20.012x
1 1.26; R2 5 0.85). In a separate study, Hiller (1969) found a weakly
significant correlation between force and surface polarity on polyethylene films modified with corona discharge (squares; y 5
20.012x 1 1.79; R2 5 0.69).

1087

arations, strong forces of adhesion can occur, but for


each power of ten increase in the distance, the force
of adhesion decreases 1,000 fold. Assuming the Hamaker constant to be of a typical value (1 3 10219 J),
and the spatular surface area to be approximately 2 3
10214 m2 (Ruibal and Ernst, 1965; Williams and Peterson, 1982; personal observation) the force of adhesion of a single seta with 100 spatulae increases
from 11 mN to 11 mN as the separation distance decreases from 1 nm to 0.1 nm. The actual force of adhesion of a single seta can reach almost 200 mN (Autumn et al., 2000). This would correspond to a separation distance of 0.380.81 nm (assuming 1001,000
spatulae/seta). The actual magnitude of this gap distance remains unknown.
Van der Waals and capillary adhesion are not mutually exclusive mechanisms. Although water at the
seta-substrate interface would increase the gap distance and therefore reduce the strength of van der
Waals adhesion, the diameter of a single water molecule (0.3 nm) remains well within the range of van der
Waals attraction (Israelachvili, 1992). Thus, in the
presence of a thin film of water, it is conceivable that
the two mechanisms are working in tandem. As more
layers of water molecules intervene, however, resistance to shear forces across the fluid is provided solely
by its viscosity (Baier et al., 1968). Water does not
have a high viscosity and therefore cannot withstand
high shear forces. Therefore, capillary forces are
strong in the normal direction and weak in the parallel
direction, whereas the opposite is true of setal adhesion. Empirically, setae have demonstrated stronger resistance to shear forces than to normal forces (Autumn
et al., 2000), suggesting that if capillary adhesion occurs, the films involved would have to be relatively
thin.
The extent to which a thin film of water will form
over the available interfacial area is dependent on the
relative vapor pressure. At extremely low humidity,
capillary adhesion will be weak due to lack of adsorbed water on substrates. At high humidity, water
begins to saturate the interstices of rough surfaces, acting as a lubricant (Israelachvili, 1992). This can be
described as the sand castle effect: one cannot build a
sand castle out of sand that is either very dry or very
wet. Dry sand does not adhere because the wetted interfacial area is too low. Very wet sand does not adhere
because the radius of the water meniscus between particles approaches the size of the particle itself, and capillary forces decline toward zero (Israelachvili, 1992).
If geckos were to rely solely on capillary adhesion for
attachment, setal function could be constrained by relative humidity of the habitat. However, since pad-bearing gecko species are found in habitats ranging from
tropical rain forests to dry, rocky desert, humidity does
not seem to have a strong influence on effective adhesion in nature.

GECKO ADHESION MECHANISMS

K. AUTUMN
AND

(2)

To test directly whether capillary adhesion or van


der Waals force is the primary mechanism of adhesion
in geckos, we measured the hydrophobicity of the setal
surface, and measured adhesion on two polarizable
semiconductor surfaces that varied greatly in hydrophobicity (Autumn et al., 2002). We measured the parallel force of single gecko toes on a gallium arsenide
(GaAs) semiconductor surface that is highly hydrophobic (u 5 1108). As a control, we measured parallel

Testing the van der Waals and capillary adhesion


hypotheses

Thus, for gL 5 72 mJ/m2 (for water) and WGG 5 60


mJ/m2 (a typical value for a nonpolar van der Waals
solid) we theoretically expect WGS to vary monotonically from WGS 93 mJ/m2 for uLS 5 08 to WGS 66
mJ/m2 for uLS 5 908 to WGS 5 0 for uLS 5 1808. This
correlation, however, is expected to hold only when
the substrate surface is hydrophobic (u . 608). A reanalysis of Hillers data (Hiller, 1968, 1969; Fig. 4)
using Equation 1 in fact supports the van der Waals
hypothesis with a strong correlation between force and
adhesion energies for u . 608.

WGS 5 gLWGG(1 1 cosuLS)

Since Hillers (1968, 1969) data correlate the


strength of setal adhesion with the polarity of the surface as indicated by the water droplet contact angle (u;
Fig. 3), it is reasonable to consider capillary action a
likely mechanism of adhesion. However, recent advances in the understanding of intermolecular surface
forces (Israelachvili, 1992) allow us to revisit Hillers
data (Hiller, 1968, 1969) and ask if they support only
a capillary adhesive mechanism. The adhesion energy
(W) between two identical solid surfaces is related to
the contact angle (u) of a liquid droplet on the surface
via the Young-Dupre equation, gL(1 1 cosu) 5 W,
where gL is the surface tension (or energy) of the liquid
(L) in units of mN/m (or mJ/m2). However, if the two
adhering surfaces are different materials, as for gecko
setae (G) on a substrate surface (S), the interfacial adhesion energy (WGS) bears no simple linear relation to
the liquid (water) contact angle. The relationships can
now be expressed as, gL(1 1 cosuLG) 5 WGG and gL(1
1 cosuLS) 5 WSS, and by the thermodynamic relation
(Israelachvili, 1992): WGS 5 1/2WGG 1 1/2WSS 2 gGS,
where gGS is the interfacial tension of the setae-substrate interface, and WGG and WSS are the cohesion energies of setal and substrate materials, respectively. An
approximate relation, valid when the two adhering surfaces are nonpolar and interact with each other
only via van der Waals dispersion forces, is (Israelachvili, 1992): WGS 5 WGGWSS. Replacing WGG and
W SS with the above equations gives: W GS 5
gL(1 1 cosuLG)(1 1 cosuLS), and a correlation between adhesion energy (WGS) and water droplet contact
angle on the substrate (uLS) may be expressed as

A reanalysis of the relationship between water


contact angle and gecko adhesion

1088

CONCLUSION AND FUTURE DIRECTIONS


Gecko setae are a novel type of adhesive in that the
strength of adhesion depends largely on geometry

force on the strongly hydrophilic (u 5 08) silicon dioxide (SiO2) semiconductor surface. We also compared the perpendicular force of single isolated gecko
setae on hydrophilic (SiO2, u 5 08) and hydrophobic
(Si, u 5 81.98) micro-electro-mechanical-systems
(MEMS) force sensors. If wet, capillary adhesive forces dominate, we expected a lack of adhesion on the
strongly hydrophobic GaAs and Si MEMS surfaces. In
contrast, if van der Waals forces dominate, we predicted large adhesive forces on the hydrophobic, but
polarizable GaAs and Si MEMS surfaces. In either
case we expected strong adhesion to the hydrophilic
SiO2 semiconductor and MEMS control surfaces.
We showed that tokay gecko setae are strongly hydrophobic (160.98; Autumn et al., 2002), probably a
consequence of the hydrophobic side groups of -keratin (Gregg and Rogers, 1984). Parallel stress of live
gecko toes on GaAs and SiO2 semiconductors was not
significantly different, and adhesion of a single gecko
seta on the hydrophobic and hydrophobic MEMS cantilevers differed by only 2%. These results reject the
hypothesis that water contact angle (u) of a surface
predicts attachment forces in gecko setae, as suggested
by Hiller (1968, 1969), and are consistent with our
reanalysis (above). Since van der Waals force is the
only mechanism that can cause hydrophobic surfaces
to adhere in air (Israelachvili, 1992), the GaAs and
hydrophobic MEMS semiconductor experiments provide direct evidence that van der Waals force is the
mechanism of adhesion in gecko setae, and that waterbased capillary forces are not significant.

FIG. 4. Reanalysis of Hillers previous measurements of geckos


adhering to surfaces of varying surface polarity (u water droplet
contact angle of surface; Hiller, 1968, 1969). Normalized force versus adhesion energy approximated by (1 1 cosu). The apparent
correlation between force and u (Fig. 3) suggested that adhesion in
geckos was a function of u. It is now known (Israelachvili, 1992)
that if one of two adhering surfaces is hydrophilic (u , 608; open
circles), their adhesion energy bears no simple relationship to the
liquid (water) contact angle of either surface. However, for hydrophobic surfaces (u . 608) adhering by van der Waals forces (see
Eqn. 2) the adhesion energy is approximately proportional to
(1 1 cosu). Thus, the correlations between force and adhesion energy for hydrophobic surfaces (closed circles: y 5 1.41x 2 1.27; R2
5 0.83; closed squares: y 5 1.60x 2 0.88; R2 5 0.73) support the
hypothesis that geckos adhere by van der Waals forces.

A. M. PEATTIE

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Alexander, R. M. N. 1992. Exploring biomechanics: Animals in motion. Scientific American, New York.
Aristotle, Historia Animalium, trans. D. A. W. Thompson. 1918. Clarendon, Oxford, http://classics.mit.edu/Aristotle/historypanim.html.
Attygalle, A. B., D. J. Aneshansley, J. Meinwald, and T. Eisner.
2000. Defense by foot adhesion in a chrysomelid beetle (Hemisphaerota cyanea): Characterization of the adhesive oil. Zoology (Jena). 103:16.
Autumn, K., S. T. Hsieh, D. M. Dudek, J. Chen, C. Chitaphan, and
R. J. Full. 1999a. Dynamics of geckos running vertically. Amer.
Zool. 38:84A.
Autumn, K., S. T. Hsieh, D. M. Dudek, J. Chen, C. Chitaphan, and
R. J. Full. 1999b. Function of feet in ascending and descending
geckos. Amer. Zool. 38:84A.
Autumn, K., Y. A. Liang, S. T. Hsieh, W. Zesch, W.-P. Chan, W. T.
Kenny, R. Fearing, and R. J. Full. 2000b. Adhesive force of a
single gecko foot-hair. Nature 405:681685.

REFERENCES

ACKNOWLEDGMENTS
The research described in this paper was the result
of collaboration between the Autumn, Full, Fearing,
and Kenny labs, and was supported under the auspices
of Alan Rudolph, director of DARPAs Controlled Biological and Biomimetic Systems program, N6600100-C-8047 and N66001-01-C-8072. Thanks to Michael Broide, Walter Federle, Valeurie Friedman,
Amanda Gassett, Wendy Hansen, Tony Russell, Simon
Sponberg, and two anonymous reviewers. We are especially grateful to Jacob Israelachvili for Equation 2.

rather than on surface chemistry. Many secrets of setal


form and function remain: The feet of geckos captured
in nature are usually clean (e.g., Stenodactylus khobarensis; Russell, 1979), and the possibility exists that
gecko setae are actually self-cleaning. We have yet to
uncover the design principles underlying the variation
in, and multiple convergent evolutions of setal, scansor, and foot structures in geckos (Russell, 1972, 1976,
1979; Bauer and Russell, 1990), and in convergent setal adhesive systems in anoles (Peterson and Williams,
1981), skinks (Williams and Peterson, 1982), and insects (Gillett and Wigglesworth, 1932; Edwards and
Tarkanian, 1970; Nachtigall, 1974; Bauchhenss and
Renner, 1977; Hill, 1977; Rovner, 1978; Stork, 1980,
1983; Walker et al., 1985; Wigglesworth, 1987; Roscoe and Walker, 1991; Betz, 1996; Gorb, 1998, 2001;
Attygalle et al., 2000; Eisner and Aneshansley, 2000;
Koelsch, 2000; Gorb et al., 2001; Gorb and Beutel,
2001; Scherge and Gorb, 2001). It is clear that gecko
setae are vastly overbuilt for adhesion to smooth ideal
surfaces under static conditions. This may not be the
case for natural surfaces, or for dynamic conditions
such as for running, jumping, and falling. Characterization of the surfaces and loading regimes seta-bearing animals experience in nature will be necessary before we can address the question of how much of a
safety factor exists in the gecko adhesive system.
Study of the setal adhesive system requires an extraordinary degree of integration across scales and disciplinesfrom quantum mechanics to biomechanics
and highlights the benefits of an integrative approach
to biology.

1089

Autumn, K., M. J. Ryan, and D. B. Wake. 2002. Integrating historical and mechanistic biology enhances the study of adaptation.
Q. Rev. Biol. (In press).
Baier, R. E., E. G. Shafrin, and W. A. Zisman. 1968. Adhesion:
Mechanisms that assist or impede it. Science 162:13601368.
Bauchhenss, E. and M. Renner. 1977. Pulvillus of Calliphora erythrocephala Meig (Diptera: Calliphoridae). Int. J. Insect Morphol. 6:225227.
Bauer, A. M. and A. P. Russell. 1990. Alternative digital scansor
design in the New Caledonian Gekkonid genera Bavayia and
Eurydactylodes. Mem. Queensland Mus. 29:299310.
Betz, O. 1996. Function and evolution of the adhesion-capture apparatus of Stenus species (Coleoptera, Staphylinidae). Zoomorphology 116:1534.
Blackwall, J. 1845. On the means by which walk various animals
on the vertical surface of polished bodies. Ann. Nat. Hist. XV:
115.
Brainerd, E. L. 1994. Adhesion force of ants on smooth surfaces.
Amer. Zool. 34:128A.
Cartier, O. 1872. Studien uber den feineren Bau der Haut bei den
Reptilien. I. Abt. Epidermis der Geckotiden. Verhandl. Wurz.
Phys.-med. Gesell. 1:8396.
Cartier, O. 1874. Studien uber den feineren Bau der Epidermis bei
ber die Wachsthumserscheinungen der
den Reptilien. II. Abt. U
Oberhaut von Schlangen und Eidechsen bei der Hautung. Arb.
zool.-zoot. Inst. Wurz. 1:239258.
Christenson, H. K. 1993. Adhesion and surface energy of mica in
air and water. J. Phys. Chem. 97:1203441.
Chui, B. W., T. W. Kenny, H. J. Mamin, B. D. Terris, and D. Rugar.
1998. Independent detection of vertical and lateral forces with
a sidewall-implanted dual-axis piezoresistive cantilever. Appl.
Phys. Lett. 72:13881390.
Dellit, W.-D. 1934. Zur anatomie und physiologie der Geckozehe.
Jena. Z. Naturw. 68:613656.
Dickinson, M. H., C. T. Farley, R. J. Full, M. A. Koehl, R. Kram,
and S. Lehman. 2000. How animals move: An integrative view.
Science 288:100106.
Dixon, A. F. G., P. C. Croghan, and R. P. Gowing. 1990. The mechanism by which aphids adhere to smoth surfaces. J. Exp. Biol.
152:243253.
Edwards, J. S. and M. Tarkanian. 1970. The adhesive pads of Heteroptera: A re-examination. Proc. R. Entomol. Soc. London 45:
15.
Eisner, T. and D. J. Aneshansley. 2000. Defense by foot adhesion in
a beetle (Hemisphaerota cyanea). Proc. Nat. Acad. Sci. U.S.A.
97:65686573.
Emerson, S. B. and D. Diehl. 1980. Toe pad morphology and mechanisms of sticking in frogs. Biol. J. Linn. Soc. 13:199216.
Gadow, H. 1901. The Cambridge natural history. Vol. 8. Amphibia
and reptiles. McMillan and Co., London.
Gay, C. 2002. StickinessSome fundamentals of adhesion. Integr.
Comp. Biol. 42:11231126.
Gay, C. and L. Leibler. 1999. Theory of tackiness. Phys. Rev. Lett.
82:936939.
Gennaro, J. G. J. 1969. The gecko grip. Nat. Hist. 78:3643.
Gillett, J. D. and V. B. Wigglesworth. 1932. The climbing organ of
an insect, Rhodnius prolixus (Hemiptera, Reduviidae). Proc. R.
Soc. London B 111:364376.
Gorb, S. N. 1998. The design of the fly adhesive pad: Distal tenent
setae are adapted to the delivery of an adhesive secretion. Proc.
R. Soc. London B 265:747752.
Gorb, S. N. 2001. Attachment devices of insect cuticle. Kluwer Academic Publishers, Dordrecht.
Gorb, S. N. and R. G. Beutel. 2001. Evolution of locomotory attachment pads of hexapods. Naturwissenschaften. 88:530534.
Gorb, S. N., E. Gorb, and V. Kastner. 2001. Scale effects on the
attachment pads and friction forces in syrphid flies (Diptera,
Syrphidae). J. Exp. Biol. 204:14211431.
Green, D. M. 1981. Adhesion and the toe-pads of treefrogs. Copeia
4:790796.
Gregg, K. and G. E. Rogers. 1984. Feather keratin: Composition,
structure, and biogenesis. In J. Bereiter-Hahn, A. G. Matoltsy,

GECKO ADHESION MECHANISMS

K. AUTUMN
AND

and K. S. Richards (eds.), Biology of the integument 2: Vertebrates, pp. 666694. Springer-Verlag, New York, New York.
Haase, A. 1900. Untersuchungen uber den Bau und die Entwicklung
der Haftlappen bei den Geckotiden. Archiv. f. Naturgesch. 66:
321345.
Hanna, G. and W. J. P. Barnes. 1991. Adhesion and detatchment of
the toe pads of tree frogs. J. Exp. Biol. 155:103125.
Hepworth, J. 1854. On the structure of the foot of the fly. Q. J.
Micros. Sci. 2:158163.
Hill, D. E. 1977. The pretarsus of salticid spiders. Zool. J. Linn.
Soc. 60:319338.
Hiller, U. 1968. Untersuchungen zum Feinbau und zur Funktion der
Haftborsten von Reptilien. Z. Morph. Tiere. 62:307362.
Hiller, U. 1969. Correlation between corona-discharge of polyethylene-films and the adhering power of Tarentola m. mauritanica
(Rept.). Forma et functio. 1:350352.
Hiller, U. 1975. Comparative studies on the functional morphology
of two gekkonid lizards. J. Bombay Nat. Hist. Soc. 73:278282.
Hora, S. L. 1923. The adhesive apparatus on the toes of certain
geckos and tree frogs. J. Proc. Asiat. Soc. Beng. 9:137145.
Irschick, D. J., C. C. Austin, K. Petren, R. Fisher, J. B. Losos, and
O. Ellers. 1996. A comparative analysis of clinging ability
among pad-bearing lizards. Biol. J. Linn. Soc. 59:2135.
Israelachvili, J. 1992. Intermolecular and surface forces. Academic
Press, New York.
Israelachvili, J. 2001. Tribology of ideal and non-ideal surfaces and
fluids. In B. Bhushan, Fundamentals of tribologybridging the
gap between the macro-, micro-, and nano-scales, pp. 631650.
Kluwer Academic Publishers, Dordrecht.
Kinloch, A. J. 1987. Adhesion and adhesives: Science and technology. Chapman and Hall, New York.
Koelsch, G. 2000. The ultrastructure of glands and the production
and function of the secretion in the adhesive capture apparatus
of Stenus species (Coleoptera: Staphylinidae). Can. J. Zool. 78:
465475.
Kunzig, R. 1999. Why does it stick? Bubbles, bubbles everywhere,
each one a small suction cup. Discover. 2729.
Lauder, G. V. 1991. Biomechanics and evolution: Integrating physical and historical biology in the study of complex systems. In
J. M. V. Rayner and R. J. W. (eds.), Biomechanics in evolution,
pp. 119. Cambridge University Press, Cambridge, Massachusetts.
Lee, Y. I., M. Kogan, and J. R. J. Larsen. 1986. Attachment of the
potato leafhopper to soybean plant surfaces as affected by morphology of the pretarsus. Entomol. Exp. Appl. 42:101107.
Lees, A. D. and J. Hardie. 1988. The organs of adhesion in the aphid
Megoura viciae. J. Exp. Biol. 136:209228.
Liang, Y. A., K. Autumn, S. T. Hsieh, W. Zesch, W.-P. Chan, R.
Fearing, R. J. Full, and T. W. Kenny. 2000. Adhesion force
measurements on single gecko setae. Technical Digest of the
2000 Solid-State Sensor and Actuator Workshop. 2000:3338.
Maderson, P. F. A. 1964. Keratinized epidermal derivatives as an aid
to climbing in gekkonid lizards. Nature 203:780781.
Mahendra, B. C. 1941. Contributions to the bionomics, anatomy,
reproduction and development of the Indian house gecko Hemidactylus flaviviridis Ruppell. Part II. The problem of locomotion. Proc. Indian Acad. Sci., Sec. B 13:288306.
Nachtigall, W. 1974. Biological mechanisms of attachment: The
comparative morphology and bioengineering of organs for linkage, suction, and adhesion. Springer-Verlag, New York.
Noy, A., C. D. Frisbie, L. F. Rozsnyai, M. S. Wrighton, and C. M.
Lieber. 1995. Chemical force microscopy: Exploiting chemically-modified tips to quantify adhesion, friction, and functional
group distributions in molecular assemblies. J. Am. Chem. Soc.
117:79437951.
Persson, B. N. J. 1999. Sliding friction. Surf. Sci. Rep. 33:83119.
Peterson, J. A. and E. E. Williams. 1981. A case study in retrograde
evolution: The onca lineage in anoline lizards. II. Subdigital
fine structure. Bull. Mus. Comp. Zool. 149:215268.
Roscoe, D. T. and G. Walker. 1991. The adhesion of spiders to
smooth surfaces. Bull. Br. Arachnol. Soc. 8:224226.
Rosenberg, H. I. and R. Rose. 1999. Volar adhesive pads of the

1090

feathertail glider, Acrobates pygmaeus (Marsupialia; Acrobatidae). Can. J. Zool. 77:233248.


Rovner, J. S. 1978. Adhesive hairs in spiders: Behavioral functions
and hydraulically mediated movement. Symp. Zool. Soc. London 42:99108.
Ruibal, R. and V. Ernst. 1965. The structure of the digital setae of
lizards. J. Morph. 117:271294.
Russell, A. P. 1972. The foot of gekkonid lizards: A study in comparative and functional anatomy. Ph.D. Diss., University of
London.
Russell, A. P. 1975. A contribution to the functional morphology of
the foot of the tokay, Gekko gecko (Reptilia, Gekkonidae). J.
Zool. London 176:437476.
Russell, A. P. 1976. Some comments concerning the interrelationships amongst gekkonine geckos. In A. d. A. Bellairs and C. B.
Cox (eds.), Morphology and biology of reptiles, pp. 217244.
Academic Press, London.
Russell, A. P. 1979. Parallelism and integrated design in the foot
structure of gekkonine and diplodactyline geckos. Copeia 1979:
121.
Russell, A. P. 1981. Descriptive and functional anatomy of the digital vascular system of the tokay, Gekko gecko. J. Morph. 169:
293323.
Russell, A. P. 1986. The morphological basis of weight-bearing in
the scansors of the tokay gecko (Reptilia: Sauria). Can. J. Zool.
64:948955.
Russell, A. P. 2002. Integrative functional morphology of the gekkotan adhesive system (Reptila: Gekkota). Integr. Comp. Biol.
42:11541163.
Ryan, M. J., K. Autumn, and D. B. Wake. 1998. Is integrative biology a luxury or necessity? A consideration from sexual selection. Int. Biol. Iss. News & Rev. 2:6872.
Savageau, M. A. 1991. Reconstructionist molecular biology. New
Biol. 3:190198.
Scherge, M. and S. N. Gorb. 2001. Biological micro- and nanotribology: Natures solutions. Springer, Berlin.
Scherge, M., X. Li, and J. A. Schaefer. 1998. The effect of water
on friction of MEMS. Tribol. Lett. 215220.
Schleich, H. H. and W. Kastle. 1986. Ultrastrukturen an GeckoZehen (Reptilia: Sauria: Gekkonidae). Amphibia-Reptilia. 7:
141166.
Schmidt, H. R. 1904. Zur Anatomie und Physiologie der Geckopfote. Jena Z. Naturw. 39:551.
Simmermacher, G. 1884. Untersuchungen uber Haftapparate an Tarsalgliedern von Insekten. Zeitschr. Wiss. Zool. 40:481556.
Sponberg, S., W. Hansen, A. Peattie, and K. Autumn. 2001. Dynamics of isolated gecko setal arrays. Amer. Zool. 41:1594A.
Stork, N. E. 1980. Experimental analysis of adhesion of Chrysolina
polita (Chrysomelidae: Coleoptera) on a variety of surfaces. J.
Exp. Biol. 88:91107.
Stork, N. E. 1983. A comparison of the adhesive setae on the feet
of lizards and arthropods. J. Nat. Hist. 17:829835.
Thomas, R. C., J. E. Houston, R. M. Crooks, K. Taisun, and T. A.
Michalske. 1995. Probing adhesion forces at the molecular
scale. J. Am. Chem. Soc. 117:38303834.
von Wittich. 1854. Der Mechanismus der Haftzehen von Hyla arborea. Arch. f. Anat. u. Phys. 180.
Wagler, J. 1830. Naturliches System der Amphibien. J. G. Cottaschen Buchhandlung, Munich.
Wainwright, S. A., W. D. Biggs, J. D. Currey, and J. M. Gosline.
1982. Mechanical design in organisms. Princeton University
Press, Princeton, New Jersey.
Walker, G., A. B. Yule, and J. Ratcliffe. 1985. The adhesive organ
of the blowfly, Calliphora vomitoria: A functional approach
(Diptera: Calliphoridae). J. Zool. London 205:297307.
Weitlaner, F. 1902. Eine Untersuchung uber den Haftfu des Gecko.
Verhdl. Zool. Bot. Ges. Wien. 52:328332.
Wigglesworth, V. B. 1987. How does a fly cling to the under surface
of a glass sheet? J.Exp. Biol. 129:373376.
Williams, E. E. and J. A. Peterson. 1982. Convergent and alternative
designs in the digital adhesive pads of scincid lizards. Science
215:15091511.

A. M. PEATTIE

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Downloaded from http://icb.oxfordjournals.org/ by guest on January 16, 2015

Potrebbero piacerti anche