This is the html version of the file http://www.jenlab.de/english/services/laser.pdf.
G o o g l e automatically generates html versions of documents as we crawl the web.
To link to or bookmark this page, use the following url: http://www.google.com/search?q=cache:mzdeSE0HlMMJ:www.jenlab.de/english/services/laser.pdf+skin+penetration+depth+of+NIR+radiation&hl=en


Google is not affiliated with the authors of this page nor responsible for its content.
These search terms have been highlighted:  skin  penetration  depth  nir  radiation 

Page 1
Abstract Near infrared (NIR) laser microscopy enables
optical micromanipulation, piconewton force determina-
tion, and sensitive fluorescence studies by laser twee-
zers. Otherwise, fluorescence images with high spatial
and temporal resolution of living cells and tissues can be
obtained via non-resonant fluorophore excitation with
multiphoton NIR laser scanning microscopes. Further-
more, NIR femtosecond laser pulses at TW/cm
2
intensi-
ties can be used to realize non-invasive contact-free sur-
gery of nanometer-sized structures within living cells
and tissues. Applications of these novel versatile NIR la-
ser-based tools for the determination of motility forces,
coenzyme and chlorophyll imaging, three-dimensional
multigene detection, non-invasive optical sectioning of
tissues (“optical biopsy”), functional protein imaging,
and nanosurgery of chromosomes are described.
Key words FISH · FRET · Laser tweezers · Multiphoton
microscopes · Nanosurgery
Introduction
Conventional light microscopy in “Life Sciences” in-
cluding confocal laser scanning microscopy is based on
the use of ultraviolet (UV) and visible radiation (Pawley
1995; Greulich 1999). One of the favorite methods in
live cell studies is fluorescence imaging. Around 80% of
applications of light microscopes in “Life Sciences” in-
volve fluorescence microscopy. Fluorophores include en-
dogenous (intrinsic) and exogenous (applied) dyes. Most
of the endogenous fluorophores in cells, such as trypto-
phan, dopamine, serotonin, NAD(P)H, and flavins,
require UV or blue excitation wavelengths. Important
exogenous fluorophores, such as the DNA markers
Hoechst 33342 and 4,6-diamidino-2-phenylindole
(DAPI) as well as the calcium indicators Fura-2 and
Indo-1 possess absorption bands in the UV only. Other
exogenous fluorophores are designed to absorb at the
visible wavelengths of the laser and mercury lamp exci-
tation radiation. UV microscopes have also been em-
ployed as microsurgery tools (Greulich 1999).
However, it must be noted that the use of UV and
short-wavelength visible radiations in conventional light
microscopy have certain disadvantages mainly because
of low light penetration depth and the potential of severe
photodamage to living cells (see Cunningham et al.
1985; Tyrell and Keyse 1990). It has been shown that
during high-resolution fluorescence microscopy with the
365-nm mercury lamp radiation results within seconds in
failed cellular reproduction, modifications in intracellu-
lar redox state, and DNA strand breaks (König et al.
1996a).
A versatile innovation in live cell microscopy is based
on the application of near infrared (NIR) laser radiation
in the spectral range of 700–1100 nm, the “optical win-
dow” of cells and tissues. Unpigmented cells appear as
nearly transparent objects in this spectral range. There-
fore, these cells can potentially experience extremely
high light intensities up to 100 gigawatt per square centi-
meter (1 GW/cm
2
=10
9
W/cm
2
) without damage. This
enormous intensity corresponds to a 12 orders higher
light intensity than sunlight reaching the surface of the
earth. When using ultrashort NIR laser pulses and the
mean intensity remains below 20 MW/cm
2
, the cellular
heating at <100 GW/cm
2
peak intensities is less than 2°C
(Liu et al. 1995; Schönle and Hell 1998).
The combination of NIR lasers and microscopes pro-
vides novel non-contact biomedical tools in “Life Sci-
ences”. These tools include not only imaging devices for
damage-free live cell imaging, but also for optical micro-
manipulation, intracellular photochemistry, and nanosur-
gery.
Robert Feulgen Prize Lecture 2000 presented at the 42nd sympo-
sium of the Society for Histochemistry in Les Diablerets, Switzer-
land, on 23 September 2000
K. König (
)
Laser Microscopy Division, Institute of Anatomy II,
Friedrich Schiller University Jena, Teichgraben 7,
07743 Jena, Germany
e-mail: kkoe@mti-n.uni-jena.de
Tel.: +49-3641-938560, Fax: +49-3641-938552
Histochem Cell Biol (2000) 114:79–92
DOI 10.1007/s004180000179
ROBERT FEULGEN PRIZE LECTURE
Karsten König
Laser tweezers and multiphoton microscopes in life sciences
Accepted: 26 June 2000 / Published online: 19 July 2000
© Springer-Verlag 2000

Page 2
When using highly focused continuous wave (cw)
NIR laser radiation, cells and organelles can be optically
confined in the focal volume of a high numerical aper-
ture (NA) objective by the radiation pressure (Ashkin
1970, 1980; Ashkin and Dziedzik 1987, 1989). These
tools are called laser tweezers (optical traps) and can be
used for optical micromanipulation, cell sorting, pico-
newton (pN) force measurements, motile cell diagnos-
tics, and as sources for two-photon excited fluorescence.
The other important NIR tool is the multiphoton laser
scanning microscope for fluorescence imaging with high
spatial and temporal resolution, photoinduced uncaging
of compounds, and nanoprocessing (Denk et al. 1990;
König et al. 1999a). Typically, these microscopes are
based on the use of ultrashort laser pulses in the femto-
second range (1 fs=10
–15
s). Multiphoton excitation
based on the simultaneous absorption of photons was
predicted in 1931 (Göppert-Meyer 1931) and first real-
ized with the availability of lasers in 1961 (Kaiser and
Garret 1961). In 1990, Denk et al. realized two-photon
excitation of fluorophores in living cells.
This paper focuses on applications of laser tweezers
and multiphoton femtosecond laser microscopes in “Life
Sciences” encompassing cell biology, biotechnology, and
medicine.
Laser tweezers
Principle
Laser tweezers, also known as (single beam gradient
force) optical traps, are based on pN force generation
during the interaction of highly focused laser beams with
dielectric particles, including cells and organelles.
Light, as a carrier of momentum, exerts pressure
which can be used to accelerate particles (Ashkin 1980).
This is the reason why the tail of a comet is directed
away from the sun. If a light beam of sufficient radiation
pressure is directed against gravity, particles can be
transduced into “gravity-free” conditions. The radiation
pressure of the sunlight on earth is of the low order of
µN/cm
2
, however, with the availability of laser sources,
laser light pressure can be used to move small objects.
As a spin-off of Ashkin’s initial studies on optical parti-
cle trapping (Ashkin 1970, 1980), optical cell microma-
nipulation by laser light became possible.
By focusing a single laser beam with a high NA ob-
jective, a gradient field of light intensity is created with
the highest intensity values in the focal volume and the
lowest in the periphery. Interaction with objects of a
higher refractive index than the surrounding results in
the formation of gradient forces. If the interaction is
primarily determined by beam refraction and when the
absorption is negligible, the net force acts toward the
focal volume. The net force (trapping force) of the or-
der of pN can be used to “pull” and to confine micro-
and nanometer-sized objects in the focal volume
(Fig. 1).
When using highly focused cw NIR laser beams (“mi-
crobeams”), pigment-free cells can be optically trapped
and can be manipulated in three dimensions without
physically touching them. In particular, contact-free cell
transport can be performed by moving the foci of the la-
ser beam in the desired direction. Alternatively, the stage
with the sample can be moved without displacement of
the trapped target.
Typical laser sources for optical trapping are the
Nd:YAG laser at 1064 nm and laser diodes. Often, the la-
ser beam is coupled by light fibers or directly to a video
microscope and focused to a diffraction-limited spot by
objectives with NA>1. Motor-driven mirrors in combi-
nation with joy-sticks allow maneuvering the position of
the tweezers.
Optical micromanipulation
In contrast to mechanical manipulation with microma-
nipulators, laser tweezers enable contact-free microma-
nipulation even through the glass windows of a closed
cell chamber. Such a versatile chamber (JenLab, Jena,
Germany) with 170-µm-thick glasses and a silicone gas-
ket is depicted in Fig. 2. Cells and medium can be easily
transferred and changed with standard needles and sy-
ringes.
We used closed cell chambers to trap human sperma-
tozoa of patients with impaired fertility and to study the
influence of the intense NIR beam on the cellular metab-
olism and vitality (König et al. 1995a). The NIR beam
traps spermatozoa by confinement of the head, the DNA-
containing region of high refractive index, in the focal
volume. The goal of the study was to prove if laser twee-
zers are an appropriate optical micromanipulation tool to
realize laser-assisted in vitro fertilization (IVF) by opti-
cal transport of a single sperm cell to the oocyte. This
method was suggested by Tadir et al. (1989) as an alter-
native to mechanical approaches.
Interestingly, laser tweezers at a wavelength of 800 or
1064 nm did not harm the cell even over extended trap-
ping periods of up to 15 min, whereas optical traps at
80
Fig. 1 Schematic representation of laser tweezers

Page 3
700–800 nm affected the cellular metabolism and led to
cell death within seconds (König et al. 1996a,b). The
reason is the phenomenon of trap-induced two-photon
excitation followed by harmful UV effects as explained
later.
Most optical trapping studies are conducted with the
1064-nm beam of an Nd:YAG laser which appears as a
safe micromanipulation tool even if the cell experiences
enormous energy densities (fluences) of some GJ/cm
2
.
When using 100-mW laser tweezers, the trap-induced
temperature increase is between 1 and 2°C. Meanwhile,
the first clinical applications of laser tweezers for laser-
assisted IVF have been reported (see, for example,
Wiedemann and Montag 1994). In addition, laser twee-
zers have been employed as micromanipulation tools in
pharmacology, biotechnology, and cell biology. For ex-
ample, Ashkin and Dziedzic (1989) studied the mechani-
cal properties of the cytoplasm in the interior of living
cells, and Chu’s group measured relaxation curves of
single DNA molecules (Perkins et al. 1994). Zahn and
Seeger (1999) used laser tweezers for drug screening.
Buican et al. (1987) are referred as pioneers of automat-
ed cell sorting by optical traps.
Force measurements
The net force (trapping force, F) depends linearly
on the laser power P and can be represented by:
F=QP/c
where c is the velocity of light in the medium and Q the
trapping efficiency parameter with values between 0 and
2. The parameter Q depends on the optical properties of
the trapped object, such as the refractive indices, as well
as on the beam profile and alignment. The knowledge of
the trapping parameter allows the calculation of the trap-
ping force and the subsequent use of the laser tweezers
as force measure (picotensometer).
Q can be calculated theoretically in the case of spheri-
cal particles with known refractive index. However, for
“real” biological objects with non-uniform morphology
and heterogeneous optical density, the trapping parame-
ter and the trapping forces, respectively, have to be de-
termined experimentally. The most common method is
the “escape force method” where the trapped object ex-
periences a flow of liquid of increasing velocity. The
flow exerts a drag force, F
drag
, which can be calculated
by Stokes’ law. This force depends on the velocity, vis-
cosity, and the target morphology and dimension. At a
certain velocity, the drag force has the same value as the
trapping force and the object can “escape”. It is at this
point that the drag force can be used to calibrate the trap-
ping force.
We determined the trapping parameter and the trap-
ping forces of human spermatozoa confined in an 800-nm
trap (König et al. 1996c). With a mean Q parameter of
0.12, the trapping force was found to be about 50 pN at
100 mW laser power. In the case of healthy spermatozoa,
a mean trapping power of about 80 mW was required to
confine these highly motile cells. Under the assumption
of linear swim motion, the mean intrinsic motility force
of human sperm can therefore be estimated to be about
45 pN (König et al. 1996c).
Trap-induced forces have been extensively used to
study motor proteins (Block et al. 1990; Kuo and
Scheetz 1993) and binding behavior, such as receptor-
mediated interactions of cells with glycoproteins (Zahn
and Seeger 1999). A review of biological applications of
optical forces was published by Svoboda and Block
(1994).
Fluorescence diagnostics of motile cells
Fluorescence studies on single motile cells are difficult
due to pN motility force driven sample movement. Cer-
tain bacteria, algae, and spermatozoa can achieve veloci-
ties higher than 50 µm/s. The problem of sample move-
ment can be overcome when the object of investigation
is confined to the focal volume of the objective by laser
tweezers. We combined sensitive imaging devices, such
as ultrafast time-gated slow-scan CCD cameras, and la-
ser tweezers to be able to image the autofluorescence of
highly motile cells with high spatial (submicron) and
temporal (picosecond, 1 ps=10
–12
s) resolution.
Figure 3 demonstrates an example of autofluores-
cence imaging of a highly motile sperm cell. The endog-
enous fluorophores have been excited with the 365-nm
radiation of a high-pressure mercury lamp and detected
in the blue/green spectral range. The only intracellular
region of autofluorescence of the motile cell was found
to be the midpiece of the cell which is the location of the
mitochondria. The fluorescence of mitochondria is main-
ly based on the reduced coenzyme NADH. Interestingly,
the sperm head including the acrosome region became
brightly fluorescent following exposure to extended
UVA radiation. The changes in the autofluorescence pat-
81
Fig. 2 Photograph of a sterile cell chamber which can be used in
optical trapping and multiphoton femtosecond laser microscopy

Page 4
tern were accompanied by reduced motility and finally
loss of vitality (König et al. 1996b).
An example of autofluorescence lifetime imaging is
depicted in Fig. 4. The fluorescence lifetime is an intrin-
sic property of the fluorophore and its microenvironment
and independent of concentration. Imaging of fluores-
cence lifetimes (τ mapping) enables spatially resolved
fluorescence lifetime determination and fluorophore sep-
aration. The figure exhibits the autofluorescence pattern
of the biflagellate green microalga Haematococcus plu-
vialis which was trapped by means of multiple traps at
1047 nm. Chlorophyll was excited with a picosecond la-
ser diode at 633 nm and imaged with an ultrafast time-
gated camera with a tunable time-delay (0–20 ns) be-
tween fluorescence excitation and detection (König et al.
1998). A mean fluorescence lifetime in the picosecond
82
Fig. 3 Autofluorescence imaging of an optically trapped single
motile sperm cell. The fluorescence emanates from the mitochon-
dria-rich region corresponding to the midpiece of the sperm
Fig. 4 Time-resolved autofluorescence imaging of an optically
trapped motile green alga. The time-delay between excitation of chlo-
rophyll and detection of fluorescence was spaced in steps of 500 ps

Page 5
range was determined which increased up to 1.4 ns when
exposed to herbicides, indicating disturbed energy trans-
fer.
Laser tweezers as non-linear diagnostic tools
Fluorescence diagnostics can also be performed with the
trapping beam as fluorescence excitation source. Surpris-
ingly, during trap experiments with dye-labeled sperma-
tozoa in 1994 we found that highly focused cw NIR
beams at 100 mW power are able to excite the visible
fluorescence of the intracellular dyes (König et al.
1995b). Excitation occurred only in a sub-femtoliter fo-
cal volume (1 fl=10
–18
m
3
), the region of highest light in-
tensity (Fig. 5). The fluorescence followed a squared re-
lation on laser power indicating a two-photon process. In
this case, the fluorescence is excited by the simultaneous
absorption of two low-energy (NIR) photons. Each of
them provides half the energy required for molecule ex-
citation (Fig. 6). As an example, a fluorophore which is
normally excited at 400 nm (for example red fluorescent
protoporphyrin IX) can therefore be excited with 800-nm
radiation.
Two-photon excitation radiation requires a high pho-
ton concentration in space and time due to the low mo-
lecular two-photon absorption cross-sections (10
–48
to
83
Fig. 5 Trap-induced visible fluorescence inside the optically con-
fined head of spermatozoa labeled with SYBR-green
Fig. 6 Principle of two-photon excitation. The simultaneous ab-
sorption of two near infrared (NIR) photons at MW/cm
2
and
GW/cm
2
light intensities induces visible fluorescence. At TW/cm
2
intensities multiphoton ionization leads to plasma-induced abla-
tion which can be used for nanoprocessing. S
n
, S
1
, S
o
represent dif-
ferent energy levels, E energy
10
–50
cm
4
s). To achieve such a high photon concentra-
tion, NIR light intensities of at least 20 MW/cm
2
are re-
quired.
Assuming a 100-mW laser beam at λ=800 nm and an
illumination spot of diameter d=λ/NA=615 nm and area
0.30 µm
2
, a high intensity of power/area=33 MW/cm
2
can
be calculated which is sufficient to induce two-photon
effects. Therefore, laser tweezers act as sources of two-
photon excitation and can be used as novel tools for non-
linear fluorescence excitation (the fluorescence intensity
has a squared, not a linear, dependence on power). One of
the interesting features of trap-induced fluorescence is the
possibility to localize the intracellular trapping spot with
high accuracy. Florin et al. (1998) use this effect to create
a novel photonic force microscope.
The two-photon excited fluorescence can also be used
to study possible harmful effects of the laser tweezers on
the cell trapped with the NIR beam. For example, when
injecting the viability indicators SYBR14 (green fluores-
cent live cell indicator) and propidium iodide (red fluo-
rescent dead cell indicator) from Molecular Probes
(Eugene, Ore., USA) into the cell chamber with sperma-
tozoa, the possible onset of trap-induced lethal effects
can be studied without external fluorescence excitation
sources. Using the methods of microspectrofluorometry
and spectrally resolved fluorescence imaging, we were
able to study the damaging effect of short-wavelength

Page 6
(<800 nm) laser tweezers (König 1998). NIR micro-
beams in the spectral range of 700–800 nm are able
to excite endogenous absorbers with electronic transi-
tions in the UV which may lead to harmful UVA effects
(Cunningham et al. 1985; Tyrell and Keyse 1990).
Using the trap-induced fluorescence we also studied
the effect of the laser beam quality and found: (1) that
the damage process is based on a two-photon process
and (2) that certain cw laser beams contain unstable pi-
cosecond laser pulses (“spikes”) which may enhance de-
structive effects (König et al. 1996d).
Multiphoton laser microscopy
Principle and setup of a multiphoton laser
scanning microscope
CW laser microbeams in combination with scanning mi-
croscopes can be used for three-dimensional (3D) imag-
ing of two-photon excited fluorophores (Hänninen et al.
1994; Booth and Hell 1998). In contrast to conventional
confocal laser scanning microscopes, no spatial filter
(pinhole) is required to obtain 3D images. This is be-
cause of the minute (sub-femtoliter) two-photon excita-
tion volume which can be used to scan the sample
(Fig. 7). Due to lack of NIR absorption outside the focal
volume, there is neither out-of-focus photobleaching nor
out-of-focus photodamage.
In two-photon microscopes, the fluorescence intensity
increases quadratically with the excitation intensity and
the power, respectively. However, trapping effects and
photothermal effects limit the use of high power cw
sources for fast fluorescence scanning microscopy. Fast
multiphoton fluorescence imaging is much more effi-
cient when using high repetition pulsed laser systems
with moderate peak power in the W and kW range but
with low mean µW and mW power. The two-photon ex-
cited fluorescence yield, I
F
, follows the following rela-
tion (Denk et al. 1990):
I
F
P
2
α/(τf
2
π
2
NA
4
/(hcλ)
2
where P is the mean power, α the molecular two-photon
absorption coefficient, τ the pulse width, f the repetition
frequency, NA the numerical aperture, h Planck’s con-
stant, c the velocity of light, and λ the wavelength. Be-
cause the fluorescence yield depends on a P
2
/τ relation,
two-photon microscopy with 1-ps laser pulses requires
threefold higher mean powers than 110 fs pulses. Both,
picosecond as well as femtosecond laser scanning micro-
scopes are now commercially available from leading mi-
croscopy suppliers.
In three-photon fluorescence microscopy (Gryczynski
et al. 1995; Wokosin et al. 1995; Hell et al. 1996; Maiti
et al. 1997), where three photons are absorbed simulta-
neously, I
F
depends on a P
3
/τ
2
relation. In this case, effi-
cient excitation requires the use of femtosecond laser mi-
croscopes. Three-photon excitation has been used to im-
age serotonin in living cells (Maiti et al. 1997).
Our multiphoton microscopes are based on the use of
femtosecond solid state laser sources. We use turn-key,
air-cooled, single-box laser systems. The first one is a ti-
tanium:sapphire laser (Vitesse 800-HP; Coherent, Santa
Clara, USA) with an output of 1 W mean power,
80 MHz repetition frequency, 800 nm wavelength, 80 fs
pulse width, and 55.5×34×18 cm
3
dimensions. We use
this laser for two-photon and three-photon fluorescence
studies in cells and tissues, and fluorescence in situ hy-
bridization (FISH) studies, as well as for nanosurgery.
The second ultracompact one (Femtolite; IMRA, Ann
Arbor, Mich., USA) is a frequency doubled erbium-
doped fiber laser at 780 nm with 40 mW mean power,
50 MHz repetition frequency, 780 nm wavelength, 180 fs
pulse width, and 19.3×10.9×8.2 cm
3
dimensions (elec-
tronic controller: 24.9×30.5×7.2 cm
3
). The power at the
sample of <7 mW is sufficient to excite a variety of flu-
orophores in cell monolayers and in FISH studies.
84
Fig. 7 Under same focusing conditions with high numerical aper-
ture objectives, two-photon excitation is confined in the minute
focal volume due to the required high intensity. By contrast, one-
photon excitation results in fluorescence along the illumination
cones. hv Energy of a photon

Page 7
The laser beam is expanded by an 1:4 Galilean tele-
scope, coupled to a modified inverted confocal laser
scanning microscope (LSM410; Zeiss, Jena, Germany)
and focused to a diffraction-limited spot by 40×, 63×, or
100× objectives of NA>1.2. Due to optical dispersion
which results in pulse broadening during transmission
through microscope optics, the pulse width at the sample
is about 150–200 fs (König 2000). At 5 mW mean pow-
er, the peak power and the peak intensities reach values
of 0.4 kW and 0.6×10
12
W/cm
2
(0.6 TW/cm
2
), respec-
tively, when assuming a full width half maximum beam
size of λ/2NA310 nm. A typical pixel dwell time of the
beam during one scan is 4 µs which results in a frame
rate of 1 s/frame. At zoom 4, 512×512 pixels cover a
sample area of 80×80 µm.
Multiphoton excited fluorescence is typically regis-
tered with a detector at the baseport of the modified mi-
croscope. A 700-nm short pass filter prevents the scat-
tered laser radiation from reaching the detector. We use
different camera systems, photomultiplier tubes (PMTs),
and a spectrometer as detectors.
The described microscope setup can also be used for
conventional one-photon confocal microscopy with an
internal He-Ne laser, pinholes, and the internal PMTs of
the microscope.
Multiphoton multicolor FISH (MM-FISH)
An interesting feature of multiphoton microscopy is the
possibility to excite a variety of fluorophores simulta-
neously at one NIR wavelength due to overlapping
two(three)-photon excitation spectra (Xu et al. 1996). We
have taken advantage of this to detect various genome
regions labeled with multiple fluorescent targets with
different emission wavelengths (König et al. 2000a). La-
beling is based on standard FISH procedure (see Speel
1999 for a review).
In contrast to the conventional detection method,
where different excitation wavelengths in the UV, blue,
and green range are required to induce the visible fluo-
rescence of the most common FISH fluorophores and of
the general DNA stain DAPI, the novel approach uses
NIR light at one excitation wavelength only. Figure 8
shows the emission spectra of FISH fluorophores which
can be non-linearly excited at 780 and 800 nm.
In addition to the advantage of using a single excita-
tion wavelength to realize multicolor FISH, multiphoton
microscopy enables optical sectioning of thick samples
(interphase nuclei, embryos, biopsies) in different
planes. There is no photobleaching of FISH fluorophores
in out-of-focus regions. Reconstruction of 3D fluores-
cence images from the optical sections provides informa-
tion on the genome architecture, such as 3D organization
of chromosomes and their well-defined domains such as
centromeres and telomeres. Such a 3D fluorescence im-
age of centromeric regions of the chromosomes 1, 3, 6,
12, and X in the interphase nucleus of an amniotic fluid
cell is seen in Fig. 9. The 3D image has been recon-
structed from a stack of 20 images which are spaced by
0.75 µm. The five FISH fluorophores have been excited
simultaneously. Also the blue fluorescence of the DNA
counterstain DAPI has been induced with NIR femtosec-
ond laser pulses. Spatially resolved DAPI fluorescence
imaging provides information about the nuclear architec-
ture and enables the determination of the intranuclear lo-
ci of the fluorescent centromeric regions.
Besides multifluorophore excitation and optical sec-
tioning, a third advantage of using excitation radiation in
the 700- to 1200-nm spectral range is the high light pen-
etration depth. We used MM-FISH for the detection of
centromeric regions in multilayer samples, such as hu-
man biopsies. In particular, we imaged two-photon excit-
ed spectrum green-labeled C-4 and spectrum orange-la-
beled C-X probes in thick kidney cryosections with high
spatial resolution. The nuclear area was visualized after
85
Fig. 8 The depicted FISH fluorophores with emission in the blue,
green, yellow, and red can be simultaneously excited at 780 and
800 nm. DAPI 4,6-diamidino-2-phenylindole, DAC diethyl-
aminocoumarine, FITC fluorescein isothiocyanate, Cy cyanine

Page 8
counterstaining with ethidium bromide and optical sec-
tioning with NIR laser pulses (König et al. 2000b).
Potential applications of this new MM-FISH tech-
nique are in the field of molecular cytogenetics, prenatal
and preimplantation diagnosis, and molecular pathology.
Live cell imaging
In addition to the in vitro studies, imaging of DNA in
living cells and tissues can be performed by two-photon
excited imaging of the DNA fluorophore Hoechst 33342.
The high penetration depth of NIR radiation also allows
the spatially resolved detection of the fluorophore
in deep tissue. NIR radiation at 800 nm has a typical
penetration depth in tissue of several millimeters in
contrast to some micrometers when using one-photon
fluorescence excitation in the UV (Chong et al. 1990).
Figure 10 shows DNA images of tumor tissue in living
mice after the topical application of a Hoechst–DMSO
mixture. The Hoechst-labeled chromatin can be clearly
visualized in the different cell layers. Considering a
120×120×30 µm
3
=0.4×10
–12
m
3
tumor volume, a mean
number of 100 nuclei in such a volume can be counted
from the stack of images. Interestingly, scattering in tur-
bid media such as tissues does not decrease the excellent
lateral resolution of <0.4 µm and the ca. 1 µm axial reso-
lution significantly within the first 100 µm tissue.
Multiphoton microscopy has been widely used in imag-
ing of single living cells, cell monolayers, and embryos
(see König 2000 for a review) including mapping of ion
channels by two-photon photochemical microscopy (Denk
1994) and neuron imaging (Denk and Svoboda 1997).
86
Fig. 10 Non-invasive optical sectioning in a living anesthetized
mouse. Depicted are Hoechst 33342-labeled tumor tissue layers
Fig. 9 Three-dimensional (dz=0.75 µm) six-color detection of in-
terphase nuclei of amniotic cells with the NIR-excited fluoropho-
res spectrum aqua (centromere chromosome X, blue), spectrum
orange (centromere chromosome 3, red), rhodamine 110 (centro-
mere chromosome 12, yellow), spectrum green (centromere chro-
mosome 6, green), DAC (centromere chromosome 1, white), and
DAPI (light blue). The image is false-color coded and reconstruct-
ed from a stack of 20 optical sections

Page 9
An interesting feature of NIR multiphoton microsco-
py is the possibility to excite endogenous fluorophores
such as the reduced coenzymes NADH and NADPH as
well as flavin coenzymes with NIR radiation in the range
of 700–800 nm (Piston et al. 1995; König et al. 1996d).
Because the oxidized forms NAD(P) do not exhibit fluo-
rescence, the NAD(P)H attributed autofluorescence can
be used to obtain information on the intracellular redox
state, modifications in the respiratory chain, and cellular
metabolism (König and Schneckenburger 1994; König et
al. 1995b).
The two-photon excitation of tissue autofluorescence
may become a useful non-invasive technology to obtain
“optical biopsies” without physical removal of tissue.
Similar to tomography with X-rays, intense NIR micro-
beams can be employed to perform optical sectioning of
the tissue.
Endogenous fluorophores and structures which can be
excited via a two-photon or a three-photon excitation
process include serotonin, dopamine, and tryptophan
with emission in the UV, NAD(P)H, collagen, elastin,
melanin, and flavins with blue/green fluorescence, lipo-
fuscin, Zn-coproporphyrin, and Zn-protoporphyrin with
yellow emission, and coproporphyrin, protoporphyrin,
and chlorophyll with red fluorescence.
The first two-photon excited autofluorescence images
of in vivo human skin have been performed (Masters et
al. 1997; König 2000). Figure 11 depicts a single in vivo
autofluorescence image from a stack of high-resolution
fluorescence images. The image was acquired at a depth
of 30 µm from the palmar surface of the forearm. The
stack was obtained up to 150 µm into the skin with a
depth increment of 5 µm. The stratum corneum with a
typical thickness of 15 µm at the investigated loci was
found to be highly fluorescent when excited at 800 nm.
In particular, the border of the hexagonal-shaped tissue
structures exhibited fluorescence. Below this tissue lay-
er, individual cells could clearly be visualized by the au-
tofluorescence of intracellular structures in the cyto-
plasm. Cell nuclei and cell membranes did not fluoresce.
The autofluorescence was stronger in the innermost layer
of the epidermis, the basal layer, than in the surround-
ings. Epidermis and dermis could be differentiated, and
elastin and collagen fibers could be visualized.
Photodamage due to multiphoton microscopy
Due to the lack of out-of-focus photodamage and photo-
bleaching, multiphoton NIR microscopy appears as a
safe novel tool without impact on cellular metabolism,
reproduction, and vitality. In fact, Chinese hamster ovary
cells can be scanned with an 800-nm femtosecond laser
beam at 2 mW for hours without damage to the exposed
cells and their derivatives (König 2000). Squirrel et al.
(1999) intermittently exposed hamster embryos for 24 h
with ultrashort laser pulses of GW/cm
2
peak intensity
without impact on embryo development in contrast to
one-photon laser scanning microscopy where the embry-
os did not survive.
However, above certain thresholds photodamage may
occur. As pointed out in the section on laser tweezers, in-
tense cw NIR beams can induce cell damage via a two-
photon effect. The same destructive effects would there-
fore occur when using ultrashort laser pulses at even
higher light intensities.
We found mainly two types of photodamage during
NIR microscopy. The first is a slow process probably
based on two-photon excitation of endogenous absorbers
and subsequent photo-oxidation processes resulting in
the formation of destructive reactive oxygen species
(ROS). This process appears at MW/cm
2
and GW/cm
2
intensities and depends strongly on wavelength. The sec-
ond damage process is of immediate effect and requires
high intensities in the TW/cm
2
range and is based on
multiphoton ionization, optical breakdown phenomena,
and intracellular plasma formation. It results in material
ablation and disruption including complete cell fragmen-
tation.
Studying the first photochemical attributed photo-
damage process we found that the mitochondria and the
Golgi apparatus are the major targets of NIR micro-
beams (Oehring et al. 2000). Above certain light intensi-
ties, laser-exposed cells either fail to divide, become gi-
ant cells, or die. Due to the P
2
/τ dependence of the slow
damage process, photodamage was found to be more
pronounced at shorter pulses (König et al. 1999b).
More recently it has been demonstrated that NIR laser
irradiation at certain higher power levels evoke genera-
87
Fig. 11 Non-invasive optical biopsy of high spatial resolution
with near infrared femtosecond laser pulses. Depicted is an in vivo
autofluorescence image of human skin showing an epidermal layer
at a depth of 30 µm. Note that individual cells as well as cytoplas-
mic structures are clearly visible

Page 10
88
Fig. 12A–C Nanosurgery with NIR femtosecond laser pulses.
Force microscopy image of cuts through human chromosomes
(A). A minimum cut size of 110 nm was achieved. Transmission
and fluorescence image of living cells labeled with rhodamine 123
before (B) and after (C) knocking out of a single mitochondrion
(asterisks)

Page 11
89
tion of ROS that can be cytochemically visualized in vi-
vo using Ni-3,3-diaminobenzidine (Ni-DAB) as well as
with a recently developed fluorescent probe Jenchrom px
blue from JenLab (Tirlapur et al. submitted). In addition,
the irradiated cells manifest membrane-barrier dysfunc-
tion, drastic alterations in the morphology of the nuclear
envelope, and fragmentation of the DNA. Because the
cytological changes observed in NIR-irradiated cells
bear striking similarities to those seen in cells undergo-
ing programmed cell death (Li and Darzynkiewicz
1999), it has been inferred that intense NIR femtosecond
laser pulses can induce apoptosis-like death (Tirlapur
and König submitted).
Femtosecond laser surgery of cellular nanostructures
When the plasma-mediated photodamage process can be
confined to a tiny intracellular volume by parking the
beam at pixels of interest, the destructive effects of in-
tense NIR laser pulses can be used to cut cellular struc-
tures or to “drill” holes in the target (König et al. 1999a).
Material processing with femtosecond laser pulses has
the advantages of minimal ablation threshold, low trans-
fer of optical energy into destructive mechanical energy,
and the absence of thermal damage to surrounding struc-
tures compared to nanosecond pulses used in conven-
tional microsurgery.
We use the novel surgery tool to perform dissections
of chromosomes and to knock out intracellular structures
at mean powers of 30–50 mW with microsecond and
millisecond beam dwell times (Fig. 12A,B). Because on-
ly the central part of the illumination spot provides suffi-
cient intensity for plasma-induced ablation, laser cuts
and holes can be generated in the target with a size be-
low the diffraction-limited spot size. We realized a mini-
mum cut-size of 110 nm into the human chromosome 1
which is to date likely be the smallest laser cut in biolog-
ical material. The topography of this laser cut was ana-
lyzed and measured with a scanning force microscope.
We were also able to perform chromosome dissections
within round living cells (König et al. 1999a). The use of
femtosecond NIR pulses at TW/cm
2
light intensities
therefore provide novel non-invasive tools to perform
nanosurgery without destructive influence to the sur-
roundings and enable nanoprocessing within living cells
and tissues.
Further potential applications of NIR microscopes
Four dimensional (4D) microscopy in space and time
An interesting feature of multiphoton laser scanning mi-
croscopes with picosecond and femtosecond pulsed la-
sers is the possibility of simultaneously performing 4D
imaging in space and time. Of particular relevance is mi-
croscopic imaging with ultrafast temporal resolution in
the range of picoseconds and nanoseconds corresponding
to the range of fluorescence lifetimes (see, for example,
Lakowicz 1983; So et al. 1998). Hence with a pulsed la-
ser excitation source it is possible to upgrade a 3D multi-
photon microscope to a versatile 4D imaging device.
Fig. 13 Time-resolved multiphoton microscopy. Fluorescence de-
cay curves after two-photon excitation of intracellular fluoropho-
res along lines of 128 pixels

Page 12
In order to realize 4D two-photon microscopy we
have incorporated a fast PMT in combination with a sin-
gle photon counting unit (SPC 730; Becker and Hickl,
Berlin, Germany). The unit enables the fast registration
of fluorescence decay curves following excitation with
the 80-MHz femtosecond laser at a pixel of interest dur-
ing single-point illumination as well as the registration
of 128×128=16384 decay curves during scanning. Ex-
amples of fluorescence decay curves of a variety of in-
tracellular fluorophores along line scans and an example
of a τ image are presented in Figs. 13 and 14.
Two-photon fluorescence resonance
energy transfer (FRET)
As yet an unexplored interesting potential feature of
multiphoton microscopes with high NA objectives is the
use of the minute sub-femtoliter excitation volume for
studying single molecules and intramolecular interac-
tions. In particular the combination of two-photon NIR
excitation and FRET allows monitoring of the spatially
resolved relationship between two macromolecules with-
in living cells (Fig. 15).
The non-radiative energy transfer (Foerster mecha-
nism) occurs in close proximity of a fluorescence donor
(D) in the excited state and a fluorescence acceptor (A).
The efficiency of the energy transfer decreases with R
–6
,
where R is the distance between D and A. Therefore,
FRET occurs mainly within intermolecular distances
with R<10 nm. It provides a variety of information on
molecular interactions such as binding behavior, inter-
molecular distance, diffusion kinetics, and association
reactions (Pollak and Heim 1999). In order to realize ef-
ficient two-photon FRET the following conditions are
essential:
1. Efficient two-photon excitation of the donor molecule
2. Spectral overlap between the emission spectrum of
the donor and that of the excitation spectrum of the
one-photon acceptor as well as appropriate relative
orientation of both molecules
3. Separation between acceptor and donor emission
4. Distance less than 10 nm.
Due to the broad two-photon excitation spectrum, a vari-
ety of fluorophores preferably with UV and blue one-
photon absorption maxima can be used as donors. In the
case of probing protein–protein interactions, GFP mu-
tants such as B(lue)FP with 445 nm emission, C(yan)FP
(505 nm), Sapphire (511 nm), eGFP (511 nm) and Y(el-
low)FP (540 nm) can be used. FRET can occur between
two GFPs or one GFP and a second fluorophore. Ac-
cording to the spectral overlap, potential FRET pairs in-
clude BFP–GFP, CFP–YFP, Sapphire–fluorescein, and
GFP–fluorescein molecules.
4D microscopy provides the possibility to perform
time-resolved energy transfer imaging which enables
studies of the static and dynamic mobility of macromole-
90
Fig. 14 Fluorescence lifetime image of a living Chinese hamster
ovary cell labeled with the DNA probe Hoechst 33342
Fig. 15 Principle of
two-photon fluorescence
resonance energy transfer

Page 13
cules. Often, the fluorescence decays are relatively sensi-
tive to the acceptor-donor distances (Lakowicz 1983).
Non-invasive nanosurgery in tissues
As pointed out earlier, nanosurgery can be performed
within a single living cell. In principle, knocking out of
cellular structures can be performed deep in living tissue
without disturbing surface layers. Such highly precise
NIR laser-based nanoprocessing of cellular structures
without compromising the vitality of cells has numerous
potential applications in cell and developmental biology
particularly in studies addressing spatial-temporal con-
trol of developmental events and functional interactions
between organelles, as well as cell–cell communication.
Initial NIR femtosecond laser experiments have been
performed in living plant tissue. Using 800-nm laser
pulses, laser-mediated dye loading (Tirlapur and König
1999) and intratissue nanosurgery of the cell wall in a
living leaf of Elodea densa has also been successfully
accomplished. Following treatment, the ultrastructural
analysis of the corresponding area revealed a clean non-
staggering cut across the cell wall that measured less
than 350 nm. More interestingly, using the same NIR
femtosecond laser pulses, a single plastid or a part of the
organelle could be completely knocked out without af-
fecting the adjacent organelles or the viability of the cell.
The vitality of the cells after nanoprocessing has been
ascertained by exclusion of propidium iodide from the
cells as well as by the presence of cytoplasmic streaming
(Tirlapur and König submitted).
Potential medical applications include the use of fem-
tosecond laser microscopes in eye- and neurosurgery, tis-
sue engineering, laser-assisted IVF, and gene therapy.
Universal NIR laser-based optical workstations
Currently available laser microscope systems utilize
Nd:YAG lasers or laser diodes for optical trapping, nitro-
gen lasers for microsurgery, and titanium:sapphire lasers
for multiphoton microscopy (Greulich 1999). In contrast
to such an elaborate systems using varied kinds of lasers,
NIR laser workstations in future can be based on a single
laser that combines the possibility of optical trapping,
non-contact cell transport, non-linear fluorescence imag-
ing with high spatial and temporal resolution, and photo-
chemical microscopy, as well as nanosurgery.
Conclusions
NIR laser microscopes can be used as versatile non-inva-
sive biomedical tools for optical micromanipulation, di-
agnostics, photochemistry, and surgery. It is therefore
conceivable that in the following years NIR microscopy
has enormous potential to become a method of choice in
biotechnology, cell biology, and medicine.
Acknowledgements I thank K.-J. Halbhuber, director of the Insti-
tute of Anatomy II, for his constant support and encouragement
and all the members of my research group, including A. Göhlert,
P. Fischer, I. Lemke, H. Oehring, C. Peuckert, I. Riemann, and
U.K. Tirlapur for their help in various ways. Thanks are also due
to M.W. Berns, B. Tromberg, T. Krasieva, Y. Liu, H. Liang, Y.
Tadir (Beckman Laser Institute and Medical Clinic, Irvine, USA),
L. Svaasand (Norwegian Institute of Technology, Trondheim, Nor-
way), and K.O. Greulich (Institute of Molecular Biotechnology,
Jena, Germany) for their support in laser tweezers experiments,
and P. So (MIT, Cambridge, USA) and E. Gratton (University of
Illinois, Urbana-Champaign, USA) for their support in the initial
femtosecond microscopy experiments. The work in the authors
laboratory is supported in part by grants from the Federal Ministry
of Education, Science, Research and Technology (BMBF), the
German Science Foundation (DFG), the Ministry of Science,
Research and Culture of the state of Thuringia (TMWFK), Zeiss,
Jena, and Coherent.
References
Ashkin A (1970) Acceleration and trapping of particles by radia-
tion pressure. Phys Rev Lett 24:156–159
Ashkin A (1980) Application of laser radiation pressure. Science
210:1081–1088
Ashkin A, Dziedzic JM (1987) Optical trapping and manipulation
of viruses and bacteria. Science 235:1517–1520
Ashkin A, Dziedzic JM (1989) Internal cell manipulation using
near infrared laser traps. Proc Nat Acad Sci USA 86:7914–
7918
Block SM, Goldstein LS, Schnapp BJ (1990) Bead movement by
single kinesin molecules studies with optical tweezers. Nature
348:348–352
Booth MJ, Hell SW (1998) Continuous wave excitation two-photon
fluorescence microscopy exemplified with the 647-nm ArKr
laser line. J Microsc 190:298–304
Buican TN, Smyth MJ, Crissman HA, Salzman GC, Stewart CC,
Martin JC (1987) Automated single-cell manipulation and
sorting by light trapping. Appl Opt 26:5311–5316
Chong WF, Prahl SA, Welch AJ. (1990) A review of the optical
properties of biological tissues. IEEE J Quantum Electron
26:2166–2185
Cunningham ML, Johnson JS, Giovanazzi SM, Peak MJ (1985)
Photosensitised production of superoxide anion by monochro-
matic (290–405 nm) ultraviolet irradiation of NADH and
NADPH coenzymes. Photochem Photobiol 42:125–128
Denk W (1994) Two-photon scanning photochemical microscopy:
mapping ligand-gated ion channel distribution. Proc Natl Acad
Sci USA 91:6629–6633
Denk W, Svoboda K (1997) Photon upmanship: why multiphoton
imaging is more than a gimmick. Neuron 18:351–357
Denk W, Strickler JH, Webb WW (1990) Two-photon laser scan-
ning microscope. Science 248:73–76
Florin EL, Pralle A, Stelzer EHK, Hörber JKH (1998) Photonic
force microscope calibration by thermal noise analysis. Appl
Phys A 66:575–578
Göppert-Meyer M (1931) Über Elementarakte mit zwei Quanten-
sprüngen. Göttinger Dissertation. Ann Phys 9:273–294
Greulich KO (1999) Micromanipulation by light in biology and
medicine. Birkhäuser, Basel
Gryczynski I, Szmaczinski H, Lakowicz JR (1995) On the possi-
bility of calcium imaging using Indo-1 with three photon exci-
tation. Photochem Photobiol 62:804–808
Hänninen PE, Soini E, Hell SW (1994) Continuous wave excita-
tion two-photon fluorescence microscopy. J Microsc 176:222–
225
Hell SW, Bahlmann K, Schrader M, Soini A, Malak H, Gryczynski
I, Lakowicz JR (1996) Three-photon excitation in fluorescence
microscopy. J Biomed Opt 1:71–74
Kaiser W, Garret C (1961) Two-photon excitation in CaF
2
:Eu
2+
.
Phys Rev Lett 7:229–231
91

Page 14
König K (1998) Laser tweezers are sources of two-photon excita-
tion. Cell Mol Biol 44:721–734
König K (2000) Invited review: multiphoton microscopy in life
sciences. J Microsc (in press)
König K, Schneckenburger H (1994) Laser-induced autofluores-
cence for medical diagnosis. J Fluorescence 4:17–40
König K, Liang H, Berns MW, Tromberg B (1995a) Cell damage
by near-IR beams. Nature 377:20–21
König K, Liu Y, Sonek GJ, Berns MW, Tromberg BJ (1995b) Au-
tofluorescence spectroscopy of optically-trapped cells during
light stress. Photochem Photobiol 62:830–835
König K, Tadir Y, Patrizio P, Berns MW, Tromberg BJ (1996a)
Effects of ultraviolet exposure and near infrared laser tweezers
on human spermatozoa. Hum Reprod 11:2162–2164
König K, Svaasand L, Liu Y, Sonek G, Patrizio P, Tadir Y, Berns
MW, Tromberg BJ (1996b) Determination of motility forces of
human spermatozoa using an 800 nm optical trap. Cell Mol
Biol 42:501–509
König K, Liang H, Berns MW, Tromberg B (1996c) Cell damage
in near infrared multimode optical traps as a result of multi-
photon absorption. Opt Lett 21:1090–1092
König K, So PTC, Mantulin WW, Tromberg BJ, Gratton E
(1996d) Two-photon excited lifetime imaging of autofluores-
cence in cells during UVA and NIR photostress. J Microsc
183:197–204
König K, Böhme S, Leclerc N, Ahuja R (1998) Time-gated auto-
fluorescence microscopy of motile green microalga in an opti-
cal trap. Cell Mol Biol 44:763–770
König K, Becker TW, Fischer P, Riemann I, Halbhuber KJ
(1999a) Pulse-length dependence of cellular response to in-
tense near-infrared laser pulses in multiphoton microscopes.
Opt Lett 24:113–115
König K, Riemann I, Fischer P, Halbhuber KJ (1999b) Intracellu-
lar nanosurgery with near infrared femtosecond laser pulses.
Cell Mol Biol 45:195–201
König K, Riemann I. Fischer P, Halbhuber KJ (2000a) Multiplex
FISH and three dimensional DNA imaging with near infrared
femtosecond laser pulses. Histochem Cell Biol (in press)
König, K, Göhlert, A, Liehr T, Loncarevic IF, Riemann, I (2000b)
Two-photon multicolour FISH: a versatile technique to detect
specific sequences within single DNA molecules in cells and
tissues. Single Mol 1:41–51
Kuo SC, Scheetz MP (1993) Force of single kinesin molecules
measured with optical tweezers. Science 260:232
Lakowicz JR (1983) Principles of fluorescence spectroscopy. Ple-
num Press, New York
Li X, Darzynkiewicz Z (1999) The Schrödinger’s cat quandary in
cell biology: integration of live functional analysis with mea-
surements of fixed cells in analysis of apoptosis. Exp Cell Res
249:404–412
Liu Y, Cheng D, Sonek GJ, Berns MW, Chapman CF, Tromberg
BJ (1995) Evidence for localised cell heating induced by near
infrared optical tweezers. Biophys J 68:2137–2144
Maiti S, Shear JB, Williams RM, Zipfel WR, Webb WW (1997)
Measuring serotonin in live cells with three-photon excitation.
Science 275:530–532
Masters BR, So PTC, Gratton E (1997) Multiphoton excitation
fluorescence microscopy and spectroscopy of in vivo human
skin. Biophys J 72:2405–2412
Oehring H, Riemann I, Fischer P, Halbhuber KJ, König K (2000)
Ultrastructure and reproduction behaviour of single CHO-K1
cells exposed to near infrared femtosecond laser pulses. Scan-
ning (in press)
Pawley JB (1995) Handbook of biological confocal microscopy,
2nd edn. Plenum Press, New York
Perkins TT, Quake SR, Smith DE, Chu S (1994) Relaxation of sin-
gle DNA molecule observed by optical microscopy. Science
264:822–826
Piston DW, Masters BR, Webb WW (1995) Three-dimensionally
resolved NAD(P)H cellular metabolic redox imaging of the in
situ cornea with two-photon excitation laser scanning micros-
copy. J Microsc 178:20–27
Pollak BA, Heim R (1999) Using GFP in FRET-based applica-
tions. Trends Cell Biol 9:57–60
Schönle A, Hell S (1998) Heating by absorption in the focus of an
objective lens. Opt Lett 23:325–327
So PTC, König K, Berland K, Dong CY, French T, Bühler C,
Ragan T, Gratton E. (1998) Review: new time-resolved tech-
niques in two-photon microscopy. Cell Mol Biol 44:771–793
Speel A (1999) Detection and amplification systems for sensitive
multiple-target DNA and RNA in situ hybridisation: looking
inside cells with a spectrum of colours. Histochem Cell Biol
112:89–113
Squirrel JM, Wokosin DL, White JG, Bavister BD (1999) Long-
term two-photon fluorescence imaging of mammalian embry-
os without compromising viability. Nat Biotechnol 17:763–
767
Svoboda K, Block SM (1994) Biological applications of optical
forces. Annu Rev Biophys Biomol Struct 23:247–285
Tadir Y, Wright WH, Vafa O (1989) Micromanipulation of sperm
by a laser-generated trap. Fertil Steril 52:870–873
Tirlapur UK, König K (1999) Near infrared femtosecond laser
pulses as a novel non-invasive means for dye-permeation and
3D imaging of localised dye-coupling in the Arabidopsis root
meristem. Plant J 20:363–370
Tyrell RM, Keyse SM (1990) The interaction of UVA radiation
with cultured cells. J Photochem Photobiol 4:349–361
Wiedemann R, Montag M (1994) Laser-assisted oocyte microin-
semination and direct transuterine tubal transfer in male infer-
tility. Poster. Intern Symposium on male factor in human infer-
tility. 21–22 April, Paris
Wokosin DL, Centonze, VE, Crittenden A, White JG (1995) Three
photon excitation of blue emitting fluorophores by laser scan-
ning microscopy. Mol Biol Cell 6:113a
Xu C, Zipfel W, Shear JB, Williams RM, Webb WW (1996)
Multiphoton fluorescence excitation: new spectral windows
for biological nonlinear microscopy. Proc Natl Acad Sci USA
93:10763–10768
Zahn M, Seeger S (1999) Optical tweezers in pharmacology. Cell
Mol Biol 44:747–761
92