Open Access
22 January 2020 Multiphoton microscopy: a personal historical review, with some future predictions
Author Affiliations +
Abstract

The historical development of multiphoton microscopy is described, starting with a review of two-photon absorption, and including two- and three-photon fluorescence microscopies, and second- and third-harmonic generation microscopies. The effects of pulse length on signal strength and breakdown are considered. Different contrast mechanisms, including use of nanoparticles, are discussed. Two new promising techniques that can be applied to multiphoton microscopy are described.

1.

Two-Photon Absorption

Perhaps the first mention of two-photon absorption was by Dirac,1 who said:

“These terms correspond to processes in which two light-quanta are emitted or absorbed simultaneously, and cannot arise in a light-quantum theory in which there are no forces between the light quanta. The effects of these terms will be found to be negligible, so that this disagreement with the light-quantum theory is not serious.”

So he seems to say that two-photon absorption is not a physical effect but an artifact of his theory, and that as this effect is small, we can neglect it. Then in the first edition of his book The Principles of Quantum Mechanics, he says, on p. 231:2

“The extra terms in (22) would correspond to transitions in which two photons are simultaneously absorbed or emitted and the possibility of such transitions requires a more complicated interaction energy than that assumed in (9). The physical effects of these terms are, however, small and unimportant, and so we shall neglect them.”

So here he again says that the two-photon effects are small and neglectable but would be physically described by a more complicated theory. Then, much later, in the fourth edition (1958),3 on p. 239:

“Similarly for a more general radiative process in which two or more photons are simultaneously emitted or absorbed, the probability is proportional to a factor Iν1 for each absorbed photon and Iν1+hν3/c2 for each emitted photon. Thus the process is stimulated by incident radiation in the same direction and with the same frequency and polarization as any of the emitted photons.”

And on p. 244:

“The matrix elements referring to simultaneous absorption or emission of two photons may be written down in the same way, but they lead to physical effects too small to be of practical importance.”

Dirac also introduced the quantum theory for Rayleigh scattering and Raman scattering.4

The full theory of two-photon absorption was developed by Göppert-Mayer.58 She received her PhD in 1930, whereas Dirac had received his in 1926. It is interesting to note that Dirac spent February to June 1927 and June 1928 in Göttingen visiting Max Born, Göppert-Meyer’s supervisor.9 During the first visit, he submitted two papers, “The quantum theory of emission and absorption of radiation,” which introduces the quantum theory of creation and annihilation of photons, and “The quantum theory of dispersion.”1,4

The influence of Dirac on Göppert-Mayer’s work was investigated by Masters:10

“The significance of this careful reconstruction of her Göttingen dissertation, together with a thorough comparison of the two papers that Dirac published in 1927, demonstrates that Göppert not only used and cited Dirac’s papers, but the extent to which she incorporated theoretical techniques from those two papers is significant. Previously, this incorporation of Dirac’s work into her Göttingen dissertation has either not been described or has been ignored in the literature on the history of quantum mechanics.”

Masters also discusses earlier related work and its influence on Göppert-Mayer.10

Various different nonlinear processes were observed experimentally soon after the invention of the laser. In particular, the first experimental observation of two-photon excitation was in 1961.11

It is often stated that two-photon absorption is much weaker than single-photon absorption because the cross-section is much smaller. But, although it is indeed much weaker, as two photons must be in the same vicinity at almost the same time, the cross-sections cannot be directly compared as they have different units (cm4s as compared with cm2). As a consequence, if the unit of distance was taken as an angstrom, for example, the cross-sections would not be so different in magnitude.

Beer’s law, for the depletion of a beam intensity with distance x, is modified for two-photon absorption.12 It becomes, rather than an exponential decay,

Eq. (1)

I=I01+βxI0,
where I0 is the original intensity and β is the two-photon absorption coefficient.

2.

Nonlinear and Multiphoton Microscopy

The first published nonlinear microscope images formed in a scanning mode, from second harmonic generation (SHG), were presented as a post-deadline paper at a conference in 197713 and subsequently in a paper in 1978.14 Our group in Oxford was led by Rudi Kompfner, who had already published images from SHG in the scanning acoustic microscope:15

“At the focus the acoustic beam converges to a diameter comparable with the acoustic wavelength. Because of this sharp convergence the intensity at the focus can be large. It, therefore, occurred to us to look for the nonlinear effects which become pronounced at high intensities.”

In another paper published by our group in 1978, the advantage of scanning using a highly focused laser beam for the production of nonlinear optical effects, in general, was highlighted as follows:16

“In the scanning optical microscope nonlinear interactions are expected to occur between the object and a highly focused beam of light, which we hope will open new ways of studying matter in microscopic detail hitherto not available. Nonlinear interactions include the generation of sum frequencies, Raman scattering, two-photon fluorescence, and others. We feel that the method will be of particular interest in studying biological materials, some of which have large second-harmonic generation coefficients, and the wide range of these coefficients should give very strong contrast in the images formed. Furthermore, frequency mixing should give information concerning the chemical structure of the object.”

In particular, we were actively working on the experimental demonstration of SHG imaging, but in addition, specifically proposed other nonlinear imaging modes, including two-photon fluorescence and coherent Raman scattering.

The normalized electric energy density We at the focus of a high numerical aperture aplanatic microscope objective for an instantaneous power P, calculated using the theory of Richards,17 is

Eq. (2)

We=k2nPπc[5(1cos3/2α)+3(1cos5/2α)]2450sin2α,
where nsinα is the numerical aperture and k=2π/λ, as shown in Fig. 1.18,19 The dashed curve gives the approximate prediction of a scalar, paraxial approximation.

Fig. 1

The electric energy density at the focus of a high numerical aperture microscope objective.

JBO_25_1_014511_f001.png

Our SHG images were acquired using a continuous-wave YAG laser, although we knew the advantages of using a short-pulsed laser and also experimented with a mode-locked (ps) argon ion laser. We looked at various inorganic crystals such as KD*P, LiNbO3, ZnO, and quartz. We demonstrated that SHG imaging exhibits an optical sectioning property without the use of a confocal pinhole:14

“Detail outside the focal plane does not interfere with the image as much as in a conventional microscope since the harmonic generated is proportional to the intensity squared and this results in the main contribution only coming from the region of focus where the intensity is very large.”

This optical sectioning property allows three-dimensional imaging of thick samples, such as tissue. It is also exhibited by two-photon fluorescence microscopy (2PFM) and is responsible, together with the reduced scattering of the longer wavelength excitation, to the improved penetration depth of two-photon fluorescence or SHG microscopy.

The SHG approach was extended to biological samples by Freund in 1986.20 The first reported coherent anti-Stokes Raman spectroscopy microscope image was by Duncan et al. in 1982.21

The first published report of 2PFM was by Denk et al. in 1990.22 They stressed that photobleaching (from two-photon absorption) is much reduced as compared with single-photon fluorescence microscopy, as a result of the optical sectioning property.

Three-photon fluorescence microscopy was reported in 1996.23,24 SHG microscopy using femtosecond pulses, and combined SHG and 2PFM, was reported by Gauderon et al.25,26 Third-harmonic generation (THG) microscopy was reported by Barad et al.27

Image formation and resolution in two-photon and three-photon microscopy have been investigated in various papers.2834 Although combining two-photon microscopy with a confocal pinhole gives an improved spatial resolution and optical sectioning,28,31,33 usually signal is low, so that as much light as possible must be detected. For this reason, nondescanned detection is often used to increase detection efficiency.35 However, there has been some experimental evidence that signal-to-noise-ratio is optimized for a particular pinhole size.34,36 Some researchers seem to think that 2PFM gives better resolution than confocal microscopy, but this is not the case when the same emission wavelength is assumed for each case.28 Partially coherent image formation in SHG microscopy has also been studied.37 Although the SHG process is itself coherent, integration over a finite-sized detector gives an overall partially coherent effect.

An important pulse-shaping technique that has been applied to multiphoton microscopy is temporal focusing.38,39 The different spectral components of the laser pulse are spread out and recombined in the focus. Then at regions far from focus, the pulse length at any point is long, so that multiphoton signal is reduced. Temporal focusing gives an optical sectioning effect, even for line-scan or widefield excitation.

An analysis of two-photon imaging in scattering media, such as biological tissue, has shown that the fundamental limitation to penetration depth is determined by background that emanates from the sample surface.40,41 The mechanism is that, as the excitation within the sample is reduced by scattering, and as two-photon absorption is a square law process, the two-photon fluorescence signal is eventually reduced below the background level.

3.

Pulse Length

Because of the nonlinearity (square-law behavior) of two-photon processes, for either two-photon fluorescence or SHG, short pulses should be used to maximize the two-photon signal. This was appreciated and analyzed in a report written by our group in 1975 (available on ResearchGate),42

“Then it can be shown that the shorter the pulses can be made, keeping the mean power constant, the higher will be the instantaneous intensity, and hence the greater will be the harmonic signal produced. It is therefore desirable to use a light source capable of producing the shortest possible pulses, consistent with adequate mean power i.e., Neodymium-YAG lasers, or dye-lasers.”

This was, of course, before the invention of the titanium sapphire laser. It was also mentioned in our paper:14

“It is possible to increase the harmonic signal for a given specimen temperature rise by pulsing the laser. Fast-scanning of the beam results in the whole specimen reaching an equilibrium temperature, in contrast to the slow-scan case where the probe is effectively stationary on the object, and allows a higher incident power.”

The report also discusses confocal microscopy and scanning near-field microscopy, extracts on which were reprinted later.43 In particular, experimental results were obtained on superesolution using a saturable absorber.

Actually, laser pulses can be too short for efficient two-photon absorption. If the frequency spread of the laser pulses is greater than the bandwidth of the absorption spectrum of the sample, absorption will be reduced.

Denk et al. filed a patent for two-photon microscopy using pulses shorter than 1 ps,44 because, as we have discussed, 2PFM itself had been proposed many years earlier. Although short pulses increase the two-photon signal, the specification of 1 ps seems rather arbitrary. Hanninen and Hell45 subsequently patented multiphoton microscopy using pulses greater than 1 ps. These patents have now expired, and several companies manufacture multiphoton microscopes.

A laser pulse of peak power P and duration t results in a two-photon energy signal

Eq. (3)

E2p=aP2t,
where a is a constant that depends on the two-photon cross-section. It is also well known (e.g., from Refs. 14, 16, and 42) that the average power that can be focused on to the sample is limited because of heating and other undesirable side effects. The average power is given as

Eq. (4)

Pav=Pt/T,
where T is the time interval between pulses. We have during the pulse

Eq. (5)

E2p=aPav2T2/t,
so that the average two-photon signal power is given as

Eq. (6)

P2p=aPav2T/t,
and thus the signal is increased as the pulse length decreases while the average power is held constant.

However, an ultimate limit to how short the pulses can be is set by the onset of optical breakdown. Docchio et al.46 showed that the power density needed to cause optical breakdown in water actually increases as the pulse length is decreased. Stern et al.47 showed that the power needed to cause surface ablation of the cornea can be expressed as

Eq. (7)

P=bA/t1/2,
where b is a constant and A is the cross-sectional area of the focused spot. At breakdown, we would thus expect a two-photon signal

Eq. (8)

P2p=ab2A2/T.

Thus, it is seen that in this case the two-photon signal at the onset of breakdown is independent of the pulse length t.

In order to obtain the maximum two-photon signal for a given average power, the pulse length can thus be decreased until close to breakdown. Breakdown occurs when

Eq. (9)

t=(PavT/bA),
which is thus independent of a, and therefore of the two-photon cross-section. Thus, one does not need to know the two-photon cross-section in order to predict the minimum pulse length that can be used to maximize the two-photon signal while avoiding breakdown. Taking typical values of breakdown given by Docchio et al.,46 and Stern et al.47 give a value of b of about 4×106  Js1/2cm2. For ordinary mode-locked lasers, T is about 1.25×108  s. We find that for an average power of 100 mW and A=2×109  cm2, t can be reduced to about 20 fs while still avoiding breakdown. For an average power of 10 mW, t is less than 1 fs for breakdown to occur. Thus, onset of breakdown occurs at subpicosecond pulse lengths for any reasonable value of average power.

However, recent results show that even these figures for breakdown are lower than those found for subpicosecond pulses. From Du et al.48 and Tien et al.49 for fused silica, the damage threshold can conservatively be taken to be constant at about 3  J/cm2 for pulses shorter than about 10 ps. Thus for subpicosecond pulses, the fluence at breakdown is inversely proportional to the pulse length. The result is that pulses of average power <100  mW breakdown never occurs for any pulse length. These predictions, together with the behavior of the fluence for constant average power, are shown in Fig. 2.

Fig. 2

Beam fluence as a function of pulse length for a given average power. T is taken as 1.25×108  s. A is taken as 2×109  cm2. The onset of breakdown is also indicated.

JBO_25_1_014511_f002.png

In addition, Docchio et al.46 also showed that the power density needed to cause breakdown increases as the size of the focused spot decreases. So for a high numerical aperture lens, one would expect that even shorter pulses could be used without causing breakdown. Although the data in Docchio et al.46 do not allow an accurate extrapolation of the effect of increasing the numerical aperture to the levels used in scanning laser microscopy, we can estimate that, for the parameters given above, average powers in the order of watts would be needed to cause breakdown.

The fluence for a given two-photon power can also be calculated from Eq. (3) or Eq. (6). We obtain

Eq. (10)

F=1A(P2pTat)1/2,
so that lines of constant two-photon signal for pulses longer than about a picosecond are parallel to the breakdown curve in Fig. 2. Thus in this regime, the two-photon signal at the onset of breakdown is independent of pulse length.

In the above, we have assumed a repetition rate T1 appropriate for an ordinary mode-locked laser. Then many pulses will arrive during the pixel dwell time, so T can be increased. From Eq. (6), the average two-photon signal power is then also increased. In fact, it can be increased so much that wide-field imaging becomes feasible.50,51 The optimum repetition rate for two-photon excitation has been investigated, but the interaction of many different effects complicates the overall behavior.52

4.

Contrast Mechanisms

Multiphoton fluorescence microscopy can be performed using autofluorescence or by labeling with dyes. An alternative for confocal microscopy is to use conjugated nanoparticles, such as 5- to 15-nm gold nanospheres.53 Nanoparticles, such as nanospheres, nanoshells, and nanorods, avoid photobleaching of fluorescent labels.

SHG occurs from noncentrosymmetric structures.54 These include as examples collagen and actin. Polarized SHG can be used to determine the nonlinear permittivity elements.55,56 Illumination using radially polarized light produces a longitudinal electric field, allowing observation and determination of additional elements.57 The surface of a centrosymmetric crystal can also generate a harmonic signal.58,59 Plasmonic enhancement from a rough surface of ×104 has been reported.60 In a centrosymmetric metallic medium, the second-order polarization P(2) is of the form61

Eq. (11)

P(2)=α(E·)E+βE·E+γE×(×E),
where α and β (electric quadrupole term) are constants giving the strength of surface terms, and γ gives the strength of a magnetic dipole, volume term. A theory for SHG from a spherical particle was developed.62 It was shown that SHG can also give rise to enhanced backscattering from rough surfaces (a peak in scattering in the reverse direction of the incident beam).63 Surface molecular monolayers have also been detected by SHG.64

SHG is a coherent signal, but an incoherent signal, hyper-Rayleigh scattering (HRS), was observed from colloids.65,66 Nappa et al.67 observed that SHG from small particles was dipolar in nature, but from larger particles, >80  nm, quadrupole scattering was observed. HRS has been observed from nanorods.68 SHG has been observed from gold nanospheres conjugated to antibodies,69 and SHG microscopy was performed with fluorescent polymer encapsulated gold nanoparticles.70

Two-photon luminescence (TPL), an “unusual broad luminescence background,” was observed from noble metals and enhanced by roughened surfaces.60,71 This led to plasmon-enhanced TPL microscopy72 and in vivo TPL microscopy using gold nanorods.36 Usually, the nanorods are excited at resonance, but Balla et al. proposed excitation of nanorods off-resonance (at 1200 nm). While the luminescence signal is reduced, this has the advantages of increased excitation wavelength and therefore penetration, allowing signal detection in the near-infrared, and of reduced heating of the nanoparticles.73 THG gives contrast from interfaces.27 THG from single gold nanoparticles has also been observed.74

5.

Two-Photon Focal Modulation Microscopy

We finish with two recently developed techniques that can be applied to multiphoton microscopy. The first is focal modulation microscopy (FMM).75 In this technique, a laser beam is divided into two beams one of which is frequency-shifted. These two beams illuminate different regions of the illuminating lens pupil, a variety of different geometries having been proposed. The beams intersect in the focal region, producing an intensity modulated at the difference frequency. The technique thus has strong similarities with temporal focusing. The generated fluorescence signal is also modulated and is detected using lock-in techniques. The result is better penetration into a scattering medium and an improved spatial resolution.76 This method can also be applied to 2PFM.77,78 Compared with conventional two-photon microscopy, transverse resolution is increased by 70%, while the axial resolution is increased twofold. Results showed that the signal-to-background ratio of 2PF-FMM can be up to five times better than in regular two-photon microscopy at the depth of 500  μm.78

6.

Two-Photon Image Scanning Microscopy

Another promising new microscopical technique is image scanning microscopy (ISM).79,80 In this approach, the pinhole of a confocal microscope is replaced by a detector array. The signals from the detector elements are reassigned to the correct image location, which is to a first approximation for single-photon fluorescence midway between the illumination and detection points, and then summed.7981 The result is a combination of improved transverse and axial resolution and an improved signal-to-noise ratio. The detected power is determined by the size of the detector array. Transverse resolution for an offset detector pixel has been found to be improved relative to the central pixel,82 so overall transverse resolution for the detector array can be even better than for confocal microscopy with a small pinhole. Our implementation uses an array of 5×5 avalanche photodiodes.83

ISM can also be successfully applied to multiphoton microscopy.8486 This improves the optical sectioning property of two-photon microscopy, as a result of an additional confocal sectioning effect. Although a confocal pinhole is rarely used for two-photon microscopy, the penetration depth for ISM can be increased because the confocal sectioning effect reduces the strength of the background from the sample surface.40,41 A theoretical treatment of two-photon ISM has also been presented.87

7.

Conclusion

We have discussed the historical development of multiphoton microscopy and some of its major features. We have also indicated two promising future developments for multiphoton microscopy, namely FMM and ISM. Much more detail can be found in the excellent books edited by Alberto Diaspro and Karsten König and the reprint collection of Barry Masters.8890

Disclosures

The author has no relevant financial interests in the manuscript and no other potential conflicts of interest to disclose.

Acknowledgments

The author acknowledges many fruitful discussions with Naveen Balla, Marco Castello, Nanguang Chen, Guy Cox, Takahiro Deguchi, Alberto Diaspro, Wei Gong, Min Gu, Rainer Heintzmann, Zhiwei Huang, Sami Koho, Karsten König, Phil Lukins, Shalin Mehta, Shakil Rehman, Ke Si, Peter So, Giorgio Tortarolo, Giuseppe Vicidomini, Elijah Yew, and many others.

References

1. 

P. A. M. Dirac, “The quantum theory of dispersion,” Proc. R. Soc. Lond., 114 710 –728 (1927). https://doi.org/10.1098/rspa.1927.0071 Google Scholar

2. 

P. A. M. Dirac, The Principles of Quantum Mechanics, 1st ed.The Clarendon Press, Oxford (1930). Google Scholar

3. 

P. A. M. Dirac, The Principles of Quantum Mechanics, 4th ed.Oxford University Press, Oxford (1958). Google Scholar

4. 

P. A. M. Dirac, “The quantum theory of the emission and absorption of radiation,” Proc. R. Soc. Lond. A, 114 243 –265 (1927). https://doi.org/10.1098/rspa.1927.0039 PRLAAZ 1364-5021 Google Scholar

5. 

M. Göppert, “Über die Wahrscheinlichkeit des Zusammenwirkens zweier Lichtquanten in einem Elementarakt,” Naturwissenschaften, 17 932 –932 (1929). https://doi.org/10.1007/BF01506585 NATWAY 0028-1042 Google Scholar

6. 

M. Göppert-Mayer, “Über Elementarakte mit zwei Quantensprüngen,” Ann. Phys., 401 273 –294 (1931). https://doi.org/10.1002/andp.19314010303 ANPYA2 0003-3804 Google Scholar

7. 

M. Göppert-Mayer, “Elementary processes with two quantum transitions,” Ann. Phys., 18 466 –479 (2009). https://doi.org/10.1002/andp.200910358 ANPYA2 0003-3804 Google Scholar

8. 

B. R. Masters, “English translation of Göppert (1929) and Göppert-Mayer (1931),” (2010). http://quantum-history.mpiwg-berlin.mpg.de/news/workshops/hq3/hq3_talks/21_masters_appendix.pdf Google Scholar

9. 

G. Farmelo, The Strangest Man. The Hidden Life of Paul Dirac, Mystic of the Atom, Basic Books, New York (2009). Google Scholar

10. 

B. R. Masters, “The Origins of Maria Göppert's Dissertation on Two-Photon Quantum Transitions at Göttingen's Institutes of Physics 1920-1933,” Traditions and Transformations in the History of Quantum Physics, 209 –230 Max Planck Research Library for the History and Development of Knowledge, Berlin (2013). Google Scholar

11. 

W. Kaiser and C. Garrett, “Two-photon excitation in CaF2:Eu2+,” Phys. Rev. Lett., 7 229 –231 (1961). https://doi.org/10.1103/PhysRevLett.7.229 Google Scholar

12. 

B. Nikolaus, D. Z. Zhang and P. E. Toschek, “Two-photon laser,” Phys. Rev. Lett., 47 171 –173 (1981). https://doi.org/10.1103/PhysRevLett.47.171 PRLTAO 0031-9007 Google Scholar

13. 

C. J. R. Sheppard et al., “The scanning harmonic optical microscope,” IEEE J. Quantum Electron., 13 912 (1977). https://doi.org/10.1109/JQE.1977.1069615 IEJQA7 0018-9197 Google Scholar

14. 

J. N. Gannaway and C. J. R. Sheppard, “Second harmonic imaging in the scanning optical microscope,” Opt. Quantum Electron., 10 435 –439 (1978). https://doi.org/10.1007/BF00620308 OQELDI 0306-8919 Google Scholar

15. 

R. Kompfner and R. A. Lemons, “Nonlinear acoustic microscope,” Appl. Phys. Lett., 28 295 –297 (1976). https://doi.org/10.1063/1.88759 APPLAB 0003-6951 Google Scholar

16. 

C. J. R. Sheppard and R. Kompfner, “Resonant scanning optical microscope,” Appl. Opt., 17 2879 –2882 (1978). https://doi.org/10.1364/AO.17.002879 APOPAI 0003-6935 Google Scholar

17. 

B. Richards, “Diffraction in systems of high relative aperture,” Astronomical Optics and Related Subjects, 352 –359 North Holland Publishing Company, Amsterdam (1955). Google Scholar

18. 

C. J. R. Sheppard, A. Choudhury and J. N. Gannaway, “Electromagnetic field near the focus of wide-angular lens and mirror systems,” IEE J. Microwaves Opt. Acoust., 1 129 (1977). https://doi.org/10.1049/ij-moa.1977.0015 Google Scholar

19. 

C. J. R. Sheppard and H. J. Matthews, “Imaging in high aperture optical systems,” J. Opt. Soc. Am. A, 4 (8), 1354 –1360 (1987). https://doi.org/10.1364/JOSAA.4.001354 JOAOD6 0740-3232 Google Scholar

20. 

I. Freund and M. Deutsch, “Second-harmonic microscopy of biological tissue,” Opt. Lett., 11 94 –96 (1986). https://doi.org/10.1364/OL.11.000094 OPLEDP 0146-9592 Google Scholar

21. 

M. D. Duncan, J. Reintjes and T. J. Manuccia, “Scanning coherent anti-stokes Raman microscope,” Opt. Lett., 7 350 –352 (1982). https://doi.org/10.1364/OL.7.000350 OPLEDP 0146-9592 Google Scholar

22. 

W. Denk, J. Strickler and W. Webb, “Two-photon laser scanning flourescence microscopy,” Science, 248 73 –76 (1990). https://doi.org/10.1126/science.2321027 SCIEAS 0036-8075 Google Scholar

23. 

C. Xu et al., “Multiphoton fluorescence excitation: new spectral windows for biological nonlinear microscopy,” Proc. Natl. Acad. Sci. U. S. A., 93 10763 –10768 (1996). https://doi.org/10.1073/pnas.93.20.10763 Google Scholar

24. 

S. Hell et al., “Three-photon excitation in fluorescence microscopy,” J. Biomed. Opt., 1 (1), 71 –74 (1996). https://doi.org/10.1117/12.229062 JBOPFO 1083-3668 Google Scholar

25. 

R. Gauderon, P. Lukins and C. J. R. Sheppard, “Three-dimensional second-harmonic generation imaging with femtosecond laser pulses,” Opt. Lett., 23 (15), 1209 –1211 (1998). https://doi.org/10.1364/OL.23.001209 OPLEDP 0146-9592 Google Scholar

26. 

R. Gauderon, P. Lukins and C. Sheppard, “Simultaneous multi-channel two-photon microscopy,” Micron, 32 685 –689 (2001). https://doi.org/10.1016/S0968-4328(00)00067-6 MICNB2 0047-7206 Google Scholar

27. 

Y. Barad et al., “Nonlinear scanning laser microscopy by third-harmonic generation,” Appl. Phys. Lett., 70 922 –924 (1996). https://doi.org/10.1063/1.118442 APPLAB 0003-6951 Google Scholar

28. 

C. J. R. Sheppard and M. Gu, “Image formation in two-photon fluorescence microscopy,” Optik, 86 104 –106 (1990). OTIKAJ 0030-4026 Google Scholar

29. 

C. J. R. Sheppard, “Image formation in three-photon fluorescence microscopy,” Bioimaging, 4 (3), 124 –128 (1996). https://doi.org/10.1002/1361-6374(199609)4:3<124::AID-BIO2>3.0.CO;2-2 BOIMEL 0966-9051 Google Scholar

30. 

M. Gu, “Resolution in three-photon fluorescence scanning microscopy,” Opt. Lett., 21 (13), 988 –990 (1996). https://doi.org/10.1364/OL.21.000988 OPLEDP 0146-9592 Google Scholar

31. 

M. Gu and C. J. R. Sheppard, “Effects of a finite-sized pinhole on 3-D image formation in confocal two-photon fluorescence microscopy,” J. Mod. Opt., 40 2009 –2024 (1993). https://doi.org/10.1080/09500349314552011 JMOPEW 0950-0340 Google Scholar

32. 

M. Gu and C. J. R. Sheppard, “Comparison of three-dimensional imaging properties between two-photon and single-photon fluorescence microscopy,” J. Microsc., 177 (2), 128 –137 (1995). https://doi.org/10.1111/j.1365-2818.1995.tb03543.x JMICAR 0022-2720 Google Scholar

33. 

R. Gauderon, P. B. Lukins and C. J. R. Sheppard, “Effect of a confocal pinhole in two-photon microscopy,” Microsc. Res. Tech., 47 210 –214 (1999). https://doi.org/10.1002/(SICI)1097-0029(19991101)47:3<210::AID-JEMT7>3.0.CO;2-H MRTEEO 1059-910X Google Scholar

34. 

R. Gauderon and C. J. R. Sheppard, “Effect of a finite-size pinhole on the noise performance in single-, two- and three-photon fluorescence microscopy,” Appl. Opt., 38 3562 –3565 (1999). https://doi.org/10.1364/AO.38.003562 APOPAI 0003-6935 Google Scholar

35. 

G. Cox and C. J. R. Sheppard, “Practical limits of resolution in confocal and non-linear microscopy,” Microsc. Res. Tech., 63 18 –22 (2004). https://doi.org/10.1002/jemt.10423 MRTEEO 1059-910X Google Scholar

36. 

H. Wang et al., “In vitro and in vivo two-photon luminescence imaging of single gold nanorods,” Proc. Natl. Acad. Sci. U. S. A., 102 15752 –15756 (2005). https://doi.org/10.1073/pnas.0504892102 Google Scholar

37. 

R. Gauderon and C. Sheppard, “Two-dimensional weak-object transfer functions in the scanning harmonic microscope,” J. Mod. Opt., 47 (7), 1195 –1202 (2000). https://doi.org/10.1080/09500340008232166 JMOPEW 0950-0340 Google Scholar

38. 

E. Tai, D. Oron and Y. Silberberg, “Improved depth resolution in video-rate line-scanning multiphoton microscopy using temporal focusing,” Opt. Lett., 30 1686 –1688 (2005). https://doi.org/10.1364/OL.30.001686 OPLEDP 0146-9592 Google Scholar

39. 

G. Zhu et al., “Simultaneous spatial and temporal focusing of ultrashort pulses,” Opt. Express, 13 2153 –2159 (2005). https://doi.org/10.1364/OPEX.13.002153 OPEXFF 1094-4087 Google Scholar

40. 

M. Oheim et al., “Two-photon microscopy in brain tissue: parameters influencing the imaging depth,” J. Neurosci. Methods, 111 29 –37 (2001). https://doi.org/10.1016/S0165-0270(01)00438-1 JNMEDT 0165-0270 Google Scholar

41. 

P. Theer and W. Denk, “On the fundamental imaging-depth limit in two-photon microscopy,” J. Opt. Soc. Am. A, 23 3139 –3149 (2006). https://doi.org/10.1364/JOSAA.23.003139 JOAOD6 0740-3232 Google Scholar

42. 

R. Kompfner et al., “Applications of quantum electronics, part 1. The scanning optical microscope,” (1975). Google Scholar

43. 

R. Kompfner et al., “The scanning optical microscope (extracts on near-field microscopy),” Scanning, 16 327 –332 (1994). https://doi.org/10.1002/sca.4950160310 SCNNDF 0161-0457 Google Scholar

44. 

W. Denk, J. P. Strickler and W. W. Webb, “Two-photon laser microscopy,” US Patent No. 5,034,613 (1991).

45. 

P. Hanninen and S. Hell, “Luminescence-scanning microscopy process and a luminescence microscope utilizing picosecond or greater pulse lasers,” US Patent No. 5,777,732 (1998).

46. 

F. Docchio, C. A. Sacchi and J. Marshall, “Experimental investigation of optical breakdown thresholds in ocular media under single pulse irradiation with different pulse durations,” Lasers Ophthalmol., 1 83 –93 (1986). Google Scholar

47. 

D. Stern et al., “Corneal ablation by nanosecond, picosecond, and femtosecond lasers at 532 nm and 625 nm,” Arch. Ophthalmol., 107 587 –592 (1989). https://doi.org/10.1001/archopht.1989.01070010601038 Google Scholar

48. 

D. Du et al., “Laser-induced breakdown by impact ionization in SiO2 with pulse widths from 7 ns to 150 fs,” Appl. Phys. Lett., 64 3071 –3073 (1994). https://doi.org/10.1063/1.111350 APPLAB 0003-6951 Google Scholar

49. 

A.-C. Tien et al., “Short-pulse laser damage in transparent materials as a function of pulse duration,” Phys. Rev. Lett., 82 3883 –3886 (1999). https://doi.org/10.1103/PhysRevLett.82.3883 PRLTAO 0031-9007 Google Scholar

50. 

L. C. Cheng et al., “Spatiotemporal focusing-based widefield multiphoton microscopy for fast optical sectioning,” Opt. Express, 20 8939 –8948 (2012). https://doi.org/10.1364/OE.20.008939 OPEXFF 1094-4087 Google Scholar

51. 

C. Macias-Romero et al., “Wide-field medium-repetition-rate multiphoton microscopy reduces photodamage of living cells,” Biomed. Opt. Express, 7 1458 –1467 (2016). https://doi.org/10.1364/BOE.7.001458 BOEICL 2156-7085 Google Scholar

52. 

K. Charan et al., “Fiber-based tunable repetition rate source for deep tissue two-photon fluorescence microscopy,” Biomed. Opt. Express, 9 2304 –2311 (2018). https://doi.org/10.1364/BOE.9.002304 BOEICL 2156-7085 Google Scholar

53. 

C. J. Cogswell, D. K. Hamilton and C. J. R. Sheppard, “Colour reflection microscopy using red, green and blue lasers,” J. Microsc., 165 103 –117 (1992). https://doi.org/10.1111/j.1365-2818.1992.tb04308.x JMICAR 0022-2720 Google Scholar

54. 

J. Ducuing and N. Bloembergen, “Observation of reflected light harmonics at the boundary of piezoelectric crystals,” Phys. Rev. Lett., 10 474 –476 (1963). https://doi.org/10.1103/PhysRevLett.10.474 PRLTAO 0031-9007 Google Scholar

55. 

S. Roth and I. Freund, “Second harmonic generation in collagen,” J. Chem. Phys., 70 1637 –1643 (1979). https://doi.org/10.1063/1.437677 JCPSA6 0021-9606 Google Scholar

56. 

P. Stoller et al., “Polarization-dependent optical second-harmonic imaging of rat-tail tendon,” J. Biomed. Opt., 7 (2), 205 –214 (2002). https://doi.org/10.1117/1.1431967 JBOPFO 1083-3668 Google Scholar

57. 

E. Y. S. Yew and C. J. R. Sheppard, “Second harmonic generation polarization microscopy with tightly focused linearly and radially polarized beams,” Opt. Commun., 275 453 –457 (2007). https://doi.org/10.1016/j.optcom.2007.03.065 OPCOB8 0030-4018 Google Scholar

58. 

F. Brown, R. E. Parks and A. M. Sleeper, “Nonlinear optical reflection from a metallic boundary,” Phys. Rev. Lett., 14 1029 –1031 (1965). https://doi.org/10.1103/PhysRevLett.14.1029 PRLTAO 0031-9007 Google Scholar

59. 

S. S. Jha, “Nonlinear optical reflection from a metal surface,” Phys. Rev. Lett., 15 412 –414 (1965). https://doi.org/10.1103/PhysRevLett.15.412 PRLTAO 0031-9007 Google Scholar

60. 

C. K. Chen, A. R. B. de Castro and Y. R. Shen, “Surface enhanced second harmonic generation,” Phys. Rev. Lett., 46 145 –148 (1981). https://doi.org/10.1103/PhysRevLett.46.145 PRLTAO 0031-9007 Google Scholar

61. 

N. Bloembergen et al., “Optical second-harmonic generation in reflection from media with inversion symmetry,” Phys. Rev. E, 174 813 –822 (1968). https://doi.org/10.1103/PhysRev.174.813 Google Scholar

62. 

G. S. Agarwal and S. S. Jha, “Theory of second harmonic generation at a metal surface with surface plasmon excitation,” Solid State Commun., 41 499 –501 (1982). https://doi.org/10.1016/0038-1098(82)90534-8 SSCOA4 0038-1098 Google Scholar

63. 

K. A. O’Donnell, R. Torre and C. S. West, “Observations of second-harmonic generation from randomly rough metal surfaces,” Phys. Rev. B, 55 7985 –7992 (1997). https://doi.org/10.1103/PhysRevB.55.7985 Google Scholar

64. 

C. K. Chen et al., “Detection of molecular monolayers by optical second-harmonic generation,” Phys. Rev. Lett., 46 1010 –1012 (1981). https://doi.org/10.1103/PhysRevLett.46.1010 PRLTAO 0031-9007 Google Scholar

65. 

F. Vance, B. Lemon and J. Hupp, “Enormous hyper-Rayleigh scattering from nanocrystalline gold particle suspensions,” J. Phys. Chem. B, 102 10091 –10093 (1998). https://doi.org/10.1021/jp984044u JPCBFK 1520-6106 Google Scholar

66. 

P. Galletto, P. Brevet and H. Girault, “Enhancement of second harmonic response by adsorbates on gold colloids: the effect of aggregation,” J. Phys. Chem. B, 103 8706 –8710 (1999). https://doi.org/10.1021/jp991937t JPCBFK 1520-6106 Google Scholar

67. 

J. Nappa et al., “Electric dipole origin of second harmonic generation of small metallic nanoparticles,” Phys. Rev. B, 71 165404 –165407 (2005). https://doi.org/10.1103/PhysRevB.71.165407 Google Scholar

68. 

J. Nappa et al., “Second harmonic generation from small gold metallic particles: from the dipolar to the quadrupolar response,” J. Chem. Phys., 125 184712 (2006). https://doi.org/10.1063/1.2375095 JCPSA6 0021-9606 Google Scholar

69. 

C. X. Zhang et al., “Hyper-Rayleigh scattering of protein-modified gold nanoparticles,” Anal. Biochem., 320 136 –140 (2003). https://doi.org/10.1016/S0003-2697(03)00353-1 ANBCA2 0003-2697 Google Scholar

70. 

H. Clark et al., “Second harmonic generation microscopy of fluorescent polymer encapsulated gold nanoparticles,” J. Am. Chem. Soc., 122 10234 –10235 (2000). https://doi.org/10.1021/ja002223v JACSAT 0002-7863 Google Scholar

71. 

G. Boyd, Z. Yu and Y. Shen, “Photoinduced luminescence from the noble metals and its enhancement on roughened surfaces,” Phys. Rev. B, 33 7923 –7936 (1986). https://doi.org/10.1103/PhysRevB.33.7923 Google Scholar

72. 

D. Yelin et al., “Multiphoton plasmon-resonance microscopy,” Opt. Express, 11 1385 –1391 (2003). https://doi.org/10.1364/OE.11.001385 OPEXFF 1094-4087 Google Scholar

73. 

N. K. Balla, C. J. R. Sheppard and P. T. C. So, “Multiphoton luminescence of gold nanorods upon excitation with wavelengths away from their absorption maxima,” Proc. SPIE, 7910 79100N (2011). https://doi.org/10.1117/12.876014 PSISDG 0277-786X Google Scholar

74. 

M. Lippitz, M. A. van Dijk and M. Orrit, “Third-harmonic generation from single gold nanoparticles,” Nanoletters, 5 799 –802 (2005). https://doi.org/10.1021/nl0502571 Google Scholar

75. 

N. G. Chen, C. H. Wong and C. J. R. Sheppard, “Focal modulation microscopy,” Opt. Express, 16 (23), 18764 –18769 (2008). https://doi.org/10.1364/OE.16.018764 OPEXFF 1094-4087 Google Scholar

76. 

W. Gong et al., “Improved spatial resolution in fluorescence focal modulation microscopy,” Opt. Lett., 34 (12), 3508 –3510 (2009). https://doi.org/10.1364/OL.34.003508 OPLEDP 0146-9592 Google Scholar

77. 

K. Si et al., “Two-photon focal modulation microscopy in turbid media,” Appl. Phys. Lett., 99 233702 (2011). https://doi.org/10.1063/1.3665936 APPLAB 0003-6951 Google Scholar

78. 

Y. Zheng et al., “Two-photon focal modulation microscopy for high-resolution imaging in deep tissue,” J. Biophotonics, 12 e201800247 (2018). https://doi.org/10.1002/jbio.201800247 Google Scholar

79. 

C. J. R. Sheppard, “Super-resolution in confocal imaging,” Optik, 80 53 –54 (1988). OTIKAJ 0030-4026 Google Scholar

80. 

C. B. Müller and J. Enderlein, “Image scanning microscopy,” Phys. Rev. Lett., 104 198101 (2010). https://doi.org/10.1103/PhysRevLett.104.198101 PRLTAO 0031-9007 Google Scholar

81. 

C. J. R. Sheppard, S. B. Mehta and R. Heintzmann, “Superresolution by image scanning microscopy using pixel reassignment,” Opt. Lett., 38 (15), 2889 –2892 (2013). https://doi.org/10.1364/OL.38.002889 OPLEDP 0146-9592 Google Scholar

82. 

I. J. Cox, C. J. R. Sheppard and T. Wilson, “Improvement in resolution by nearly confocal microscopy,” Appl. Opt., 21 (5), 778 –781 (1982). https://doi.org/10.1364/AO.21.000778 APOPAI 0003-6935 Google Scholar

83. 

M. Castello et al., “A robust and versatile platform for image scanning microscopy enabling super-resolution FLIM,” Nat. Methods, 16 175 –178 (2019). https://doi.org/10.1038/s41592-018-0291-9 1548-7091 Google Scholar

84. 

M. Ingaramo et al., “Two-photon excitation improves multifocal structured illumination microscopy in thick scattering tissue,” Proc. Natl. Acad. Sci. U. S. A., 111 5254 –5259 (2014). https://doi.org/10.1073/pnas.1314447111 Google Scholar

85. 

I. Gregor et al., “Rapid nonlinear image scanning microscopy,” Nat. Methods, 14 1087 –1089 (2017). https://doi.org/10.1038/nmeth.4467 1548-7091 Google Scholar

86. 

S. V. Koho et al., “Easy two-photon image scanning microscopy with SPAD array and blind image reconstruction,” bioRXiv, (2019). https://doi.org/10.1101/563288 Google Scholar

87. 

C. J. R. Sheppard et al., “Image formation in image scanning microscopy, including the case of two-photon excitation,” J. Opt. Soc. Am. A, 34 1339 –1350 (2017). https://doi.org/10.1364/JOSAA.34.001339 JOAOD6 0740-3232 Google Scholar

88. 

A. Diaspro, Confocal and Two-Photon Microscopy, Wiley-Liss, New York (2002). Google Scholar

89. 

K. König, Multiphoton Microscopy and Fluorescence Lifetime Imaging, de Gruyter, Berlin (2017). Google Scholar

90. 

B. R. Masters, Selected Papers on Multiphoton Excitation Microscopy, 175 SPIE Press, Bellingham, Washington (2003). Google Scholar

Biography

Colin J. R. Sheppard is a external collaborator and visiting scientist at the Istituto Italiano di Tecnologia, Genoa, Italy; honorary professorial fellow at the University of Wollongong, Australia; and supernumerary fellow of Pembroke College, Oxford, United Kingdom. Previously, he was a professor and head in the Department of Bioengineering at the National University of Singapore; professor of physics at the University of Sydney; and university lecturer in engineering science at the University of Oxford. He received his PhD in engineering from the University of Cambridge in 1973 and his DSc degree in physical sciences from the University of Oxford in 1986. He has held visiting positions at many different institutions, including UC-Berkeley, MIT, EPFL, TU-Delft, and Tokyo University. He developed an early laser microscope in 1975, patented scanning microscopy using Bessel beams in 1977, published the first demonstration of scanning two-photon microscopy (by SHG) in 1977, proposed two-photon fluorescence and CARS microscopy in 1978, launched the first commercial confocal microscope in 1982, and developed the first confocal microscope with computer control and storage in 1983.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Colin J. R. Sheppard "Multiphoton microscopy: a personal historical review, with some future predictions," Journal of Biomedical Optics 25(1), 014511 (22 January 2020). https://doi.org/10.1117/1.JBO.25.1.014511
Received: 31 August 2019; Accepted: 23 December 2019; Published: 22 January 2020
Lens.org Logo
CITATIONS
Cited by 27 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Multiphoton microscopy

Microscopy

Second-harmonic generation

Absorption

Confocal microscopy

Luminescence

Signal detection

Back to Top